text
stringlengths
4
2.78M
meta
dict
--- abstract: 'We consider one loop QCD corrections and renormalization group running of the neutrinoless double beta decay amplitude focusing on the short-range part of the amplitude (without the light neutrino exchange) and find that these corrections can be sizeable. Depending on the operator under consideration, there can be moderate to large cancellations or significant enhancements. We discuss several specific examples in this context. Such large corrections will lead to significant shifts in the half-life estimates which currently are known to be plagued with the uncertainties due to nuclear physics inputs to the physical matrix elements.' author: - Namit Mahajan title: Neutrinoless double beta decay and QCD corrections --- =========ł==ø==================Ł=Ø=¶====== It is now experimentally well established that neutrinos have mass and they mix with each other (see [@Tortola:2012te] for the best fit values of the parameters). Being electrically neutral allows the possibility of them to be Majorana particles [@Majorana:1937vz]. The observation of neutrinoless double beta () decay, $(A,Z)\rightarrow (A,Z+2) + 2e^-$, will establish the Majorana nature and lepton number violation beyond any doubt [@Furry:1939qr]. Therefore, the search for neutrinoless double beta decay continues to be an important area. Theoretically as well, decay is heralded as a useful probe of physics beyond the standard model (SM). can potentially discriminate between the two hierarchies of the neutrino masses, and this, in turn can be used to rule out specific models of neutrino mass generation. In the context of models which involve TeV scale particles, like low scale seesaw models or low energy supersymmetric models including models with R-parity violation, imposes stringent constraints on the model parameters. The same set of diagrams, with appropriate changes in the momentum flow, can lead to interesting signatures at LHC. Constraints from thus can prove rather useful for phenomenological studies (see e.g. [@Keung:1983uu] for an incomplete list discussing various aspects). The decay amplitude can be split into the so called long-range and short-range parts (for a review of theoretical and experimental issues and the sources of uncertainties and errors, see [@Rodejohann:2011mu] and references therein ). Here, the long-range refers to the fact that there is an intermediate light neutrino involved. This should be contrasted with the short-range part of the amplitude in which the intermediate particles are all much much heavier that the relevant scale of the process $\sim {\mathcal{O}}$(GeV). In such a case, the heavier degrees of freedom can be systematically integrated out leaving behind a series of operators built out of low energy fields weighted by coefficients, called Wilson coefficients (denoted by $C_i$ below), which are functions of the parameters of the large mass degrees of freedom that have been integrated out (see e.g [@Georgi:1994qn]). This provides a very convenient framework to evaluate the decay amplitude in terms of short distance coefficients which encode all the information about the high energy physics one may be trying to probe via a low energy process. This also neatly separates the particle physics input from the nuclear physics part which enters via the nuclear matrix elements (NMEs) of the quark level operators sandwiched between the nucleon states. In what follows, the discussion will be centered around the short range part though we believe that many of the arguments and results may also apply to the long range part. More care may be needed in the latter case though. Given a specific model it is straightforward to write down the amplitude for the quark level process and compute the short distance coefficient. The complete amplitude then involves NMEs. At present, the biggest source of uncertainty stems from the NMEs, and theoretical predictions show a marked sensitivity on the NMEs used (see [@Simkovic:2007vu] for some of the recent NME calculations and predictions for rates). On the experimental side, studies have been carried out on several nuclei. Only one of the experiments, the Heidelberg-Moskow collaboration (HM collab.) [@KlapdorKleingrothaus:2006ff] has claimed observation of signal in $^{76}{\mathrm Ge}$. The half-life at $68\%$ confidence level is: $T^{0\nu}_{1/2}(^{76}{\mathrm Ge}) = 2.23^{+0.44}_{-0.31}\times 10^{25}\, {\mathrm yr}$. A combination of the Kamland-Zen [@Gando:2012zm] and EXO-200 [@Auger:2012ar] results, both using $^{136}{\mathrm Xe}$, yields a lower limit on the half-life $T^{0\nu}_{1/2}(^{136}{\mathrm Xe}) > 3.4 \times 10^{25}\, {\mathrm yr}$ which is at variance with the HM claim. Very recently GERDA experiment reported the lower limit on the half-life based on the first phase of the experiment [@Agostini:2013mzu]: $T^{0\nu}_{1/2}(^{76}{\mathrm Ge}) > 2.1 \times 10^{25}\, {\mathrm yr}$. A combination of all the previous limits results in a lower limit $T^{0\nu}_{1/2}(^{76}{\mathrm Ge}) > 3.0 \times 10^{25}\, {\mathrm yr}$ at $90\%$ confidence level. The new GERDA result (and the combination) is (are) again at odds with the positive claim of HM collab. The GERDA results have been challenged [@Klapdor-Kleingrothaus:2013cja] on account of low statistics and poorer resolution. Very clearly, there is some tension among the experimental results and higher statistics in future will shed more light. To reduce the dependence (or sensitivity) on NMEs, predictions for for various nuclei can be compared. Further, it is necessary to establish if the long-range contribution, coming from the light neutrino exchange, can saturate the experimental limits (or positive claims). This is investigated in [@Dev:2013vxa], and the conclusion drawn is that the light neutrino exchange falls short of saturating the current limits. Also, for some choices of NMEs, the $^{76}{\mathrm Ge}$ positive result can be consistent with $^{136}{\mathrm Xe}$ limits when considered individually but not when combined. In view of the immense importance of , both experimentally and theoretically, it is important to ensure that theoretical calculations are very precise. In the present article, we consider dominant one loop QCD corrections and renormalization group effects to the amplitude. To the best of our knowledge, this has not been studied before and as we show below, QCD corrections can have a significant impact on the rate, thereby impacting the constraints on the model parameters. We begin by recapitulating the essential steps in arriving at the final amplitude for . Using the Feynman rules for a given model, all possible terms can be easily written. Since the momentum flowing through any of the internal lines is far smaller than the masses of the respective particles and can be neglected, this leads to the low energy amplitude at the quark level. Parts of the amplitude may require Fierz rearrangement (for example in supersymmetric theories) to express it in colour singlet form which can then be sandwiched between the nucleon states after taking the non-relativistic limit. This last step results in NMEs. We shall not be concerned with the issue of uncertainties creeping in due to NME calculations here. We shall, rather, choose to work with a particular set of NMEs and focus on the impact of perturbative QCD corrections. As an example, consider a heavy right handed neutrino and SM gauge group. The resulting amplitude is of the form $$\begin{aligned} {\mathcal{A}} &\sim& \frac{1}{M_W^4M_N}\bar{u}\gamma_{\mu}(1-\gamma_5)d\,\bar{e}\gamma^{\mu}\gamma^{\nu}(1+\gamma_5)e^c\, \bar{u}\gamma_{\nu}(1-\gamma_5)d \nonumber \\ &=& \underbrace{\frac{1}{M_W^4M_N}}_{G}\underbrace{\bar{u}\gamma_{\mu}(1-\gamma_5)d\, \bar{u}\gamma^{\mu}(1-\gamma_5)d}_{{\mathcal{J}}_{q,\mu}{\mathcal{J}}_q^{\mu}}\, \underbrace{\bar{e}(1+\gamma_5)e^c}_{j_l}\end{aligned}$$ where we used $\gamma_{\mu}\gamma_{\nu} = g_{\mu\nu}-2i\sigma_{\mu\nu}$ and the fact that $\bar{e}\sigma_{\mu\nu}(1+\gamma_5)e^c $ vanishes identically. So does $\bar{e}\gamma_{\mu}e^c$. This was noted in [@Prezeau:2003xn]. These, thus, restrict the form of the leptonic current. $G$ denotes the analogue of Fermi constant. The exact form of $G$ will be model dependent. The physical amplitude is written as $${\mathcal{A}}_{0\nu 2\beta} = \langle f\vert i{\mathcal{H}}_{\mathrm eff}\vert i\rangle \sim G\, \underbrace{\langle f\vert {\mathcal{J}}_{q,\mu}{\mathcal{J}}_q^{\mu}\vert i\rangle}_{\boldmath NME}\,j_l$$ This clearly illustrates how the short distance or high energy physics separates from the low energy matrix elements. The effective Hamiltonian for a given model is expressed as a sum of operators, $O_i$ weighted by the Wilson coefficents $C_i$: ${\mathcal{H}}_{\mathrm eff} = G_i C_i O_i$, where we have allowed for more than one $G$ for more complicated theories. In the above case, there is only one operator $O_1 = {\mathcal{J}}_{q,\mu}{\mathcal{J}}_q^{\mu}\,j_l = \bar{u_i}\gamma_{\mu}(1-\gamma_5)d_i\,\bar{u_j}\gamma^{\mu}(1-\gamma_5)d_j\, \bar{e}(1+\gamma_5)e^c$ ($i,j$ denoting the colour indices) and the corresponding Wilson coefficient $C_1=1$. In other models like SUSY with R-parity violation [@Mohapatra:1986su] or leptoquarks [@Hirsch:1996qy], Fierz transformations have to be employed to bring the operators in the colour matched form. The specific NME that finally enters the rate depends on the Lorentz and Dirac structure of the quark level operator involved. This is not the entire story. From the effective field theory point of view, the integrating out of the heavier degrees of freedom happens at the respective thresholds and then the obtained effective Lagrangian/Hamiltonian has to be properly evolved down to the relevant physical scale of the problem ($\sim {\mathcal{O}}$(GeV) in the present case). This is similar to what happens in non-leptonic meson decays (see for example [@Buchalla:1995vs]). For simplicity, we assume that the heavy particles are all around the electroweak (EW) scale and in obtaining the numerical values, we shall put $M_W$ as the scale for all. This facilitates one step integrating out of all the heavy degrees of freedom. Therefore, the above statement about $C_1$ being unity should now be written as $C_1(M_W) = 1$. Next consider one loop QCD corrections. The full amplitude is evaluated with one gluon exchange ($O(\alpha_s) $) and matched with the amplitude at the same order in $\alpha_s$ in the effective theory. Fig.\[fig1\] shows representative diagrams in the full and effective theory. This has two effects: (i) $C_1$ gets corrected and reads $C_1(M_W) = 1 + \frac{\alpha_s}{4\pi}{\mathcal{N}} \ln\left(\frac{M_W^2}{\mu_W^2}\right)$, where $\mu_W$ is the renormalization scale and ${\mathcal{N}}$ is a calculable quantity. This coefficient is then evolved down to the ${\mathcal{O}}$(GeV) using the renormalization group (RG) equations; (ii) QCD corrections induce the colour mismatched operator $O_2 = \bar{u_i}\gamma_{\mu}(1-\gamma_5)d_j\,\bar{u_j}\gamma^{\mu}(1-\gamma_5)d_i\,\bar{e}(1+\gamma_5)e^c$ with coefficient $C_2 = \frac{\alpha_s}{4\pi}{\mathcal{N}}' \ln\left(\frac{M_W^2}{\mu_W^2}\right)$. When evaluating the quark level matrix element in the effective theory, both the operators contribute and in fact lead to mixing. This approach is a consistent one and also reduces the scale dependence of the physical matrix elements. Without following the above steps, the short distance coefficient would have been evaluated at the high scale while the physical matrix elements at a low scale, leading to large scale dependence, which is not a physical effect but rather an artifact of the calculation. Armed with this machinery, we now consider specific examples to bring out the impact of QCD corrections. As mentioned above, to simplify the discussion, we assume all the heavy particles beyond the SM to be around the EW scale. The technical details and explicit expressions for some of the models leading to neutrinoless double beta decay and related phenomenology will be presented elsewhere. Here we provide approximate numerical values of the Wilson coefficients of the operators considered. For the time being, we neglect the mixing of operators under renormalization. This can have a large impact on some of the coefficients but their inclusion is beyond the scope of the present work. 0.32cm 1.35cm First we consider left-right symmetric model and focus our attention on operators generated due to $W_L$ and $W_R$ exchange: $$\begin{aligned} O^{LL}_1 &=& \bar{u_i}\gamma_{\mu}(1-\gamma_5)d_i\,\bar{u_j}\gamma^{\mu}(1-\gamma_5)d_j\,\bar{e}(1+\gamma_5)e^c \nonumber \\ O^{LL}_2 &=& \bar{u_i}\gamma_{\mu}(1-\gamma_5)d_j\,\bar{u_j}\gamma^{\mu}(1-\gamma_5)d_i\,\bar{e}(1+\gamma_5)e^c \nonumber \\ O^{RR}_1 &=& \bar{u_i}\gamma_{\mu}(1+\gamma_5)d_i\,\bar{u_j}\gamma^{\mu}(1+\gamma_5)d_j\,\bar{e}(1+\gamma_5)e^c \nonumber \\ O^{RR}_2 &=& \bar{u_i}\gamma_{\mu}(1+\gamma_5)d_j\,\bar{u_j}\gamma^{\mu}(1+\gamma_5)d_i\,\bar{e}(1+\gamma_5)e^c \nonumber \\ O^{LR}_1 &=& \bar{u_i}\gamma_{\mu}(1-\gamma_5)d_i\,\bar{u_j}\gamma^{\mu}(1+\gamma_5)d_j\,\bar{e}(1+\gamma_5)e^c \nonumber \\ O^{LR}_2 &=& \bar{u_i}\gamma_{\mu}(1-\gamma_5)d_j\,\bar{u_j}\gamma^{\mu}(1+\gamma_5)d_i\,\bar{e}(1+\gamma_5)e^c\end{aligned}$$ Following the general steps outlined above, the Wilson coefficents can be evaluated at the high scale and run down to $\mu \sim {\mathcal{O}}$(GeV) (see also [@Cho:1993zb]). Their approximate values read: $$\begin{aligned} C^{LL,RR}_1 \sim 1.3 &,& C^{LL,RR}_2 \sim -0.6 \nonumber\\ C^{LR,RL}_1 \sim 1.1 &,& C^{LR,RL}_2 \sim 0.7\end{aligned}$$ To evaluate the physical matrix elements, the colour mismatched operators $O^{AB}_2$ have to be Fierz transformed. Under Fierz rearrangement, $(V-A)\otimes (V-A)$ and $(V+A)\otimes (V+A)$ retain their form while $(V-A)\otimes (V+A) \rightarrow -2 (S-P)\otimes (S+P)$. With this rearrangement, $LL,\,RR$ operators effectively yield $C^{LL,RR}_1+C^{LL,RR}_2$ as the effective couplings with the same NMEs involved, implying substantial cancellation (by about a factor of two). The $LR$ operator Fierz transformed brings in a different combination of NMEs. Explicitly, following for example the last reference in [@Rodejohann:2011mu], we have the following (not showing the lepton current explicitly): $$\langle {\mathcal{J}}^{(V\pm A)}{\mathcal{J}}_{(V\pm A)}\rangle \propto \frac{m_A}{m_Pm_e} ({\mathcal{M}}_{GT,N}\, \mp \alpha^{SR}_3 {\mathcal{M}}_{F,N})$$ where $\vert{\mathcal{M}}_{GT,N}\vert \sim (2-4)\vert{\mathcal{M}}_{F,N}\vert$ for all the nuclei considered, and $\alpha^{SR}_3 \sim 0.63$. Thus, to a good accuracy the above matrix element is essentially governed by ${\mathcal{M}}_{GT,N}$. In the above equation, the relative negative sign between the two terms on the right hand side corresponds to $(V+A)\otimes (V+A)$ and $(V-A)\otimes (V-A)$ structures on the left hand side, while for the $(V+A)\otimes (V-A)$ structure, the relative sign is positive. On the other hand, $$\langle {\mathcal{J}}^{(S\pm P)}{\mathcal{J}}_{(S\pm P)}\rangle \propto - \alpha^{SR}_1 {\mathcal{M}}_{F,N}$$ with $\alpha^{SR}_1 \sim 0.145\frac{m_A}{m_Pm_e}$. Clearly, the Fierz transformed operator in this case turns out to be subdominant. This simple exercise illustrates the large impact and importance of QCD corrections in the context of . As obtained above, QCD corrections can lead to substantial shift in the rate for specific models, thereby changing the limits on the model parameters significantly. As our next example, consider theories where the interactions are $S\pm P$ form, like SUSY with R-parity violation or leptoquarks etc. In such cases, the operators have the structure: $$\begin{aligned} O^{SP\pm\pm}_1 &=& \bar{u_i}(1\pm\gamma_5)d_i\, \bar{u_j}(1\pm\gamma_5)d_j\, \bar{e}(1+\gamma_5)e^c \nonumber \\ O^{SP\pm\pm}_2 &=& \bar{u_i}(1\pm\gamma_5)d_j\, \bar{u_j}(1\pm\gamma_5)d_i\, \bar{e}(1+\gamma_5)e^c \nonumber \\ O^{SP+-}_1 &=& \bar{u_i}(1+\gamma_5)d_i\, \bar{u_j}(1-\gamma_5)d_j\, \bar{e}(1+\gamma_5)e^c \nonumber \\ O^{SP+-}_2 &=& \bar{u_i}(1+\gamma_5)d_j\, \bar{u_j}(1-\gamma_5)d_i\, \bar{e}(1+\gamma_5)e^c \end{aligned}$$ The Wilson coefficients of the colour mismatched operators are about 0.1-0.5 times those of the colour allowed operators in magnitude. This could be argued from the $1/N_c(\,\sim 0.3\,\,\textrm{for} N_c=3)$ counting rules for the colour mismatched structures, up to factors of order unity. Following the same chain of arguments, the colour mismatched operators need to be Fierz transformed before computing the physical matrix elements. Under Fierz transformations we have: $(S+P)\otimes (S-P) \rightarrow \frac{1}{2}(V+A)\otimes (V-A)$ implying that the colour mismatched operator, after Fierz transformation, may provide the dominant contribution (see Eq.(5) and Eq.(6)). Consequently the amplitudes, and therefore the limits on parameters may change by a factor of five or so. That the colour mismatched operator can provide a large contribution is again something we are familiar with from $K\to\pi\pi$ decays where the QCD (and electroweak) penguin operator after Fierz transformation gives the dominant contribution, though QCD and electroweak penguin contributions tend to cancel each other in this case. The most interesting and the largest effect in the examples considered above comes about when considering the $O^{SP++,--}$ operators. $(S\pm P)\otimes (S\pm P) \rightarrow \frac{1}{4}[2(S\pm P)\otimes (S\pm P) - (S\pm P)\sigma_{\mu\nu}\otimes\sigma^{\mu\nu}]$ under Fierz rearrangement. The tensor-pseudotensor structure yields the following NME: $$\langle {\mathcal{J}}^{\mu\nu}{\mathcal{J}}_{\mu\nu}\rangle \propto -\alpha^{SR}_2 {\mathcal{M}}_{GT,N}$$ with $\alpha^{SR}_2 \sim 9.6\frac{m_A}{m_Pm_e}$ which is about $200$ times larger than $\langle {\mathcal{J}}^{(S\pm P)}{\mathcal{J}}_{(S\pm P)}\rangle$. Conservatively taking the corresponding Wilson coefficient to be $0.1$ of the colour allowed operator, the relative contributions are: $$\vert\frac{O^{SP++}_2}{O^{SP++}_1}\vert \geq 10$$ The above discussion makes it very clear that the QCD corrections to are rather important and should be included systematically. These corrections can be as large as or in fact larger than in most cases than the uncertainty due to NMEs and are independent of the particular set of NMEs considered. As eluded to above, we have considered only pairs of operators $O^{AB}_1,\, O^{AB}_2$ while obtaining the approximate values of $C's$ at the low scale. The effect of mixing with other operators has been ignored at this stage. This could further lead to significant corrections for some of the operators. We plan to systematically investigate these issues elsewhere. This (and the shift above) is rather large and can completely change the phenomenological constraints. In theories with many contributions to , it is essential to understand the interplay between different competing amplitudes to set limits on the couplings and masses of the particles. In such cases, the discussion above becomes even more important. Low (TeV) scale models appear to be attractive due to plausible signatures at LHC, where QCD corrections will be inevitable. It is therefore important to include the dominant QCD corrections at the very least in order to set meaningful limits on model parameters.\ In this article we have investigated the impact of one loop QCD corrections to the amplitude. This, to the best of our knowledge, is the first time this issue has been discussed. We found that QCD corrections can have a large impact, ranging from near cancellation to huge enhancement of the rate. Since is an important process to search experimentally and has the potential to link seemingly unrelated processes, particularly in the context of TeV scale models, it is rather important to ensure that theoretical predictions are precise enough to be compared to the experimental results. As such, the calculations suffer from large uncertainties due to the choice of NMEs, which are non-perturbative in nature. What we have found is that even perturbative corrections have the potential to shift the predictions by a large amount. This by itself is a rather important aspect and such corrections need to be systematically computed for various models of interest. The shift in the limits on model parameters also implies that the related phenomenology at say the LHC (in specific models) will also get modified. There are other issues related to operator mixing which have not been incorporated here. These may also become important in the context of specific theories and should be consistently included. Furthermore, QCD corrections need to be evaluated for the light neutrino exchange contribution as well. As mentioned in the beginning, the light neutrino contribution is unable to saturate the present experimental limits. It remains to be seen if including the radiative corrections eases out this tension, and to what extent [@nm]. [99]{} D. V. Forero, M. Tortola and J. W. F. Valle, Phys. Rev. D [**86**]{}, 073012 (2012) \[arXiv:1205.4018 \[hep-ph\]\]; G. L. Fogli, E. Lisi, A. Marrone, D. Montanino, A. Palazzo and A. M. Rotunno, Phys. Rev. D [**86**]{}, 013012 (2012) \[arXiv:1205.5254 \[hep-ph\]\]; M. C. Gonzalez-Garcia, M. Maltoni, J. Salvado and T. Schwetz, JHEP [**1212**]{} (2012) 123 \[arXiv:1209.3023 \[hep-ph\]\]. E. Majorana, Nuovo Cim.  [**14**]{}, 171 (1937). W. H. Furry, Phys. Rev.  [**56**]{}, 1184 (1939); J. Schechter and J. W. F. Valle, Phys. Rev. D [**25**]{}, 2951 (1982). W. -Y. Keung and G. Senjanovic, Phys. Rev. Lett.  [**50**]{}, 1427 (1983); H. V. Klapdor-Kleingrothaus and U. Sarkar, Mod. Phys. Lett. A [**16**]{}, 2469 (2001) \[hep-ph/0201224\]; H. V. Klapdor-Kleingrothaus and U. Sarkar, Mod. Phys. Lett. A [**18**]{}, 2243 (2003) \[hep-ph/0304032\]; M. Blennow, E. Fernandez-Martinez, J. Lopez-Pavon and J. Menendez, JHEP [**1007**]{}, 096 (2010) \[arXiv:1005.3240 \[hep-ph\]\]; A. Ibarra, E. Molinaro and S. T. Petcov, JHEP [**1009**]{}, 108 (2010) \[arXiv:1007.2378 \[hep-ph\]\]; V. Tello, M. Nemevsek, F. Nesti, G. Senjanovic and F. Vissani, Phys. Rev. Lett.  [**106**]{}, 151801 (2011) \[arXiv:1011.3522 \[hep-ph\]\]; M. Nemevsek, F. Nesti, G. Senjanovic and Y. Zhang, Phys. Rev. D [**83**]{}, 115014 (2011) \[arXiv:1103.1627 \[hep-ph\]\]; J. Bergstrom, A. Merle and T. Ohlsson, JHEP [**1105**]{}, 122 (2011) \[arXiv:1103.3015 \[hep-ph\]\]; M. Nemevsek, F. Nesti, G. Senjanovic and V. Tello, arXiv:1112.3061 \[hep-ph\]; J. C. Helo, M. Hirsch, H. Pas and S. G. Kovalenko, arXiv:1307.4849 \[hep-ph\]; P. S. B. Dev, C. -H. Lee and R. N. Mohapatra, arXiv:1309.0774 \[hep-ph\]. W. Rodejohann, Int. J. Mod. Phys. E [**20**]{}, 1833 (2011) \[arXiv:1106.1334 \[hep-ph\]\]; J. J. Gomez-Cadenas, J. Martin-Albo, M. Mezzetto, F. Monrabal and M. Sorel, Riv. Nuovo Cim.  [**35**]{}, 29 (2012) \[arXiv:1109.5515 \[hep-ex\]\]; J. D. Vergados, H. Ejiri and F. Simkovic, Rept. Prog. Phys.  [**75**]{}, 106301 (2012) \[arXiv:1205.0649 \[hep-ph\]\]; F. F. Deppisch, M. Hirsch and H. Pas, J. Phys. G [**39**]{}, 124007 (2012) \[arXiv:1208.0727 \[hep-ph\]\]. H. Georgi, Ann. Rev. Nucl. Part. Sci.  [**43**]{}, 209 (1993). F. Simkovic, A. Faessler, V. Rodin, P. Vogel and J. Engel, Phys. Rev. C [**77**]{}, 045503 (2008) \[arXiv:0710.2055 \[nucl-th\]\]; A. Faessler, G. L. Fogli, E. Lisi, V. Rodin, A. M. Rotunno and F. Simkovic, Phys. Rev. D [**79**]{}, 053001 (2009) \[arXiv:0810.5733 \[hep-ph\]\]; J. Menendez, A. Poves, E. Caurier and F. Nowacki, Nucl. Phys. A [**818**]{}, 139 (2009) \[arXiv:0801.3760 \[nucl-th\]\]; P. K. Rath, R. Chandra, K. Chaturvedi, P. K. Raina and J. G. Hirsch, Phys. Rev. C [**82**]{}, 064310 (2010) \[arXiv:1104.3965 \[nucl-th\]\]; P. K. Rath, R. Chandra, P. K. Raina, K. Chaturvedi and J. G. Hirsch, Phys. Rev. C [**85**]{}, 014308 (2012) \[arXiv:1106.1560 \[nucl-th\]\]; A. Meroni, S. T. Petcov and F. Simkovic, JHEP [**1302**]{}, 025 (2013) \[arXiv:1212.1331\]; J. Barea, J. Kotila and F. Iachello, Phys. Rev. C [**87**]{}, 014315 (2013) \[arXiv:1301.4203 \[nucl-th\]\]; M. T. Mustonen and J. Engel, arXiv:1301.6997 \[nucl-th\]; F. Simkovic, V. Rodin, A. Faessler and P. Vogel, Phys. Rev. C [**87**]{}, 045501 (2013) \[arXiv:1302.1509 \[nucl-th\]\]. H. V. Klapdor-Kleingrothaus and I. V. Krivosheina, Mod. Phys. Lett. A [**21**]{}, 1547 (2006). A. Gando [*et al.*]{} \[KamLAND-Zen Collaboration\], Phys. Rev. Lett.  [**110**]{}, no. 6, 062502 (2013) \[arXiv:1211.3863 \[hep-ex\]\]. M. Auger [*et al.*]{} \[EXO Collaboration\], Phys. Rev. Lett.  [**109**]{}, 032505 (2012) \[arXiv:1205.5608 \[hep-ex\]\]. M. Agostini [*et al.*]{} \[GERDA Collaboration\], Phys. Rev. Lett.  [**111**]{}, 122503 (2013) \[arXiv:1307.4720 \[nucl-ex\]\]. H. V. Klapdor-Kleingrothaus, I. V. Krivosheina and S. N. Karpov, arXiv:1308.2524 \[hep-ex\]. P. S. Bhupal Dev, S. Goswami, M. Mitra and W. Rodejohann, arXiv:1305.0056 \[hep-ph\]. G. Prezeau, M. Ramsey-Musolf and P. Vogel, Phys. Rev. D [**68**]{}, 034016 (2003) \[hep-ph/0303205\]. R. N. Mohapatra, Phys. Rev. D [**34**]{}, 3457 (1986). M. Hirsch, H. V. Klapdor-Kleingrothaus and S. G. Kovalenko, Phys. Lett. B [**378**]{}, 17 (1996) \[hep-ph/9602305\]. G. Buchalla, A. J. Buras and M. E. Lautenbacher, Rev. Mod. Phys.  [**68**]{}, 1125 (1996) \[hep-ph/9512380\]. D. Binosi and L. Theussl, Comput. Phys. Commun.  [**161**]{}, 76 (2004) \[arXiv:hep-ph/0309015\]; D. Binosi, J. Collins, C. Kaufhold and L. Theussl, Comput. Phys. Commun.  [**180**]{}, 1709 (2009) \[arXiv:0811.4113 \[hep-ph\]\]. P. L. Cho and M. Misiak, Phys. Rev. D [**49**]{}, 5894 (1994) \[hep-ph/9310332\]. N. Mahajan, [*in preparation*]{}.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We present some clues to the study of the renormalization group, at graduate level, as well as some bibliographical pointers to classical resources. Just the kind of things one had liked to hear when starting to study the subject.' author: - 'A. Rivero [^1]' title: A short lecture on Divergences --- This was going to be a notice on “recent advances on renormalization group theory”, from which advanced students can get bibliography to please their teachers. But I happened to visit Jacques Gabay’s printing house, and I decided to take a wider view. At least, wider than usual treatises on Quantum Field Theory. Gabay’s mission is to keep in print old mathematics texts from the late XIXth and early XXth, and his work helps to keep the perspective. Fact is, all generations of physicists since Euler times have been used to live with divergences. Students are supposed to become exposed to the subject gradually, but this gradation varies strongly across schools and faculties, and the balance keeps more in the room of Cauchy than in Borel quarters. There is even a darker side, about if everything which is legal in Mathematics should be legal in Physics, but this question keeps usually in the philosophical level. Still, one should point that classical mechanics, the science of newtonian limits, is mathematically legal but physically ruled out! The usual scenario for divergences is: we have a differential equation. We look for a solution from power series expansion. Most times, the solution is known to exist, say from Picard’s fixed point method, say from other convergence theorem. But we are forced to pick up a power series expansion on the “wrong” parameter, so that the convergence radius is zero. Or (change $x-> 1/z$ if necesary) we could to know only a asymptotic series around infinity. Poincare treatises are the first ones showing all of this. Note, still, that we are here in a purely classical matter. We can do perturbation theory with a small coupling constant in a perturbation potential, and then expand the solution as a power series of the coupling constant. The series diverges, and then we have only an asymptotic series. Borel transform, or other resummation techniques, can be invoked to get a better expansion. Sometimes even a convergent series is known, for instance for the three-body problem (which, remember, is chaotical respect to small perturbations), but even then it can happen that the divergent series is faster than the convergent one for a given level of precision! This kind of problems when solving perturbation series is thus expected to appear also in quantum mechanics and quantum field theory. In fact, an argument from Dyson [@dyson] shows that the perturbative expansion of quantum electrodynamics will give place to a divergent series of this kind. This is not to be confused with renormalizability questions, which will refer to each term in this expansion. A interesting detail of the quantum approach is that it does not start from the differential equation, but from the action principle. This is a puzzling thing, because the action principle is an integral between two conditions, and it must be connected to a local differential equation. It seems that quantum theory has an alternate way to go from the lagrangian, in the action law, to the hamiltonian, in Schroedinger equation; it is very suggestive to read this derivation in the original article of Feynman[@feynman]. Later, Schwinger and Dyson produced a way to deduce, from variation of a quantum action, the canonical equations of quantum theory. In any case, calculation efficiency suggest to start always from perturbations of the action, and this is the engine who drives to diagrammatics. Also, we have a richer set of expansions when going to quantum theory. Besides the perturbative method around a coupling constant, we have the possibility to expand on Plank constant. In QM, the Schroedinger equation is written in Ricatti form, $$i \hbar {d\xi \over dx} + 2 m (E-V) - \xi^2 =0$$ and then we expand $$\xi = \sum \xi_n(x) (\hbar / i)^n$$ AFC [@AFC] refer to this as the eikonal expansion (This change of variables is characteristic also of the Hoft algebra of diffeomorfisms used by Connes and Kreimer). And the diagrams give still other expansion; if we are in QFT with only a coupling constant, then a series on Plank constant coincides with the expansion on number of loops. But for the general case, we have still this third possibility, on the number of loops of each Feynman diagram. Besides perturbations, one could expect also divergences for bad shaped, singular, potentials, but their study has been bypassed by history. Still, it is worth to mention that the gravitating n-body problem can be formulated -and solved- including collisions throught the singularities. Well, this was the world of our grand-grandfathers in the first decades of XXth century. Then it came a more painful source of divergences. Not only happened that the perturbation expansion was around bad places, but also that every term in the expansion happened to contain a divergence of his own. As the effect relates to the indeterminate creation of virtual particles, this is the thing we are expected to learn in any QFT course. Still, it is worth to mention the three steps in taking control of this source of problems: –First, Feynman, Dyson and Schwinger got to recognize how the divergence can be absorbed in the coupling constant in a consistent form. This comes from a lot of previous experience on considering the cloud of virtual particles around a singular charge, and the real surprise is that just a finite number of coupling constants can adjust for all the divergences. –Second, Gellman and Low[@gl], in a paper of compulsory reading, identify a general method to predict the renormalized values without entering in detailed calculation: the renormalization group. Asymptotical analysis get here a completely new meaning. –Third, Kadanoff, Wilson and Kogut recognize the renormalization group as a question of scaling between different orders of magnitude. The study of fixed points, and perturbations around them, conects this approach with the previous results. In the first step, the problem is traced to an limit when $x-y$ goes to zero. In the Second step, the bug is transformed in a feature, and general properties of this $x-y$ divergence are systematised. In the third step, the question of existence of scales is related with the problem of continuum limit of a lattice. A renormalization “semigroup”, decimating degrees of freedom, is sketched, and Wilson shows how it can be related to the previous G-L group. A fourth step could be the finding of asymptotically free theories; this is standard lore in your textbooks. Wilson’s approach has the merit of making very explicit the role of the cut-off that regularises the theory. In a renormalisable theory, this cut-off is removed after the manipulations and then the Gellman-Low equation (and Callan-Simanzik) become apparent. It is possible to keep the cut-off and to do calculations with an “effective” theory, even if it is not renormalisable. Feelings about this can be mixed. One paralell which can be invoked is the origin of quantum mechanics: The Plank constand did appear in physics as a mean to regularise the energy distribution, then getting finite results where the Rayleigh-Jeans theory was divergent. Also other cut-off techniques are natural in thermodinamics; in any case it is very delicate to extend these similtudes, as one must distinguish between ultraviolet and infrared divergences; in this presentation we are almost forgetting the IR ones. After recognising the role of $x-y$ in the divergences, it is more natural to ask “what about QM?” in the straighforward formulation as $0+1$ dimensional field theory. Of course, power counting in a $0+1$ theory shows that any polynomial interaction will be free of divergences (in the sense of renormalization). But it is very interesting to study the role of $t_1-t_0$ from the point of view of scaling and renormalization theory. This exercise was published by Polonyi[@polonyi] in 1993[^2]. It is not only QM, by the way. The ideas coming from renormalization group, and perhaps the renormalization group itself, descent back to the theory of classical evolution of differential equations. There, dimensional analysis and scaling got a fresh airh from both Wilson and GellMann-Low techniques. We could mention the works of Goldenfeld-Chen- Oono[@gco], as well as the ones of Kunihiro, while AFC describes some previous attemps middle way between classical and quantum. Also the original question of resummation techniques (by changing the coefficients in the perturbative expansion) get some help from this, even if it sounds strange, it seems it is possible to mimic QFT and to use the techniques of control of coefficients (the RG) to control the whole expansion. In the mean time, since Wilson’s age, the renormalization group had been rigorized from the point of view of mathematical physics. This task was accomplished by Bogoliugov, who worked out both the distributional character of the theory and the mechanics of the substraction process. The approach summarized into the BPHZ method, and additionaly some more analytical part of this research drove to the formulation of Epstein and Glaser. This was all the rigour a physicist could even desire.The surprise come recently when Connes and Kreimer found a rigorization from pure mathematics. Kreimer got a method to map Feynman diagrams to rooted trees, and it happened that this method formulated a Hoft algebra, nice enough to hold the group structure of renormalization. And this surprise came with a message from calculus, when it was noticed, by Brouder[@brouder], that the Hoft algebra of trees was the same technique that had been used by Butcher in the seventies, to classify the numerical Runge Kutta methods! We can understand why if we look from a very general point of view, stratification of compactified configuration spaces, as Kreimer explains lately. Or we can go back in the time, when the series expansion of the solutions of a differential equation was classified by labelling each term with a Cayley tree. Start from a differential equation $x'=F(x)$. One want to get the series $x(t)= x_0 + x_1 t + x_2 t^2 + ...$, and it is straighforward that we can proceed by derivating the original equation to get first $$x''= F'(x) x' = F'(x) F(x)$$ then $$x'''=F''(x) (x')^2 + F'(x) x'' = F'' F^2 + F' F' F$$ and so on. The increasing complexity of the expansion can be tamed by using rooted trees to describe each term[@conn.l]. The same technique is used to tame the terms in a QFT expansion and to recursively apply the renormalization, so it seems we are again approaching in the domain of the theory of differential calculus. If the circle finally closes, we will have learn some very deep lessons on the structure of calculus and its interaction with physics. [30]{} G.A Arteca, F.M. Fernandez, E.A. Castro, [*Large Order Perturbation Theory and Summation Methods in Quantum Mechanics*]{}, Lecture Notes In Chemistry, 53, Springer E. Borel, [*Lessons sur las series divergentes*]{}, ed. Gautier-Villars, 1928 (reprinted editions Jacques Gabay) E. Brezin, J.C. LeGuillou, J. Zinn-Justin, [ *Perturbation theory at large Order, I et II*]{}, Phys Rev D, v. 15, p. 1544 and Phys Rev D v. 15, p. 1558 Ch. Brouder, [Runge-Kutta methods and renormalization]{}, European Physical Journal C v. 12, p. 521-534 J.C. Butcher, [*An Algebraic Theory of Integration Methods*]{} Math. Comp. v. 26, p. 79 For instance A. Connes and D. Kreimer hep-th/9912092, as well as D. Kreimer q-alg/9707029 and hep-th/0010059 Alain Connes and Dirk Kreimer, [*Lessons from Quantum Field Theory*]{}, hep-th/9904044 F.J.Dyson, Phys Rev 85, p. 631 LY Chen, N. Goldenfeld, Y. Oono, Phys. Rev. E, v. 54 p. 376 Feynman, [*Space-Time Approach to NR Quantum Mechanics*]{} Rev Mod Phys 20, p. 367 M. Gell-Mann and F.E. Low [*Quantum Electrodinamics at Small Distances*]{}, Phys. Rev. v. 95, p. 1300 G t’Hooft, Nucl Phys B 35, p. 167; G. t’Hooft and M.Veltman, Nucl. Phys. B44 p. 189 D.J. Broadhurst and D. Kreimer, [*Renormalization tamed: 30 loop Pade Borel Resumation*]{}, hep-th/9912093 T. Kunihiro, for instance hep-th/9505166 and hep-th/9801196 Polonyi, arxiv:hep-th/9409004, hep-th/9412042 and hep-th/9711061 Tim R. Morris, hep-th/9802039 [^1]: EU Politecnica de Teruel, Universidad de Zaragoza [^2]: Tarrach in Barcelona, as well as Boya and myself in Zaragoza, did took some interest on different views of the renormalization process of QM
{ "pile_set_name": "ArXiv" }
--- abstract: 'In this paper we develop a theory of Grothendieck’s six operations of lisse-étale constructible sheaves on Artin stacks locally of finite type over an affine regular noetherian scheme of dimension $\leq 1$. We also generalize the classical base change theorems and Kunneth formula to stacks, and prove new results about cohomological descent for unbounded complexes.' address: - 'École Polytechnique CMLS UMR 7640 CNRS F-91128 Palaiseau Cedex France' - 'University of Texas at Austin Department of Mathematics 1 University Station C1200 Austin, TX 78712-0257, USA' author: - Yves Laszlo and Martin Olsson title: 'The six operations for sheaves on Artin stacks I: Finite Coefficients' --- Introduction ============ We denote by $\Lambda$ a Gorenstein local ring of dimension $0$ and characteristic $l$. Let $S$ be an affine regular, noetherian scheme of dimension $\leq 1$ and assume $l$ is invertible on $S$. We assume that all $S$-schemes of finite type $X$ satisfy $\mathrm{cd}_{l}(X)<\infty$ (see \[cohdimension\] for more discussion of this). For an algebraic stack ${\mathcal}X$ locally of finite type over $S$ and $*\in \{+, -, b, \emptyset, [a,b]\}$ we write $\operatorname{{D}}_c^*({\mathcal}X)$ for the full subcategory of the derived category $\operatorname{{D}}^*({\mathcal}X)$ of complexes of $\Lambda $–modules on the lisse-étale site of ${\mathcal}X$ with constructible cohomology sheaves. In this paper we develop a theory of Grothendieck’s six operations of lisse-étale constructible sheaves on Artin stacks locally of finite type over $S$[^1]. In forthcoming papers, we will also develop a theory of adic sheaves and perverse sheaves for Artin stacks. In addition to being of basic foundational interest, we hope that the development of these six operations for stacks will have a number of applications. Already the work done in this paper (and the forthcoming ones) provides the necessary tools needed in several papers on the geometric Langland’s program (e.g. [@LN], [@Lau], [@FGV]). We hope that it will also shed further light on the Lefschetz trace formula for stacks proven by Behrend ([@Ber03]), and also to versions of such a formula for stacks not necessarily of finite type. We should also remark that recent work of Toen should provide another approach to defining the six operations for stacks, and in fact should generalize to a theory for $n$–stacks. Let us describe more precisely the contents of this papers. For a morphism $f:{\mathcal}X\rightarrow {\mathcal}Y$ of such $S$–stacks we define functors $$Rf_*:\operatorname{{D}}_c^{+}({\mathcal}X)\rightarrow \operatorname{{D}}_c^{+}({\mathcal}Y), \ \ Rf_!:\operatorname{{D}}_c^-({\mathcal}X)\rightarrow \operatorname{{D}}_c^-({\mathcal}Y),$$ $$Lf^*:\operatorname{{D}}_c({\mathcal}Y)\rightarrow \operatorname{{D}}_c({\mathcal}X), \ \ Rf^!:\operatorname{{D}}_c({\mathcal}Y)\rightarrow \operatorname{{D}}_c({\mathcal}X),$$ $$\operatorname{\mathcal{R}\it{hom}}:\operatorname{{D}}_c^{-}({\mathcal}X)^{\text{op}}\times \operatorname{{D}}_c^+({\mathcal}X)\rightarrow \operatorname{{D}}_c^+({\mathcal}X),$$ and $$(-){\operatorname{\stackrel{\mathbf L}{\otimes}}}(-):\operatorname{{D}}_c^-({\mathcal}X)\times \operatorname{{D}}_c^-({\mathcal}X)\rightarrow \operatorname{{D}}_c^-({\mathcal}X)$$ satisfying all the usual adjointness properties that one has in the theory for schemes[^2]. The main tool is to define $Rf_!,f^!$, even for unbounded constructible complexes, by duality. One of the key points is that, as observed by Laumon, the dualizing complex is a local object of the derived category and hence has to exist for stacks by glueing (see \[1.2\]). Notice that this formalism applies to non-separated schemes, giving a theory of cohomology with compact supports in this case. Previously, Laumon and Moret-Bailly constructed the truncations of dualizing complexes for Bernstein-Lunts stacks (see [@Lau-Mor2000]). Our constructions reduces to theirs in this case. Another approach using a dual version of cohomological descent has been suggested by Gabber but seems to be technically much more complicated. \[cohdimension\] The cohomological dimension hypothesis on schemes of finite type over $S$ is achieved for instance if $S$ is the spectrum of a finite field or of a separably closed field. In dimension $1$, it will be achieved for instance for the spectrum of a complete discrete valuation field with residue field either finite or separably closed, or if $S$ is a smooth curve over ${\mathbf{C}},{\mathbf{F}}_q$ (cf. [@SGA43], exp. X and [@Ser94]). In these situations, $\mathrm{cd}_l(X)$ is bounded by a function of the dimension $\dim(X)$. Notice that, as pointed out by Illusie, recent results of Gabber enables one to dramatically weaken the hypothesis on $S$. Unfortunately no written version of these results seems to be available at this time. Conventions ----------- Recall that for any ring ${\mathcal{O}}$ of a topos, the category of complexes of ${\mathcal{O}}$-modules has enough $K$-injective (or homotopically injective). Recall that a complex $I$ is $K$-injective if for any acyclic complex of ${\mathcal{O}}$-modules $A$, the complex ${\mathcal}Hom(A,I)$ is acyclic (see [@Spa88]). For instance, any injective resolution in the sense of Cartan-Eileberg of a bounded below complex is $K$-injective. This result is due, at least for sheaves on a topological space to [@Spa88] and enables him to extend the formalism of direct images and $\operatorname{\mathcal{R}\it{hom}}$ to unbounded complexes. But this result is true for any Grothendieck category ([@Serp03]). Notice that the category of ${\mathcal{O}}$-modules has enough K-flat objects, enabling one to define $\operatorname{\stackrel{\mathbf L}{\otimes}}$ for unbounded objects ([@Spa88]). All the stacks we will consider will be locally of finite type over $S$. As in [@Lau-Mor2000], lemme 12.1.2, the lisse-étale topos ${\mathcal{X}}_{\textup{lis-\'et}}$ can be defined using the site ${{\textup{Lisse-Et}}}({\mathcal{X}})$ whose objects are $S$-morphisms $u:U\operatorname{\rightarrow}{\mathcal{X}}$ where $U$ is an algebraic space which is *separated and of finite type over $S$*. The topology is generated by the pretopology such that the covering families are finite families $(U_i,u_i)\operatorname{\rightarrow}(U,u)$ such that $\bigsqcup U_i\operatorname{\rightarrow}U$ is surjective and étale (use the comparison theorem [@SGA41], III.4.1 remembering ${\mathcal{X}}$ is locally of finite type over $S$). Notice that products over ${\mathcal{X}}$ are representable in ${{\textup{Lisse-Et}}}({\mathcal{X}})$, simply because the diagonal morphism ${\mathcal{X}}\operatorname{\rightarrow}{\mathcal{X}}\times_S{\mathcal{X}}$ is representable by definition ([@Lau-Mor2000]). If $C$ is a complex of sheaves and $d$ a locally constant valued function $C(d)$ is the Tate twist and $C[d]$ the shifted complex. We denote $C(d)[2d]$ by $C{\langle}d{\rangle}$. Let $\Omega=\Lambda{\langle}\dim(S){\rangle}$ be the dualizing complex of $S$ ([@SGA4.5], “Dualité”). Homological algebra =================== Existence of $K$–injectives --------------------------- Let $({\mathcal}S, {{\mathcal{O}}} )$ denote a ringed site, and let ${\mathcal}C$ denote a full subcategory of the category of ${{\mathcal{O}}} $–modules on ${\mathcal}S$. Let $M$ be a complex of ${{\mathcal{O}}} $–modules on ${\mathcal}S$. By ([@Spa88], 3.7) there exists a morphism of complexes $f:M\rightarrow I$ with the following properties: 1. $I = \varprojlim I_n$ where each $I_n$ is a bounded below complex of flasque ${{\mathcal{O}}} $–modules. 2. The morphism $f$ is induced by a compatible collection of quasi–isomorphisms $f_n:\tau _{\geq -n}M\rightarrow I_n.$ 3. For every $n$ the map $I_n\rightarrow I_{n-1}$ is surjective with kernel $K_n$ a bounded below complex of flasque ${{\mathcal{O}}} $–modules. 4. For any pair of integers $n$ and $i$ the sequence $$0\rightarrow K_n^i\rightarrow I_n^i\rightarrow I_{n-1}^i\rightarrow 0$$ is split. In fact ([@Spa88], 3.7) shows that we can choose $I_n$ and $K_n$ to be complexes of injective ${{\mathcal{O}}} $–modules (in which case (iv) follows from (iii)). However, for technical reasons it is sometimes useful to know that one can work just with flasque sheaves. We make the following finiteness assumption, which is the analog of [@Spa88], 3.12 (1). \[assumption1\] For any object $U\in {\mathcal}S$ there exists a covering $\{U_i\rightarrow U\}_{i\in I}$ and an integer $n_0$ such that for any sheaf of ${{\mathcal{O}}} $–modules $F\in {\mathcal}C$ we have $H^n(U_i, F) = 0$ for all $n\geq n_0$. Let ${\mathcal}S={{\textup{Lisse-Et}}}({\mathcal}X)$ be the lisse-étale site of an algebraic $S$-stack locally of finite type ${\mathcal}X$ and ${{\mathcal{O}}}$ a constant local Artinian ring of characteristic $\Lambda$ invertible on $S$. Then the class ${\mathcal}C$ of all ${{\mathcal{O}}}$-sheaves, cartesian or not, satisfies the assumption. Indeed, if $U\in{\mathcal}S$ is of finite type over $S$ and $F\in{\mathcal}S$, one has $H^n(U,F)=H^n(U_{{\textup{\'et}}},F_U)$[^3] which is zero for $n$ bigger than a constant depending only on $U$ (and not on $F$). Therefore, one can take the trivial covering in this case. We could also take ${{\mathcal{O}}}={\mathcal}O_{{\mathcal}X}$ and ${\mathcal}C$ to be the class of quasi-coherent sheaves. With hypothesis \[assumption1\], one has the following criterion for $f$ being a quasi-isomorphism (cf. [@Spa88], 3.13). \[1.4\] Assume that ${\mathcal}H^j(M)\in {\mathcal}C$ for all $j$. Then the map $f$ is a quasi–isomorphism. In particular, if each $I_n$ is a complex of injective ${{\mathcal{O}}} $–modules then by [[@Spa88], 2.5,]{} $f:M\rightarrow I$ is a $K$–injective resolution of $M$. For a fixed integer $j$, the map ${\mathcal}H^j(M)\rightarrow {\mathcal}H^j(I_n)$ is an isomorphism for $n$ sufficiently big. Since this isomorphism factors as $${\mathcal}H^j(M)\rightarrow {\mathcal}H^j(I)\rightarrow {\mathcal}H^j(I_n)$$ it follows that the map ${\mathcal}H^j(M)\rightarrow {\mathcal}H^j(I)$ is injective. To see that ${\mathcal}H^j(M)\rightarrow {\mathcal}H^j(I)$ is surjective, let $U\in {\mathcal}S$ be an object and $\gamma \in \Gamma (U, I^j)$ an element with $d\gamma = 0$ defining a class in ${\mathcal}H^j(I)(U)$. Since $I = \varprojlim I_n$ the class $\gamma $ is given by a compatible collection of sections $\gamma _n\in \Gamma (U, I_n^j)$ with $d\gamma _n = 0$. Let $({\mathcal}U = \{U_i\operatorname{\rightarrow}U\},n_0)$ be the data provided by \[assumption1\]. Let $N$ be an integer greater than $n_0-j$. For $m>N$ and $U_i\in {\mathcal}U$ the sequence $$\Gamma (U_i, K_m^{j-1})\rightarrow \Gamma (U_i, K_m^{j})\rightarrow \Gamma (U_i, K_m^{j+1})\rightarrow \Gamma (U_i, K_m^{j+2})$$ is exact. Indeed $K_m$ is a bounded below complex with ${\mathcal}H^j(K_m)\in {\mathcal}C$ for every $j$ and ${\mathcal}H^j(K_m) = 0$ for $j\geq -m+2$. It follows that $H^j(U_i, K_m) = 0$ for $j\geq n_0-m+2$. Since the maps $\Gamma (U_i, I_m^r)\rightarrow \Gamma (U_i, I_{m-1}^r)$ are also surjective for all $m$ and $r$, it follows from ([@Spa88], 0.11) applied to the system $$\Gamma (U_i, I_m^{j-1})\rightarrow \Gamma (U_i, I_m^i)\rightarrow \Gamma (U_i, I_m^{j+1})\rightarrow \Gamma (U_i, I_m^{j+2})$$ that the map $$H^j(\Gamma (U_i, I))\rightarrow H^j(\Gamma (U_i, I_m))$$ is an isomorphism. Then since the map ${\mathcal}H^j(M)\rightarrow {\mathcal}H^j(I_m)$ is an isomorphism it follows that for every $i$ the restriction of $\gamma $ to $U_i$ is in the image of ${\mathcal}H^j(M)(U_i)$. $\square$ Next consider a fibred topos ${\mathcal}T\operatorname{\rightarrow}D$ with corresponding total topos ${\mathcal}T_\bullet$ ([@SGA42], VI.7). We call ${\mathcal}T_\bullet$ a *$D$-simplicial topos*. Concretely, this means that for each $i\in D$ the fiber ${\mathcal}T_i$ is a topos and that any $\delta\in\operatorname{\mathrm{Hom}}_D(i,j)$ comes together with a morphism of topos $\delta:{\mathcal}T_i\operatorname{\rightarrow}{\mathcal}T_j$ such that $\delta^{-1}$ is the inverse image functor of the fibred structure. The objects of the total topos are simply collections $(F_i\in E_i)_{i\in D}$ together with functorial transition morphisms $\delta^{-1}F_j\operatorname{\rightarrow}F_i$ for any $\delta\in \operatorname{\mathrm{Hom}}_D(i,j)$. We assume furthermore that ${\mathcal}T_\bullet$ is ringed by a ${{\mathcal{O}}} _\bullet$ and that for any $\delta\in\operatorname{\mathrm{Hom}}_D(i,j)$, the morphism $\delta:({\mathcal}T_i,{{\mathcal{O}}}_i)\operatorname{\rightarrow}({\mathcal}T_j,{{\mathcal{O}}}_j)$ is flat. Let $\Delta ^+$ be the category whose objects are the ordered sets $[n] = \{0, \dots , n\}$ ($n\in \mathbb{N}$) and whose morphisms are injective order-preserving maps. Let $D$ be the opposite category of $\Delta ^+$. In this case ${\mathcal}T_.$ is called a *strict simplicial topos*. For instance, if $U\operatorname{\rightarrow}{\mathcal}X$ is a presentation, the simplicial algebraic space $U_\bullet=\mathrm{cosq}_0(U/{\mathcal}X)$ defines a strict simplicial topos $U_{\bullet{\textup{lis-\'et}}}$ whose fiber over $[n]$ is $U_{n{\textup{lis-\'et}}}$. For a morphism $\delta :[n]\rightarrow [m]$ in $\Delta ^+$ the morphism $\delta:{\mathcal}T_{m}\operatorname{\rightarrow}{\mathcal}T_n$ is induced by the (smooth) projection $U_m\operatorname{\rightarrow}U_n$ defined by $\delta\in \operatorname{\mathrm{Hom}}_{\Delta^{+{\textup{opp}}}}([m],[n])$. Let $\mathbf N$ be the natural numbers viewed as a category in which $\operatorname{\mathrm{Hom}}(n, m)$ is empty unless $m\geq n$ in which case it consists of a unique element. For a topos $T$ we can then define an $\mathbf N$-simplicial topos $T^{\mathbf N}$. The fiber over $n$ of $T^{\mathbf{N}}$ is $T$ and the transition morphisms by the identity of $T$. The topos $T^{\mathbf N}$ is the category of projective systems in $T$. If ${{\mathcal{O}}} _\bullet $ is a constant projective system of rings then the flatness assumption is also satisfied, or more generally if $\delta ^{-1}{{\mathcal{O}}} _n\rightarrow {{\mathcal{O}}} _m$ is an isomorphism for any morphism $\delta :m\rightarrow n$ in $\mathbb{N}$ then the flatness assumption holds. Let ${\mathcal}C_\bullet $ be a full subcategory of the category of ${{\mathcal{O}}} _\bullet $–modules on a ringed $D$-simplicial topos $({\mathcal}T_\bullet,{{\mathcal{O}}}_\bullet)$. For $i\in D$, let $e_i:{\mathcal}T_n\operatorname{\rightarrow}{\mathcal}T_\bullet$ the morphism of topos defined by $e_i^{-1}F_\bullet=F_n$ (cf. [@SGA42], Vbis, 1.2.11). Recall that the family $e_i^{-1},i\in D$ is conservative. Let ${{\mathcal}C}_i$ denote the essential image of ${{\mathcal}C}_\bullet$ under $e_{i}^{-1}$ (which coincides with $e_i^*$ on $\text{Mod}({\mathcal}T_\bullet,{{\mathcal{O}}}_\bullet)$ because $e_i^{-1}{{\mathcal{O}}}_\bullet={{\mathcal{O}}}_i$). \[assume2\] For every $i\in D$ the ringed topos $({\mathcal}T_i, {{\mathcal{O}}} _i)$ is isomorphic to the topos of a ringed site satisfying \[assumption1\] with respect to ${\mathcal}C_i$. Let ${\mathcal}T_\bullet$ be the topos $({\mathcal}X_{{\textup{lis-\'et}}})^{\mathbf N}$ of a $S$-stack locally of finite type. Then, the full subcategory ${\mathcal}C_{\bullet}$ of $\text{Mod}({\mathcal}T_\bullet,{{\mathcal{O}}}_\bullet)$ whose objects are families $F_i$ of *cartesian* modules satisfies the hypothesis. Let $M$ be a complex of ${{\mathcal{O}}} _\bullet $–modules on ${\mathcal}T_\bullet $. Again by ([@Spa88], 3.7) there exists a morphism of complexes $f:M\rightarrow I$ with the following properties: 1. $I = \varprojlim I_n$ where each $I_n$ is a bounded below complex of injective modules. 2. The morphism $f$ is induced by a compatible collection of quasi–isomorphisms $f_n:\tau _{\geq -n}M\rightarrow I_n.$ 3. For every $n$ the map $I_n\rightarrow I_{n-1}$ is surjective with kernel $K_n$ a bounded below complex of injective ${{\mathcal{O}}} $–modules. 4. For any pair of integers $n$ and $i$ the sequence $$\label{extseq} 0\rightarrow K_n^i\rightarrow I_n^i\rightarrow I_{n-1}^i\rightarrow 0$$ is split. \[1.7\] Assume that ${\mathcal}H^j(M)\in {\mathcal}C_\bullet $ for all $j$. Then the morphism $f$ is a quasi–isomorphism and $f:M\rightarrow I$ is a $K$–injective resolution of $M$. By [@Spa88], 2.5, it suffices to show that $f$ is a quasi–isomorphism. For this in turn it suffices to show that for every $i\in D $ the restriction $e_i^*f:e_i^*M\rightarrow e_i^*I$ is a quasi–isomorphism of complexes of ${{\mathcal{O}}}_i$-modules since the family $e_i^* = e_i^{-1}$ is conservative. But $e_i^*:\text{Mod}({\mathcal}T_\bullet,{{\mathcal{O}}}_\bullet)\operatorname{\rightarrow}\text{Mod}({\mathcal}T_i,{{\mathcal{O}}}_i)$ has a left adjoint $e_{i!}$ defined by $$[e_{i!}(F)]_j=\oplus_{\delta\in\operatorname{\mathrm{Hom}}_D(j,i)}\delta^*F$$ with the obvious transition morphisms. It is exact by the flatness of the morphisms $\delta$. It follows that $e_i^*$ takes injectives to injectives and commutes with direct limits. We can therefore apply \[1.4\] to $e_i^*M\operatorname{\rightarrow}e_i^*I$ to deduce that this map is a quasi–isomorphism. $\square$ In what follows we call a $K$–injective resolution $f:M\rightarrow I$ obtained from data (i)-(iv) as above a *Spaltenstein resolution*. The main technical lemma is the following. \[key-truncation\] Let $\epsilon :({\mathcal}T_\bullet , {{\mathcal{O}}} _\bullet )\rightarrow (S, \Psi )$ be a morphism of ringed topos, and let $C$ be a complex of ${{\mathcal{O}}} _\bullet $–modules. Assume that 1. ${\mathcal}H^n(C)\in {\mathcal}C_\bullet $ for all $n$. 2. There exists $i_0$ such that $R^i\epsilon_*{\mathcal}H^n(C)=0$ for any $n$ and any $i> i_0$. Then, if $j\geq -n+i_0$, we have $R^j\epsilon_*C=R^j\epsilon_*\tau_{\geq -n}C$. By \[1.7\] and assumption (1), there exists a Spaltenstein resolution $f:C\rightarrow I$ of $C$. Let $J_n := \epsilon _*I_n$ and $D_n :=\epsilon _*K_n$. Since the sequences \[extseq\] are split, the sequences $$0\rightarrow D_n\rightarrow J_n\rightarrow J_{n-1}\rightarrow 0$$ are exact. The exact sequence \[extseq\] and property (S ii) defines a distinguished triangle $$K_n\rightarrow\tau_{\geq -n}C\rightarrow\tau_{\geq -n+1}C$$ showing that $K_n$ is quasi–isomorphic to ${\mathcal}H^{-n}(C)[n]$. Because $K_n$ is a bounded below complex of injectives, one gets $$R\epsilon_*{\mathcal}H^{-n}(C)[n]=\epsilon_*K_n$$ and accordingly $$R^{j+n}\epsilon_*{\mathcal}H^{-n}(C)={\mathcal}H^j(\epsilon_*K_n)={\mathcal}H^j(D_n).$$ By assumption (2), we have therefore $${\mathcal}H^j(D_n) = 0\textup{ for }j> -n+i_0.$$ By ([@Spa88], 0.11) this implies that $${\mathcal}H^j(\varprojlim J_n) \rightarrow {\mathcal}H^j(J_{n})$$ is an isomorphism for $j\geq -n+i_0$. But, by adjunction, $\epsilon_*$ commutes with projective limit. In particular, one has $$\varprojlim J_n=\epsilon_*I,$$ and by (S i) and (S ii) $$R\epsilon_*C=\epsilon_*I\textup{ and }R\epsilon_*\tau_{\geq -n}C=\epsilon_*J_n.$$ Thus for any $n$ such that $j\geq -n+i_0$ one has $$R^j\epsilon _*C = {\mathcal}H^j(\epsilon_*I) = {\mathcal}H^j(J_{n}) = R^j\epsilon _*\tau_{\geq -n}C.$$ $\square$ The descent theorem ------------------- Let $({\mathcal}T_\bullet ,{{\mathcal{O}}}_{\bullet})$ be a simplicial or strictly simplicial [^4] ringed topos ($D=\Delta^{\textup{opp}}$ or $D=\Delta^{+{\textup{opp}}}$), let $(S, \Psi)$ be another ringed topos, and let $\epsilon :({\mathcal}T_\bullet ,{{\mathcal{O}}}_{\bullet})\rightarrow (S, \Psi )$ be an augmentation. Assume that $\epsilon $ is a flat morphism (i.e. for every $i\in D $, the morphism of ringed topos $({\mathcal}T_i, {{\mathcal{O}}}_i)\rightarrow (S, \Psi )$ is a flat morphism). Let ${\mathcal}C$ be a full subcategory of the category of $\Psi $–modules, and assume that ${\mathcal}C$ is closed under kernels, cokernels and extensions (one says that ${\mathcal}C$ is a *Serre subcategory*). Let $\operatorname{{D}}(S)$ denote the derived category of $\Psi $–modules, and let $\operatorname{{D}}_{{\mathcal}C}(S)\subset \operatorname{{D}}(S)$ be the full subcategory consisting of complexes whose cohomology sheaves are in ${\mathcal}C$. Let ${\mathcal}C_\bullet $ denote the essential image of ${\mathcal}C$ under the functor $\epsilon ^*:\text{\rm Mod}(\Psi )\rightarrow \text{\rm Mod}({{\mathcal{O}}} _\bullet )$. We assume the following condition holds: \[2.13\] Assumption \[assume2\] holds (with respect to ${\mathcal}C_\bullet $), and $\epsilon^*:{\mathcal}C\operatorname{\rightarrow}{\mathcal}C_\bullet$ is an equivalence of categories with quasi-inverse $R\epsilon_*$. The full subcategory ${\mathcal}C_{\bullet }\subset \text{\rm Mod}({{\mathcal{O}}} _\bullet )$ is closed under extensions, kernels and cokernels. Consider an extension of sheaves of ${{\mathcal{O}}} _\bullet $–modules $$\begin{CD} 0@>>> \epsilon ^*F_1@>>> E@>>> \epsilon ^*F_2@>>> 0, \end{CD}$$ where $F_1, F_2\in {\mathcal}C$. Since $R^1\epsilon _*\epsilon ^*F_1 = 0$ and the maps $F_i\rightarrow R^0\epsilon _*\epsilon ^*F_i$ are isomorphisms, we obtain by applying $\epsilon ^* \epsilon _*$ a commutative diagram with exact rows $$\begin{CD} 0@>>> \epsilon ^*F_1@>>> \epsilon ^*\epsilon _*E@>>> \epsilon ^*F_2@>>> 0\\ @. @V\text{id}VV @V\alpha VV @VV\text{id}V @. \\ 0@>>> \epsilon ^*F_1@>>> E@>>> \epsilon ^*F_2@>>> 0. \end{CD}$$ It follows that $\alpha $ is an isomorphism. Furthermore, since ${\mathcal}C$ is closed under extensions we have $\epsilon _*E\in {\mathcal}C$. Let $f\in\operatorname{\mathrm{Hom}}(\epsilon^*F_1,\epsilon^*F_2)$. There exists a unique $\varphi\in\operatorname{\mathrm{Hom}}(F_1,F_2)$ such that $f=\epsilon^*\varphi$. Because $\epsilon^*$ is exact, it maps the kernel and cokernel of $\varphi$, which are objects of ${\mathcal}C$, to the kernel and cokernel of $f$ respectively. Therefore, the latter are objects of ${\mathcal}C_\bullet$. $\square$ Let $\operatorname{{D}}({\mathcal}T_\bullet )$ denote the derived category of ${{\mathcal{O}}} _\bullet $–modules, and let $\operatorname{{D}}_{{\mathcal}C_\bullet }({\mathcal}T_\bullet )\subset \operatorname{{D}}({\mathcal}T_\bullet )$ denote the full subcategory of complexes whose cohomology sheaves are in ${\mathcal}C_\bullet $. Since $\epsilon $ is a flat morphism, we obtain a morphism of triangulated categories (the fact that these categories are triangulated comes precisely from the fact that both ${\mathcal}C$ and ${\mathcal}C_\bullet$ are Serre categories [@Grivel]). $$\label{pullback} \epsilon ^*:\operatorname{{D}}_{{\mathcal}C}(S)\rightarrow \operatorname{{D}}_{{\mathcal}C_\bullet }({\mathcal}T_\bullet ).$$ \[mainthm\] The functor $\epsilon^*$ of \[pullback\] is an equivalence of triangulated categories with quasi–inverse given by $R\epsilon _*$. Note first that if $M_\bullet \in \operatorname{{D}}_{{\mathcal}C_\bullet }({\mathcal}T_\bullet )$, then by lemma \[key-truncation\], for any integer $j$ there exists $n_0$ such that $R^j\epsilon _*M_\bullet = R^j\epsilon _*\tau _{\geq n_0}M_\bullet $. In particular, we get by induction $R^j\epsilon _*M_\bullet \in {\mathcal}C$. Thus $R\epsilon _*$ defines a functor $$R\epsilon _*:\operatorname{{D}}_{{\mathcal}C_\bullet }({\mathcal}T_\bullet )\rightarrow \operatorname{{D}}_{{\mathcal}C}(S).$$ To prove \[mainthm\] it suffices to show that for $M_\bullet \in \operatorname{{D}}_{{\mathcal}C_\bullet }({\mathcal}T_\bullet )$ and $F\in \operatorname{{D}}_{{\mathcal}C}(S)$ the adjunction maps $$\epsilon ^*R\epsilon _*M_\bullet \rightarrow M_\bullet , \ \ \ F\rightarrow R\epsilon _*\epsilon ^*F.$$ are isomorphisms. For this note that for any integers $j$ and $n$ there are commutative diagrams $$\begin{CD} \epsilon ^*R^j\epsilon _*M_\bullet @>>> {\mathcal}H^j(M_\bullet )\\ @VVV @VVV \\ \epsilon ^*R^j\epsilon _*\tau _{\geq n}M_\bullet @>>> {\mathcal}H^j(\tau _{\geq n}M_\bullet ), \end{CD}$$ and $$\begin{CD} {\mathcal}H^j(F)@>>> R^j\epsilon _*\epsilon ^*F\\ @VVV @VVV \\ {\mathcal}H^j(\tau _{\geq n}F)@>>> R^j\epsilon _*\epsilon ^*\tau _{\geq n}F. \end{CD}$$ By the observation at the begining of the proof, there exists an integer $n$ so that the vertical arrows in the above diagrams are isomorphisms. This reduces the proof \[mainthm\] to the case of a bounded below complex. In this case one reduces by devissage to the case when $M_\bullet \in {\mathcal}C_\bullet $ and $F\in {\mathcal}C$ in which case the result holds by assumption. $\square$ The Theorem applies in particular to the following examples. Let $S$ be an algebraic space and $X_\bullet \rightarrow S$ a flat hypercover by algebraic spaces. We then obtain an augmented simplicial topos $\epsilon :(X_{\bullet , {\textup{\'et}}}, {\mathcal}O_{X_{\bullet, {\textup{\'et}}}})\rightarrow (S_{\textup{\'et}}, {\mathcal}O_{{\textup{\'et}}}).$ Note that this augmentation is flat. Let ${\mathcal}C$ denote the category of quasi–coherent sheaves on $S_{\textup{\'et}}$. Then the category ${\mathcal}C_\bullet $ is the category of cartesian sheaves of ${\mathcal}O_{X_{\bullet, {\textup{\'et}}}}$–modules whose restriction to each $X_n$ is quasi–coherent. Let $\operatorname{{D}}_{\text{qcoh}}(X_\bullet )$ denote the full subcategory of the derived category of ${\mathcal}O_{X_\bullet, {\textup{\'et}}}$–modules whose cohomology sheaves are quasi–coherent, and let $\operatorname{{D}}_{\text{qcoh}}(S)$ denote the full subcategory of the derived category of ${\mathcal}O_{S_{\textup{\'et}}}$–modules whose cohomology sheaves are quasi–coherent. Theorem \[mainthm\] then shows that the pullback functor $$\epsilon ^*:\operatorname{{D}}_{\text{qcoh}}(S)\rightarrow \operatorname{{D}}_{\text{qcoh}}(X_\bullet )$$ is an equivalence of triangulated categories with quasi–inverse $R\epsilon _*$. Let ${\mathcal}X$ be an algebraic stack and let $U_\bullet \rightarrow {\mathcal}X$ be a smooth hypercover by algebraic spaces. Let $\operatorname{{D}}({\mathcal}X)$ denote the derived category of sheaves of ${\mathcal}O_{{\mathcal}X_{{{\textup{lis-\'et}}}}}$–modules in the topos ${\mathcal}X_{{{\textup{lis-\'et}}}}$, and let $\operatorname{{D}}_{\text{qcoh}}({\mathcal}X)\subset \operatorname{{D}}({\mathcal}X)$ be the full subcategory of complexes with quasi–coherent cohomology sheaves. Let $U_{\bullet }^+$ denote the strictly simplicial algebraic space obtained from $U_\bullet $ by forgetting the degeneracies. Since the Lisse-Étale topos is functorial with respect to smooth morphisms, we therefore obtain a strictly simplicial topos $U_{\bullet {{\textup{lis-\'et}}}}$ and a flat morphism of ringed topos $$\epsilon :(U_{\bullet {{\textup{lis-\'et}}}}, {\mathcal}O_{U_{\bullet {{\textup{lis-\'et}}}}})\rightarrow ({\mathcal}X_{{{\textup{lis-\'et}}}}, {\mathcal}O_{{\mathcal}X_{{{\textup{lis-\'et}}}}}).$$ Then \[2.13\] holds with ${\mathcal}C$ equal to the category of quasi–coherent sheaves on ${\mathcal}X$. The category ${\mathcal}C_\bullet $ in this case is the category of cartesian ${\mathcal}O_{U_{\bullet {{\textup{lis-\'et}}}}}$–modules $M_\bullet $ such that the restriction $M_n$ is a quasi–coherent sheaf on $U_n$ for all $n$. By \[mainthm\] we then obtain an equivalence of triangulated categories $$\operatorname{{D}}_{\text{qcoh}}({\mathcal}X)\rightarrow \operatorname{{D}}_{\text{qcoh}}(U_{\bullet , {{\textup{lis-\'et}}}}),$$ where the right side denotes the full subcategory of the derived category of ${\mathcal}O_{U_{\bullet {{\textup{lis-\'et}}}}}$–modules with cohomology sheaves in ${\mathcal}C_{\bullet }$. On the other hand, there is also a natural morphism of ringed topos $$\pi :(U_{\bullet {{\textup{lis-\'et}}}}, {\mathcal}O_{U_{\bullet {{\textup{lis-\'et}}}}})\rightarrow (U_{\bullet {\textup{\'et}}}, {\mathcal}O_{U_{\bullet {\textup{\'et}}}})$$ with $\pi _*$ and $\pi ^*$ both exact functors. Let $\operatorname{{D}}_{\text{qcoh}}(U_{\bullet {\textup{\'et}}})$ denote the full subcategory of the derived category of ${\mathcal}O_{U_{\bullet {\textup{\'et}}}}$–modules consisting of complexes whose cohomology sheaves are quasi–coherent (i.e. cartesian and restrict to a quasi–coherent sheaf on each $U_{n {\textup{\'et}}}$). Then $\pi $ induces an equivalence of triangulated categories $\operatorname{{D}}_{\text{qcoh}}(U_{\bullet {\textup{\'et}}})\simeq \operatorname{{D}}_{\text{qcoh}}(U_{\bullet {{\textup{lis-\'et}}}})$. Putting it all together we obtain an equivalence of triangulated categories $\operatorname{{D}}_{\text{qcoh}}({\mathcal}X_{{{\textup{lis-\'et}}}})\simeq \operatorname{{D}}_{\text{qcoh}}(U_{\bullet {\textup{\'et}}})$. \[etexample\] Let ${\mathcal}X$ be an algebraic stack locally of finite type over $S$ and ${{\mathcal{O}}} $ be a constant local Artinian ring $\Lambda$ of characteristic invertible on $S$. Let $U_\bullet \rightarrow {\mathcal}X$ be a smooth hypercover by algebraic spaces, and ${\mathcal}T_\bullet $ the localized topos ${\mathcal}X_{\text{{{\textup{lis-\'et}}}}}|_{U_\bullet }$. Take ${\mathcal}C$ to be the category of constructible sheaves of ${{\mathcal{O}}} $–modules. Then \[mainthm\] gives an equivalence $\operatorname{{D}}_c({\mathcal}X_{\text{{{\textup{lis-\'et}}}}})\simeq \operatorname{{D}}_c({\mathcal}T_\bullet , \Lambda )$. On the other hand, there is a natural morphism of topos $\lambda :{\mathcal}T_\bullet \rightarrow U_{\bullet , et}$ and one sees immediately that $\lambda _*$ and $\lambda ^*$ induce an equivalence of derived categories $\operatorname{{D}}_c({\mathcal}T_\bullet , \Lambda )\simeq \operatorname{{D}}_c(U_{\bullet , et}, \Lambda )$. It follows that $\operatorname{{D}}_c({\mathcal}X_{{{\textup{lis-\'et}}}})\simeq \operatorname{{D}}_c(U_{\bullet , et})$. The BBD gluing lemma -------------------- The purpose of this section is to explain how to modify the proof of the gluing lemma [@BBD82 3.2.4] for unbounded complexes. Let $\widetilde {\Delta } $ denote the strictly simplicial category of finite ordered sets with injective order preserving maps, and let $\Delta ^+\subset \widetilde {\Delta } $ denote the full subcategory of nonempty finite ordered sets. For a morphism $\alpha $ in $\widetilde {\Delta } $ we write $s(\alpha )$ (resp. $b(\alpha )$) for its source (resp. target). Let $T$ be a topos and $U_\cdot \rightarrow e$ a strictly simplicial hypercovering of the initial object $e\in T$. For $[n]\in \widetilde {\Delta } $ write $U_n$ for the localized topos $T|_{U_n}$ where by definition we set $U_\emptyset = T$. Then we obtain a strictly simplicial topos $U_\cdot $ with an augmentation $\pi :U_\cdot \rightarrow T$. Let $\Lambda $ be a sheaf of rings in $T$ and write also $\Lambda $ for the induced sheaf of rings in $U_\cdot $ so that $\pi $ is a morphism of ringed topos. Let $\mathcal C_\cdot $ denote a full substack of the fibered and cofibered category over $\widetilde {\Delta } $ $$[n]\mapsto (\text{category of sheaves of $\Lambda $--modules in $U_n$})$$ such that each $\mathcal C_n$ is a Serre subcategory of the category of $\Lambda $–modules in $U_n$. For any $[n]$ we can then form the derived category $D_{\mathcal C}(U_n, \Lambda )$ of complexes of $\Lambda $–modules whose cohomology sheaves are in $\mathcal C_n$. The categories $D_{\mathcal C}(U_n, \Lambda )$ form a fibered and cofibered category over $\widetilde {\Delta } $. We make the following assumptions on $\mathcal C$: \[assumption1b\] (i) For any $[n]$ the topos $U_n$ is equivalent to the topos associated to a site $\mathcal S_n$ such that for any object $V\in \mathcal S_n$ there exists an integer $n_0$ and a covering $\{V_j\rightarrow V\}$ in $\mathcal S_n$ such that for any $F\in \mathcal C_n$ we have $H^n(V_j, F) = 0$ for all $n\geq n_0$. \(ii) The natural functor $$\mathcal C_{\emptyset }\rightarrow (\text{cartesian sections of $\mathcal C|_{\Delta ^+}$ over $\Delta ^+$})$$ is an equivalence of categories. \(iii) The category $D(T, \Lambda )$ is compactly generated. The case we have in mind is when $T$ is the lisse-étale topos of an algebraic stack $\mathcal X$ locally of finite type over an affine regular, noetherian scheme of dimension $\leq 1$, $U_\cdot $ is given by a hypercovering of $\mathcal X$ by schemes, $\Lambda $ is a Gorenstein local ring of dimension $0$ and characteristic $l$ invertible on $\mathcal X$, and $\mathcal C$ is the category of constructible $\Lambda $–modules. In this case the category $D_{c}(\mathcal X_{\text{lis-et}}, \Lambda )$ is compactly generated. Indeed a set of generators is given by sheaves $j_!\Lambda [i]$ for $i\in \mathbb{Z}$ and $j:U\rightarrow \mathcal X$ an object of the lisse-étale site of $\mathcal X$. There is also a natural functor $$\label{Apullback} D_{\mathcal C}(T, \Lambda )\rightarrow (\text{cartesian sections of $[n]\mapsto D_{\mathcal C}(U_n, \Lambda )$ over $\Delta ^+$}).$$ \[1.2\] Let $[n]\mapsto K_n\in D_{\mathcal C}(U_n, \Lambda )$ be a cartesian section of $[n]\mapsto D_{\mathcal C}(U_n, \Lambda )$ over $\Delta ^+$ such that $\mathcal Ext^i(K_0, K_0) = 0$ for all $i<0$. Then $(K_n)$ is induced by a unique object $K\in D_{\mathcal C}(T, \Lambda )$ via the functor \[Apullback\]. The uniqueness is the easy part: \[D1\] Let $K, L\in D(T, \Lambda )$ and assume that $\mathcal Ext^i(K, L) = 0$ for $i<0$. Then $U\mapsto \text{\rm Hom}_{D(U, \Lambda )}(K|_U, L|_U)$ is a sheaf. Let $\mathcal H$ denote the complex $\mathcal Rhom(K, L)$. By assumption the natural map $\mathcal H\rightarrow \tau _{\geq 0}\mathcal H$ is an isomorphism. It follows that $\text{Hom}_{D(U, \Lambda )}(K|_U, L|_U)$ is equal to the value of $\mathcal H^0(\mathcal H)$ on $U$ which implies the lemma. The existence part is more delicate. Let $\mathcal A$ denote the fibered and cofibered category over $\widetilde {\Delta } $ whose fiber over $[n]\in \widetilde {\Delta } $ is the category of $\Lambda $–modules in $U_n$. For a morphism $\alpha :[n]\rightarrow [m]$, $F\in \mathcal A(n)$ and $G\in \mathcal A(m)$ we have $$\text{Hom}_\alpha (F, G) = \text{Hom}_{\mathcal A(m)}(\alpha ^*F, G) = \text{Hom}_{\mathcal A(n)}(F, \alpha _*G).$$ We write $\mathcal A^+$ for the restriction of $\mathcal A$ to $\Delta ^+$. Define a new category $\text{tot}(\mathcal A^+)$ as follows: 1. The objects of $\text{tot}(\mathcal A^+)$ are collections of objects $(A^n)_{n\geq 0}$ with $A^n\in \mathcal A(n)$. 2. For two objects $(A^n)$ and $(B^n)$ we define $$\text{Hom}_{\text{tot}(\mathcal A^+)}((A^n), (B^n)):= \prod _{\alpha }\text{Hom}_\alpha (A^{s(\alpha )}, B^{b(\alpha )}),$$ where the product is taken over all morphisms in $\Delta ^+$. 3. If $f=(f_\alpha )\in \text{Hom}((A^n), (B^n))$ and $g=(g_\alpha )\in \text{Hom}((B^n), (C^n))$ are two morphisms then the composite is defined to be the collection of morphisms whose $\alpha $ component is defined to be $$(g\circ f)_\alpha := \sum _{\alpha = \beta \gamma }g_\beta f_\gamma$$ where the sum is taken over all factorizations of $\alpha $. The category $\text{tot}(\mathcal A^+)$ is an additive category. Let $(K, d)$ be a complex in $\text{tot}(\mathcal A^+)$ so for every degree $n$ we are given a family of objects $(K^n)^m\in \mathcal A(m)$. Set $$K^{n, m}:= (K^{n+m})^n.$$ For $\alpha :[n]\rightarrow [m]$ in $\Delta ^+$ let $d(\alpha )$ denote the $\alpha $–component of $d$ so $$d(\alpha )\in \text{Hom}_\alpha ((K^p)^n, (K^{p+1})^m) = \text{Hom}_\alpha (K^{n, p-n}, K^{m, m-p-1})$$ or equivalently $d(\alpha )$ is a map $K^{n, p}\rightarrow K^{m, p+n-m+1}$. In particular, $d(\text{id}_{[n]})$ defines a map $K^{n, m}\rightarrow K^{n, m+1}$ and as explained in [@BBD82 3.2.8] this map makes $K^{n, *}$ a complex. Furthermore for any $\alpha $ the map $d(\alpha )$ defines an $\alpha $–map of complexes $K^{n, *}\rightarrow K^{m, *}$ of degree $n-m+1$. The collection of complexes $K^{n, *}$ can also be defined as follows. For an integer $p$ let $L^pK$ denote the subcomplex with $(L^pK)^{n, m}$ equal to $0$ if $n<p$ and $K^{n, m}$ otherwise. Note that for any $\alpha :[n]\rightarrow [m]$ which is not the identity map $[n]\rightarrow [n]$ the image of $d(\alpha )$ is contained in $L^{p+1}K$. Taking the associated graded of $L$ we see that $$\text{gr}_L^nK[n] = (K^{n, *}, d^{\prime \prime })$$ where $d^{\prime \prime }$ denote the differential $(-1)^nd(\text{id}_{[n]})$. Note that the functor $(K, d)\mapsto K^{n, *}$ commutes with the formation of cones and with shifting of degrees. As explained in [@BBD82 3.2.8] a complex in $\text{tot}(\mathcal A^+)$ is completely characterized by the data of a complex $K^{n, *}\in C(\mathcal A^+)$ for every $[n]\in \Delta ^+$ and for every morphism $\alpha :[n]\rightarrow [m] $ an $\alpha $–morphism $d(\alpha ):K^{n, *}\rightarrow K^{m, *}$ of degree $n-m+1$, such that $d(\text{id}_{[n]})$ is equal to $(-1)^n$ times the differential of $K^{n, *}$ and such that for every $\alpha $ we have $$\sum _{\alpha = \beta \gamma }d(\beta )d(\gamma ) = 0.$$ Via this dictionary, a morphism $f:K\rightarrow L$ in $C(\text{tot}(\mathcal A^+))$ is given by an $\alpha $–map $f(\alpha ):K^{n, *}\rightarrow K^{m, *}$ of degree $n-m$ for every morphism $\alpha :[n]\rightarrow [m]$ in $\Delta ^+$ such that for any morphism $\alpha $ we have $$\sum _{\alpha = \beta \gamma }d(\beta )f(\gamma ) = \sum _{\alpha = \beta \gamma }f(\beta )d(\gamma ).$$ Let $K(\text{tot}(\mathcal A^+))$ denote the category whose objects are complexes in $\text{tot}(\mathcal A^+)$ and whose morphisms are homotopy classes of morphisms of complexes. The category $K(\text{tot}(\mathcal A^+))$ is a triangulated category. Let $L\subset K(\text{tot}(\mathcal A^+))$ denote the full subcategory of objects $K$ for which each $K^{n, *}$ is acyclic for all $n$. The category $L$ is a localizing subcategory of $K(\text{tot}(\mathcal A^+))$ in the sense of [@Bok-Nee93 1.3] and hence the localized category $D(\text{tot}(\mathcal A^+))$ exists. The category $D(\text{tot}(\mathcal A^+))$ is obtained from $K(\text{tot}(\mathcal A^+))$ by inverting quasi–isomorphisms. Recall that an object $K\in K(\text{tot}(\mathcal A^+))$ is called *$L$–local* if for any object $X\in L$ we have $\text{Hom}_{K(\text{tot}(\mathcal A^+))}(X, K) = 0$. Note that the functor $K\mapsto K^{n, *}$ descends to a functor $$D(\text{tot}(\mathcal A^+))\rightarrow D(U_n, \Lambda ).$$ We define $D^+(\text{tot}(\mathcal A^+))\subset D(\text{tot}(\mathcal A^+))$ to be the full subcategory of objects $K$ for which there exists an integer $N$ such that $\mathcal H^j(K^{n, *}) = 0$ for all $n$ and all $j\leq N$. Recall [@Bok-Nee93 4.3] that a *localization* for an object $K\in K(\text{tot}(\mathcal A^+))$ is a morphism $K\rightarrow I$ with $I$ an $L$–local object such that for any $L$–local object $Z$ the natural map $$\label{des1} \text{Hom}_{K(\text{tot}(\mathcal A^+))}(I, Z)\rightarrow \text{Hom}_{K(\text{tot}(\mathcal A^+))}(K, Z)$$ is an isomorphism. A morphism $K\rightarrow I$ is a localization if and only if $I$ is $L$–local and for every $n$ the map $K^{n, *}\rightarrow I^{n, *}$ is a quasi–isomorphism. By [@Bok-Nee93 2.9] the morphism \[des1\] can be identified with the natural map $$\label{des2} \text{Hom}_{D(\text{tot}(\mathcal A^+))}(I, Z)\rightarrow \text{Hom}_{D(\text{tot}(\mathcal A^+))}(K, Z).$$ If $K\rightarrow I$ is a localization it follows that this map is a bijection for every $L$–local $Z$. By Yoneda’s lemma applied to the full subcategory of $D(\text{tot}(\mathcal A^+))$ of objects which can be represented by $L$–local objects, it follows that this holds if and only if $K\rightarrow I$ induces an isomorphism in $D(\text{tot}(\mathcal A^+))$ which is the assertion of the lemma. \[L-local\] Let $K\in C(\text{\rm tot}(\mathcal A^+))$ be an object with each $K^{n, *}$ homotopically injective. Then $K$ is $L$–local. Let $X\in L$ be an object. We have to show that any morphism $f:X\rightarrow K$ in $C(\text{tot}(\mathcal A^+))$ is homotopic to zero. Such a homotopy $h$ is given by a collection of maps $h(\alpha )$ such that $$f(\alpha ) = - \sum _{\alpha = \beta \gamma }d(\beta )h(\gamma )+h(\beta )d(\gamma ).$$ We usually write just $h$ for $h(\text{id}_{[n]})$. We construct these maps $h(\alpha )$ by induction on $b(\alpha )-s(\alpha )$. For $s(\alpha ) = b(\alpha )$ we choose the $h(\alpha )$ to be any homotopies between the maps $f(\text{id}_{[n]})$ and the zero maps. For the inductive step, it suffices to show that $$\Psi (\alpha ) = f(\alpha )+d(\alpha )h+hd(\alpha ) + \sum ^{}_{\alpha = \beta \gamma }{}^\prime d(\beta )h(\gamma )+h(\beta )d(\gamma )$$ commutes with the differentials $d$, where $\Sigma _{\alpha = \beta \gamma }'$ denotes the sum over all possible factorizations with $\beta $ and $\gamma $ not equal to the identity maps. For then $\Psi (\alpha )$ is homotopic to zero and we can take $ h(\alpha )$ to be a homotopy between $\Psi (\alpha )$ and $0$. Define $$A(\alpha ) = \sum _{\alpha = \beta \gamma }{}^\prime d(\beta )h(\gamma )+h(\gamma )d(\beta )$$ and $$B(\alpha ) = d(\alpha )h+hd(\alpha )+A(\alpha ).$$ $$\sum _{\alpha = \beta \gamma }{}^\prime A(\beta )d(\gamma )-d(\beta )A(\gamma ) = \sum _{\alpha = \beta \gamma }{}^\prime h(\beta )S(\gamma )-S(\beta )h(\gamma ),$$ where $S(\alpha )$ denotes $\sum _{\alpha = \beta \gamma }{}^\prime d(\beta )d(\gamma ).$ $$\begin{aligned} \sum _{\alpha = \beta \gamma }{}^\prime A(\beta )d(\gamma )-d(\beta )A(\gamma ) & = & \sum _{\alpha = \epsilon \rho \gamma }{}^\prime d(\epsilon )h(\rho )d(\gamma )+h(\epsilon )d(\rho )d(\gamma ) - d(\epsilon )h(\rho )d(\gamma )-d(\epsilon )d(\rho )h(\gamma )\\ & = & \sum _{\alpha = \beta \gamma }{}^\prime h(\beta )S(\gamma )-S(\beta )h(\gamma ),\end{aligned}$$ where $\Sigma _{\alpha = \epsilon \rho \gamma }'$ denotes the sum over all possible factorizations with $\epsilon $, $\rho $, and $\gamma $ not equal to the identity maps. \[Alem1\] $$\sum _{\alpha = \beta \gamma }{}^\prime B(\beta )d(\gamma )-d(\beta )B(\gamma )= -h(d(\alpha )d+dd(\alpha ))+(d(\alpha )d+dd(\alpha ))h+\sum _{\alpha = \beta \gamma }{}^\prime h(\beta )S(\gamma )-S(\beta )h(\gamma ).$$ $$\begin{aligned} &&\sum _{\alpha = \beta \gamma }{}^\prime B(\beta )d(\gamma )-d(\beta )B(\gamma )\\ & = & \sum _{\alpha =\beta \gamma }{}^\prime d(\beta )hd(\gamma )+hd(\beta )d(\gamma )+A(\beta )d(\gamma )-d(\beta )d(\gamma )h-d(\beta )hd(\gamma ) - d(\beta )A(\gamma )\\ & = & -h(d(\alpha )d+dd(\alpha ))+(d(\alpha )d+dd(\alpha ))h+\sum _{\alpha = \beta \gamma }{}^\prime h(\beta )S(\gamma )-S(\beta )h(\gamma ).\end{aligned}$$ We can now prove \[L-local\]. We compute $$\begin{aligned} &&dA(\alpha )-A(\alpha )d \\ & = & \sum _{\alpha = \beta \gamma }{}^\prime dd(\beta )h(\gamma )+dh(\beta )d(\gamma )-d(\beta )h(\gamma )d-h(\beta )d(\gamma )d\\ &=& \sum _{\alpha = \beta \gamma }{}^\prime dd(\beta )h(\gamma )+(-f(\beta )-B(\beta )-h(\beta )d)d(\gamma )-d(\beta )(-f(\gamma )-B(\gamma )-dh(\gamma ))-h(\beta )d(\gamma )d\\ &=& \sum _{\alpha = \beta \gamma }{}^\prime dd(\beta )h(\gamma )-f(\beta )d(\gamma )-B(\beta )d(\gamma )\\&-&h(\beta )dd(\gamma ) + d(\beta )f(\gamma )+d(\beta )B(\gamma )+d(\beta )dh(\gamma )-h(\beta )d(\gamma )d\\ & = & [\sum _{\alpha = \beta \gamma }{}^\prime (-S(\beta )h(\gamma ))+h(\beta )S(\gamma )-f(\beta )d(\gamma )+d(\beta )f(\gamma )]\\ & + &h(d(\alpha )d+dd(\alpha ))-(d(\alpha )d+dd(\alpha ))h-\sum _{\alpha = \beta \gamma }{}^\prime h(\beta )S(\gamma )-S(\beta )h(\gamma )\\ & = & f(\alpha )d-df(\alpha )+fd(\alpha )-d(\alpha )f+h(d(\alpha )d+dd(\alpha ))-(d(\alpha )d+dd(\alpha ))h.\end{aligned}$$ So finally $$\begin{aligned} && d\Psi (\alpha )-\Psi (\alpha )d \\ &=& df(\alpha )+dd(\alpha )h+dhd(\alpha )+dA(\alpha )-f(\alpha )d-d(\alpha )hd-hd(\alpha )d-A(\alpha )d\\ & = & df(\alpha )+dd(\alpha )h+dhd(\alpha )-f(\alpha )d -d(\alpha )hd-hd(\alpha )d\\ & + &f(\alpha )d-df(\alpha )+fd(\alpha )-d(\alpha )f+hd(\alpha )d + hdd(\alpha )-d(\alpha )dh-dd(\alpha )h\\ & = & 0.\end{aligned}$$ This completes the proof of \[L-local\]. Let $$\epsilon ^*:C(\mathcal A(\emptyset ))\rightarrow C(\text{tot}(\mathcal A^+))$$ be the functor sending a complex $K$ to the object of $C(\text{tot}(\mathcal A^+))$ with $\epsilon ^*K^{n, *} = K$ with maps $d(\text{id}_{[n]})$ equal to $(-1)^n$ times the differential, for $\partial _i:[n]\rightarrow [n+1]$ the map $d(\partial _i)$ is the canonical map of complexes, and all other $d(\alpha )$’s are zero. The functor $\epsilon ^*$ takes quasi–isomorphisms to quasi–isomorphisms and hence induces a functor $$\epsilon ^*:D(\mathcal A(\emptyset))\rightarrow D(\text{tot}(\mathcal A^+)).$$ The functor $\epsilon ^*$ has a right adjoint $R\epsilon _*:D(\text{\rm tot}(\mathcal A^+))\rightarrow D(\mathcal A(\emptyset ))$ and $R\epsilon _*$ is a triangulated functor. We apply the adjoint functor theorem [@Neeman 4.1]. By our assumptions the category $D(\mathcal A(\emptyset ))$ is compactly generated. Therefore it suffices to show that $\epsilon ^*$ commutes with coproducts (direct sums) which is immediate. More concretely, the functor $R\epsilon _*$ can be computed as follows. If $K$ is $L$–local and there exists an integer $N$ such that for every $n$ we have $K^{n, m}=0$ for $m<N$, then $R\epsilon _*K$ is represented by the complex with $$(\epsilon _*K)^p = \oplus _{n+m=p}\epsilon _{n*}K^{n, m}$$ with differential given by $\sum d(\alpha )$. This follows from Yoneda’s lemma and the observation that for any $F\in D(\mathcal A(\emptyset))$ we have $$\begin{aligned} \text{Hom}_{D(\mathcal A(\emptyset))}(F, R\epsilon _*K) & = & \text{Hom}_{D(\text{tot}(\mathcal A^+))}(\epsilon ^*F, K)\\ & = & \text{Hom}_{K(\text{tot}(\mathcal A^+))}(\epsilon ^*F, K) \text{ \ \ since $K$ is $L$-local}\\ & = & \text{Hom}_{K(\mathcal A(\emptyset ))}(F, \epsilon _*K) \text{ \ \ by \cite[3.2.12]{BBD82}}.\end{aligned}$$ \[A9\] For any $F\in D^+(\mathcal A(\emptyset ))$ the natural map $F\rightarrow R\epsilon _*\epsilon ^*F$ is an isomorphism. Represent $F$ by a complex of injectives. Then $\epsilon ^*F$ is $L$–local by \[L-local\]. The result then follows from cohomological descent. \[A10\] Let $K\in D^+_{\text{\rm cart}}(\text{\rm tot}(\mathcal A^+))$ be an object. Then $\epsilon ^*R\epsilon _*K\rightarrow K$ is an isomorphism. In particular, $R\epsilon _*$ and $\epsilon ^*$ induce an equivalence of categories between $D^+_{\text{\rm cart}}(\text{\rm tot}(\mathcal A^+))$ and $D^+(\mathcal A(\emptyset ))$. For any integer $s$ and system $(K^{n, *}, d(\alpha ))$ defining an object of $C(\text{tot}(\mathcal A^+))$ we obtain a new object by $(\tau _{\leq s}K^{n, *}, d(\alpha ))$ since for any $\alpha $ which is not the identity morphism the map $d(\alpha )$ has degree $\leq 0$. We therefore obtain a functor $\tau _{\leq s}:C(\text{tot}(\mathcal A^+))\rightarrow C(\text{tot}(\mathcal A^+))$ which takes quasi–isomorphisms to quasi–isomorphisms and hence descends to a functor $$\tau _{\leq s}:D(\text{tot}(\mathcal A^+))\rightarrow D(\text{tot}(\mathcal A^+)).$$ Furthermore, there is a natural morphism of functors $\tau _{\leq s}\rightarrow \tau _{\leq s+1}$ and we have $$K\simeq \text{hocolim }\tau _{\leq s}K.$$ Note that the functor $\epsilon ^*$ commutes with homotopy colimits since it commutes with direct sums. If we show the proposition for the $\tau _{\leq s}K$ then we see that the natural map $$\epsilon ^*(\text{hocolim}R\epsilon _*\tau _{\leq s}K)\simeq \text{hocolim}\epsilon ^*R\epsilon _*\tau _{\leq s}K\rightarrow \text{hocolim} \tau _{\leq s}K\simeq K$$ is an isomorphism. In particular $K$ is in the essential image of $\epsilon ^*$. Write $K = \epsilon ^*F$. Then by \[A9\] $R\epsilon _*K\simeq F$ whence $\epsilon ^*R\epsilon _*K\rightarrow K$ is an isomorphism. It therefore suffices to prove the proposition for $K$ bounded above. Considering the distinguished triangles associated to the truncations $\tau _{\leq s}K$ we further reduce to the case when $K$ is concentrated in just a single degree. In this case, $K$ is obtained by pullback from an object of $\mathcal A(\emptyset )$ and the proposition again follows from \[A9\]. For an object $K\in K(\text{tot}(\mathcal A^+))$, we define $\tau _{\geq s}K$ to be the cone of the natural map $\tau _{\leq s-1}K\rightarrow K$. Observe that the category $K(\text{tot}(\mathcal A^+))$ has products and therefore also homotopy limits. Let $K\in K_{\mathcal C}(\text{\rm tot}(\mathcal A^+))$ be an object. By \[A10\], for each $s$ we can find a bounded below complex of injectives $I_s\in C(\mathcal A(\emptyset ))$ and a quasi–isomorphism $\sigma _s:\tau _{\geq s}K\rightarrow \epsilon ^*I_s$. Since $\epsilon ^*I_s$ is $L$–local and $\epsilon ^*:D^+(\mathcal A(\emptyset ))\rightarrow D(\text{tot}(\mathcal A^+))$ is fully faithful by \[A10\], the maps $\tau _{\geq s-1}K\rightarrow \tau _{\geq s}K$ induce a unique morphism $t_s:I_{s-1}\rightarrow I_s$ in $K(\mathcal A(\emptyset ))$ such that the diagrams $$\begin{CD} \tau _{\geq s-1}K@>>> \tau _{\geq s}K\\ @V\sigma _{s-1}VV @VV\sigma _sV \\ \epsilon ^*I_{s-1}@>t_s>> \epsilon ^*I_s \end{CD}$$ commutes in $K(\text{tot}(\mathcal A^+))$. The natural map $K\rightarrow \text{\rm holim}\epsilon ^*I_s$ is a quasi–isomorphism. It suffices to show that for all $n$ the map $K^{n, *}\rightarrow \text{holim} \epsilon _n^*I_s$ is a quasi-isomorphism, where $\epsilon _n:U_n\rightarrow T$ is the projection. Let $\mathcal S_n$ be a site inducing $U_n$ as in \[assumption1b\]. We show that for any integer $i$ the map of presheaves on the subcategory of $\mathcal S_n$ satisfying the finiteness assumption in \[assumption1b\] (i) $$(V\rightarrow U_n)\mapsto H^i(V, K^{n, *})\rightarrow H^i(V, \text{holim} \epsilon _n^*I_s)$$ is an isomorphism. For this note that for every $s$ there is a distinguished triangle $$\mathcal H^s(K^{n, *})[s]\rightarrow \epsilon _n^*I_s\rightarrow \epsilon _n^*I_{s-1}$$ and hence by the assumption \[assumption1b\] (i) the map $$\label{transmap} H^i(V, \epsilon _n^*I_s)\rightarrow H^i(V, \epsilon _n^*I_{s-1})$$ is an isomorphism for $s<i-n_0$. Since each $\epsilon _n^*I_s$ is a complex of injectives, the complex $\prod _s \epsilon _n^*I_s$ is also a complex of injectives. Therefore $$H^i(V, \prod _s\epsilon _n^*I_s)=H^i(\prod _s\epsilon _n^*I_s(V)) = \prod _sH^i(\epsilon _n^*I_s(V)).$$ It follows that there is a canonical long exact sequence $$\begin{CD} \cdots @>>> \prod _sH^i(\epsilon _n^*I_s(V))@>1-\text{shift}>> \prod _sH^i(\epsilon _n^*I_s(V))@>>> H^i(V, \text{holim}\epsilon _n^*I_s)@>>> \cdots . \end{CD}$$ From this and the fact that the maps \[transmap\] are isomorphisms for $s$ sufficiently big it follows that the cohomology group $H^i(V, \text{holim}\epsilon _n^*I_s)$ is isomorphic to $H^i(V, K^{n, *})$ via the canonical map. Passing to the associated sheaves we obtain the proposition. \[A13\] Every object $K\in D_{\mathcal C}(\text{\rm tot}(\mathcal A^+))$ is in the essential image of the functor $$\epsilon ^*:D_{\mathcal C}(\mathcal A(\emptyset ))\rightarrow D_{\mathcal C}(\text{\rm tot}(\mathcal A^+)).$$ Since $\epsilon ^*$ also commutes with products and hence also homotopy limits we find that $ K\simeq \epsilon ^*(\text{holim }I_s) $ in $D_{\mathcal C}(\text{tot}(\mathcal A^+))$ (note that $\mathcal H^i(\text{holim}I_s)$ is in $\mathcal C$ since this can be checked after applying $\epsilon ^*$). Let $[n]\mapsto K^n$ be a cartesian section of $[n]\mapsto D(U_n, \Lambda )$ such that $\mathcal Ext^i(K^n, K^n) = 0$ for all $i<0$. Then $(K^n)$ is induced by an object of $D(\text{\rm tot}(\mathcal A^+))$. Represent each $K^n$ by a homotopically injective complex (denoted by the same letter) in $C(U_n, \Lambda )$ for every $n$. For each morphism $\partial _i:[n]\rightarrow [n+1]$ (the unique morphism whose image does not contain $i$) choose a $\partial _i$-map of complexes $\partial _i^*:K^n\rightarrow K^{n+1}$ inducing the given map in $D(U_{n+1}, \Lambda )$ by the strictly simplicial structure. The proof then proceeds by the same argument used to prove [@BBD82 3.2.9]. Combining this with \[A13\] we obtain \[1.2\]. Dualizing complex ================= Dualizing complexes on algebraic spaces --------------------------------------- Let $W$ be an algebraic space and $w:W\operatorname{\rightarrow}S$ be a separated[^5] morphism of finite type with $W$ an algebraic space. We’ll define ${\Omega}_w$ by glueing as follows. By the comparison lemma ([@SGA41], III.4.1), the étale topos $W_{\textup{\'et}}$ can be defined using the site ${{\textup{\'Etale}}}(W)$ whose objects are étale morphisms $A:U\operatorname{\rightarrow}W$ where $a:U\operatorname{\rightarrow}S$ is affine of finite type. The localized topos $W_{{\textup{\'et}}|U}$ coincides with $U_{\textup{\'et}}$. Unless otherwise explicitly stated, we will ring the various étale or lisse-étale topos which will be appear by the constant Gorenstein ring $\Lambda$ of dimension $0$ of the introduction. Notice that this is not true for the corresponding lisse-étale topos. This fact will cause some difficulties below. Let $\Omega $ denote the dualizing complex of $S$, and let $\alpha :U\rightarrow S$ denote the structural morphism. We define $$\label{def-omega} {\Omega}_A=\alpha^!\Omega\in {\operatorname{{D}}}(U_{\textup{\'et}},\Lambda)=\operatorname{{D}}({W_{\textup{\'et}}}_{|U}).$$ which is the (relative) dualizing complex of $U$, and therefore one gets by biduality ([@SGA4.5], Th. finitude 4.3) $$\label{OAOA}\operatorname{\mathcal{R}\it{hom}}({\Omega}_A,{\Omega}_A)=\Lambda$$ implying at once $$\label{extW} \operatorname{\mathcal{E}\it{xt}}^i_{{W_{{\textup{\'et}}}}_{|U}}({\Omega}_A,{\Omega}_A)=0{\textup{ if }}i<0.$$ We want to apply the glueing theorem \[1.2\]. Let us therefore consider a diagram $$\xymatrix{V\ar[rr]^\sigma\ar@{..>}@/_1pc/[rdd]^ \beta\ar[rd]_B&&U\ar@{..>}@/^1pc/[ldd]_\alpha\ar[ld]^A\\ &W\ar[d]\\&S}$$ with a commutative triangle and $A,B\in{{\textup{\'Etale}}}(W)$. \[foncto-KW\] There is a functorial isomorphism $$\sigma^{*}{\Omega}_A={\Omega}_B.$$ Let $\tilde W=U\times_W V$ : it is an affine scheme, of finite type over $S$, and étale over both $U,V$. In fact, we have a cartesian diagram $$\xymatrix{U\times_WV\ar[r]\ar[d]_\delta&W\ar[d]^\Delta\\U\times_SV\ar[r]&W\times_S W}$$ where $\Delta$ is a closed immersion ($W/S$ separated) showing that $\tilde W=U\times_WV$ is a closed subscheme of $U\times_SV$ which is affine. Looking at the graph diagram with cartesian square $$\xymatrix{\tilde W\ar[r]^b\ar[d]_a&U\ar[d]^A\\V\ar@/^1pc/[u]^s \ar[r]_B\ar[ru]^\sigma&W},$$ we get that $a,b$ are étale and separated like $A,B$. One deduces a commutative diagram $$\xymatrix{\tilde W\ar[r]^b\ar[d]_a&U\ar[d]^\alpha\\V\ar@/^1pc/[u]^s \ar[r]_\beta\ar[ru]^\sigma&S}.$$ We claim that $$b^*\alpha^!\Omega=a^*\beta^!\Omega.$$ Indeed, $a,b$ being smooth of relative dimension $0$, one has $$b^*\alpha^!\Omega=b^!\alpha^!\Omega$$ and analogously $$a^*\beta^!\Omega=a^!\beta^!\Omega.$$ Because $\alpha b=\beta a$, one gets $b^!\alpha^!=a^!\beta^!$. Pulling back by $s$ gives the result.$\square$ Therefore $({\Omega}_A)_{A\in{{\textup{\'Etale}}}(W)}$ defines locally an object ${\Omega_w}$ of $\operatorname{{D}}(W)$ with vanishing negative $\operatorname{\mathcal{E}\it{xt}}$’s (recall that $w:W\rightarrow S$ is the structural morphism). By \[1.2\], we get \[KW\] There exists a unique ${\Omega}_w\in\operatorname{{D}}(W_{\textup{\'et}})$ such that ${\Omega}_{w|U}={\Omega}_A$. We need functoriality for smooth morphisms. \[fonctW\] If $f:W_1\operatorname{\rightarrow}W_2$ is a smooth $S$-morphism of relative dimension $d$ between algebraic space separated and of finite type over $S$ with dualizing complexes ${\Omega}_1,{\Omega}_2$, then $$f^*{\Omega}_2={\Omega}_1{\langle}-d{\rangle}.$$ Start with $U_2\operatorname{\rightarrow}W_2$ étale and surjective with $U_2$ affine say. Then, $\tilde W_1=W_1\times_{U_2}W_2$ is an algebraic space separated and of finite type over $S$. Let $U_1\operatorname{\rightarrow}\tilde W_1$ be a surjective étale morphism with $U_1$ affine and let $g:U_1\operatorname{\rightarrow}U_2$ be the composition $U_1\operatorname{\rightarrow}\tilde W_1\operatorname{\rightarrow}U_2$. It is a smooth morphism of relative dimension $d$ between affine schemes of finite type from which follows the formula $g^!(-)=g^*(-){\langle}d{\rangle}$. Therefore, the pull-backs of $L_1={\Omega}_1{\langle}-d{\rangle}$ and $f^*{\Omega}_2$ on $U_1$ are the same, namely ${\Omega}_{U_1}$. One deduces that these complexes coincide on the covering sieve ${W_{1{\textup{\'et}}}}_{|U_1}$ and therefore coincide by \[D1\] (because the relevant negative $\operatorname{\mathcal{E}\it{xt}}^i$’s vanish.$\square$ Étale dualizing data -------------------- Let ${\mathcal{X}}\operatorname{\rightarrow}S$ be an algebraic $S$-stack locally of finite type. Let $A:U\operatorname{\rightarrow}{\mathcal{X}}$ in ${{\textup{Lisse-Et}}}({\mathcal{X}})$ and $\alpha:U\operatorname{\rightarrow}S$ the composition $U\operatorname{\rightarrow}{\mathcal{X}}\operatorname{\rightarrow}S$. We define $$\label{def-KA} K_A=\Omega_\alpha{\langle}-d_A{\rangle}\in {\operatorname{{D}}}_c(U_{\textup{\'et}},\Lambda)$$ where $d_A$ is the relative dimension of $A$ (which is locally constant). Up to shift and Tate torsion, $K_A$ is the (relative) dualizing complex of $U$ and therefore one gets by biduality $$\label{KAKA}\operatorname{\mathcal{R}\it{hom}}(K_A,K_A)=\Lambda\textup{ and }\operatorname{\mathcal{E}\it{xt}}^i_{U_{\textup{\'et}}}(K_A,K_A)=0{\textup{ if }}i<0.$$ We need again a functoriality property of $K_A$. Let us consider a diagram $$\xymatrix{V\ar[rr]^\sigma\ar@{..>}@/_1pc/[rdd]^ \beta\ar[rd]_B&&U\ar@{..>}@/^1pc/[ldd]_\alpha\ar[ld]^A\\ &{\mathcal{X}}\ar[d]\\&S}$$ with a $2$-commutative triangle and $A,B\in{{\textup{Lisse-Et}}}({\mathcal{X}})$. \[foncto-K\] There is a functorial identification $$\sigma^{*}K_A=K_B.$$ Let $W=U\times_{\mathcal{X}}V$ which is an algebraic space. One has a commutative diagram with cartesian square $$\xymatrix{W\ar[r]^b\ar[d]_a&U\ar[d]^A\\V\ar@/^1pc/[u]^s \ar[r]_B\ar[ru]^\sigma&{\mathcal{X}}}.$$ In particular, $a,b$ are smooth and separated like $A,B$. One deduces a commutative diagram $$\xymatrix{W\ar[r]^b\ar[d]_a&U\ar[d]^\alpha\\V\ar@/^1pc/[u]^s\ar[r]_\beta\ar[ru]^\sigma&S}.$$ I claim that $$b^*K_A=a^*K_B=K_w.$$ where $w$ denotes the structural morphism $W\operatorname{\rightarrow}S$. Indeed, $a,b$ being smooth of relative dimensions $d_A,d_B$, one has \[fonctW\] $$b^*K_A=b^*\Omega_\alpha{\langle}-d_A{\rangle}=\Omega_\alpha{\langle}-d_A-d_B{\rangle}$$ and analogously $$a^*K_B=a^*\Omega{\langle}-d_B{\rangle}=\Omega_\alpha{\langle}-d_B-d_A{\rangle}.$$ Because $\alpha b=\beta a$, one gets $b^!\alpha^!=a^!\beta^!$. Pulling back by $s$ gives the result. $\square$ \[finitude-inj\_W\] Because all $S$-schemes of finite type satisfy $\textrm{cd}_\Lambda(X)<\infty$, we know that $K_X$ is not only of finite quasi-injective dimension but of finite injective dimension ([@SGA5], I.1.5). By construction this implies that $K_A$ is of finite injective dimension for $A$ as above. Lisse-étale dualizing data -------------------------- In order to define $\Omega_{\mathcal{X}}\in\operatorname{{D}}({\mathcal{X}}_{\textup{lis-\'et}})$ by glueing, we need glueing data $\kappa_A\in\operatorname{{D}}({\mathcal{X}}_{{\textup{lis-\'et}}|U}), U\in{{\textup{Lisse-Et}}}({\mathcal{X}})$. The inclusion $${{\textup{\'Etale}}}(U)\hookrightarrow{{\textup{Lisse-Et}}}({\mathcal{X}})_{|U}$$ induces a continuous morphism of sites. Since finite inverse limits exist in ${{\textup{\'Etale}}}(U)$ and this morphism of sites preserves such limits, it defines by ([@SGA41], 4.9.2) a morphism of topos (we abuse notation slightly and omit the dependence on $A$ from the notation) $$\epsilon:{\mathcal{X}}_{{\textup{lis-\'et}}|U}\operatorname{\rightarrow}U_{\textup{\'et}}.$$ \[LETdata\] Let us describe more explicitely the morphism $\epsilon$. Let ${{\textup{Lisse-Et}}}({\mathcal}X)_{|U}$ denote the category of morphisms $V\rightarrow U$ in ${{\textup{Lisse-Et}}}({\mathcal}X)$. The category ${{\textup{Lisse-Et}}}({\mathcal}X)_{|U}$ has a Grothendieck topology induced by the topology on ${{\textup{Lisse-Et}}}({\mathcal}X)$, and the resulting topos is canonicallly isomorphic to the localized topos ${\mathcal{X}}_{{\textup{lis-\'et}}|U}$. Note that there is a natural inclusion ${{\textup{Lisse-Et}}}(U)\hookrightarrow {{\textup{Lisse-Et}}}({\mathcal}X)_{|U}$ but this is not an equivalence of categories since for an object $(V\rightarrow U)\in {{\textup{Lisse-Et}}}({\mathcal}X)_{|U}$ the morphism $V\rightarrow U$ need not be smooth. It follows that an element of ${\mathcal{X}}_{{\textup{lis-\'et}}|U}$ is equivalent to giving for every $U$–scheme of finite type $V\rightarrow U$, such that the composite $V\rightarrow U\rightarrow {\mathcal{X}}$ is smooth, a sheaf ${\mathcal{F}}_V\in V_{{\textup{\'et}}}$ together with morphisms $f^{-1}{\mathcal{F}}_V\rightarrow {\mathcal{F}}_{V'}$ for $U$–morphisms $f:V'\rightarrow V$. Furthermore, these morphisms satisfy the usual compatibility with compositions. Viewing ${\mathcal{X}}_{{\textup{lis-\'et}}|U}$ in this way, the functor $\epsilon^{-1}$ maps ${\mathcal{F}}$ on $U_{\textup{\'et}}$ to ${\mathcal{F}}_V=\pi^{-1}{\mathcal{F}}\in V_{\textup{\'et}}$ where $\pi:V\operatorname{\rightarrow}U\in{{\textup{Lisse-Et}}}({\mathcal{X}})_{|U}$. For a sheaf $F\in {\mathcal{X}}_{{\textup{lis-\'et}}|U}$ corresponding to a collection of sheaves ${\mathcal{F}}_V$, the sheaf $\epsilon _*F$ is simply the sheaf ${\mathcal{F}}_U$. In particular, the functor $\epsilon _*$ is exact and, accordingly, that $H^*(U,F)=H^*(U_{{\textup{\'et}}},F_U)$ for any shaf of $\Lambda$ modules of ${\mathcal{X}}$. A morphism $f:U\operatorname{\rightarrow}V$ of ${{\textup{Lisse-Et}}}({\mathcal{X}})$ induces a diagram $$\label{main-diag-loc} \begin{CD} {\mathcal}X_{{\textup{lis-\'et}}}|_U@>\epsilon >> U_{{\textup{\'et}}}\\ @VfVV @VVV \\ {\mathcal}X_{{\textup{lis-\'et}}}|_{V}@>\epsilon >> V_{{\textup{\'et}}} \end{CD}$$ where ${\mathcal{X}}_{{\textup{lis-\'et}}|U}\operatorname{\rightarrow}{\mathcal{X}}_{{\textup{lis-\'et}}|V}$ is the localization morphism ([@SGA41], IV.5.5.2) which we still denote by $f$ slightly abusively. For a sheaf ${\mathcal{F}}\in V_{{\textup{\'et}}}$, the pullback $f^{-1}\epsilon^{-1}{\mathcal{F}}$ is the sheaf corresponding to the system which to any $p:U'\rightarrow U$ associates $p^{-1}f^{-1}{\mathcal{F}}$. In particular, $f^{-1}\circ \epsilon ^{-1} = \epsilon ^{-1}\circ f^{-1}$ which implies that \[main-diag-loc\] is a commutative diagram of topos. We define $$\label{def-dual-loc} \kappa_A=\epsilon^*K_A\in\operatorname{{D}}({\mathcal{X}}_{{\textup{lis-\'et}}|U}).$$ By the preceding discussion, if $$\xymatrix{U\ar[rr]^f\ar[rd]_A&&V\ar[ld]^B\\&{\mathcal{X}}}$$ is a morphism in ${{\textup{Lisse-Et}}}({\mathcal{X}})$, we get $$f^*\kappa_B=\kappa_A$$ showing that the family $(\kappa_A)$ defines locally an object of $\operatorname{{D}}({\mathcal{X}}_{\textup{lis-\'et}})$. Glueing the local dualizing data -------------------------------- Let $A\in{{\textup{Lisse-Et}}}({\mathcal{X}})$ and $\epsilon:{\mathcal{X}}_{{\textup{lis-\'et}}|U}\operatorname{\rightarrow}U_{\textup{\'et}}$ be as above. We need first the vanishing of $\operatorname{\mathcal{E}\it{xt}}^i(\kappa_A,\kappa_A), i<0$. \[tauto-lisse-etale\] Let ${\mathcal{F}},{\mathcal{G}}\in\operatorname{{D}}(U_{\textup{\'et}})$. One has - $\operatorname{\mathrm{Ext}}^i({\epsilon}^{*}{{\mathcal{F}}},{\epsilon}^{*}{{\mathcal{G}}})=\operatorname{\mathrm{Ext}}^i({{\mathcal{F}}},{{\mathcal{G}}})$. - The étale sheaf $\operatorname{\mathcal{E}\it{xt}}^i({\epsilon}^{*}{{\mathcal{F}}},{\epsilon}^{*}{{\mathcal{G}}})_U$ on $U_{\textup{\'et}}$ is $\operatorname{\mathcal{E}\it{xt}}^i_{U_{\textup{\'et}}}({{\mathcal{F}}},{{\mathcal{G}}})$. Since $\epsilon _*$ is exact and for any sheaf $F\in U_{\textup{\'et}}$ one has $F=\epsilon _*\epsilon ^*F$, the adjunction map $F\rightarrow R\epsilon _*\epsilon ^*F$ is an isomorphism for any $F\in \operatorname{{D}}(U_{\textup{\'et}})$. By trivial duality, one gets $${\epsilon}_*\operatorname{\mathcal{R}\it{hom}}({\epsilon}^{*}{{\mathcal{F}}},{\epsilon}^{*}{{\mathcal{G}}})=\operatorname{\mathcal{R}\it{hom}}({{\mathcal{F}}},{\epsilon}_*{\epsilon}^{*}{{\mathcal{G}}})=\operatorname{\mathcal{R}\it{hom}}({{\mathcal{F}}},{{\mathcal{G}}}).$$ Taking ${{\mathcal}H}^iR\Gamma$ gives *(i)*. By construction, $\operatorname{\mathcal{E}\it{xt}}^i(\epsilon^*{\mathcal{F}},\epsilon^*{\mathcal{F}})_U$ is the sheaf associated to the presheaf on $U_{{\textup{\'et}}}$ which to any étale morphism $\pi :V\rightarrow U$ associates $\operatorname{\mathrm{Ext}}^i(\pi^*\epsilon^*{\mathcal{F}},\pi^*\epsilon^*{\mathcal{G}})$ where $\pi^*$ is the the pull-back functor associated to the localization morphism $$({\mathcal{X}}_{\textup{lis-\'et}}|_U)_{|V}={\mathcal{X}}_{{\textup{lis-\'et}}|V}\operatorname{\rightarrow}{\mathcal{X}}_{{\textup{lis-\'et}}|U}$$ ([@SGA42], V.6.1). By the commutativity of the diagram \[main-diag-loc\], one has $\pi^*\epsilon^*=\epsilon^*\pi^*$. Therefore $$\operatorname{\mathrm{Ext}}^i(\pi^*\epsilon^*{\mathcal{F}},\pi^*\epsilon^*{\mathcal{G}})=\operatorname{\mathrm{Ext}}^i(\epsilon^*\pi^*{\mathcal{F}},\epsilon^*\pi^*{\mathcal{G}})= \operatorname{\mathrm{Ext}}^i_{V_{\textup{\'et}}}(\pi^*{\mathcal{F}},\pi^*{\mathcal{G}}),$$ the last equality is by [*(i)*]{}. Since $\operatorname{\mathcal{E}\it{xt}}_{U_{\textup{\'et}}}({\mathcal{F}}, {\mathcal{G}})$ is also the sheaf associated to this presheaf we obtain [*(ii)*]{}.$\square$ Using \[KAKA\], one obtains \[kAkA\] One has $\operatorname{\mathcal{R}\it{hom}}(\kappa_A,\kappa_A)=\Lambda$ and therefore $\operatorname{\mathcal{E}\it{xt}}^i(\kappa_A,\kappa_A)=0$ if $i<0$. The discussion above shows that we can apply \[1.2\] to $(\kappa_A)$ to get \[exis-dual\] There exists ${\Omega}_{\mathcal{X}}(p)\in {\operatorname{{D}}}^b({\mathcal{X}}_{\textup{lis-\'et}})$ inducing $\kappa_A$ for all ${A\in{{\textup{Lisse-Et}}}({\mathcal{X}})_{|X}}$. It is well defined up to unique isomorphism. The independence of the presentation is straightforward and is left to the reader : \[indep-K\]Let $p_i:X_i\operatorname{\rightarrow}{\mathcal{X}},i=1,2$ two presentations as above. There exists a canonical, functorial isomorphism ${\Omega}_{\mathcal{X}}(p_1){\xrightarrow{\sim}}{\Omega}_{\mathcal{X}}(p_2)$. \[def-dualisant\] The *dualizing complex of ${\mathcal{X}}$* is the “essential” value $\Omega _{{\mathcal{X}}}\in \operatorname{{D}}^b({\mathcal{X}}_{{\textup{lis-\'et}}})$ of ${\Omega_{\mathcal{X}}}(p)$, where $p$ runs over presentations of ${\mathcal{X}}$. It is well defined up to canonical functorial isomorphism and is characterized by $\Omega_{{\mathcal{X}}|U}=\epsilon^*K_A$ for any $A:U\operatorname{\rightarrow}{\mathcal{X}}$ in ${{\textup{Lisse-Et}}}({\mathcal{X}})$. Biduality --------- For $A,B$ any abelian complexes of some topos, there is a biduality *morphism* $$\label{bidual} A\operatorname{\rightarrow}\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(A,B),B)$$ (replace $B$ by some homotopically injective complex isomorphic to it in the derived catgory). In general, it is certainly not an isomorphism. \[bidualnonborne\] Let $u:U\operatorname{\rightarrow}S$ be a separated $S$-scheme (or algebraic space) of finite type and $A\in {\operatorname{{D}}}_c(U_{\textup{\'et}},\Lambda)$. Then the biduality morphism $$A\operatorname{\rightarrow}\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(A,K_U),K_U)$$ is an isomorphism (where $K_U$ is -up to shift and twist- the dualizing complex of $U_{\textup{\'et}}$). If $A$ is moreover bounded, it is the usual theorem of  [@SGA4.5]. Let us denote by $\tau_n$ the two-sides truncation functor $$\tau_{\geq -n}\tau_{\leq n}.$$ We know that $K_U$ is a dualizing complex ([@SGA5], exp. I), and is of *finite injective dimension* (\[finitude-inj\_W\]); the homology in degree $n$ of the biduality morphism $A\operatorname{\rightarrow}DD(A)$ is therefore the same as the homology in degree $n$ of the biduality morphism $\tau_mA\operatorname{\rightarrow}DD(\tau_m A)$ for $m$ large enough and the lemma follows.$\square$ We will be interested in a commutative diagram $$\xymatrix{V\ar[rr]^f\ar[rd]_B&&\ar[ld]^AU\\&{\mathcal{X}}}$$ as above. \[tec1\] Let ${\mathcal{F}}\in\operatorname{{D}}_c({\mathcal{X}}_{{\textup{lis-\'et}}})$ and let ${\mathcal{F}}_U\in D_c(U_{\textup{\'et}})$ be the object obtained by restriction. - One has $f^*\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}}_U,K_A)=\operatorname{\mathcal{R}\it{hom}}(f^*{\mathcal{F}}_U,f^*K_A)=\operatorname{\mathcal{R}\it{hom}}(f^*{\mathcal{F}}_U,K_B)$. - Moreover, $\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}}_U,K_A)$ is constructible. Let’s prove (*i*). By \[foncto-K\], one has $f^*K_A=K_B$, therefore one has a morphism $$f^*\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}}_U,K_A)\operatorname{\rightarrow}\operatorname{\mathcal{R}\it{hom}}(f^*{\mathcal{F}}_U,K_B).$$ To prove that it is an isomorphism, consider first the case when $f$ is smooth. Because both $K_A$ and $K_B$ are of finite injective dimension (\[finitude-inj\_W\]), one can assume that $F$ is bounded where it is obviously true by reduction to $F$ the constant sheaf (or use [@SGA5], I.7.2). Therefore the result holds when $f$ is smooth. From the case of a smooth morphism, one reduces the proof in general to the case when ${\mathcal{X}}$ is a scheme. Let ${\mathcal{F}}_{{\mathcal{X}}}\in D_c({\mathcal{X}}_{\textup{\'et}})$ denote the complex obtained by restricting ${\mathcal{F}}$. By the smooth case already considered, we have $$\begin{aligned} f^*\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}}_U,K_A)& \simeq & f^*A^*\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}}_{\mathcal{X}}, K_{\mathcal{X}})\\ & = & B^*\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}}_{\mathcal{X}}, K_{\mathcal{X}})\\ & \simeq & \operatorname{\mathcal{R}\it{hom}}(B^*{\mathcal{F}}_X, B^*K_{\mathcal{X}})\\ & \simeq &\operatorname{\mathcal{R}\it{hom}}(f^*{\mathcal{F}}_U, f^*K_A).\end{aligned}$$ For (*ii*), one can also assume ${\mathcal{F}}$ bounded and one uses [@SGA5], I.7.1.$\square$ \[tec2\] Let ${\mathcal{F}}\in\operatorname{{D}}_c({\mathcal{X}}_{\textup{lis-\'et}})$. Then, $$\epsilon^*\operatorname{\mathcal{R}\it{hom}}_{U_{\textup{\'et}}}({\mathcal{F}}_U,K_A)=\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}},{\Omega_{\mathcal{X}}})_{|U}$$ where ${\mathcal{F}}_U=\epsilon_*{\mathcal{F}}_{|U}$ is the restriction of ${\mathcal{F}}$ to ${{\textup{\'Etale}}}(U)$. By definition of constructibility, ${{\mathcal}H}^i({\mathcal{F}})$ are cartesian sheaves. In other words, $\epsilon_*$ being exact, the adjunction morphism $$\epsilon^*{\mathcal{F}}_U=\epsilon^*\epsilon_*{\mathcal{F}}_{|U}\operatorname{\rightarrow}{\mathcal{F}}_{|U}$$ is an isomorphism. We therefore have$$\begin{aligned} \operatorname{\mathcal{R}\it{hom}}({\mathcal{F}},\Omega)_{|U} &=& \operatorname{\mathcal{R}\it{hom}}({\mathcal{F}}_{|U},\Omega_{|U})\\ &=& \operatorname{\mathcal{R}\it{hom}}(\epsilon^*{\mathcal{F}}_U,\epsilon^* K_A) \\\end{aligned}$$ Therefore, we get a morphism $$\epsilon^*\operatorname{\mathcal{R}\it{hom}}_{U_{\textup{\'et}}}({\mathcal{F}}_U,K_A)\operatorname{\rightarrow}\operatorname{\mathcal{R}\it{hom}}(\epsilon^*{\mathcal{F}}_U,\epsilon^* K_A)=\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}},{\Omega_{\mathcal{X}}})_{|U}.$$ By \[tauto-lisse-etale\], one has $$\operatorname{\mathcal{E}\it{xt}}^i({\epsilon}^{*}{{\mathcal{F}}_U},{\epsilon}^{*}K_A)_V= \operatorname{\mathcal{E}\it{xt}}^i_{V_{\textup{\'et}}}(f^{*}{{\mathcal{F}}_U},f^{*}K_A).$$ But, one has $${{\mathcal}H}^i(\epsilon^*\operatorname{\mathcal{R}\it{hom}}_{U_{\textup{\'et}}}({\mathcal{F}}_U,K_A))_V=f^*\operatorname{\mathcal{E}\it{xt}}^i_{U_{\textup{\'et}}}({\mathcal{F}}_U,K_A))$$ and the lemma follows from \[tec1\].$\square$ One gets immediately (cf. [@SGA5], I.1.4) \[diminffinie\] ${\Omega_{\mathcal{X}}}$ is of finite quasi-injective dimension. It seems over-optimistic to think that ${\Omega_{\mathcal{X}}}$ would be of finite injective dimension even if ${\mathcal{X}}$ is a scheme. \[dula-construct\] If $A\in D_c({\mathcal{X}})$, then $\operatorname{\mathcal{R}\it{hom}}(A,{\Omega_{\mathcal{X}}})\in D_c({\mathcal{X}})$. Immediate consequence of \[tec1\] and \[tec2\].$\square$ \[Dinvolutif\] The (contravariant) functor $$D_{\mathcal{X}}:\left\{\begin{array}{ccc}{\operatorname{{D}}}_c({\mathcal{X}})&\operatorname{\rightarrow}&{\operatorname{{D}}}_c({\mathcal{X}})\\{\mathcal{F}}&\mapsto&\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}},{\Omega_{\mathcal{X}}})\end{array}\right.$$ is an involution. More precisely, the morphism $$\iota:\operatorname{Id}\operatorname{\rightarrow}{\operatorname{{D}}}_{\mathcal{X}}\circ {\operatorname{{D}}}_{\mathcal{X}}$$ induced by \[bidual\] is an isomorphism. We have to prove that the cone $C$ of the biduality morphism is zero in the derived category, that is to say $$C_U=\epsilon_*C_{|U}=0\text{ in }\operatorname{{D}}_c(U_{\textup{\'et}}).$$ But we have$$\begin{aligned} \epsilon_*(\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}},{\Omega_{\mathcal{X}}}),{\Omega_{\mathcal{X}}}))_{|U} &=& \epsilon_*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}},{\Omega_{\mathcal{X}}})_{|U},\Omega_{{\mathcal{X}}|U}) \\ &\stackrel{\ref{tec2}}{=}& \epsilon_*\operatorname{\mathcal{R}\it{hom}}(\epsilon^*\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}}_U,K_A),\Omega_{{\mathcal{X}}|U})\\ &=&\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}}_U,K_A),\epsilon_*\epsilon^*K_A)\text{ by trivial duality} \\ &=&\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}({\mathcal{F}}_U,K_A),K_A)\\ &\stackrel{\ref{bidualnonborne}}{=}&{\mathcal{F}}_U\end{aligned}$$ $\square$ Verdier duality $D_{\mathcal{X}}$ identifies $D^a_c$ and $D_c^{-a}$ with $a=\varnothing,\pm 1,b$ and the usual conventions $-\varnothing=\varnothing$ and $-b=b$. \[Danti\] One has a canonical (bifunctorial) morphism $$\operatorname{\mathcal{R}\it{hom}}(A,B)=\operatorname{\mathcal{R}\it{hom}}(D(B),D(A))$$ for all $A,B\in D_c({\mathcal{X}})$. Let us prove first a well-known formula \[bilder\] Let $A,B,C$ be complexes of $\Lambda$ modules on ${\mathcal{X}}_{\textup{lis-\'et}}$. One has canonical identifications $$\operatorname{\mathcal{R}\it{hom}}(A,\operatorname{\mathcal{R}\it{hom}}(B,C))=\operatorname{\mathcal{R}\it{hom}}(A\operatorname{\stackrel{\mathbf L}{\otimes}}B,C)=\operatorname{\mathcal{R}\it{hom}}(B,\operatorname{\mathcal{R}\it{hom}}(A,C)).$$ One can assume $A,B$ homotopically flat and $C$ homotopically injective. Let $X$ be an acyclic complex. One has $$\operatorname{\mathrm{Hom}}(X,\operatorname{\mathcal{H}om}(B,C))=\operatorname{\mathrm{Hom}}(X\otimes B,C).$$ Because $B$ est homotopically flat, $X\otimes B$ is acyclic. Moreover, $C$ being homotopically injective, the abelian complex $\operatorname{\mathrm{Hom}}(X\otimes B,C)$ is acyclic. Therefore, $\operatorname{\mathcal{H}om}(B,C)$ homotopically injective. One gets therefore $$\operatorname{\mathcal{R}\it{hom}}(A,\operatorname{\mathcal{R}\it{hom}}(B,C))=\operatorname{\mathcal{H}om}(A,\operatorname{\mathcal{H}om}(B,C))=\operatorname{\mathcal{H}om}(A\otimes B,C)=\operatorname{\mathcal{R}\it{hom}}(A\operatorname{\stackrel{\mathbf L}{\otimes}}B,C).$$$\square$ One gets then $$\operatorname{\mathcal{R}\it{hom}}(D(B),D(A))=\operatorname{\mathcal{R}\it{hom}}(D(B),\operatorname{\mathcal{R}\it{hom}}(A,\Omega _{\mathcal{X}}))\stackrel{\ref{bilder}}{=}\operatorname{\mathcal{R}\it{hom}}(A,D\circ D(B))=\operatorname{\mathcal{R}\it{hom}}(A,B).$$$\square$ The $6$ operations ================== The functor $\operatorname{\mathcal{R}\it{hom}}(-, -)$ ------------------------------------------------------ Let ${\mathcal{X}}$ be an $S$–stack locally of finite type. As in any topos, one can define internal hom $\operatorname{\mathcal{R}\it{hom}}_{{\mathcal{X}}_{{\textup{lis-\'et}}}}(F, G)$ for any $F\in \operatorname{{D}}^-({\mathcal{X}})$ and $G\in \operatorname{{D}}^+({\mathcal}X)$. Let $F\in \operatorname{{D}}_c^-({\mathcal{X}})$ and $G\in \operatorname{{D}}_c^+({\mathcal{X}})$, and let $j$ be an integer. Then the restriction of the sheaf ${\mathcal}H^j(\operatorname{\mathcal{R}\it{hom}}_{{\mathcal{X}}_{\textup{lis-\'et}}}(F, G))$ to the étale topos of any object $U\in {{\textup{Lisse-Et}}}({\mathcal{X}})$ is canonically isomorphic to $\operatorname{\mathcal{E}\it{xt}}^j_{U_{\textup{\'et}}}(F_U, G_U)$, where $F_U$ and $G_U$ denote the restrictions to $U_{\textup{\'et}}$. The sheaf ${\mathcal}H^j(\operatorname{\mathcal{R}\it{hom}}_{{\mathcal}X_{{\textup{lis-\'et}}}}(F, G))$ is the sheaf associated to the presheaf which to any smooth affine ${\mathcal}X$–scheme $U$ associates $\text{Ext}^j_{{\mathcal}X_{{\textup{lis-\'et}}|U}}(F, G)$, where ${\mathcal}X_{{\textup{lis-\'et}}|U}$ denotes the localized topos. Let $\epsilon :{\mathcal}X_{{\textup{lis-\'et}}|U}\rightarrow U_{{\textup{\'et}}}$ be the morphism of topos induced by the inclusion of ${{\textup{\'Etale}}}(U)$ into ${{\textup{Lisse-Et}}}({\mathcal}X)_{|U}$. Then since $F$ and $G$ have constructible cohomology, the natural maps $\epsilon ^*\epsilon _*F\rightarrow F$ and $\epsilon ^*\epsilon _*G\rightarrow G$ are isomorphisms in $\operatorname{{D}}({\mathcal}X_{{\textup{lis-\'et}}|U})$. By the projection formula it follows that $$\text{Ext}^j_{{\mathcal}X_{{\textup{lis-\'et}}|U}}(F, G)\simeq \text{Ext}^j_{{\mathcal}X_{{\textup{lis-\'et}}|U}}(\epsilon ^*\epsilon _*F, \epsilon ^*\epsilon _*G)\simeq \text{Ext}^j_{U_{{\textup{\'et}}}}(\epsilon _*F, \epsilon _*G).$$ Sheafifying this isomorphism we obtain the isomorphism in the lemma. $\square$ If $F\in \operatorname{{D}}^-_c({\mathcal{X}})$ and $G\in \operatorname{{D}}^+_c({\mathcal{X}})$, the complex $\operatorname{\mathcal{R}\it{hom}}_{{\mathcal}X_{\textup{lis-\'et}}}(F, G)$ lies in $\operatorname{{D}}_c^+({\mathcal}X)$. By the previous lemma and the constructibility of the cohomology sheaves of $F$ and $G$, it suffices to prove the following statement: Let $f:V\rightarrow U$ be a smooth morphism of schemes of finite type over $S$, and let $F\in \operatorname{{D}}_c^-(U_{\textup{\'et}})$ and $G\in \operatorname{{D}}_c^+(U_{\textup{\'et}})$. Then the natural map $f^*\operatorname{\mathcal{R}\it{hom}}_{U_{\textup{\'et}}}(F, G)\rightarrow \operatorname{\mathcal{R}\it{hom}}_{V_{{\textup{\'et}}}}(f^*F, f^*G)$ is an isomorphism as we saw in the proof of \[tec1\] (see [@SGA5], I.7.2). $\square$ \[homdescription\] Let $X/S$ be an $S$–scheme locally of finite type and $X\rightarrow {\mathcal{X}}$ be a smooth surjection. Let $X_\bullet \rightarrow {\mathcal{X}}$ be the resulting strictly simplicial space. Then for $F\in \operatorname{{D}}_c^-({\mathcal{X}}_{{\textup{lis-\'et}}})$ and $G\in \operatorname{{D}}_c^+({\mathcal{X}}_{{\textup{lis-\'et}}})$ there is a canonical isomorphism $$\operatorname{\mathcal{R}\it{hom}}_{{\mathcal{X}}_{{\textup{lis-\'et}}}}(F, G)|_{X_\bullet, {\textup{\'et}}}\simeq \operatorname{\mathcal{R}\it{hom}}_{X_{\bullet {\textup{\'et}}}}(F|_{X_{\bullet, {\textup{\'et}}}}, G|_{X_{\bullet, {\textup{\'et}}}}).$$ In particular, $\operatorname{\mathcal{R}\it{hom}}_{X_{\bullet {\textup{\'et}}}}(F|_{X_{\bullet, {\textup{\'et}}}}, G|_{X_{\bullet, {\textup{\'et}}}})$ maps under the equivalence of categories $\operatorname{{D}}_c(X_{\bullet, {\textup{\'et}}})\simeq \operatorname{{D}}_c({\mathcal}X_{{\textup{lis-\'et}}})$ to $\operatorname{\mathcal{R}\it{hom}}_{{\mathcal{X}}_{{\textup{lis-\'et}}}}(F, G)$. Let ${\mathcal}X_{{\textup{lis-\'et}}|X_\bullet }$ denote the strictly simplicial localized topos and consider the morphisms of topos $$\begin{CD} {\mathcal{X}}_{{\textup{lis-\'et}}}@<\pi << {\mathcal{X}}_{{\textup{lis-\'et}}|X_\bullet }@>\epsilon >> X_{\bullet , {\textup{\'et}}}. \end{CD}$$ Let $F_{\textup{\'et}}:= \epsilon _*\pi ^*F$ and $G_{\textup{\'et}}:= \epsilon _*\pi ^*G$. Since $F,G\in \operatorname{{D}}_c({\mathcal}X_{{\textup{lis-\'et}}})$, the natural maps $F \simeq R\pi _*\epsilon ^*F_{\textup{\'et}}$ and $G \simeq R\pi _*\epsilon ^*G$ are isomorphisms (\[mainthm\]). Using the projection formula we then obtain $$\begin{aligned} \operatorname{\mathcal{R}\it{hom}}_{{\mathcal{X}}_{{\textup{lis-\'et}}}}(F, G)|_{X_\bullet, {\textup{\'et}}} & \simeq & \epsilon _*\pi ^*\operatorname{\mathcal{R}\it{hom}}_{{\mathcal{X}}_{{\textup{lis-\'et}}}}(F, G)\\ & \simeq & \epsilon _*\pi ^*\pi _*\operatorname{\mathcal{R}\it{hom}}_{{\mathcal{X}}_{{\textup{lis-\'et}}|X_\bullet }}(\epsilon ^*F_{\textup{\'et}}, \epsilon ^*G_{\textup{\'et}})\\ & \simeq & \epsilon _*\operatorname{\mathcal{R}\it{hom}}_{{\mathcal{X}}_{{\textup{lis-\'et}}|X_\bullet }}(\epsilon ^*F_{\textup{\'et}}, \epsilon ^*G_{\textup{\'et}})\\ & \simeq & \operatorname{\mathcal{R}\it{hom}}_{X_{\bullet, {\textup{\'et}}}}(F_{\textup{\'et}}, G_{\textup{\'et}}).\end{aligned}$$ $\square$ The functor $f^*$ ----------------- The lisse-étale site is not functorial (cf. [@Ber03], 5.3.12): a morphism of stacks does not induce a general a morphism between corresponding lisse-étale topos. In [@Ols05], a functor $f^*$ is constructed on $\operatorname{{D}}_c^+$ using cohomological descent. Using the results of \[mainthm\] which imply that we have cohomological descent also for unbounded complexes, the construction of [@Ols05] can be used to define $f^*$ on the whole category ${\mathcal}D_c$. Let us review the construction here. Let $f:{\mathcal{X}}\operatorname{\rightarrow}{\mathcal{Y}}$ be a morphism of algebraic $S$–stacks locally of finite type. Choose a commutative diagram $$\xymatrix{X\ar[r]\ar[d]&{\mathcal{X}}\ar[d]\\Y\ar[r]&{\mathcal{Y}}}$$ where the horizontal lines are presentations inducing a commutative diagram of strict simplicial spaces $$\xymatrix{X_{\bullet}\ar[r]^{\eta_X}\ar[d]_{f_\bullet}& {\mathcal{X}}\ar[d]^f\\Y_{\bullet}\ar[r]^{\eta_Y}&{\mathcal{Y}}.}$$ We get a diagram of topos $$\xymatrix{X_{\bullet,{\textup{\'et}}}\ar[d]_{f_\bullet}&{\mathcal}X_{{\textup{lis-\'et}}}|_{X_{\bullet }}\ar[r]^{\eta_X}\ar[l]_-{\Phi_X}& {\mathcal{X}}_{\textup{lis-\'et}}\\Y_{\bullet,{\textup{\'et}}}&\ar[l]_-{\Phi_Y}{\mathcal}Y_{{\textup{lis-\'et}}}|_{Y_{\bullet }}\ar[r]^{\eta_Y}&{\mathcal{Y}}_{\textup{lis-\'et}}.}$$ By \[etexample\] the horizontal morphisms induce equivalences of topos $$\operatorname{{D}}_c({\mathcal}X_{{\textup{lis-\'et}}})\simeq \operatorname{{D}}_c(X_{\bullet , {\textup{\'et}}}), \ \ \ \operatorname{{D}}_c({\mathcal}Y_{{\textup{lis-\'et}}})\simeq \operatorname{{D}}_c(Y_{\bullet, {\textup{\'et}}}).$$ We define the functor $f^*:\operatorname{{D}}_c({\mathcal}Y_{{\textup{lis-\'et}}})\rightarrow \operatorname{{D}}_c({\mathcal}X_{{\textup{lis-\'et}}})$ to be the composite $$\begin{CD} \operatorname{{D}}_c({\mathcal}Y_{{\textup{lis-\'et}}})\simeq \operatorname{{D}}_c(Y_{\bullet , {\textup{\'et}}})@>f_\bullet ^*>> \operatorname{{D}}_c(X_{\bullet, {\textup{\'et}}})\simeq \operatorname{{D}}_c({\mathcal}X_{{\textup{lis-\'et}}}), \end{CD}$$ where $f_\bullet ^*$ denotes the derived pullback functor induced by the morphism of topos $f_\bullet :X_{\bullet, {\textup{\'et}}}\rightarrow Y_{\bullet , {\textup{\'et}}}$. Note that $f^*$ takes distinguished triangles to distinguished triangles since this is true for $f_{\bullet }^*$. \[adjoint\*\] Let $A\in \operatorname{{D}}_c^-({\mathcal{Y}})$ and let $B\in \operatorname{{D}}_c^+({\mathcal{X}})$. Then there is a canonical isomorphism $$f_*\operatorname{\mathcal{R}\it{hom}}(f^*A, B)\simeq \operatorname{\mathcal{R}\it{hom}}(A, f_*B).$$ where we write $f_*$ for $Rf_*$. By \[homdescription\] and [@Ols05], we have $$Rf_*\operatorname{\mathcal{R}\it{hom}}(f^*A, B)|_{Y_{\bullet, {\textup{\'et}}}}\simeq Rf_{\bullet *}\operatorname{\mathcal{R}\it{hom}}_{X_{\bullet , {\textup{\'et}}}}(f_{\bullet }^*A|_{Y_{\bullet, {\textup{\'et}}}}, B|_{X_{\bullet, {\textup{\'et}}}}).$$ The result therefore follows from the usual adjunction $$Rf_{\bullet *}\operatorname{\mathcal{R}\it{hom}}_{X_{\bullet , {\textup{\'et}}}}(f_\bullet ^*(A|_{Y_{\bullet , {\textup{\'et}}}}), B|_{X_{\bullet {\textup{\'et}}}})\simeq \operatorname{\mathcal{R}\it{hom}}_{Y_{\bullet, {\textup{\'et}}}}(A|_{\bullet, {\textup{\'et}}}, f_*B|_{X_{\bullet {\textup{\'et}}}}).$$ $\square$ Its definitely hopeless to generalize \[adjoint\*\] to $B\in\operatorname{{D}}_c({\mathcal{X}})$ because in general $Rf_*$ does not map $\operatorname{{D}}_c$ to itself (for example consider $B\mathbb{G}_m\operatorname{\rightarrow}\operatorname{\mathrm{Spec}}(k)$ and $B=\oplus_{i\geq 0}\Lambda[i]$). One can even show that \[adjoint\*\] still holds for arbitrary $A\in\operatorname{{D}}_c({\mathcal{Y}})$, but the geometric significance is unclear because it is an equality of non constructible complexes. Definition of $Rf_!,f^!$ {#4.3} ------------------------ Let $f:{\mathcal{X}}\operatorname{\rightarrow}{\mathcal{Y}}$ be a morphism of stacks (locally of finite type over $S$) of finite type. Recall ([@Lau-Mor2000], corollaire 18.4.4) that $Rf_*$ maps $D_c^+({\mathcal{X}}_{\textup{lis-\'et}})$ to $D_c^+({\mathcal{Y}}_{\textup{lis-\'et}})$. \[deff!\] We define $$Rf_!: D_c^-({\mathcal{X}}_{\textup{lis-\'et}})\operatorname{\rightarrow}D_c^-({\mathcal{Y}}_{\textup{lis-\'et}})$$ by the formula $$Rf_!=D_{\mathcal{Y}}\circ Rf_*\circ D_{\mathcal{X}},$$ and $$f^!: D_c^-({\mathcal{Y}}_{\textup{lis-\'et}})\operatorname{\rightarrow}D_c^-({\mathcal{X}}_{\textup{lis-\'et}})$$ by the formula $$f^!=D_{\mathcal{X}}\circ f^{*}\circ D_{\mathcal{Y}}.$$ By construction, one has$$\label{fonct-ext-dual} f^!\Omega_{\mathcal{Y}}=\Omega_{\mathcal{X}}.$$ \[dual\] Let $A\in D_c^-({\mathcal{X}}_{\textup{lis-\'et}})$ and $B\in D_c^-({\mathcal{Y}}_{\textup{lis-\'et}})$. Then there is a (functorial) adjunction formula $$Rf_*\operatorname{\mathcal{R}\it{hom}}(A,f^!B)=\operatorname{\mathcal{R}\it{hom}}(Rf_!A,B).$$ We write $D$ for $D_{\mathcal{X}},D_{\mathcal{Y}}$ and $A'=D(A)\in D_c^+({\mathcal{X}})$. One has $$\begin{aligned} \operatorname{\mathcal{R}\it{hom}}(Rf_!D(A'),B) & =\operatorname{\mathcal{R}\it{hom}}(D(Rf_*A'),B) \\ & =\operatorname{\mathcal{R}\it{hom}}(D(B),Rf_*A')\ (\ref{Danti})\\ &=Rf_*\operatorname{\mathcal{R}\it{hom}}(f^*D(B),A')\ (\ref{adjoint*})\\ &=Rf_*\operatorname{\mathcal{R}\it{hom}}(D(A'),f^!B)\ (\ref{Danti})\end{aligned}$$ $\square$ Projection formula ------------------ \[otimes=rhom\] Let $A,B\in D_c({\mathcal{X}})$. 1. One has $$\operatorname{\mathcal{R}\it{hom}}(A,B)=D_{\mathcal{X}}(A\operatorname{\stackrel{\mathbf L}{\otimes}}D_{\mathcal{X}}(B)).$$ 2. If $A,B\in D^-_c({\mathcal{X}})$, then $A\operatorname{\stackrel{\mathbf L}{\otimes}}B\in D^-_c({\mathcal{X}})$. 3. If $A\in D^-_c({\mathcal{X}}),B\in D^+_c({\mathcal{X}})$, then $\operatorname{\mathcal{R}\it{hom}}(A,B)\in D^+_c({\mathcal{X}})$. Let $\Omega _{\mathcal{X}}$ be the dualizing complex of ${\mathcal{X}}$. $$\begin{aligned} \operatorname{\mathcal{R}\it{hom}}(A,B) &=\operatorname{\mathcal{R}\it{hom}}(D_{\mathcal{X}}(B),\operatorname{\mathcal{R}\it{hom}}(A,\Omega _{\mathcal{X}}))\ (\ref{Danti}) \\ & =\operatorname{\mathcal{R}\it{hom}}(D_{\mathcal{X}}(B)\operatorname{\stackrel{\mathbf L}{\otimes}}A,\Omega _{\mathcal{X}})\ (\ref{bilder})\\ &=D_{\mathcal{X}}(A\operatorname{\stackrel{\mathbf L}{\otimes}}D_{\mathcal{X}}(B) )\end{aligned}$$ proving (i). For (ii), using truncations, one can assume that $A,B$ are sheaves : the result is obvious in this case. Statement (iii) follows from the two previous points. $\square$ \[corkunneth\] Let $f:{\mathcal{X}}\rightarrow {\mathcal{Y}}$ be a morphism as in \[4.3\], and let $B\in D^-_c({\mathcal{Y}}), A\in D^-_c({\mathcal{X}})$. One has the projection formula $$R f_!(A\operatorname{\stackrel{\mathbf L}{\otimes}}f^*B)=R f_!A\operatorname{\stackrel{\mathbf L}{\otimes}}B.$$ Notice that the left-hand side is well defined by \[otimes=rhom\]. One has $$\begin{aligned} R f_!(A\operatorname{\stackrel{\mathbf L}{\otimes}}f^*B) & =D_{\mathcal{Y}}\circ R f_*\circ D_{\mathcal{X}}(A\operatorname{\stackrel{\mathbf L}{\otimes}}D_{\mathcal{X}}f^!D_{\mathcal{Y}}B)\\ & =D_{\mathcal{Y}}\circ R f_*(\operatorname{\mathcal{R}\it{hom}}(A,f^!D_{\mathcal{Y}}B))\ (\ref{otimes=rhom})\\ &=D_{\mathcal{Y}}(\operatorname{\mathcal{R}\it{hom}}(R f_!A,D_{\mathcal{Y}}B))\ (\ref{dual})\\ &=R f_!A\operatorname{\stackrel{\mathbf L}{\otimes}}B\ (\ref{otimes=rhom})\text{ and}\ (\ref{Dinvolutif}).\end{aligned}$$ $\square$ \[f\*f!\] For all $A\in D_c^+({\mathcal{Y}}),B\in D^-_c({\mathcal{Y}})$, one has $f^!\operatorname{\mathcal{R}\it{hom}}(A,B)=\operatorname{\mathcal{R}\it{hom}}(f^*A,f^!B)$. By lemma \[otimes=rhom\] and biduality, the formula reduces to the formula $$f^*(A\operatorname{\stackrel{\mathbf L}{\otimes}}D(B))=f^*A\operatorname{\stackrel{\mathbf L}{\otimes}}f^*D(B).$$ Using suitable presentation, one is reduced to the obvious formula $$f_\bullet^*(A_\bullet\operatorname{\stackrel{\mathbf L}{\otimes}}B_\bullet)=f^*_\bullet A_\bullet\operatorname{\stackrel{\mathbf L}{\otimes}}f^*_\bullet B_\bullet$$ for a morphism $f_\bullet$ of stricltly simplicial étale topos. $\square$ Computation of $f^!$ for $f$ smooth ----------------------------------- Let $f:{\mathcal{X}}\operatorname{\rightarrow}{\mathcal{Y}}$ be a smooth morphism of stacks of relative dimension $d$. Using \[D1\], one gets immediately the formula $$f^*\Omega_{\mathcal{Y}}=\Omega_{\mathcal{X}}\langle -d\rangle$$ (choose a presentation of $Y\operatorname{\rightarrow}{\mathcal{Y}}$ and then a presentation $X\operatorname{\rightarrow}{\mathcal{X}}_Y$; the morphism $X\operatorname{\rightarrow}Y$ being smooth, one checks that these two complexes coincide on ${\mathcal{X}}_{{\textup{lis-\'et}}|X}$ and have zero negative $\operatorname{\mathcal{E}\it{xt}}$’s). \[Rhom-im-inv\] Let $A\in D_c({\mathcal{Y}})$. Then, the canonical morphism $$f^*\operatorname{\mathcal{R}\it{hom}}(A,\Omega_{\mathcal{Y}})\operatorname{\rightarrow}\operatorname{\mathcal{R}\it{hom}}(f^*A,f^*\Omega_{\mathcal{Y}})$$ is an isomorphism. Using \[tauto-lisse-etale\], one is reduced to the usual statement for étale sheaves on algebraic spaces. Because, in this case, both $\Omega_{\mathcal{Y}}$ and $f^*\Omega_{\mathcal{Y}}$ are of finite injective dimension, one can assume that $A$ is bounded or even a sheaf. The assertion is well-known in this case (by dévissage, one reduces to $A=\Lambda_{\mathcal{Y}}$ in which case the assertion is trivial, cf. [@SGA5], exp. I).$\square$ \[f!lisse\] Let $f:{\mathcal{X}}\operatorname{\rightarrow}{\mathcal{Y}}$ be a smooth morphism of stacks of relative dimension $d$. One has $f^!=f^*\langle d\rangle$. Let $j:\operatorname{\mathcal{U}}\operatorname{\rightarrow}{\mathcal{X}}$ be an open immersion. Let us denote for a while $j_{{\textup{!`}}}$ the extension by zero functor : it is an exact functor on the category sheaves preserving constructibility and therefore passes to the derive category $\operatorname{{D}}_c$. \[!=!!\] One has $j^!=j^*$ and $j_!=j_{{\textup{!`}}}$. The first equality is a particular case of \[f!lisse\]. Because $j^*$ has a left adjoint $j_{{\textup{!`}}}$ which is exact, it preserves (homotopical) injectivity. Let $A,B$ be constructible complexes on $\operatorname{\mathcal{U}},{\mathcal{X}}$ respectively and assume that $B$ is homotopically injective. One has $$\begin{aligned} \operatorname{\mathrm{Rhom}}(j_{{\textup{!`}}}A,B) &=& \operatorname{\mathrm{Hom}}(j_{{\textup{!`}}}A,B) \\ &=& \operatorname{\mathrm{Hom}}(A,j^*B)\textup{ (adjunction)} \\ &=&\operatorname{\mathrm{Rhom}}(A,j^*B)\end{aligned}$$ Taking ${{\mathcal}H}^0$, one obtains that $j^*$ is the right adjoint of $j_{{\textup{!`}}}$ proving the lemma because $j^!=j^*$ is the right adjoint of $j_!$.$\square$ Computation of $Ri_!$ for $i$ a closed immersion ------------------------------------------------ Let $i:{\mathcal{X}}\hookrightarrow{\mathcal{Y}}$ be a closed immersion and ${\operatorname{\mathcal{U}}}={\mathcal{Y}}-{\mathcal{X}}\hookrightarrow {\mathcal{Y}}$ the open immersion of the complement : both are representable. We define the cohomology with support on ${\mathcal{X}}$ for any $F\in{\mathcal{X}}_{{\textup{lis-\'et}}}$ as follows. First, for any $Y\operatorname{\rightarrow}{\mathcal{Y}}$ in ${{\textup{Lisse-Et}}}({\mathcal{Y}})$, the pull-back $Y_{\operatorname{\mathcal{U}}}\operatorname{\rightarrow}{\operatorname{\mathcal{U}}}$ is in ${{\textup{Lisse-Et}}}({\operatorname{\mathcal{U}}})$ and $Y_{\operatorname{\mathcal{U}}}\operatorname{\rightarrow}{\operatorname{\mathcal{U}}}\operatorname{\rightarrow}{\mathcal{Y}}$ is in ${{\textup{Lisse-Et}}}({\mathcal{Y}})$. Then, we define $\underline{H}^0_{\mathcal{X}}(F)$ $$\Gamma(Y,\underline{H}^0_{\mathcal{X}}(F))=\operatorname{\mathrm{ker}}(\Gamma(Y,F)\operatorname{\rightarrow}\Gamma(Y_{\operatorname{\mathcal{U}}},F))$$ and $R\Gamma_{\mathcal{X}}$ is the total derived functor of the left exact functor $\operatorname{{\mathcal}{H}}_{\mathcal{X}}^0$. \[dual-imm-ferme\] One has $\Omega_{\mathcal{X}}=i^*R\Gamma_{\mathcal{X}}(\Omega_{\mathcal{Y}})$. If $i$ is a closed immersion of schemes (or algebraic spaces), one has a canonical (and functorial) isomorphism, simply because $i^*\underline{H}^0_{\mathcal{X}}$ is the right adjoint of $i_*$. If $K$ denotes one of the objects on the two sides of the equality to be proven, one has therefore $\operatorname{\mathcal{E}\it{xt}}^i(K,K)=0$ for $i<0$. Therefore, these isomorphisms glue (use theorem 3.2.2 of [@BBD82] as before).$\square$ \[prop-dual-imm-ferme\] The functor $B\mapsto i^*R\Gamma_{\mathcal{X}}(B)$ is the right adjoint of $i_*$, and therefore coincides with $i^!$. More generally, one has $$\operatorname{\mathcal{R}\it{hom}}(i_*A,B)=i_*\operatorname{\mathcal{R}\it{hom}}(A,i^*R\underline{H}^0_{\mathcal{X}}(B))$$ for all $A\in D({\mathcal{X}}),B\in D({\mathcal{Y}}).$ Moreover, one has one has $i_!=i_*$ and has a right adjoint, the sections with support on ${\mathcal{X}}$. If $A,B$ are sheaves, one has the usual adjunction formula $$\hom(i_*A,B)=i_*\hom(A,i^*\underline{H}^0_{\mathcal{X}}(B)).$$ Because $i_*$ is exact, it’s right adjoint sends homotopically injective complexes to homotopically injective complexes. The derived version follows. One gets therefore $$\begin{array}{ccll} i_!A &=&\operatorname{\mathcal{R}\it{hom}}(i_*\operatorname{\mathcal{R}\it{hom}}(A,\Omega_{\mathcal{X}}),\Omega_{\mathcal{Y}}) \\ &=& i_*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(A,\Omega_{\mathcal{X}}),i^*R\underline{H}^0_{\mathcal{X}}(\Omega_{\mathcal{Y}})) \\ &=& i_*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(A,\Omega_{\mathcal{X}}),\Omega_X)&(\ref{dual-imm-ferme} ) \\ &=&i_*A &(\ref{bidualnonborne}) \end{array}$$ $\square$ Computation of $f^!$ for a universal homeomorphism -------------------------------------------------- By universal homeomorphism we mean a representable, radiciel and surjective morphism. By Zariski’s main theorem, such a morphism is finite. In the schematic situation, we know that such a morphism induces an isomorphism of the étale topos ([@SGA42], VIII.1.1). In particular, $f^*$ is also a right adjoint of $f_*$. Being exact, one gets in this case an identification $f^*=f^!$. In particular, $f^*$ identifies the corresponding dualizing complexes. Exactly as in the proof of \[dual-imm-ferme\], one gets \[dual-uni-homeo\] Let $f:{\mathcal{X}}\operatorname{\rightarrow}{\mathcal{Y}}$ be a universal homeomorphism of stacks. One has $f^*\Omega_{\mathcal{X}}=\Omega_{\mathcal{Y}}$. One gets therefore \[cor-uni-homeo\] Let $f:{\mathcal{X}}\operatorname{\rightarrow}{\mathcal{Y}}$ be a universal homeomorphism of stacks. One has one has $f^!=f^*$ and $Rf_!=Rf_*$. One has $$\begin{array}{ccll} f^!A&=&\operatorname{\mathcal{R}\it{hom}}(f^*\operatorname{\mathcal{R}\it{hom}}(A,\Omega_{\mathcal{Y}}),\Omega_{\mathcal{X}}) \\ &=& \operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(f^*A,f^*\Omega_{\mathcal{Y}}),\Omega_{\mathcal{X}})&(\ref{Rhom-im-inv}) \\ &=& \operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(f^*A,\Omega_{\mathcal{X}}),\Omega_{\mathcal{X}})&(\ref{dual-uni-homeo} ) \\ &=&f^*A &(\ref{bidualnonborne}). \end{array}$$ The last formula follows by adjunction. $\square$ Computation of $Rf_!$ via hypercovers ------------------------------------- Let $Y$ be an $S$–scheme of finite type and $f:{\mathcal{X}}\rightarrow Y$ a morphism of finite type from an algebraic stack ${\mathcal{X}}$. Let $X_\bullet \rightarrow {\mathcal{X}}$ be a smooth hypercover by algebraic spaces, and for each $n$ let $d_n$ denote the locally constant function on $X_n$ which is the relative dimension over ${\mathcal{X}}$. By the construction, the restriction of the dualizing complex $\Omega _{{\mathcal{X}}}$ to each $X_{n, {\textup{\'et}}}$ is canonically isomorphic to the dualizing complex $K_{X_n}= \Omega _{{\mathcal{X}}_n}\langle -d_n\rangle$ of $X_n$. Let $K _{X_\bullet }$ denote the restriction of $\Omega _{{\mathcal{X}}}$ to $X_{\bullet , {\textup{\'et}}}$. Let $L\in D_c^-({\mathcal{X}})$, and let $L|_{X_\bullet }$ denote the restriction of $L$ to $X_{\bullet , {\textup{\'et}}}$. Then $D_{{\mathcal{X}}}(L)|_{X_\bullet }$ is isomorphic to $D_{X_\bullet }(L|_{X_\bullet }):= \operatorname{\mathcal{R}\it{hom}}_{X_{\bullet , {\textup{\'et}}}}(L|_{X_\bullet }, K_{X_\bullet })$. In particular, the restriction of $Rf_!L$ to $Y_{{\textup{\'et}}}$ is canonically isomorphic to $$\operatorname{\mathcal{R}\it{hom}}_{Y_{{\textup{\'et}}}}(Rf_{\bullet *}D_{X_\bullet }(L|_{X_\bullet }), K_{Y})\in D_c(Y_{{\textup{\'et}}}),$$ where $f_{\bullet }:X_{{\textup{\'et}}}\rightarrow Y_{{\textup{\'et}}}$ denotes the morphism of topos induced by $f$. Let $Y_{\bullet , {\textup{\'et}}}$ denote the simplicial topos obtained by viewing $Y$ as a constant simplicial scheme. Let $\epsilon :Y_{\bullet, {\textup{\'et}}}\rightarrow Y_{{\textup{\'et}}}$ denote the canonical morphism of topos, and let $\tilde f:X_{\bullet , {\textup{\'et}}}\rightarrow Y_{\bullet , {\textup{\'et}}}$ be the morphism of topos induced by $f$. We have $f_\bullet = \epsilon \circ \tilde f$. As in [@Ols05], 2.7, it follows that there is a canonical spectral sequence $$E_1^{pq} = R^qf_{p*}D_{X_p}(L|_{X_p})\implies R^{p+q}f_{\bullet *}D_{X_\bullet }(L_{X_\bullet }).$$ On the other hand, we have $$R^qf_{p*}D_{X_p}(L|_{X_p}) = R^qf_{p*}\operatorname{\mathcal{R}\it{hom}}(L|_{X_p}, \Omega _{X_p}\langle -d_p\rangle ) \simeq {\mathcal}H^q(D_Y(Rf_{p!}(L|_{X_p}\langle d_p\rangle )),$$ where the second isomorphism is by biduality \[Dinvolutif\]. Combining all this we obtain \[5.16\] There is a canonical spectral sequence $$\label{5.16.1} E_1^{pq} = {\mathcal}H^q(D_{Y_{\text{\rm et}}}(Rf_{p!}L|_{X_p}\langle d_p\rangle ))\implies {\mathcal}H^{p+q}(D_{Y_{\text{\rm et}}}(Rf_!L|_{Y_{\text{\rm et}}})).$$ Let $k$ be an algebraically closed field and $G$ a finite group. We can then compute $H^*_c(BG, \Lambda )$ as follows. We first compute $\operatorname{\mathcal{R}\it{hom}}(R\Gamma _!(BG, \Lambda ), \Lambda )$. Let $\text{Spec}(k)\rightarrow BG$ be the surjection corresponding to the trivial $G$–torsor, and let $X_\bullet \rightarrow BG$ be the $0$–coskeleton. Note that each $X_n$ isomorphic to $G^n$ and in particular is a discrete collection of points. Therefore $Rf_{!p}\Lambda \simeq \text{Hom}(G^n, \Lambda )$. From this it follows that $\operatorname{\mathcal{R}\it{hom}}(R\Gamma _!(BG, \Lambda ), \Lambda )$ is represented by the standard cochain complex computing the group cohomology of $\Lambda $, and hence $R\Gamma _!(BG, \Lambda )$ is the dual of this complex. In particular, this can be nonzero in infinitely many negative degrees. For example if $G = \mathbb{Z}/\ell $ for some prime $\ell $ and $\Lambda = \mathbb{Z}/\ell $ since in this case the group cohomology $H^i(G, \mathbb{Z}/\ell )\simeq \mathbb{Z}/\ell $ for all $i\geq 0$. Let $k$ be an algebraically closed field and $P$ the affine line $\mathbb{A}^1$ with the origin doubled. By definition $P$ is equal to two copies of $\mathbb{A}^1$ glued along $\mathbb{G}_m$ via the standard inclusions $\mathbb{G}_m\subset \mathbb{A}^1$. We can then compute $R\Gamma _!(P, \Lambda )$ as follows. Let $j_i:\mathbb{A}^1\hookrightarrow P$ ($i=1,2$) be the two open immersions, and let $h:\mathbb{G}_m\hookrightarrow P$ be the inclusion of the overlaps. We then have an exact sequence $$0\rightarrow h_!\Lambda \rightarrow j_{1!}\Lambda \oplus j_{2!}\Lambda \rightarrow \Lambda \rightarrow 0.$$ From this we obtain a long exact sequence $$\cdots \rightarrow H^i_c(\mathbb{G}_m, \Lambda )\rightarrow H^i_c(\mathbb{A}^1, \Lambda )\oplus H^i_c(\mathbb{A}^1, \Lambda )\rightarrow H^i_c(P, \Lambda )\rightarrow \cdots.$$ From this sequence one deduces that $H^0_c(P, \Lambda )\simeq \Lambda $, $H^2_c(P, \Lambda )\simeq \Lambda (1)$, and all other cohomology groups vanish. In particular, the cohomology of $P$ is isomorphic to the cohomology of $\mathbb{P}^1$. Purity and the fundamental distinguished triangle ------------------------------------------------- We consider the usual situation of a closed immersion $i:{\mathcal{X}}\operatorname{\rightarrow}{\mathcal{Y}}$ of stacks, the open immersion of the complement of ${\mathcal{Y}}$ being $j:\operatorname{\mathcal{U}}={\mathcal{Y}}-{\mathcal{X}}\operatorname{\rightarrow}{\mathcal{Y}}$. For any (complex) of sheaves $A$ on ${\mathcal{Y}}$, one has the exact sequence $$0\operatorname{\rightarrow}j_{!}j^*A\operatorname{\rightarrow}A \operatorname{\rightarrow}i_*i^*A\operatorname{\rightarrow}0.$$ Therefore, for any $A\in D_c({\mathcal{Y}})$, one has the distinguished triangle (\[!=!!\]) $$\label{T1} j_{!}j^*A\operatorname{\rightarrow}A \operatorname{\rightarrow}i_*i^*A$$ which by duality gives the distinguished triangle $$\label{T2} i_{*}i^!A\operatorname{\rightarrow}A \operatorname{\rightarrow}j_*j^*A.$$ Recall (\[prop-dual-imm-ferme\]) the formula $i^!=R\underline{H}^0_{\mathcal{X}}$. The usual purity theorem for $S$-schemes gives \[purity\] Assume moreover that $i$ is a closed immersion of smooth $S$-stacks of codimension $c$ (a locally constant function on ${\mathcal{Y}}$). Then, one has $i^!A=i^*A(-c)[-2c]$. Let $d$ be the relative dimension of ${\mathcal{Y}}\operatorname{\rightarrow}S$ and $s$ the dimension of $S$. The relative dimension of ${\mathcal{X}}$ is (the restriction to ${\mathcal{X}}$ of) $d-c$. By \[f!lisse\], one has $$\Omega_{\mathcal{Y}}=\Lambda(d+s)[2d+2s]\textup{ and }\Omega_{\mathcal{X}}=\Lambda(d-c+s)[2d-2c-2e].$$ The identity $i^!\Omega_{\mathcal{Y}}=\Omega_{\mathcal{X}}$ gives therefore the formula $$\label{i!c} i^!\Lambda=\Lambda(-c)[-2c].$$ By \[dual\], one has $$i_*\operatorname{\mathcal{R}\it{hom}}(i^!\Lambda,i^!A)=\operatorname{\mathcal{R}\it{hom}}(i_!i^!\Lambda,A)$$ which by adjunction for $i_*$ gives a map $$i^*\operatorname{\mathcal{R}\it{hom}}(i_!i^!\Lambda,A)\operatorname{\rightarrow}\operatorname{\mathcal{R}\it{hom}}(i^!\Lambda,i^!A).$$ But the adjunction map (for $i_!$) $i_!i^!\Lambda\operatorname{\rightarrow}\Lambda$ dualizes to $$\operatorname{\mathcal{R}\it{hom}}(\Lambda,A)\operatorname{\rightarrow}\operatorname{\mathcal{R}\it{hom}}(i_!i^!\Lambda,A)$$ which gives by composition a morphism $$i^*A=i^*\operatorname{\mathcal{R}\it{hom}}(\Lambda,A)\operatorname{\rightarrow}\operatorname{\mathcal{R}\it{hom}}(i^!\Lambda,i^!A)=i^!A(c)[2c]$$ which is the usual morphism for closed immersion of schemes. This morphism is compatible with the duality in an obvious sense. The usual purity theorem gives then the proposition, at least for $A\in D^+_c({\mathcal{Y}})$. By duality, one gets the proposition for $A\in D^-_c({\mathcal{Y}})$, and therefore for $A\in D_c({\mathcal{Y}})$ using the distinguished triangle $\tau_{>0}A\operatorname{\rightarrow}A\operatorname{\rightarrow}\tau_{\leq 0}A$.$\square$ Base change =========== We start with a cartesian diagram of stacks $$\xymatrix{{\mathcal{X}}'\ar[r]^\pi\ar[d]_\phi\ar@{}[rd]|\Box&{\mathcal{X}}\ar[d]^f\\{\mathcal{Y}}'\ar[r]^p&{\mathcal{Y}}}$$ and we would like to prove a natural base change isomorphism $$\label{bc!} p^*Rf_!=R\phi_!\pi^*$$ of functors $D_c({\mathcal{X}})\operatorname{\rightarrow}D_c({\mathcal{Y}}')$. Though technically not needed, before proving the general base change Theorem we consider first some simpler cases where one can prove a dual version: $$\label{bc} p^!Rf_*=R\phi_*\pi^!.$$ Smooth base change {#5.1} ------------------ In this subsection we prove the base change isomorphism in the case when $p$ (and hence also $\pi $) is smooth. Because the relative dimension of $p$ and $\pi$ are the same, by \[f!lisse\], one reduces the formula \[bc\] to $$p^*Rf_*=R\phi_*\pi^*.$$ By adjunction, one has a morphism $p^*Rf_*\operatorname{\rightarrow}R\phi_*\pi^*$ which we claim is an isomorphism (for complexes bounded below this follows immediately from the smooth base change theorem). To prove that this map is an isomorphism, we consider first the case when ${\mathcal}Y'$ is algebraic space and show that our morphism restricts to an isomorphism on ${\mathcal}Y'_{{\textup{\'et}}}$. Since $p$ is representable ${\mathcal}X'$ represents a sheaf on ${\mathcal{X}}_{{\textup{lis-\'et}}}$. Let ${\mathcal{X}}_{{\textup{lis-\'et}}|{\mathcal{X}}'}$ denote the localized topos, $w:{\mathcal{X}}_{{\textup{lis-\'et}}|{\mathcal{X}}'}\rightarrow {\mathcal{Y}}_{\textup{\'et}}$ the projection, and let $A\in \operatorname{{D}}_c({\mathcal{X}})$ be a complex. Let $X\rightarrow {\mathcal{X}}$ be a smooth surjection with $X$ a scheme, and let $X_\bullet \rightarrow {\mathcal{X}}$ denote the associated simplicial space. Let $X'_\bullet $ denote the base change of $X_\bullet $ to ${\mathcal}Y'$. Then $X_\bullet '$ defines a hypercover of the initial object in the topos ${\mathcal{X}}_{{\textup{lis-\'et}}|{\mathcal{X}}'}$ and hence we have an equivalence of topos ${\mathcal{X}}_{{\textup{lis-\'et}}, {\mathcal{X}}'}\simeq {\mathcal{X}}_{{\textup{lis-\'et}}, X_{\bullet }'}$. Let $w_\bullet :{\mathcal{X}}_{{\textup{lis-\'et}}|X_{\bullet '}}\rightarrow {\mathcal}Y'_{\textup{\'et}}$ be the projection. Since the restriction functor from ${\mathcal{X}}_{{\textup{lis-\'et}}}$ to ${\mathcal{X}}_{{\textup{lis-\'et}}|X_\bullet '}$ takes homotopically injective complexes to homotopically injective complexes (since it has an exact left adjoint), $p^*Rf_*A|_{{\mathcal}Y'_{\textup{\'et}}}$ is equal to $Rw_{\bullet *}(A|_{{\mathcal{X}}_{{\textup{lis-\'et}}|X_\bullet '}})$. On the other hand, $w_\bullet $ factors as $$\begin{CD} {\mathcal}X_{{\textup{lis-\'et}}|X'_{\bullet }}@>\alpha >> X'_{\bullet , {\textup{\'et}}} @>\phi _\bullet >> {\mathcal}Y'_{{\textup{\'et}}}, \end{CD}$$ where $\phi _{\bullet }:X'_{\bullet }\rightarrow {\mathcal}Y'_{\textup{\'et}}$ is the projection. Since $\alpha _*$ is exact, we find that $Rw_{\bullet *}(A|_{{\mathcal{X}}_{{\textup{lis-\'et}}|X_\bullet '}})$ is isomorphic to $R\phi _{\bullet *}(A|_{X_{\bullet , {\textup{\'et}}}'}).$ Similarly, factoring the morphism of topos ${\mathcal}X_{{\textup{lis-\'et}}}'\simeq {\mathcal}X'_{{\textup{lis-\'et}}|X'_{\bullet }}\rightarrow {\mathcal}Y'_{\textup{\'et}}$ as $$\begin{CD} {\mathcal}X'_{{\textup{lis-\'et}}|X'_\bullet }@>>> X'_{\bullet, {\textup{\'et}}}@>\phi _\bullet >> {\mathcal}Y'_{\textup{\'et}}\end{CD}$$ we see that $R\phi_*(A|_{{\mathcal}X'_{\textup{lis-\'et}}})$ is isomorphic to $R\phi _{\bullet *}(A|_{X'_{\bullet {\textup{\'et}}}})$. We leave to the reader that the resulting isomorphism $p^*Rf_*(A)|_{{\mathcal}Y'_{\textup{\'et}}}\rightarrow R\phi_*\pi ^*A|_{{\mathcal}Y'_{\textup{\'et}}}$ agrees with the morphism defined above. Thus this proves the case when $p$ is representable. For the general case, let $Y'\rightarrow {\mathcal}Y'$ denote a smooth surjection with $Y'$ a scheme, so we have a commutative diagram $$\begin{CD} X'@>\sigma >> {\mathcal{X}}'@>\pi >> {\mathcal{X}}\\ @VgVV @VV\phi V @VVfV \\ Y'@>q>> {\mathcal{Y}}'@>p>> {\mathcal{Y}}\end{CD}$$ with cartesian squares. Let $A\in D_c({\mathcal{X}})$. To prove that the morphism $p^*Rf_*A\rightarrow R\phi_*\pi ^*A$ is an isomorphism, it suffices to show that the induced morphism $q^*p^*Rf_*A\rightarrow q^*R\phi_*\pi ^*A$ is an isomorphism. By the representable case, $q^*R\phi_*\pi ^*A\simeq Rg_*\sigma ^*\pi ^*A$ so it suffices to prove that the composite $$q^*p^*Rf_*A\rightarrow Rg_*\sigma ^*\pi ^*A$$ is an isomorphism. By the construction, this map is equal to the base change morphism for the diagram $$\begin{CD} X'@>\pi \circ \sigma >> {\mathcal{X}}\\ @VgVV @VVfV \\ Y'@>p\circ q>> {\mathcal}Y \end{CD}$$ and hence it is an isomorphism by the representable case. $\square$ Computation of $Rf_*$ for proper representable morphisms -------------------------------------------------------- \[f!=f\*propre\] Let $f:{\mathcal{X}}\rightarrow {\mathcal{Y}}$ be a proper representable morphism of $S$–stacks. Then the functor $Rf_!:D_c^-({\mathcal{X}})\rightarrow D_c^-({\mathcal{Y}})$ is canonically isomorphic to $Rf_*:D_c^-({\mathcal{X}})\rightarrow D_c^-({\mathcal{Y}})$. The key point is the following lemma. There is a canonical morphism $Rf_*\Omega _{\mathcal{X}}\rightarrow \Omega _{\mathcal{Y}}$. Using \[D1\] and smooth base change, it suffices to construct a functorial morphism in the case of schemes, and to show that $\operatorname{\mathcal{E}\it{xt}}^i(Rf_*\Omega _{{\mathcal{X}}}, \Omega _{\mathcal{Y}}) = 0$ for $i<0$. Now if ${\mathcal{X}}$ and ${\mathcal{Y}}$ are schemes, we have $\Omega _{\mathcal{X}}= f^!\Omega _{{\mathcal{Y}}}$ so we obtain by adjunction and the fact that $Rf_! = Rf_*$ a morphism $Rf_*\Omega _{\mathcal{X}}\rightarrow \Omega _{\mathcal{Y}}$. For the computation of $\operatorname{\mathcal{E}\it{xt}}$’s note that $$\operatorname{\mathcal{R}\it{hom}}(Rf_*\Omega_{\mathcal{X}},\Omega_{\mathcal{Y}})=\operatorname{\mathcal{R}\it{hom}}(\Omega_{\mathcal{X}},f^!\Omega_{\mathcal{Y}})=\Lambda.$$ $\square$ We define a map $Rf_*\circ D_{\mathcal{X}}\operatorname{\rightarrow}D_{\mathcal{Y}}\circ f_*$ by taking the composite $$Rf_*\operatorname{\mathcal{R}\it{hom}}(-,\Omega_{\mathcal{X}})\operatorname{\rightarrow}\operatorname{\mathcal{R}\it{hom}}(Rf_*(-),Rf_*\Omega_{\mathcal{X}})\operatorname{\rightarrow}\operatorname{\mathcal{R}\it{hom}}(Rf_*(-), \Omega _{\mathcal{Y}}).$$ To verify that this map is an isomorphism we may work locally on ${\mathcal{Y}}$. This reduces the proof to the case when ${\mathcal{X}}$ and ${\mathcal{Y}}$ are algebraic spaces in which case the result is standard. $\square$ Base change by an immersion {#5.3} --------------------------- In this subsection we consider the case when $p$ is an immersion. By replacing ${\mathcal{Y}}$ by a suitable open substack, one is reduced to the case when $p$ is a closed immersion. Then,  \[bc!\] follows from the projection formula \[corkunneth\] as in  [@Fre-Kie], p.81. Let us recall the argument. Let $A\in D_c({\mathcal{X}})$. Because $p$ is a closed immersion, one has $p^*p_*=\operatorname{Id}$. One has (projection formula \[corkunneth\] for $p$) $$p_*p^*Rf_!A=p_*\Lambda\operatorname{\stackrel{\mathbf L}{\otimes}}Rf_!A.$$ One has then $$Rf_!A\operatorname{\stackrel{\mathbf L}{\otimes}}p_*\Lambda=Rf_!(A\operatorname{\stackrel{\mathbf L}{\otimes}}f^*p_*\Lambda)$$ (projection formula \[corkunneth\] for $f$). But, we have trivially the base change for $p$, namely $$f^*p_*=\pi_*\phi^*.$$ Therefore, one gets$$\begin{aligned} Rf_!(A\operatorname{\stackrel{\mathbf L}{\otimes}}f^*p_*\Lambda) &=& Rf_!(A\operatorname{\stackrel{\mathbf L}{\otimes}}\pi_*\phi^*\Lambda) \\ &=& Rf_!\pi_*(\pi^*A\operatorname{\stackrel{\mathbf L}{\otimes}}\phi^*\Lambda)\textup{ projection for }\pi \\ &=& p_*\phi_!\pi^*A\textup{ because }\pi_*=\pi_!\ (\ref{f!=f*propre}).\end{aligned}$$ Applying $p^*$ gives the base change isomorphism. One can prove \[bc\], [at least for $A$ bounded below]{}, more directly as follows. Start with $A$ on ${\mathcal{X}}$ an injective complex. Because $R^0f_*A_i$ is flasque, it is $\Gamma_{{\mathcal{Y}}'}$-acyclic. Then, $p^!Rf_*A$ can be computed using the complex ${{\mathcal}H}^0_{{\mathcal{Y}}'}(R^0f_*A_i)$. On the other hand, $\pi^!A$ can be computed by the complex ${{\mathcal}H}^0_{{\mathcal{X}}'}(A_i)$ which is a flasque complex (formal, or [@SGA42], V.4.11). Therefore, the direct image by $\phi$ is just $R^0\phi_*{{\mathcal}H}^0_{{\mathcal{X}}'}(A_i)$. One is reduced to the formula $$R^0\phi_*{{\mathcal}H}^0_{{\mathcal{X}}'}={{\mathcal}H}^0_{{\mathcal{Y}}'}(R^0f_*).$$ Base change by a universal homeomorphism {#5.4} ---------------------------------------- If $p$ is a universal homeomorphism, then $p^! = p^*$ and $\pi ^! = \pi ^*$. Thus in this case \[bc\] is equivalent to an isomorphism $p^*Rf_*\rightarrow R\phi _*\pi ^*$. We define such a morphism by taking the usual base change morphism (adjunction). Let $A\in \operatorname{{D}}_c({\mathcal}X)$. Using a hypercover of ${\mathcal{X}}$ as in \[5.1\], one sees that to prove that the map $p^*Rf_*A\rightarrow R\phi _*\pi ^*A$ is an isomorphism it suffices to consider the case when ${\mathcal{X}}$ is a scheme. Furthermore, by the smooth base change formula already shown, it suffices to prove that this map is an isomorphism after making a smooth base change $Y\rightarrow {\mathcal{Y}}$. We may therefore assume that ${\mathcal{Y}}$ is also a scheme in which case the result follows from the classical corresponding result for étale topology (see [@SGA1], IV.4.10). Base change morphism in general {#5.5} ------------------------------- Before defining the base change morphism we need a general construction of strictly simplicial schemes and algebraic spaces. Fix an algebraic stack ${\mathcal}X$. In the following construction all schemes and morphisms are assumed over ${\mathcal}X$ (so in particular products are taken over ${\mathcal}X$). Let $X_\bullet $ be a strictly simplicial scheme, $[n]\in \Delta ^+ $ an object, and $a:V\rightarrow X_n$ a surjective morphism. We then construct a strictly simplicial scheme $M(X_\bullet , a)$ (sometimes written $M_{{\mathcal}X}(X_\bullet , a)$ if we want to make clear the reference to ${\mathcal}X$) with a morphism $M(X_\bullet , a)\rightarrow X_\bullet $ such that the following hold: 1. For $i<n$ the morphism $M(X_\bullet , a)_i\rightarrow X_i$ is an isomorphism. 2. $M(X_\bullet , a)_n$ is equal to $V$ with the projection to $X_n$ given by $a$. The construction of $M(X_\bullet , a)$ is a standard application of the skeleton and coskeleton functors ([@SGA42], exp. Vbis). Let us review some of this because the standard references deal only with simplicial spaces whereas we consider strictly simplicial spaces. To construct $M(X_\bullet , a)$, let $\Delta ^+ _n\subset \Delta ^+ $ denote the full subcategory whose objects are the finite sets with cardinality $\leq n$. Denote by $\text{Sch}^{\Delta ^{+{\textup{opp}}} _n}$ the category of functors from $\Delta ^{+{\textup{opp}}} _n$ to schemes (so $\text{Sch}^{\Delta ^{+{\textup{opp}}}}$ is the category of strictly simplicial schemes). Restriction from $\Delta ^{+{\textup{opp}}}$ to $\Delta ^{+{\textup{opp}}}_n$ defines a functor (the *$n$-skeleton functor*) $$\text{sq}_n:\text{Sch}^{\Delta ^{+{\textup{opp}}}}\rightarrow \text{Sch}^{\Delta ^{+{\textup{opp}}}_n}$$ which has a right adjoint $$\text{cosq}_n:\text{Sch}^{\Delta ^{+{\textup{opp}}}_n}\rightarrow \text{Sch}^{\Delta ^{+{\textup{opp}}}}$$ called the *$n$-th coskeleton functor*. For $X_\bullet \in \text{Sch}^{\Delta ^{+{\textup{opp}}}_n}$, the coskeleton $\text{cosq}_nX$ in degree $i$ is equal to $$\label{cosqdef} (\text{cosq}_nX)_i = \varprojlim _{\substack{[k]\rightarrow [i]\\ k\leq n}}X_k,$$ where the limit is taken over the category of morphisms $[k]\rightarrow [i]$ in $\Delta ^+$ with $k\leq n$. Note in particular that for $i\leq n$ we have $(\text{cosq}_nX)_i = X_i$ since the category of morphisms $[k]\rightarrow [i]$ has an initial object $\text{id}:[i]\rightarrow [i]$. \[easylem\] For any $X_\bullet \in \text{\rm Sch}^{\Delta ^{+{\textup{opp}}}_n}$ and $i>n$ the morphism $$(\text{\rm cosq}_nX)_i\rightarrow (\text{\rm cosq}_{i-1}\text{\rm sq}_{i-1}\text{\rm cosq}_nX)_i$$ is an isomorphism. Using the formula \[cosqdef\] the morphism can be identified with the natural map $$\varprojlim _{\substack{[k]\rightarrow [i]\\ k\leq n}}X_k\rightarrow \varprojlim _{\substack{[k]\rightarrow [i]\\ k\leq i-1}}(\varprojlim _{\substack{[w]\rightarrow [k]\\ w\leq n}}X_w)$$ which is clearly an isomorphism. $\square$ The functors $\text{\rm sq}_n$ and $\text{\rm cosq}_n$ commute with fiber products. The functor $\text{sq}_n$ commutes with fiber products by construction, and the functor $\text{cosq}_n$ commutes with fiber products by adjunction. $\square$ To construct $M(X_\bullet , a)$, we first construct an object $M'(X_\bullet , a)\in \text{Sch}^{\Delta ^{+{\textup{opp}}}_n}$. The restriction of $M'(X_\bullet , a)$ to $\Delta ^{+{\textup{opp}}} _{n-1}$ will be equal to $\text{sq}_{n-1}X$, and $M'(X_\bullet , a)_n$ is defined to be $V$. For $0\leq j\leq n$ define $\delta _j:M'(X_\bullet , a)_n\rightarrow M'(X_\bullet , a)_{n-1} = X_{n-1}$ to be the composite $$\begin{CD} V@>a>> X_n@>\delta _{X, j}>> X_{n-1}, \end{CD}$$ where $\delta _{j, X}$ denotes the map obtained from the strictly simplicial structure on $X_\bullet $. There is an obvious morphism $$M'(X_\bullet ,a)\operatorname{\rightarrow}\text{sq}_n(X_\bullet )\text{ inducing }\text{cosq}_nM'(X_\bullet , a))\operatorname{\rightarrow}{\text{cosq}_n\text{sq}_nX_\bullet }.$$ We then define $$M(X_\bullet , a):= (\text{cosq}_nM'(X_\bullet , a))\times _{\text{cosq}_n\text{sq}_nX_\bullet }X_\bullet ,$$ where the map $X_\bullet \rightarrow \text{cosq}_n\text{sq}_nX_\bullet $ is the adjunction morphism. The map $M(X_\bullet , a)\rightarrow X_\bullet $ is defined to be the projection. The properties (i) and (ii) follow immediately from the construction. Let ${\mathcal}X$ be an algebraic stack and $X_\bullet \rightarrow {\mathcal}X$ a hypercover by schemes. Let $n$ be a natural number and $a:V\rightarrow X_n$ a surjection. Then $M_{{\mathcal}X} (X_\bullet , a)\rightarrow {\mathcal}X$ is also a hypercover. If $X_\bullet $ is a smooth hypercover and $a$ is smooth and surjective, then $M_{{\mathcal}X}(X_\bullet , a)$ is also a smooth hypercover. By definition of a hypercover, we must verify that for all $i$ the map $$M(X_\bullet , a)_i\rightarrow (\text{cosq}_{i-1}\text{sq}_{i-1}M(X_\bullet , a))_i$$ is surjective. Note that this is immediate for $i\leq n$. For $i>n$ we compute $$\begin{aligned} (\text{cosq}_{i-1}\text{sq}_{i-1}M(X_\bullet , a))_i &\simeq & (\text{cosq}_{i-1}\text{sq}_{i-1}(\text{cosq}_nM'(X_\bullet , a)\times _{\text{cosq}_n\text{sq}_nX_\bullet }X_\bullet ))_i\\ &\simeq & (\text{cosq}_{i-1}\text{sq}_{i-1}(\text{cosq}_nM'(X_\bullet , a)))_i\times _{(\text{cosq}_i\text{sq}_{i-1}\text{cosq}_n\text{sq}_nX_\bullet )_i}(\text{cosq}_{i-1}\text{sq}_{i-1}X_\bullet )_i\\ &\simeq & (\text{cosq}_nM'(X_\bullet , a))_i\times _{(\text{cosq}_n\text{sq}_nX_\bullet )_i}(\text{cosq}_{i-1}\text{sq}_{i-1}X_\bullet )_i.\end{aligned}$$ Here the second isomorphism is because $\text{sq}_n$ and $\text{cosq}_n$ commute with products, and the third isomorphism is by \[easylem\]. Hence it suffices to show that the natural map $$X_i\rightarrow (\text{cosq}_{i-1}\text{sq}_{i-1}X_\bullet )_i$$ is surjective, which is true since $X_\bullet $ is a hypercover. This also proves that if $X_\bullet $ is a smooth hypercover and $a$ is smooth, then $M_{{\mathcal}X}(X_\bullet , a)$ is a smooth hypercover. $\square$ The construction of $M_{{\mathcal}X}(X_\bullet , a)$ is functorial. Precisely, let $f:{\mathcal}X\rightarrow {\mathcal}Y$ be a morphism of algebraic stacks, $X_\bullet \rightarrow {\mathcal}X$ a strictly simplicial scheme over ${\mathcal}X$, $Y_\bullet \rightarrow {\mathcal}Y$ a strictly simplicial scheme over ${\mathcal}Y$, and $f_\bullet :X_\bullet \rightarrow Y_\bullet $ a morphism over $f$. Then for any commutative diagram of schemes $$\begin{CD} V@>\tilde f>> W\\ @VaVV @VVbV \\ X_n@>f_n>> Y_n \end{CD}$$ there is an induced morphism of strictly simplicial schemes $M_{{\mathcal}X}(X_\bullet , a)\rightarrow M_{{\mathcal}Y}(Y_\bullet , b)$ over $f_\bullet $. \[5.4.4\] Let $f:{\mathcal}X\rightarrow {\mathcal}Y$ be a morphism of finite type between algebraic $S$-stacks locally of finite type. Then there exists smooth hypercovers $p:X_\bullet \rightarrow {\mathcal}X$ and $q:Y_\bullet \rightarrow {\mathcal}Y$ by schemes and a commutative diagram $$\label{simpdiagram} \begin{CD} X_\bullet @>f_\bullet >> Y_\bullet \\ @VpVV @VVqV \\ {\mathcal}X@>f>> {\mathcal}Y, \end{CD}$$ where each morphism $f_n:X_n\rightarrow Y_n$ is a closed immersion. We construct inductively hypercovers $X_\bullet ^{(n)}\rightarrow {\mathcal}X$ and $Y_\bullet ^{(n)}\rightarrow {\mathcal}Y$ and a commutative diagram $$\begin{CD} X_\bullet ^{(n)}@>>> Y_\bullet ^{(n)}\\ @VVV @VVV \\ {\mathcal}X@>>> {\mathcal}Y \end{CD}$$ together with a commutative diagram $$\begin{CD} X_\bullet ^{(n)}@>>> Y_\bullet ^{(n)}\\ @VVV @VVV \\ X_\bullet ^{(n-1)}@>>> Y_\bullet ^{(n-1)} \end{CD}$$ over $f$. We further arrange so that the following hold: 1. For $i<n$ the maps $X^{(n)}_i\rightarrow X^{(n-1)}_i$ and $Y^{(n)}_i\rightarrow Y^{(n-1)}_i$ are isomorphisms. 2. For $i\leq n$ the maps $X^{(n)}_i\rightarrow Y^{(n)}_i$ are closed immersions. This suffices for we can then take $X_\bullet = \varprojlim X^{(n)}$ and $Y_\bullet = \varprojlim Y^{(n)}$. For the base case $n=0$, choose any $2$–commutative diagram $$\begin{CD} X=\sqcup X_i@>\tilde f=\sqcup\tilde f_i>> Y=\sqcup Y_i\\ @V{p=\sqcup p_i}VV @VV{q=\sqcup q_i}V \\ {\mathcal}X@>>> {\mathcal}Y, \end{CD}$$ with $p_i$ and $q_i$ smooth, surjective, and of finite type, and $X_i$ and $Y_i$ affine schemes. Then $\tilde f_i$ are also of finite type, so there exists a closed immersion $X_i\hookrightarrow \mathbb{A}^{r_i}_{Y_i}$ for some integer $r$ over $X_i\rightarrow Y_i$. Replacing $Y_i$ by $\mathbb{A}^{r_i}_{Y_i}$ we may assume that $\tilde f$ is a closed immersions. We then obtain $X_\bullet ^{(0)}\rightarrow Y_\bullet ^{(0)}$ by taking the coskeletons of $p$ and $q$. Now assume that $X_\bullet ^{(n-1)}\rightarrow Y_\bullet ^{(n-1)}$ has been constructed. Choose a commutative diagram $$\begin{CD} V@>j>> W\\ @VaVV @VVbV \\ X_n^{(n-1)}@>>> Y_n^{(n-1)}, \end{CD}$$ with $a$ and $b$ smooth and surjective, and $j$ a closed immersion. Then define $X_\bullet ^{(n)}\rightarrow Y_\bullet ^{(n)}$ to be $$M_{{\mathcal}X}(X^{(n-1)}_\bullet , a)\rightarrow M_{{\mathcal}Y}(Y_\bullet ^{(n-1)}, b).$$ $\square$ \[filteringremark\] The same argument used in the proof shows that for any commutative diagram $$\begin{CD} X_{\bullet }@>f_\bullet >> Y_\bullet \\ @VpVV @VVqV \\ {\mathcal}X@>f>> {\mathcal}Y, \end{CD}$$ where $p$ and $q$ are smooth hypercovers, there exists a morphism of simplicial schemes $g:\widetilde X_\bullet \rightarrow \widetilde Y_\bullet $ over $f_\bullet $ with each $g_n:\widetilde X_n\rightarrow \widetilde Y_n$ an immersion such that $\widetilde X_\bullet $ (resp. $\widetilde Y_\bullet $) is a hypercover of ${\mathcal}X$ (resp. ${\mathcal}Y$). In other words, the category of diagrams \[simpdiagram\] is connected. Let $f:{\mathcal}X\rightarrow {\mathcal}Y$ be a morphism of algebraic stacks over $S$. For $F\in \operatorname{{D}}_c^-({\mathcal}X)$ we can compute $Rf_!F$ as follows. Let $Y_\bullet \rightarrow {\mathcal}Y$ be a smooth hypercover, and let $\pi :{\mathcal}X_{Y_\bullet }\rightarrow {\mathcal}X$ be the base change of ${\mathcal}X$ to $Y_\bullet $. Let $f_\bullet :{\mathcal}X_{Y_\bullet }\rightarrow Y_\bullet $ be the projection. Let $\omega _{{\mathcal}X_{Y_\bullet }}$ denote the pullback of the dualizing sheaf $\Omega _{{\mathcal}X}$ to ${\mathcal}X_{Y_\bullet }$, and let $D_{{\mathcal}X_{Y_\bullet }}$ denote the functor $\operatorname{\mathcal{R}\it{hom}}(-, \omega _{{\mathcal}X_{Y_\bullet }})$. Similarly let $\omega _{Y_\bullet }$ denote the pullback of $\Omega _{{\mathcal}Y}$ to $Y_\bullet $, and let $D_{Y_\bullet }$ denote $\operatorname{\mathcal{R}\it{hom}}(-, \omega _{Y_\bullet })$. If $d_n$ (resp. $d_n'$) denotes the relative dimension of $Y_n$ over ${\mathcal{Y}}$ (resp. $Y_n'$ over ${\mathcal{Y}}'$), then $d_n$ (resp. $d_n'$) is also equal to the relative dimension of ${\mathcal{X}}_{Y_n}$ over ${\mathcal{X}}$ (resp. ${\mathcal{X}}'_{Y_n'}$ over ${\mathcal{X}}'$). From \[f!lisse\] it follows that the restriction of $\omega _{{\mathcal{X}}_{Y_\bullet }}$ to ${\mathcal{X}}_{Y_n}$ is canonically isomorphic to $\Omega _{{\mathcal{X}}_{Y_n}}\langle -d_n\rangle $. Similarly the restriction of $\omega _{Y_\bullet }$ to $Y_n$ is canonically isomorphic to $\Omega _{Y_n}\langle -d_n\rangle $. Note that this combined with \[Dinvolutif\] shows that $D_{Y_\bullet }\circ D_{Y_\bullet } = \text{id}$ (resp. $D_{{\mathcal{X}}_{Y_\bullet }}\circ D_{{\mathcal{X}}_{Y_\bullet }} = \text{id}$) on the category $\operatorname{{D}}_c(Y_\bullet )$ (resp. $\operatorname{{D}}_c({\mathcal{X}}_{Y_\bullet })$). For $F\in D_c({\mathcal{X}})$, we can then consider $$D_{Y_\bullet }Rf_{\bullet *}D_{{\mathcal}X_{Y_\bullet }}(\pi ^*F)\in \operatorname{{D}}(Y_{\bullet , {\textup{\'et}}}).$$ The sheaf $D_{{\mathcal}X_{Y_\bullet }}(\pi ^*F)$ is just the restriction of $D_{{\mathcal}X}(F)$ to ${\mathcal}X_{Y_\bullet }$. It follows from this that $Rf_{\bullet *}D_{{\mathcal}X_{Y_\bullet }}(\pi ^*F)$ is equal to the restriction of $Rf_*D_{{\mathcal}X}(F)$ to $Y_\bullet $, and this in turn implies that $D_{Y_\bullet }Rf_{\bullet *}D_{{\mathcal}X_{Y_\bullet }}(\pi ^*F)$ is isomorphic to the restriction of $Rf_!F$ to $Y_{\bullet , {\textup{\'et}}}$. From this we conclude that $Rf_!F$ is equal to the sheaf obtained from $D_{Y_\bullet }Rf_{\bullet *}D_{{\mathcal}X_{Y_\bullet }}(\pi ^*F)$ and the equivalence of categories (\[etexample\]) $\operatorname{{D}}_c({\mathcal}Y)\simeq \operatorname{{D}}_c(Y_\bullet ).$ \[5.5.6\] Let $$\label{5.5.6.1} \begin{CD} {\mathcal}X'@>a>> {\mathcal}X\\ @Vf'VV @VVfV \\ {\mathcal}Y'@>b>> {\mathcal}Y \end{CD}$$ be a cartesian square of stacks over $S$. Then there is a natural isomorphism of functors $$b^*Rf_!\rightarrow Rf'_!a^*.$$ By \[5.4.4\], there exists a commutative diagram $$\label{5.5.6.3} \begin{CD} Y_\bullet '@>j>> Y_\bullet \\ @VpVV @VVqV \\ {\mathcal}Y'@>\rho >> {\mathcal}Y, \end{CD}$$ where $p$ and $q$ are smooth hypercovers and $j$ is a closed immersion. Let ${\mathcal}X'_{Y'_\bullet }$ denote the base change ${\mathcal}X'\times _{{\mathcal}Y'}Y_{\bullet }'$ and ${\mathcal}X_{Y_\bullet }$ the base change ${\mathcal}X\times _{{\mathcal}Y}Y_\bullet $. Then there is a cartesian diagram $$\begin{CD} {\mathcal}X'_{Y'_\bullet }@>i>> {\mathcal}X_{Y_\bullet }\\ @Vg'VV @VVgV \\ Y'_\bullet @>j>> Y_\bullet , \end{CD}$$ where $i$ and $j$ are closed immersions. As before let $\omega _{{\mathcal{X}}' _{Y'_\bullet }}$ (resp. $\omega _{{\mathcal{X}}_{Y_\bullet }}$, $\omega _{Y'_\bullet }$, $\omega _{Y_\bullet }$) denote the pullback of $\Omega _{{\mathcal{X}}'}$ (resp. $\Omega _{\mathcal{X}}$, $\Omega _{{\mathcal{Y}}'}$, $\Omega _{{\mathcal{Y}}}$) to ${\mathcal{X}}'_{Y'_\bullet }$ (resp. ${\mathcal{X}}_{Y_\bullet }$, $Y_\bullet '$, $Y_\bullet $), and let $D_{{\mathcal{X}}'_{Y'_\bullet }}$ (resp. $D_{{\mathcal{X}}_{Y_\bullet }}$, $D_{Y'_\bullet }$, $D_{Y_\bullet }$) denote the functor $\operatorname{\mathcal{R}\it{hom}}(-, \omega _{X'_{Y'_\bullet }})$ (resp. $\operatorname{\mathcal{R}\it{hom}}(-, \omega _{{\mathcal{X}}_{Y_\bullet }})$, $\operatorname{\mathcal{R}\it{hom}}(-, \omega _{Y'_\bullet })$, $\operatorname{\mathcal{R}\it{hom}}(-, \omega _{Y_\bullet })$). \[5.5.7\] Let ${\mathcal{T}}$ be a topos and $\Lambda $ a sheaf of rings in ${\mathcal{T}}$. Then for any $A, B, C\in \operatorname{{D}}({\mathcal{T}}, \Lambda )$ there is a canonical morphism $$A{\operatorname{\stackrel{\mathbf L}{\otimes}}}\operatorname{\mathcal{R}\it{hom}}(B, C)\rightarrow \operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(A, B), C).$$ We have $$\label{adeq} \operatorname{\mathrm{Rhom}}(A{\operatorname{\stackrel{\mathbf L}{\otimes}}}\operatorname{\mathcal{R}\it{hom}}(B, C), \operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(A, B), C)) \simeq \operatorname{\mathrm{Rhom}}(A{\operatorname{\stackrel{\mathbf L}{\otimes}}}\operatorname{\mathcal{R}\it{hom}}(B, C){\operatorname{\stackrel{\mathbf L}{\otimes}}}\operatorname{\mathcal{R}\it{hom}}(A, B), C).$$ Let $$a:A{\operatorname{\stackrel{\mathbf L}{\otimes}}}\operatorname{\mathcal{R}\it{hom}}(A, B)\rightarrow B, \ \ b:B {\operatorname{\stackrel{\mathbf L}{\otimes}}}\operatorname{\mathcal{R}\it{hom}}(B, C)\rightarrow C$$ be the evaluation morphisms. Then the morphism $$A\operatorname{\stackrel{\mathbf L}{\otimes}}\operatorname{\mathcal{R}\it{hom}}(B, C)\operatorname{\stackrel{\mathbf L}{\otimes}}\operatorname{\mathcal{R}\it{hom}}(A, B)\xrightarrow{a} B\operatorname{\stackrel{\mathbf L}{\otimes}}\operatorname{\mathcal{R}\it{hom}}(B,C)\xrightarrow{b}C$$ and the isomorphism \[adeq\] give the Lemma. $\square$ Let ${\mathcal}F$ denote the functor $$D_{Y'_\bullet }j^*D_{Y_\bullet }g_*D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}:\operatorname{{D}}_c({\mathcal{X}}'_{Y'_\bullet })\rightarrow \operatorname{{D}}(Y'_\bullet ).$$ \[5.4.8\] There is an isomorphism of functors ${\mathcal}F\simeq Rg'_*$. Consider first the functor ${\mathcal}F':= {\mathcal}F\circ g^{\prime *}$ and let $A\in \operatorname{{D}}_c^-(Y'_\bullet )$. Then $$\begin{aligned} {\mathcal}F'(A) & = & D_{Y'_\bullet }j^*D_{Y_\bullet }g_*\operatorname{\mathcal{R}\it{hom}}(i_*\operatorname{\mathcal{R}\it{hom}}(g^{\prime *}A, \omega _{{\mathcal{X}}'_{Y'_\bullet }}), \omega _{{\mathcal{X}}_{Y_\bullet }}) {{ \ \ \ }}(\text{definition})\\ & \simeq & D_{Y'_\bullet }j^*D_{Y_\bullet }g_*\operatorname{\mathcal{R}\it{hom}}(i_*\operatorname{\mathcal{R}\it{hom}}(i^*i_*g^{\prime *}A, \omega _{{\mathcal{X}}'_{Y'_\bullet }}), \omega _{{\mathcal{X}}_{Y_\bullet }}) {{ \ \ \ }}(i^*i_* = \text{id})\\ & \simeq & D_{Y'_\bullet }j^*D_{Y_\bullet }g_*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(i_*g^{\prime *}A, i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}), \omega _{{\mathcal{X}}_{Y_\bullet }}) {{ \ \ \ }}(\text{adjunction for $(i^*, i_*)$})\\ & \leftarrow & D_{Y'_\bullet }j^*D_{Y_\bullet }g_*(i_*g^{\prime *}A{\operatorname{\stackrel{\mathbf L}{\otimes}}}\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}, \omega _{{\mathcal{X}}_{Y_\bullet }})){{ \ \ \ }}(\ref{5.5.7})\\ & \simeq & D_{Y'_\bullet }j^*D_{Y_\bullet }g_*(g^*j_*A{\operatorname{\stackrel{\mathbf L}{\otimes}}}\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}, \omega _{{\mathcal{X}}_{Y_\bullet }})){{ \ \ \ }}(i_*g'^* = g^*j_* \text{ by proper base change})\\ & \simeq & D_{Y'_\bullet }j^*D_{Y_\bullet }(j_*A{\operatorname{\stackrel{\mathbf L}{\otimes}}}Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}, \omega _{{\mathcal{X}}_{Y_\bullet }})) {{ \ \ \ }}(\text{projection formula})\\ & \simeq & \operatorname{\mathcal{R}\it{hom}}(j^*\operatorname{\mathcal{R}\it{hom}}(j_*A{\operatorname{\stackrel{\mathbf L}{\otimes}}}Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}, \omega _{{\mathcal{X}}_{Y_\bullet }}), \omega _{Y_\bullet }), \omega _{Y'_\bullet }){{ \ \ \ }}(\text{definition})\\ & \simeq & j^*j_*\operatorname{\mathcal{R}\it{hom}}(j^*\operatorname{\mathcal{R}\it{hom}}(j_*A{\operatorname{\stackrel{\mathbf L}{\otimes}}}Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}, \omega _{{\mathcal{X}}_{Y_\bullet }}), \omega _{Y_\bullet }), \omega _{Y'_\bullet }){{ \ \ \ }}(j^*j_* = \text{id})\\ & \simeq & j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(j_*A{\operatorname{\stackrel{\mathbf L}{\otimes}}}Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}, \omega _{{\mathcal{X}}_{Y_\bullet }}), \omega _{Y_\bullet }), j_*\omega _{Y'_\bullet }){{ \ \ \ }}(\text{adjunction for $(j^*, j_*)$})\\ & \simeq & j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(j_*A, \operatorname{\mathcal{R}\it{hom}}(Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}, \omega _{{\mathcal{X}}_{Y_\bullet }}), \omega _{Y_\bullet })), j_*\omega _{Y'_{\bullet }})\\ & \leftarrow & A{\operatorname{\stackrel{\mathbf L}{\otimes}}}j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}, \omega _{{\mathcal{X}}_{Y_\bullet }}), \omega _{Y_\bullet }), j_*\omega _{Y'_{\bullet }}) {{ \ \ \ }}(\text{\ref{5.5.7}}).\end{aligned}$$ The following Lemma therefore shows that there is a canonical morphism $A\rightarrow {\mathcal{F}}'(A)$ functorial in $A$. For all $s\in \mathbb{Z}$ there is a canonical isomorphism $${\mathcal}H^s(j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}, \omega _{{\mathcal{X}}_{Y_\bullet }}), \omega _{Y_\bullet }), j_*\omega _{Y'_{\bullet }})\simeq R^sg'_*\Lambda .$$ In particular, $\tau _{\leq 0}j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}, \omega _{{\mathcal{X}}_{Y_\bullet }}), \omega _{Y_\bullet }), j_*\omega _{Y'_{\bullet }})\simeq g'_*\Lambda $, so the composite $$\Lambda \rightarrow g'_*\Lambda \simeq \tau _{\leq 0}j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}, \omega _{{\mathcal{X}}_{Y_\bullet }}), \omega _{Y_\bullet }), j_*\omega _{Y'_{\bullet }})$$ induces a canonical morphism $$\Lambda \rightarrow j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}, \omega _{{\mathcal{X}}_{Y_\bullet }}), \omega _{Y_\bullet }), j_*\omega _{Y'_{\bullet }}).$$ It suffices to construct such a canonical isomorphism over each $Y'_n$. Let $d_n$ (resp. $d_n'$) denote the relative dimension of $Y_n$ (resp. $Y_n'$) over ${\mathcal{Y}}$ (resp. ${\mathcal{Y}}'$). Note that $d_n$ (resp. $d_n'$) is also equal to the relative dimension of ${\mathcal{X}}_{Y_n}$ (resp. ${\mathcal{X}}'_{Y'_n}$) over ${\mathcal{X}}$ (resp. ${\mathcal{X}}'$). As mentioned above we therefore have $$\omega _{{\mathcal{X}}'_{Y'_n}}\simeq \Omega _{{\mathcal{X}}'_{Y'_n}}\langle -d_n'\rangle, \ \ \omega _{{\mathcal{X}}_{Y_n}}\simeq \Omega _{{\mathcal{X}}_{Y_n}}\langle -d_n\rangle , \ \ \omega _{Y_n}\simeq \Omega _{Y_n}\langle -d_n\rangle , \ \ \omega _{Y'_n}\simeq \Omega _{Y'_n}\langle -d_n'\rangle.$$ From this and an elementary manipulation using the identity $$\operatorname{\mathcal{R}\it{hom}}(A\langle n\rangle , B\langle m\rangle )\simeq \operatorname{\mathcal{R}\it{hom}}(A, B)\langle m-n\rangle$$ we get $$\label{bigmess} \begin{matrix} && j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\omega _{{\mathcal{X}}'_{Y'_n }}, \omega _{{\mathcal{X}}_{Y_n }}), \omega _{Y_n }), j_*\omega _{Y'_{n }})\\ &\simeq & j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\Omega _{{\mathcal{X}}'_{Y'_n }}\langle -d_n'\rangle , \Omega _{{\mathcal{X}}_{Y_n }}\langle -d_n\rangle ), \Omega _{Y_n }\langle -d_n\rangle ), j_*\Omega _{Y'_{n }}\langle -d_n'\rangle )\\ & \simeq & j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\Omega _{{\mathcal{X}}'_{Y'_n }}, \Omega _{{\mathcal{X}}_{Y_n }}), \Omega _{Y_n }), j_*\Omega _{Y'_{n }}). \end{matrix}$$ We then get $$\begin{aligned} \operatorname{\mathcal{R}\it{hom}}(i_*\Omega _{{\mathcal{X}}'_{Y'_n}}, \Omega _{{\mathcal{X}}_{Y_n}})& \simeq & \operatorname{\mathcal{R}\it{hom}}(i_!\Omega _{{\mathcal{X}}'_{Y'_n}}, \Omega _{{\mathcal{X}}_{Y_n}}){{ \ \ \ }}(\text{\ref{prop-dual-imm-ferme}})\\ & \simeq & i_*\operatorname{\mathcal{R}\it{hom}}(\Omega _{{\mathcal{X}}'_{Y'_n}}, i^!\Omega _{{\mathcal{X}}_{Y_n}}){{ \ \ \ }}(\text{\ref{dual}})\\ & \simeq & i_*\operatorname{\mathcal{R}\it{hom}}(\Omega _{{\mathcal{X}}'_{Y'_n}}, \Omega _{{\mathcal{X}}'_{Y'_n}}){{ \ \ \ }}(i^!\Omega _{{\mathcal{X}}_{Y_n}} = \Omega _{{\mathcal{X}}'_{Y'_n}})\\ & \simeq & i_*\Lambda {{ \ \ \ }}(\text{\ref{Dinvolutif}}).\end{aligned}$$ Therefore \[bigmess\] is equal to $$j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(Rg_*i_*\Lambda , \Omega _{Y_n}), j_*\Omega _{Y'_n})\simeq j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(j_*Rg'_*\Lambda , \Omega _{Y_n}), j_*\Omega _{Y'_n}).$$ Then $$\begin{aligned} j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(j_*Rg'_*\Lambda , \Omega _{Y_n}), j_*\Omega _{Y'_n})& \simeq & j^*\operatorname{\mathcal{R}\it{hom}}(j_*\operatorname{\mathcal{R}\it{hom}}(Rg'_*\Lambda , j^!\Omega _{Y_n}), j_*\Omega _{Y'_n}) {{ \ \ \ }}(\text{\ref{dual}})\\ & \simeq & j^*j_*\operatorname{\mathcal{R}\it{hom}}(j^*j_*\operatorname{\mathcal{R}\it{hom}}(Rg'_*\Lambda , \Omega _{Y'_n}), \Omega _{Y'_n}) \\ & \simeq & \operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(Rg'_*\Lambda , \Omega _{Y'_n}), \Omega _{Y'_n}){{ \ \ \ }}(j^*j_* = \text{id})\\ & \simeq & Rg'_*\Lambda {{ \ \ \ }}(\text{\ref{Dinvolutif}}).\end{aligned}$$ $\square$ The functor $\text{id}\rightarrow {\mathcal}F'$ induces for any $A\in \operatorname{{D}}_c(Y'_\bullet )$ and $B\in \operatorname{{D}}_c({\mathcal{X}}'_{Y'_\bullet })$ a morphism $$\operatorname{\mathcal{R}\it{hom}}(A, Rg'_*B)\rightarrow \operatorname{\mathcal{R}\it{hom}}(A, {\mathcal}F'(Rg'_*B))\simeq \operatorname{\mathcal{R}\it{hom}}(A, {\mathcal}F(g^{\prime *}Rg'_*B))\rightarrow \operatorname{\mathcal{R}\it{hom}}(A, {\mathcal}F(B)),$$ where the last morphism is induced by adjunction $g^{\prime *}Rg'_*B\rightarrow B$. This map is functorial in $A$, so by Yoneda’s Lemma we get a canonical morphism $Rg'_*B\rightarrow {\mathcal}F(B)$. To prove \[5.4.8\] we show that this map is an isomorphism for all $B\in \operatorname{{D}}_c({\mathcal{X}}'_{Y'_\bullet })$. For this we can restrict the map to any ${\mathcal{X}}'_{Y'_n}$. Noting that the shifts and Tate twists cancel as in \[bigmess\], we get $$\begin{aligned} {\mathcal}F(B) |_{Y'_n} & \simeq & \operatorname{\mathcal{R}\it{hom}}(j^*\operatorname{\mathcal{R}\it{hom}}(Rg_*\operatorname{\mathcal{R}\it{hom}}(i_*\operatorname{\mathcal{R}\it{hom}}(B, \Omega _{{\mathcal{X}}'_{Y'_n}}), \Omega _{{\mathcal{X}}_{Y_n}}), \Omega _{Y_n}), \Omega _{Y'_n})\\ & \simeq & \operatorname{\mathcal{R}\it{hom}}(j^*\operatorname{\mathcal{R}\it{hom}}(Rg_*i_*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(B, \Omega _{{\mathcal{X}}'_{Y'_n}}), Ri^!\Omega _{{\mathcal{X}}_{Y_n}}),\Omega _{Y_n}),\Omega _{Y'_n}){{ \ \ \ }}(\ref{dual})\\ & \simeq & \operatorname{\mathcal{R}\it{hom}}(j^*\operatorname{\mathcal{R}\it{hom}}(j_*Rg'_*B, \Omega _{Y_n}), \Omega _{Y'_n}){{ \ \ \ }}(Ri^!\Omega _{{\mathcal{X}}_{Y_n}} = \Omega _{{\mathcal{X}}'_{Y'_n}}, \ref{Dinvolutif}, \text{and} j_*Rg'_* = Rg_*i_*)\\ & \simeq & \operatorname{\mathcal{R}\it{hom}}(j^*j_*\operatorname{\mathcal{R}\it{hom}}(Rg'_*B, Rj^!\Omega _{Y_n}), \Omega _{Y_n'}){{ \ \ \ }}(\ref{dual})\\ & \simeq & \operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(Rg'_*B, \Omega _{Y'_n}), \Omega _{Y'_n}){{ \ \ \ }}(j^*j_* = \text{id}, j^!\Omega _{Y_n} = \Omega _{Y'_n})\\ & \simeq & Rg'_*B {{ \ \ \ }}(\ref{Dinvolutif}).\end{aligned}$$ We leave to the reader the task of verifying that this isomorphism agrees with the map obtained by restriction from the morphism ${\mathcal}F(B)\rightarrow Rg'_*B$ constructed above, thereby completing the proof of \[5.4.8\]. $\square$ Let $\pi :{\mathcal{X}}_{Y_\bullet }\rightarrow {\mathcal{X}}$ (resp. $\pi ':{\mathcal{X}}'_{Y'_\bullet }\rightarrow {\mathcal{X}}'$) denote the projection. The isomorphism ${\mathcal}F\simeq Rg'_*$ induces a morphism of functors $$\label{5.5.9.2} \begin{matrix} j^*D_{Y_\bullet }Rg_*D_{{\mathcal{X}}_{Y_\bullet }} & \rightarrow & j^*D_{Y_\bullet }Rg_*D_{{\mathcal{X}}_{Y_\bullet }}i_*i^* {{ \ \ \ }}(\text{id}\rightarrow i_*i^*)\\ & \simeq & D_{Y'_\bullet }D_{Y'_\bullet }j^*D_{Y_\bullet }Rg_*D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}D_{{\mathcal{X}}'_{Y'_\bullet }}i^*{{ \ \ \ }}(\ref{Dinvolutif})\\ & \simeq & D_{Y'_\bullet }{\mathcal}FD_{{\mathcal{X}}'_{Y'_\bullet }}i^*{{ \ \ \ }}(\text{definition})\\ & \simeq & D_{Y'_\bullet }Rg'_*D_{{\mathcal{X}}'_{Y'_\bullet }}i^*{{ \ \ \ }}(\ref{5.4.8}). \end{matrix}$$ This map induces a morphism $$\label{basearrow2} \begin{matrix} \rho ^*Rf_! & \simeq & \rho ^*Rq_*D_{Y_\bullet }Rg_*D_{{\mathcal{X}}_{Y_\bullet }}\pi ^*{{ \ \ \ }}(\text{cohomological descent})\\ & \rightarrow & Rp_*j^*D_{Y_\bullet }Rg_*D_{{\mathcal{X}}_{Y_\bullet }}\pi ^*{{ \ \ \ }}(\text{base change morphism})\\ &\rightarrow & Rp_*D_{Y'_\bullet }Rg'_*D_{{\mathcal{X}}'_{Y'_\bullet }}i^*\pi ^*{{ \ \ \ }}(\ref{5.5.9.2})\\ & \simeq & Rp_*D_{Y'_\bullet }Rg'_*D_{{\mathcal{X}}'_{Y'_\bullet }}\pi ^{\prime *}a^*{{ \ \ \ }}(i^*\pi ^* = \pi ^{\prime *}a^*)\\ & \simeq & Rf'_!a^*{{ \ \ \ }}(\text{cohomological descent}). \end{matrix}$$ which we call the *base change morphism*. By construction this morphism is compatible with smooth base change on ${\mathcal}Y$ and ${\mathcal}Y'$. It follows that in order to verify that \[basearrow2\] is an isomorphism it suffices to consider the case when ${\mathcal}Y'$ and ${\mathcal}Y$ are schemes. Furthermore, by construction if $X_\bullet \rightarrow {\mathcal}X$ is a smooth hypercover and $X_\bullet '$ the base change to ${\mathcal}Y'$, then the base change arrow \[basearrow2\] is compatible with the spectral sequences \[5.16\]. It follows that to verify that \[basearrow2\] is an isomorphism it suffices to consider the case of schemes which is [@SGA43], XVII, 5.2.6. Finally the independence of the choices follows by a standard argument from \[filteringremark\]. This completes the proof of \[5.5.6\]. $\square$ Equivalence of different definitions of base change morphism ------------------------------------------------------------ In this subsection we show that the base changed morphism defined in the previous subsection agrees with the morphism defined earlier for smooth morphisms, immersions, and universal homeomorphisms. ### The case when $\rho $ is smooth. {#5.6.1} Choose a diagram as in \[5.5.6.3\], and let $d$ denote the locally constant function on ${\mathcal{Y}}'$ which is the relative dimension of $\rho $. For any morphism ${\mathcal}Z\rightarrow {\mathcal{Y}}'$ we also write $d$ for the pullback of the function $d$ to ${\mathcal}Z$. Note that $$\label{twisteq} j^*\omega _{Y_\bullet }\simeq \omega _{Y'_\bullet }\langle -d\rangle, \ \ i^*\omega _{{\mathcal{X}}_{Y_\bullet }}\simeq \omega _{{\mathcal{X}}'_{Y_\bullet '}}\langle -d\rangle .$$ For any $A\in D_c({\mathcal{X}}_{Y_\bullet })$ (resp. $B\in D_c(Y_\bullet )$) there is a natural isomorphism $D_{{\mathcal{X}}'_{Y'_\bullet }}(i^*A\langle d\rangle )\simeq i^*D_{{\mathcal{X}}_{Y_\bullet }}(A)$ (resp. $D_{Y'_\bullet }j^*(B\langle d\rangle )\simeq j^*D_{Y_\bullet }(B)$). Consider the natural map $$\label{pullmorphism} \begin{matrix} i^*\operatorname{\mathcal{R}\it{hom}}(A, \omega _{{\mathcal{X}}_{Y_\bullet }})&\rightarrow & \operatorname{\mathcal{R}\it{hom}}(i^*A, i^*\omega _{{\mathcal{X}}_{Y_\bullet }})\\ &\simeq & \operatorname{\mathcal{R}\it{hom}}(i^*A, \omega _{{\mathcal{X}}'_{Y'_\bullet }})\langle -d \rangle\\ & \simeq & \operatorname{\mathcal{R}\it{hom}}(i^*A\langle d\rangle, \omega _{{\mathcal{X}}'_{Y'_\bullet }}). \end{matrix}$$ We claim that this map is an isomorphism. This can be verified over each ${\mathcal{X}}'_{Y'_n}$. Let $\pi _n:{\mathcal{X}}_{Y_n}\rightarrow {\mathcal{X}}$ (resp. $\pi _n':{\mathcal{X}}'_{Y'_n}\rightarrow {\mathcal{X}}'$) be the projection. By the equivalence of triangulated categories $D_c({\mathcal{X}}_{Y_\bullet })\simeq D_c({\mathcal{X}})$, there exists an object $A'\in D_c({\mathcal{X}})$ so that the restriction of $A$ to ${\mathcal{X}}_{Y_n}$ is isomorphic to $\pi _n^*A'$. The morphism \[pullmorphism\] is then identified with the isomorphism $$\begin{matrix} i^*\operatorname{\mathcal{R}\it{hom}}(A, \omega _{{\mathcal{X}}_{Y_\bullet }})& \simeq & i^*\pi _n^*\operatorname{\mathcal{R}\it{hom}}(A', \Omega _{{\mathcal{X}}}) & { \ \ \ \ } (\ref{Rhom-im-inv})\\ & \simeq & \pi _n^{\prime *}\rho ^*\operatorname{\mathcal{R}\it{hom}}(A', \Omega _{{\mathcal{X}}}) & \\ & \simeq & \pi _n^{\prime *}\operatorname{\mathcal{R}\it{hom}}(\rho ^*A', \rho ^*\Omega _{{\mathcal{X}}})& { \ \ \ \ } (\ref{Rhom-im-inv})\\ & \simeq & \operatorname{\mathcal{R}\it{hom}}(\pi _n^{\prime *}\rho ^*A', \pi _n^*\rho ^*\Omega _{{\mathcal{X}}}) & { \ \ \ \ } (\ref{Rhom-im-inv})\\ & \simeq & \operatorname{\mathcal{R}\it{hom}}(i^*A\langle d\rangle, \omega _{{\mathcal{X}}'_{Y'_\bullet }}). \end{matrix}$$ The same argument proves the statement $D_{Y'_\bullet }j^*(B\langle d\rangle )\simeq j^*D_{Y_\bullet }(B)$. $\square$ For any $A\in D_c(Y_\bullet )$, let $\alpha _A$ denote the isomorphism $$\begin{aligned} j^*A\langle d\rangle & \simeq & j^*D_{Y_\bullet }D_{Y_\bullet }(A)\langle d\rangle \\ & \simeq & D_{Y'_\bullet }j^*D_{Y_\bullet }(A).\end{aligned}$$ For $B\in \operatorname{{D}}_c({\mathcal{X}}_{Y_\bullet })$ let $\beta _B$ denote the isomorphism $$\begin{aligned} i^*B\langle d \rangle & \simeq & i^*D_{{\mathcal{X}}_{Y_\bullet }}D_{{\mathcal{X}}_{Y_\bullet }}(B) \\ & \simeq & D_{{\mathcal{X}}'_{Y'_\bullet }}i^*D_{{\mathcal{X}}_{Y_\bullet }}(B).\end{aligned}$$ Also for $C\in \operatorname{{D}}_c({\mathcal{X}}'_{Y'_\bullet })$ let $\gamma _C$ be the isomorphism $$\begin{aligned} C\langle -d\rangle & \simeq & D_{{\mathcal{X}}'_{Y'_\bullet }}D_{{\mathcal{X}}'_{Y'_\bullet }}(C\langle -d\rangle )\\ & \simeq & D_{{\mathcal{X}}'_{Y'_\bullet }}i^*i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(C\langle -d\rangle )\\ & \simeq & i^*D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(C),\end{aligned}$$ and let $\gamma ':D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(C)\rightarrow i_*C\langle -d\rangle $ denote the map obtained by adjunction. This map also induces for every $E\in \operatorname{{D}}_c({\mathcal{X}}_{Y_\bullet })$ a morphism $\delta _E$ given by $$\begin{CD} D_{{\mathcal{X}}_{Y_\bullet }}i_*i^*D_{{\mathcal{X}}_{Y_\bullet }}(E) @>>> D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(i^*E)\langle d\rangle @>\gamma '>> i_*i^*E. \end{CD}$$ The map $\alpha _A$ is a special case of a more general class of morphisms. For $A, M\in \operatorname{{D}}_c(Y_\bullet )$ let $S_{A, M}:j^*A\operatorname{\stackrel{\mathbf L}{\otimes}}D_{Y'_\bullet }j^*D_{Y_\bullet }(M)\rightarrow D_{Y'_\bullet }j^*D_{Y_\bullet }(A\operatorname{\stackrel{\mathbf L}{\otimes}}M)$ denote the composite $$\begin{aligned} j^*A\operatorname{\stackrel{\mathbf L}{\otimes}}\operatorname{\mathcal{R}\it{hom}}(j^*Rhom (M, \omega _{Y_\bullet }), \omega _{Y'_\bullet })& \simeq & j^*(A\operatorname{\stackrel{\mathbf L}{\otimes}}\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(M, \omega _{Y_\bullet }), j_*\omega _{Y'_\bullet })\\ & \rightarrow & j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(A, \operatorname{\mathcal{R}\it{hom}}(M, \omega _{Y_\bullet })), j_*\omega _{Y'_\bullet })\\ & \simeq & j^*\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(A\operatorname{\stackrel{\mathbf L}{\otimes}}M, \omega _{Y_\bullet }), j_*\omega _{Y'_\bullet })\\ & \simeq & D_{Y'_\bullet }j^*D_{Y_\bullet }(A\operatorname{\stackrel{\mathbf L}{\otimes}}M).\end{aligned}$$ Here the second morphism is given by \[5.5.7\] and the third morphism is by the adjunction property of $\otimes $. For any $A\in \operatorname{{D}}_c(Y_\bullet )$ the map $\alpha _A$ is equal to the composite $$\begin{CD} j^*A\langle d\rangle \simeq j^*A\operatorname{\stackrel{\mathbf L}{\otimes}}j^*\Lambda \langle d\rangle @>\alpha _{\Lambda }>> j^*A\operatorname{\stackrel{\mathbf L}{\otimes}}\operatorname{\mathcal{R}\it{hom}}(j^*Rhom (\Lambda , \omega _{Y_\bullet }), \omega _{Y'_\bullet })@>S_{A, \Lambda }>> D_{Y'_\bullet }j^*D_{Y_\bullet }(A). \end{CD}$$ This follows from the definitions. $\square$ For any $A, B, M\in \operatorname{{D}}_c(Y_\bullet )$, the diagram $$\begin{CD} j^*A\operatorname{\stackrel{\mathbf L}{\otimes}}j^*B\operatorname{\stackrel{\mathbf L}{\otimes}}D_{Y'_\bullet }j^*D_{Y_\bullet }(M)@>S_{B, M}>> j^*A\operatorname{\stackrel{\mathbf L}{\otimes}}D_{Y'_\bullet }j^*D_{Y_\bullet }(B\operatorname{\stackrel{\mathbf L}{\otimes}}M)\\ @V\simeq VV @VVS_{A, B\operatorname{\stackrel{\mathbf L}{\otimes}}M}V \\ j^*(A\operatorname{\stackrel{\mathbf L}{\otimes}}B)\operatorname{\stackrel{\mathbf L}{\otimes}}D_{Y'_\bullet }j^*D_{Y_\bullet }(M)@>S_{A\operatorname{\stackrel{\mathbf L}{\otimes}}B, M}>> D_{Y'_\bullet }j^*D_{Y_\bullet }(A\operatorname{\stackrel{\mathbf L}{\otimes}}B\operatorname{\stackrel{\mathbf L}{\otimes}}M) \end{CD}$$ commutes. Consider the diagram $$\begin{CD} A\operatorname{\stackrel{\mathbf L}{\otimes}}B\operatorname{\stackrel{\mathbf L}{\otimes}}j_*D_{Y'_\bullet }j^*D_{Y_\bullet }(M)@>>> A\operatorname{\stackrel{\mathbf L}{\otimes}}[[B, [M, \omega _{Y_\bullet }]], j_*\omega _{Y'_\bullet }]@>>> A\operatorname{\stackrel{\mathbf L}{\otimes}}[[B\operatorname{\stackrel{\mathbf L}{\otimes}}M, \omega _{Y_\bullet }], j_*\omega _{Y'_\bullet }]\\ @| @VVV @VVV \\ A\operatorname{\stackrel{\mathbf L}{\otimes}}B\operatorname{\stackrel{\mathbf L}{\otimes}}j_*D_{Y'_\bullet }j^*D_{Y_\bullet }(M)@. [[A, [B, [M, \omega _{Y_\bullet }]]], j_*\omega _{Y'_\bullet }]@>>> [[A, [B\operatorname{\stackrel{\mathbf L}{\otimes}}M, \omega _{Y_\bullet }]], j_*\omega _{Y_\bullet '}]\\ @| @VVV @VVV \\ A\operatorname{\stackrel{\mathbf L}{\otimes}}B\operatorname{\stackrel{\mathbf L}{\otimes}}j_*D_{Y'_\bullet }j^*D_{Y_\bullet }(M)@>>> [[A\operatorname{\stackrel{\mathbf L}{\otimes}}B, [M, \omega _{Y_\bullet }]], j_*\omega _{Y'_\bullet }]@>>> [[A\operatorname{\stackrel{\mathbf L}{\otimes}}B\operatorname{\stackrel{\mathbf L}{\otimes}}M, \omega _{Y_\bullet }], j_*\omega _{Y'_\bullet }] \end{CD}$$ where to ease the notation we write simply $[-,-]$ for $\operatorname{\mathcal{R}\it{hom}}(-, -)$. An elementary verification shows that each of the small inside diagrams commute, and hence the big outside rectangle also commutes. Applying $j^*$ we obtain the lemma. $\square$ Similarly, for $A, M\in \operatorname{{D}}_c({\mathcal{X}}'_{Y'_\bullet })$, let $R_{A, M}:i_*A\operatorname{\stackrel{\mathbf L}{\otimes}}D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(M)\rightarrow D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(A\operatorname{\stackrel{\mathbf L}{\otimes}}M)$ be the map $$\begin{aligned} i_*A\operatorname{\stackrel{\mathbf L}{\otimes}}D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(M) & \simeq & i_*A\operatorname{\stackrel{\mathbf L}{\otimes}}\operatorname{\mathcal{R}\it{hom}}(i_*\operatorname{\mathcal{R}\it{hom}}(M, \omega _{{\mathcal{X}}'_{Y'_\bullet }}), \omega _{{\mathcal{X}}_{Y_\bullet }})\\ & \simeq & i_*A\operatorname{\stackrel{\mathbf L}{\otimes}}\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(i_*M, i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}), \omega _{{\mathcal{X}}_{Y_\bullet }})\\ & \rightarrow & \operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(i_*A, \operatorname{\mathcal{R}\it{hom}}(i_*M, i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }})), \omega _{{\mathcal{X}}_{Y_\bullet }})\\ & \simeq &\operatorname{\mathcal{R}\it{hom}}(\operatorname{\mathcal{R}\it{hom}}(i_*(A\operatorname{\stackrel{\mathbf L}{\otimes}}M), i_*\omega _{{\mathcal{X}}'_{Y'_\bullet }}),\omega _{{\mathcal{X}}_{Y_\bullet }})\\ & \simeq &\operatorname{\mathcal{R}\it{hom}}(i_*\operatorname{\mathcal{R}\it{hom}}(A\operatorname{\stackrel{\mathbf L}{\otimes}}M, \omega _{{\mathcal{X}}'_{Y'_\bullet }}), \omega _{{\mathcal{X}}_{Y_\bullet }})\\ & = & D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(A\operatorname{\stackrel{\mathbf L}{\otimes}}M),\end{aligned}$$ where the third morphism is provided by \[5.5.7\]. As above, one verifies that for $A, B, M\in \operatorname{{D}}_c({\mathcal{X}}'_{Y'_\bullet })$ the diagram $$\begin{CD} i_*A\operatorname{\stackrel{\mathbf L}{\otimes}}i_*B\operatorname{\stackrel{\mathbf L}{\otimes}}D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(M)@>R_{B, M}>> i_*A\operatorname{\stackrel{\mathbf L}{\otimes}}D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(B\operatorname{\stackrel{\mathbf L}{\otimes}}M)\\ @V\simeq VV @VVR_{A, B\operatorname{\stackrel{\mathbf L}{\otimes}}M}V\\ i_*(A\operatorname{\stackrel{\mathbf L}{\otimes}}B)\operatorname{\stackrel{\mathbf L}{\otimes}}D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(M)@>R_{A\operatorname{\stackrel{\mathbf L}{\otimes}}B, M}>> D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(A\operatorname{\stackrel{\mathbf L}{\otimes}}B\operatorname{\stackrel{\mathbf L}{\otimes}}M) \end{CD}$$ commutes. From this it follows that if $\varphi _A:A\rightarrow {\mathcal}F'(A)$ denotes the morphism constructed in the proof of \[5.4.8\], then the diagram $$\begin{CD} A@= A\\ @V\varphi _AVV @VV\text{adjunction}V \\ D_{Y'_\bullet }j^*D_{Y_\bullet }g_*D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{\bullet }}g^{\prime *}A @. g^{\prime }_*g^{\prime *} A\\ @V\alpha _{g_*D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{\bullet }}g^{\prime *}}VV @AA\gamma A \\ j^*g_*D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}_{Y'_\bullet }'}g^{\prime *}A @>\text{base change} >> g'_*i^*D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{\bullet }}g^{\prime *}A \end{CD}$$ commutes. Note that the base change morphism in the above diagram is an isomorphism. This can be verified over each ${\mathcal{X}}'_{Y'_n}$. Here the functor $D_{{\mathcal{X}}_{Y_n}}i_*D_{{\mathcal{X}}'_{Y'_n}}$ is up to shift and Tate torsion isomorphic to $i_!=i_*$. The base change morphism is therefore induced by the isomorphism $$j^*g_*i_*\simeq j^*j_*g'_*\simeq g'_*i^*i_*.$$ By the definition of the morphism in \[5.4.8\] this implies that for any $A\in \operatorname{{D}}_c({{\mathcal{X}}'_{Y'_\bullet }})$ the diagram $$\begin{CD} D_{Y'_\bullet }j^*D_{Y_\bullet }g_*D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(A)@>\alpha _{g_*D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(A)}>> j^*g_*D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{\bullet }}(A)\langle d\rangle \\ @V\ref{5.4.8}VV @VV\text{base change} V \\ g'_*(A)@<\gamma _A<< g'_*i^*D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{\bullet }}(A)\langle d\rangle \end{CD}$$ commutes. Combining the commutativity of this diagram with the commutativity of the diagram (verification left to the reader) $$\begin{CD} D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}D_{{\mathcal{X}}'_{Y'_\bullet }}i^*D_{{\mathcal{X}}_{Y_\bullet }}@>D_{{\mathcal{X}}'_{Y'_\bullet }}^2 = \text{id}>> D_{{\mathcal{X}}_{Y_\bullet }}i_*i^*D_{{\mathcal{X}}_{Y_\bullet }}\\ @V\gamma 'VV @VV\delta V \\ i_*D_{{\mathcal{X}}'_{Y'_\bullet }}i^*D_{{\mathcal{X}}_{Y_\bullet }}\langle -d\rangle @>\beta>> i_*i^*. \end{CD}$$ one sees that the diagram $$\label{square} \begin{CD} D_{Y'_\bullet }j^*D_{Y_\bullet }g_*D_{{\mathcal{X}}_{Y_\bullet }}D_{{\mathcal{X}}_{Y_\bullet }}@>\ref{5.5.9.2}>> D_{Y'_\bullet }D_{Y'_\bullet }g'_*D_{{\mathcal{X}}_{Y_\bullet '}'}i^*D_{{\mathcal{X}}_{Y_\bullet }}\\ @V\alpha VV @VVD_{Y'_\bullet }^2 = \text{id} V \\ j^*g_*D_{{\mathcal{X}}_{Y_\bullet }}D_{{\mathcal{X}}_{Y_\bullet }}@. g'_*D_{{\mathcal{X}}'_{Y'_\bullet }}i^*D_{{\mathcal{X}}_{Y_\bullet }}\\ @VD_{{\mathcal{X}}_{Y_\bullet }}^2 = \text{id} VV @VV\beta V \\ j^*g_*\langle d\rangle @>\text{base change}>> g'_*i^*\langle d\rangle \end{CD}$$ commutes. We are now ready to prove the equivalences of the two definitions of the base change morphism. The morphism constructed in \[5.5\] is the composite $$\begin{matrix} \rho ^*f_! & = & \rho ^*q_*D_{Y_\bullet }g_*D_{{\mathcal{X}}_{Y_\bullet }}\pi ^*&\\ & \simeq & p_*j^*D_{Y_\bullet }g_*D_{{\mathcal{X}}_{Y_\bullet }}\pi ^*& \\ & \rightarrow & p_*D_{Y'_\bullet }g'_*D_{{\mathcal{X}}'_{Y'_\bullet }}i^*\pi ^* & { \ \ \ \ }(\ref{5.5.9.2})\\ & \simeq & p_*D_{Y'_\bullet }g'_*D_{{\mathcal{X}}'_{Y'_\bullet }}\pi ^{\prime *}a^*&\\ & \simeq & f'_!a^*. \end{matrix}$$ The dual version of this morphism is given by $$\begin{matrix} \rho ^!f_*& = & D_{{\mathcal{Y}}'}\rho ^*q_*D_{Y_\bullet }g_*D_{{\mathcal{X}}_{Y_\bullet }}\pi ^*D_{{\mathcal{X}}}& \\ & \simeq & p_*D_{Y'_\bullet }j^*D_{Y_\bullet }g_*D_{{\mathcal{X}}_{Y_\bullet }}D_{{\mathcal{X}}_{Y_\bullet }}\pi ^*& \\ & \rightarrow & p_*D_{Y'_\bullet }D_{Y'_\bullet }g'_*D_{{\mathcal{X}}'_{Y'_\bullet }}i^*D_{{\mathcal{X}}_{Y_\bullet }}\pi ^* & { \ \ \ \ } (\ref{5.5.9.2}) \\ & \simeq & p_*g'_*i^*\pi ^*\langle d\rangle . & \\ & \simeq & f'_*a^!.& \end{matrix}$$ By the commutativity of \[square\] this is the same as the composite $$\begin{matrix} \rho ^!f_* & = & D_{{\mathcal{Y}}'}\rho ^*q_*D_{Y_\bullet }g_*D_{{\mathcal{X}}_{Y_\bullet }}\pi ^*D_{{\mathcal{X}}}& \\ & \simeq & p_*D_{Y'_\bullet }j^*D_{Y_\bullet }g_*D_{{\mathcal{X}}_{Y_\bullet }}D_{{\mathcal{X}}_{Y_\bullet }}\pi ^*& \\ & \simeq & p_*D_{Y'_\bullet }j^*D_{Y_\bullet }g_*\pi ^*& { \ \ \ \ } D_{{\mathcal{X}}_{Y_\bullet }}^2 = \text{id} \\ & \simeq & p_*j^*g_*\pi ^* \langle d\rangle & { \ \ \ \ } D_{Y'_\bullet }j^*D_{Y_\bullet }\simeq j^*\langle d\rangle \\ & \rightarrow & p_*g'_*i^* \pi ^* \langle d \rangle & { \ \ \ \ } (\text{base change morphism }) \\ & \simeq & f'_*a^!.& \end{matrix}$$ From this it follows that the morphism defined in \[5.1\] agrees with the one defined in \[5.5\]. ### The case when $\rho $ is a universal homeomorphism The same argument used in the previous section shows the agreement of the base change morphism in \[5.5\] with the base change morphism in \[5.4\]. Indeed the only property of smooth morphisms used in the previous section is that the dualizing sheaves can be described as in \[twisteq\]. This also holds when $\rho $ is a universal homeomorphism (with $d = 0$). ### The case when $\rho $ is an immersion With notation as in \[5.3\], note first that to prove that the two base change morphisms agree it suffices to show that they agree on sheaves of the form $\pi _*A$ with $A\in D_c({\mathcal{X}}')$. Indeed for any $B\in D_c({\mathcal{X}})$ either base change isomorphism factors as $$\begin{CD} p^*f_!B@>\text{id}\rightarrow \pi _*\pi ^*>> p^*f_!\pi _*\pi ^*B@>>> \phi _!\pi ^*\pi _*\pi ^*B = \phi _!\pi ^*B. \end{CD}$$ In order to prove that the two base change morphisms agree, it is useful to first give an alternate description of the morphism defined in \[5.3\]. With notation as in \[5.3\], there is for any $A\in D_c({\mathcal{X}}')$ a canonical isomorphism $$\begin{aligned} D_{{\mathcal{X}}'}(A) & \simeq & \operatorname{\mathcal{R}\it{hom}}(A, \pi ^!\Omega _{{\mathcal{X}}})\\ & \simeq & \pi ^*\pi _*\operatorname{\mathcal{R}\it{hom}}(A, \pi ^!\Omega _{{\mathcal{X}}})\\ & \simeq & \pi ^*\operatorname{\mathcal{R}\it{hom}}(\pi _*A, \Omega _{{\mathcal{X}}})\\ & \simeq & \pi ^*D_{{\mathcal{X}}}(\pi _*A),\end{aligned}$$ and similarly $D_{{\mathcal{Y}}'}\simeq p^*D_{{\mathcal{Y}}}p_*$. We can also write these isomorphisms as $\pi _*D_{{\mathcal{X}}'} \simeq D_{\mathcal{X}}\pi _*$ and $p_*D_{{\mathcal{Y}}'} \simeq D_{{\mathcal{Y}}}p_*$. We therefore obtain a morphism $$\label{alternate2} \begin{matrix} p^*Rf_!(\pi _*A) &=& p^*D_{{\mathcal{Y}}}f_*D_{{\mathcal{X}}}(\pi _*A) & \\ & \simeq & p^*D_{{\mathcal{Y}}}f_*\pi _*D_{{\mathcal{X}}'}(A) & \\ & \simeq & p^*D_{{\mathcal{Y}}}p_*\phi _*D_{{\mathcal{X}}'}(A) & \\ & \simeq & p^*p_*D_{{\mathcal{Y}}'}\phi _*D_{{\mathcal{X}}'}(A)& \\ & \simeq & \phi _!A. & \end{matrix}$$ This morphism agrees with the one defined in \[5.3\]. In particular the morphism \[alternate2\] is an isomorphism. Chasing through the definitions this amounts to the commutativity of the following diagram $$\begin{CD} p^*f_!(\pi _*A)@>\simeq >> p^*(p_*\Lambda \operatorname{\stackrel{\mathbf L}{\otimes}}f_!\pi _*A)\\ @AAA @VV\text{projection formula}V \\ p^*D_{{\mathcal{Y}}}p_*p^*f_*D_{{\mathcal{X}}}(\pi _*A) @. p^*(f_!(f^*p_*\Lambda \operatorname{\stackrel{\mathbf L}{\otimes}}\pi _*A))\\ @V\simeq VV @VV\simeq V \\ D_{{\mathcal{Y}}'}p^*f_*D_{{\mathcal{X}}}(\pi _*A) @. p^*f_!(\pi _*A\operatorname{\stackrel{\mathbf L}{\otimes}}\pi _*\phi ^*\Lambda )\\ @A\text{base change}AA @VV\simeq V \\ D_{{\mathcal{Y}}'}\phi _*\pi ^*D_{{\mathcal{X}}}(\pi _*A) @. p^*f_!\pi _*(A\operatorname{\stackrel{\mathbf L}{\otimes}}\phi ^*\Lambda )\\ @A\simeq AA @VV\simeq V \\ D_{{\mathcal{Y}}'}\phi _*D_{{\mathcal{X}}'}A@>\simeq >> \phi _!\pi ^*(A). \end{CD}$$ We leave to the reader this verification. $\square$ In particular, since the map \[alternate2\] is an isomorphism we can define the base change morphism for $B\in D_c({\mathcal{X}})$ as the composite $$\label{alternate} \begin{CD} p^*f_!B @>>> p^*f_!(\pi _*\pi ^*B)@>\ref{alternate2}>> \phi _!\pi ^*B. \end{CD}$$ Using this alternate description of the base change morphism in \[5.3\], we can prove the equivalence with that given in \[5.5\]. By a standard reduction it suffices to consider the case of a closed immersion. So fix the diagram \[5.5.6.1\] with $\rho $ a closed immersion, and choose a diagram as in \[5.5.6.3\]. Since $\rho $ is a closed immersion we may without loss of generality assume that \[5.5.6.3\] is cartesian. \[pulllem\] The functors $j_*:D(Y_\bullet ')\rightarrow D(Y'_\bullet )$ and $i_*:D({{\mathcal{X}}'_{Y'_\bullet }})\rightarrow D({\mathcal{X}}_{Y_\bullet })$ have right adjoints $j^!$ and $i^!$ respectively. In fact $j^! = j^*R\Gamma _{Y'_\bullet }$ and $i^! = i^*R\Gamma _{{\mathcal{X}}'_{Y'_\bullet }}$. $\square$ Note that for any $[n]\in \Delta $, the restriction of $j^!$ (resp. $i^!$) to a functor $D(Y_n)\rightarrow D(Y'_n)$ (resp. $D({\mathcal{X}}_{Y_n})\rightarrow D({\mathcal{X}}'_{Y'_n})$) agrees with the usual extraordinary inverse image. This follows for example from the explicit description of these functors in the proof of \[pulllem\]. There are canonical isomorphisms $\omega _{Y'_\bullet }\simeq j^!\omega _{Y_\bullet }$ and $\omega _{{\mathcal{X}}'_{Y'_\bullet}}\simeq i^!\omega _{{\mathcal{X}}_{Y_\bullet }}$. By the glueing lemma \[1.2\], it suffices to construct an isomorphism over each $Y_n'$ (resp. ${\mathcal{X}}'_{Y'_n}$). Let $d$ denote the relative dimension of $Y_n$ over ${\mathcal{Y}}$. Then $d$ is also equal to the relative dimension of $Y'_n$ over ${\mathcal{Y}}'$, the relative dimension of ${\mathcal{X}}_{Y_n}$ over ${\mathcal{X}}$, and the relative dimension of ${\mathcal{X}}'_{Y_n'}$ over ${\mathcal{X}}'$. We therefore have $$i^!\omega _{{\mathcal{X}}_{Y_\bullet }}|_{{\mathcal{X}}'_{Y'_n}} = i^!\Omega _{{\mathcal{X}}_{Y_n}}\langle -d\rangle \simeq \Omega _{{\mathcal{X}}'_{Y'_n}}\langle -d\rangle \simeq \omega _{{\mathcal{X}}'_{Y_\bullet }}|_{{\mathcal{X}}'_{Y'_n}}$$ and $$j^!\omega _{Y_\bullet }|_{Y_n.} = j^!\Omega _{Y_n}\langle -d\rangle \simeq \Omega _{Y'_n}\langle -d\rangle \simeq \omega _{Y'_\bullet }|_{Y_n'}.$$ $\square$ For any $A\in D(Y'_\bullet )$ and $B\in D(Y_\bullet )$ (resp. $C\in D({\mathcal{X}}'_{Y'_\bullet })$ and $E\in D({\mathcal{X}}_{Y_\bullet })$) we have $$j_*\operatorname{\mathcal{R}\it{hom}}(A, j^!B)\simeq \operatorname{\mathcal{R}\it{hom}}(j_*A, B), \ \ i_*\operatorname{\mathcal{R}\it{hom}}(C, i^!E)\simeq \operatorname{\mathcal{R}\it{hom}}(i_*C, E).$$ Since $j^!$ is right adjoint to $j_*$, there is an adjunction morphism $j^!j_*\rightarrow \text{id}$. This map induces a morphism $$j_*\operatorname{\mathcal{R}\it{hom}}(A, j^!B)\simeq \operatorname{\mathcal{R}\it{hom}}(j_*A, j_*j_!B)\rightarrow \operatorname{\mathcal{R}\it{hom}}(j_*A, B).$$ That this map is an isomorphism can be verified after restricting to each $Y_n$ in which case it follows from the theory for schemes [@SGA43], XVIII, 3.1.10. The same argument gives the second isomorphism in the Lemma. $\square$ For any $A\in D_c(Y'_\bullet )$ there is a natural isomorphism $D_{Y_\bullet }j_*A\simeq j_*D_{Y'}(A)$, and for $B\in D_c({\mathcal{X}}'_{Y'_\bullet })$ there is a canonical isomorphism $D_{{\mathcal{X}}_{Y_\bullet }}i_*B\simeq i_*D_{{\mathcal{X}}'_{Y'_\bullet }}B$. For $A\in \operatorname{{D}}_c(Y'_\bullet )$, let $\alpha _A$ denote the isomorphism $$\begin{aligned} j_*A & \simeq & D_{Y'_\bullet }j^*j_*D_{Y'_\bullet }A\\ & \simeq & D_{Y'_\bullet }j^*D_{Y_\bullet }j_*A,\end{aligned}$$ and for $B\in D_{{\mathcal{X}}'_{Y'_\bullet }}$ let $\beta_B$ denote the isomorphism $$\begin{aligned} i_*B & \simeq & i_*D_{{\mathcal{X}}'_{Y'_\bullet }}D_{{\mathcal{X}}'_{Y'_\bullet }}(B)\\ & \simeq & i_*D_{{\mathcal{X}}'_{Y'_\bullet }}i^*i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(B)\\ & \simeq & i_*D_{{\mathcal{X}}'_{Y'_\bullet }}i^*D_{{\mathcal{X}}_{Y_\bullet }}(i_*B).\end{aligned}$$ Define $\gamma _B'$ to be the isomorphism $$\begin{matrix} i_*B & \simeq & D_{{\mathcal{X}}_{Y_\bullet }}D_{{\mathcal{X}}_{Y_\bullet }}i_*B\\ & \simeq & D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}B, \end{matrix}$$ and let $\gamma _B:i^*D_{{\mathcal{X}}_{Y_\bullet }}i_*D_{{\mathcal{X}}'_{Y'_\bullet }}(B)\rightarrow B$ be the isomorphism obtained by adjunction. Following the same outline used in \[5.6.1\] (replacing the $\alpha $’s, $\beta $’s, and $\gamma $’s by the above defined morphisms), one sees that the morphism \[5.5.9.2\] in the case of a closed immersion is given by the composite $$\begin{matrix} j^*D_{Y_\bullet }g_*D_{{\mathcal{X}}_{Y_\bullet }}i_* & \simeq & j^*D_{Y_\bullet }g_*i_*D_{{\mathcal{X}}'_{Y'_\bullet }}\\ & \simeq & j^*D_{Y_\bullet }j_*g'_*D_{{\mathcal{X}}'_{Y'_\bullet }}\\ & \simeq & j^*j_*D_{Y_\bullet '}g'_*D_{{\mathcal{X}}'_{Y'_\bullet }}\\ & \simeq & D_{Y_\bullet '}g'_*D_{{\mathcal{X}}'_{Y'_\bullet }}. \end{matrix}$$ From this it follows that the sequence of morphisms in \[basearrow2\] is identified via cohomological descent with the sequence of morphisms \[alternate\], and hence the two base change morphisms are the same. Kunneth formula --------------- Let ${\mathcal}Y_1$ and ${\mathcal}Y_2$ be stacks, and set ${\mathcal}Y:= {\mathcal}Y_1\times {\mathcal}Y_2$. Let $p_i:{\mathcal}Y\rightarrow {\mathcal}Y_i$ ($i=1,2$) be the projection and for two complexes $L_i\in \operatorname{{D}}^-_c({\mathcal}Y_i)$ let $L_1{\operatorname{\stackrel{\mathbf L}{\otimes}}_S}L_2\in \operatorname{{D}}({\mathcal}Y)$ denote $p_1^*L_1{\operatorname{\stackrel{\mathbf L}{\otimes}}}_{\Lambda }p_2^*L_2$. \[dualdescription\] There is a natural isomorphism $K_{{\mathcal}Y}\simeq K_{{\mathcal}Y_1}{\operatorname{\stackrel{\mathbf L}{\otimes}}_S}K_{{\mathcal}Y_2}$. By ([@SGA5], III.1.7.6) there is for any smooth morphisms $U_i\rightarrow {\mathcal}Y_i$ ($i=1,2$) with $U_i$ a scheme, a canonical isomorphism $$K_{{\mathcal}Y}|_{U_1\times _SU_2}\simeq K_{{\mathcal}Y_1}|_{U_1}{\operatorname{\stackrel{\mathbf L}{\otimes}}_S}K_{{\mathcal}Y_2}|_{U_2}.$$ Furthermore, this isomorphism is functorial with respect to morphisms $V_i\rightarrow U_i$. It follows that the sheaf $K_{{\mathcal}Y_1}{\operatorname{\stackrel{\mathbf L}{\otimes}}_S}K_{{\mathcal}Y_2}$ also satisfies the ${\mathcal}Ext$–condition (\[1.2\]), and hence to give an isomorphism as in the Lemma it suffices to give an isomorphism in the derived category of $U_1\times _SU_2$ for all smooth morphisms $U_i\rightarrow {\mathcal}Y_i$. $\square$ \[hommap\] Let $({\mathcal{T}}, \Lambda )$ be a ringed topos. Then for any $P_1, P_2, M_1, M_2\in \operatorname{{D}}({\mathcal{T}}, \Lambda )$, there is a canonical morphism $$\operatorname{\mathcal{R}\it{hom}}(P_1, M_1){\operatorname{\stackrel{\mathbf L}{\otimes}}}\operatorname{\mathcal{R}\it{hom}}(P_2, M_2)\rightarrow \operatorname{\mathcal{R}\it{hom}}(P_1{\operatorname{\stackrel{\mathbf L}{\otimes}}}P_2, M_1{\operatorname{\stackrel{\mathbf L}{\otimes}}}M_2).$$ It suffices to give a morphism $$\operatorname{\mathcal{R}\it{hom}}(P_1, M_1){\operatorname{\stackrel{\mathbf L}{\otimes}}}\operatorname{\mathcal{R}\it{hom}}(P_2, M_2){\operatorname{\stackrel{\mathbf L}{\otimes}}}P_1{\operatorname{\stackrel{\mathbf L}{\otimes}}}P_2\rightarrow M_1{\operatorname{\stackrel{\mathbf L}{\otimes}}}M_2.$$ This we get by tensoring the two evaluation morphisms $$\operatorname{\mathcal{R}\it{hom}}(P_i, M_i){\operatorname{\stackrel{\mathbf L}{\otimes}}}P_i\rightarrow M_i.$$ $\square$ For the definition and standard properties of homotopy colimits we refer to [@Bok-Nee93]. \[reduc-tenseur\] Let $A,B\in\operatorname{{D}}({\mathcal{X}})$.Then we have 1. $\operatorname{\mathrm{hocolim}}\tau_{\leq n}A=A$; 2. $A\operatorname{\stackrel{\mathbf L}{\otimes}}B=\operatorname{\mathrm{hocolim}}\tau_{\leq n}A\operatorname{\stackrel{\mathbf L}{\otimes}}\tau_{\leq n}B$. Consider the triangle $$\oplus\tau_{\leq n}A\xrightarrow{1-\text{shift}}\oplus\tau_{\leq n}A\operatorname{\rightarrow}A\leqno(*)$$ If $C=\operatorname{\mathrm{hocolim}}\tau_{\leq n}A$ is the cone of $1-\text{shift}$, one gets a morphism $C\operatorname{\rightarrow}A$. By construction, one has $${\mathcal}H(C)=\varinjlim{\mathcal}H(\tau_{\leq n}A)={\mathcal}H(A)$$ proving that $C\operatorname{\rightarrow}A$ is an isomorphism. Tensoring (\*) by $B$ we get therefore a distinguished triangle $$\oplus\tau_{\leq n}A\operatorname{\stackrel{\mathbf L}{\otimes}}B\xrightarrow{1-\text{shift}}\oplus\tau_{\leq n}A \operatorname{\stackrel{\mathbf L}{\otimes}}B\operatorname{\rightarrow}A\operatorname{\stackrel{\mathbf L}{\otimes}}B$$ proving $$\operatorname{\mathrm{hocolim}}\tau_{\leq n}A\operatorname{\stackrel{\mathbf L}{\otimes}}B=A\operatorname{\stackrel{\mathbf L}{\otimes}}B.$$ Applying this process again we find $$\operatorname{\mathrm{hocolim}}\tau_{\leq n}\operatorname{\stackrel{\mathbf L}{\otimes}}\tau_{\leq m}B=A\operatorname{\stackrel{\mathbf L}{\otimes}}B.$$ Because the diagonal is cofinal in ${\mathbf{N}}\times {\mathbf{N}}$, the lemma follows. $\square$ For $L_i\in \operatorname{{D}}^-_c({\mathcal}Y_i)$ ($i=1,2$), there is a canonical isomorphism $$D_{{\mathcal}Y_1}(L_1){\operatorname{\stackrel{\mathbf L}{\otimes}}}_SD_{{\mathcal}Y_2}(L_2)\simeq D_{{\mathcal}Y}(L_1{\operatorname{\stackrel{\mathbf L}{\otimes}}_S}L_2).$$ By \[dualdescription\] and \[hommap\] there is a canonical morphism (note here we also use that $K_{{\mathcal}Y_i}$ has finite injective dimension) $$D_{{\mathcal}Y_1}(L_1){\operatorname{\stackrel{\mathbf L}{\otimes}}}_SD_{{\mathcal}Y_2}(L_2)\rightarrow D_{{\mathcal}Y}(L_1{\operatorname{\stackrel{\mathbf L}{\otimes}}_S}L_2).$$ To verify that this map is an isomorphism, it suffices to show that for every $j\in Z$ the map $${\mathcal}H^j(D_{{\mathcal}Y_1}(L_1){\operatorname{\stackrel{\mathbf L}{\otimes}}}_SD_{{\mathcal}Y_2}(L_2))\simeq {\mathcal}H^j(D_{{\mathcal}Y}(L_1{\operatorname{\stackrel{\mathbf L}{\otimes}}_S}L_2)).$$ Because $\operatorname{\stackrel{\mathbf L}{\otimes}}$ commutes with homotopy colimits (\[reduc-tenseur\]), we deduce from $D(A)=\operatorname{\mathrm{hocolim}}D(\tau_\geq m A)$ (use \[reduc-tenseur\]) that to prove this we may replace $L_i$ by $\tau _{\geq m}L_i$ for $m$ sufficiently negative, and therefore it suffices to consider the case when $L_i\in \operatorname{{D}}^b_c({\mathcal}Y_i)$. Furthermore, we may work locally in the smooth topology on ${\mathcal}Y_1$ and ${\mathcal}Y_2$, and therefore it suffices to consider the case when the stacks ${\mathcal}Y_i$ are schemes. In this case the result is [@SGA43], XVII, 5.4.3. $\square$ Now consider morphisms of $S$–stacks $f_i:{\mathcal}X_i\rightarrow {\mathcal}Y_i$ ($i=1,2$), and let $f:{\mathcal}X:= {\mathcal}X_1\times _S{\mathcal}X_2\rightarrow {\mathcal}Y:= {\mathcal}Y_1\times _S{\mathcal}Y_2$ be the morphism obtained by taking fiber products. Let $L_i\in \operatorname{{D}}^-_c({\mathcal}X)$. There is a canonical isomorphism in $\operatorname{{D}}_c({\mathcal}Y)$ $$\label{Kunnethmap} Rf_!(L_1{\operatorname{\stackrel{\mathbf L}{\otimes}}_S}L_2)\rightarrow Rf_{1!}(L_1){\operatorname{\stackrel{\mathbf L}{\otimes}}_S}Rf_{2!}(L_2).$$ We define the morphism \[Kunnethmap\] as the composite $$\begin{CD} Rf_!(L_1{\operatorname{\stackrel{\mathbf L}{\otimes}}_S}L_2)@>\simeq >> D_{{\mathcal}Y}(f_*D_{{\mathcal}X}(L_1{\operatorname{\stackrel{\mathbf L}{\otimes}}_S}L_2))\\ @>\simeq >> D_{{\mathcal}Y}(f_*(D_{{\mathcal}X_1}(L_1){\operatorname{\stackrel{\mathbf L}{\otimes}}_S}D_{{\mathcal}X_2}(L_2)))\\ @>>> D_{{\mathcal}Y}(f_{1*}D_{{\mathcal}X_1}(L_1){\operatorname{\stackrel{\mathbf L}{\otimes}}_S}(f_{2*}D_{{\mathcal}X_2}(L_2)))\\ @>\simeq >> D_{{\mathcal}Y_1}(f_{1*}D_{{\mathcal}X_1}(L_1)){\operatorname{\stackrel{\mathbf L}{\otimes}}_S}D_{{\mathcal}Y_2}(f_{2*}D_{{\mathcal}X_2}(L_2))\\ @>\simeq >> Rf_{1!}(L_1){\operatorname{\stackrel{\mathbf L}{\otimes}}_S}Rf_{2!}(L_2). \end{CD}$$ That this map is an isomorphism follows from a standard reduction to the case of schemes using hypercovers of ${\mathcal}X_i$, biduality, and the spectral sequences \[5.16\]. $\square$ [10]{} . Société Mathématique de France 2003. Séminaire de Géométrie Algébrique du Bois-Marie 1960–1961 (SGA 1), Dirigé par A. Grothendieck. Augmenté de deux exposés de Mme M. Raynaud, Documents Mathématiques, Vol. 3. . Springer-Verlag, Berlin, 1972. Séminaire de Géométrie Algébrique du Bois-Marie 1963–1964 (SGA 4), Dirigé par M. Artin, A. Grothendieck, et J. L. Verdier. Avec la collaboration de N. Bourbaki, P. Deligne et B. Saint-Donat, Lecture Notes in Mathematics, Vol. 269. . Springer-Verlag, Berlin, 1972. Séminaire de Géométrie Algébrique du Bois-Marie 1963–1964 (SGA 4), Dirigé par M. Artin, A. Grothendieck et J. L. Verdier. Avec la collaboration de N. Bourbaki, P. Deligne et B. Saint-Donat, Lecture Notes in Mathematics, Vol. 270. . Springer-Verlag, Berlin, 1973. Séminaire de Géométrie Algébrique du Bois-Marie 1963–1964 (SGA 4), Dirigé par M. Artin, A. Grothendieck et J. L. Verdier. Avec la collaboration de P. Deligne et B. Saint-Donat, Lecture Notes in Mathematics, Vol. 305. . Springer-Verlag, Berlin, 1977. Séminaire de Géometrie Algébrique du Bois-Marie 1965–1966 (SGA 5), Edité par Luc Illusie, Lecture Notes in Mathematics, Vol. 589. Kai A. Behrend. Derived [$l$]{}-adic categories for algebraic stacks. , 163(774), 2003. A. A. Be[ĭ]{}linson, J. Bernstein, and P. Deligne. Faisceaux pervers. In [*Analysis and topology on singular spaces, I (Luminy, 1981)*]{}, volume 100 of [*Astérisque*]{}, pages 5–171. Soc. Math. France, Paris, 1982. Marcel Bökstedt and Amnon Neeman. Homotopy limits in triangulated categories. , 86(2):209–234, 1993. P. Deligne. . Springer-Verlag, Berlin, 1977. Séminaire de Géométrie Algébrique du Bois-Marie SGA 4${1\over 2}$, Avec la collaboration de J. F. Boutot, A. Grothendieck, L. Illusie et J. L. Verdier, Lecture Notes in Mathematics, Vol. 569. Eberhard Freitag and Reinhardt Kiehl. , volume 13 of [*Ergebnisse der Mathematik und ihrer Grenzgebiete (3) \[Results in Mathematics and Related Areas (3)\]*]{}. Springer-Verlag, Berlin, 1988. Translated from the German by Betty S. Waterhouse and William C. Waterhouse, With an historical introduction by J. A. Dieudonné. Edward Frenkel, Dennis Gaitsgory, and Kari Vilonen, On the geometric Langland’s conjecture, *J. Amer. Math. Soc.*, 15, 367-417, 2002. Grivel, P.-P., Catégories dérivées et foncteurs dérivés, In *Algebraic [$D$]{}-modules* edited by Borel, A., [Perspectives in Mathematics]{}, [**2**]{}, [Academic Press Inc.]{}, [Boston, MA]{}, [1987]{} Y. Laszlo and M. Olsson, *The six operations for sheaves on Artin stacks II: Adic Coefficients*, in preparation. Y. Laszlo and M. Olsson, *Perverse sheaves on Artin stacks*, in preparation. Gérard Laumon, Transformation de [F]{}ourier homogène. *Bull. Soc. Math. France*, [131]{} [(4)]{},[527–551]{}, [2003]{}. Gérard Laumon and Laurent Moret-Bailly. , volume 39 of [*Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics*]{}. Springer-Verlag, Berlin, 2000. Gérard Laumon and Bao Chau Ngo, Le lemme fondamental pour les groupes unitaires, *preprint*, 2004. A. Neeman, *The Grothendieck duality theorem via Bousfield’s techniques and Brown representability*, J. Amer. Math. Soc. **9** (1996), 205–236. Martin Olsson. Sheaves on artin stacks. , 2005. C. Serpét Resolution of unbounded complexes in [G]{}rothendieck categories. , 177(1):103–112, 2003. Jean-Pierre Serre. , volume 5 of [*Lecture Notes in Mathematics*]{}. Springer-Verlag, Berlin, fifth edition, 1994. N. Spaltenstein. Resolutions of unbounded complexes. , 65(2), 1988. [^1]: In fact our method could apply to other situations like analytic stacks or non separated analytic varieties. [^2]: We will often write $f^*,f^!$, $f_*$, $f_!$ for $Lf^*,Rf^!$, $Rf_*$, $Rf_!$. [^3]: Cf. \[LETdata\] below [^4]: One could replace simplicial by multisimplicial [^5]: Probably one can assume only that $w$ quasi-separated, cf. [@SGA43], XVII.7; but we do not need this more general version.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Perhaps the most intriguing result of Planck’s dust-polarization measurements is the observation that the power in the E-mode polarization is twice that in the B mode, as opposed to pre-Planck expectations of roughly equal dust powers in E and B modes. Here we show how the E- and B-mode powers depend on the detailed properties of the fluctuations in the magnetized interstellar medium. These fluctuations are classified into the slow, fast, and Alfvén magnetohydrodynamic (MHD) waves, which are determined once the ratio $\beta$ of gas to magnetic-field pressures is specified. We also parametrize models in terms of the power amplitudes and power anisotropies for the three types of waves. We find that the observed EE/BB ratio (and its scale invariance) and positive TE correlation cannot be easily explained in terms of favored models for MHD turbulence. The observed power-law index for temperature/polarization fluctuations also disfavors MHD turbulence. We thus speculate that the $\sim$0.1–30 pc length scales probed by these dust-polarization measurements are not described by MHD turbulence but, rather, probe the large-scale physics that drives ISM turbulence. We develop a simple phenomenological model, based on random displacements of the magnetized fluid, that produces EE/BB $\simeq2$ and a positive TE cross-correlation. According to this model, the EE/BB and TE signals are due to longitudinal, rather than transverse, modes in the random-displacement field, providing, perhaps, some clue to the mechanism that stirs the ISM. Future investigations involving the spatial dependence of the EE/BB ratio, TE correlation, and local departures from statistical isotropy in dust-polarization maps, as well as further tests of some of the assumptions in this analysis, are outlined. This work may also aid in the improvement of foreground-separation techniques for studies of cosmic microwave background polarization.' author: - 'Robert R. Caldwell$^1$, Chris Hirata$^2$, and Marc Kamionkowski$^3$' title: 'Dust-polarization maps and interstellar turbulence' --- Introduction {#sec:intro} ============ The Planck satellite has provided an extraordinary trove of detailed information on polarized emission from dust in the interstellar medium (ISM) of the Milky Way [@Ade:2014gna], with precise power spectra measured over the multipole-moment range $30 \lesssim \ell \lesssim 600$ [@Adam:2014bub]. Since the polarization of the dust emission arises from the alignment of spinning dust grains with the magnetic field [@Chandrasekhar:1953zza; @Stein:1966; @Dolginov:1972; @Dolginov:1976; @Draine:1996nd; @Draine:1996hn; @Finkbeiner:2004je; @Draine:2008hu; @Andersson:2015], the measurements are particularly important for the magnetic-field structure of the ISM. Perhaps the most surprising result from Planck is the discovery that the E-mode power in the dust polarization is twice the B-mode power [@Adam:2014bub]. (Something similar was noticed in WMAP, albeit with less significance, with synchrotron polarization, @Page:2006hz). The linear-polarization pattern can be decomposed geometrically into two rotational invariants, the E (gradient) modes and B (curl) modes [@Kamionkowski:1996ks; @Zaldarriaga:1996xe]. A randomly oriented polarization map should have equal E- and B-mode powers. Likewise, if polarization fluctuations arise as amplitude fluctuations with a fixed orientation, then the E- and B-mode powers should be equal [@Zaldarriaga:2001st; @Kamionkowski:2014wza]. The state-of-the-art pre-Planck dust-polarization models [@ODea:2011kx; @Delabrouille:2012ye] therefore all had equal E- and B-mode powers. The observed EE/BB $\simeq2$ ratio thus comes as quite a surprise. Planck also finds a cross-correlation (of positive sign) between the temperature and the E-mode component of polarization, an empirical fact that we will also employ below. Here we show how the observed EE/BB $\simeq2$ ratio depends on the detailed properties of magnetized-fluid fluctuations in the ISM. Fluctuations in a magnetized plasma are described most generally by the slow, fast, and Alfvén MHD waves; there is one for each Fourier wavevector ${\ensuremath {\mathbf{k}}}$. Models of MHD turbulence predict the power spectra for these different types of modes as a function of the magnitude and orientation (with respect to the background magnetic field) of the wavevector ${\ensuremath {\mathbf{k}}}$ [@Cho:2002qn; @Elmegreen:2004wj; @Brandenburg:2013vya; @Schekochihin:2007mw]. A vigorous effort, based on analytic arguments and numerical simulations, is afoot to nail down these predictions, with much of the effort tracing back to classic work by @Iroshnikov and @Kraichnan:1965zz and later @Shebalin:1983zz, and more recently, for example, @Goldreich:1994zz, @Lithwick:2001it, and @Cho:2002qi. The Planck Collaboration observed that correlations of filamentary structures [@Ade:2015mbc; @Adam:2014gaa] with fluctuations in the magnetic-field orientation could account for the observed ratio. The Planck Collaboration made further contact with MHD-turbulence models for the ISM in @Ade:2014hna and @Aghanim:2016uao through measurement of distributions of polarization magnitudes and orientation angles. This work does not, however, explain how the relevant density–magnetic-field correlations arise in terms of the fundamental modes of fluctuations in the magnetized fluid. There is thus room to make clearer contact with theoretical models for a magnetized fluid. Below we calculate the E- and B-mode amplitudes induced by slow, fast, and Alfvén waves for different directions of the background magnetic field with respect to the line of sight and for different wavevectors ${\ensuremath {\mathbf{k}}}$. Since the EE/BB $\simeq2$ ratio seems to be relatively generic across the sky, it must arise after averaging over all magnetic-field orientations. We thus then calculate the E and B power-spectrum amplitudes, as well as the temperature-polarization cross-correlation, obtained after averaging over all magnetic-field and ${\ensuremath {\mathbf{k}}}$ orientations. We provide results as a function of the ratio $\beta\equiv P_g/P_H$ of the gas and magnetic-field pressures, $P_g$ and $P_H$, respectively, and for a parameter $\lambda$ that describes the anisotropy of the slow, fast, and Alfvén waves. These calculations can then be used to assess the validity of any particular model for MHD turbulence specified by the power in the slow, fast, and Alfvén waves, and the anisotropy of that power. Our results suggest that for $\beta \gtrsim 1$, the observed EE/BB ratio and temperature-polarization cross correlation can be explained only if the power in fast waves greatly exceeds that in slow/Alfvén waves, and moreover, only if those fast waves have a nearly isotropic spectrum. The observations can also be explained in a low-$\beta$ (strong-field) plasma with an additional contribution from an anisotropic spectrum of Alfvén waves, but only if the slow waves are very anisotropic or somehow suppressed. We thus infer that the oberved EE/BB and TE are in tension with expectations from MHD turbulence. The apparent scale invariance of the EE/BB ratio over the range $\ell \simeq 30-600$ and the spectral index of the fluctuations—which disagrees with that expected from turbulence and that seen in electron-density fluctuations on smaller scales [@Armstrong:1995]—are also not easily accommodated by current MHD-turbulence models. We thus speculate that the $\sim$ 0.1–30 pc length scales probed by Planck may overlap the outer scale of turbulence, the largest distance scale on which turbulence is driven. (Alternatively, there may be new physics—e.g., associated with the multiphase nature of the ISM [@Norman:1996ba; @Kritsuk:2001cm]—that is not included in the MHD-turbulence models.) We then develop a simple phenomenological model, based on random displacements of a magnetized fluid, that accounts for EE/BB $\simeq2$ and TE $>0$. We further show that the TE correlation and large EE/BB are a consequence primarily of the longitudinal, rather than transverse, modes in the random-displacement field. We surmise that this may indicate something about the physics—perhaps stellar winds, protostellar outflows, supernovae [@Lacki:2013nda; @Padoan:2016], or Galactic spiral shocks [@Kim:2006ny]—that drives small-scale turbulence in the ISM. Directions for future related research include improved measurement of the Planck TE cross-correlation coefficient calculated here; studies of the variation of EE/BB and TE (that arise from variations in the background-magnetic-field orientation) across the sky; searches for local departures from statistical isotropy that arise for the same reason; and more precise measurements of the $\ell$ dependence of the dust power spectra. Moreover, as discussed below, we assume here that the dust density traces the plasma density, a hypothesis that we argue is reasonable, although one whose validity requires further investigation. There are thus further studies that should be done–including the frequency dependence of the E/B/T maps, cross-correlation with synchrotron-polarization maps, and perhaps cross-correlation with polarized-starlight surveys—to test further this hypothesis. Finally, a better understanding of the physics responsible for polarized dust emission may also aid in the development of algorithms to separate the CMB-polarization signal from polarized dust emission [@Dunkley:2008am] and thus help advance the quest for inflationary gravitational waves [@Kamionkowski:1996zd; @Seljak:1996gy; @Kamionkowski:2015yta]. Such developments must not necessarily await the next flagship satellite mission: there are prospects for considerable improvements in dust-polarization maps on small patches of sky with suborbital experiments [@Kovetz:2015pia] such as BLASTPol [@Fissel:2010aa], BFORE [@Niemack:2015qta], TOLTEC [@Wilson:2016], or PILOT [@Misawa:2014hka]. Measurements of Galactic synchrotron and/or dust polarization on larger angular scales will be improved, for example, with CLASS [@Essinger-Hileman:2014pja] or LiteBird [@Matsumura:2013aja]. Analyses similar to those we discuss can also be applied to maps of starlight polarization [@Goodman:1990; @Heiles:1996; @Fosalba:2001wr] or neutral-hydrogen filaments [@Clark:2015cpa], although the polarization strength is small, and the sparse sampling and the range of distances to stars complicates the E/B mode analysis. Moreover, similar analyses may be employed to understand, with dust-polarization maps, magnetic-field structure in specific molecular clouds [@Pelkonen:2006fp; @Kataoka:2012ci; @Koch:2013rea; @Soler:2013kga]. This paper is organized as follows: In Section \[sec:EB\] we review the E/B decomposition of a polarization map. We review the relevant properties of MHD waves in Section \[sec:mhd\]. Section \[sec:EBmodes\] calculates the E and B amplitudes that arise from slow, fast, and Alfvén waves. Section \[sec:powerspectra\] discusses calculation of the power spectra. Section \[sec:results\] presents the results of the calculations. In Section \[sec:interpretation\] we provide some possible interpretations of the data in terms of MHD-turbulence models and also discuss the tension with expectations from favored MHD-turbulence models. We therefore consider, in Section \[sec:rand\], a simple phenomenological model of random displacements in a magnetized fluid that results in EE/BB $\simeq2$ and TE $>0$. We then conclude and enumerate several further research directions in Section \[sec:conclusions\]. To avoid confusion with the E/B decomposition of polarization maps, we use ${\ensuremath {\mathbf{H}}}$ to denote the magnetic field. The c.g.s. system of units is used. Review of the E/B decomposition of a polarization map and projection effects {#sec:EB} ============================================================================ Here we recall some basic properties of the decomposition of a polarization map into E and B modes, and the way in which 3-dimensional emitting structures appear on the 2-dimensional sky. We consider a map of the linear polarization on a patch of sky sufficiently small to be assumed flat, and of solid angle $\Omega$. We assume the emission to be optically thin, which is a good approximation at microwave frequencies. The polarization is specified in terms of Stokes parameters $Q({\ensuremath {\boldsymbol{\theta}}})$ and $U({\ensuremath {\boldsymbol{\theta}}})$, measured with respect to some ${\ensuremath {\boldsymbol{\hat \theta}}}_x$-${\ensuremath {\boldsymbol{\hat \theta}}}_y$ axes in the plane of the sky, which can then be written as a complex polarization $\Pi({\ensuremath {\boldsymbol{\theta}}})=Q({\ensuremath {\boldsymbol{\theta}}})+i U({\ensuremath {\boldsymbol{\theta}}})$.[^1] The map is equivalently represented by the Fourier transform, $$\tilde \Pi({\ensuremath {\boldsymbol{\ell}}}) = \int_\Omega \, d^2{\ensuremath {\boldsymbol{\theta}}}\, \Pi({\ensuremath {\boldsymbol{\theta}}}) e^{- i {\ensuremath {\boldsymbol{\ell}}}\cdot {\ensuremath {\boldsymbol{\theta}}}}. \label{eqn:Fourier}$$ The density of Fourier modes in the 2-dimensional ${\ensuremath {\boldsymbol{\ell}}}$-plane is $\Omega/(2\pi)^2$. The Stokes parameters, and the complex polarization, are not rotational invariants; under a rotation of the coordinate axes by an angle $\alpha$, the polarization transforms as $\Pi \to \Pi e^{2 i \alpha}$. The polarization field can be represented in terms of rotational invariants $E$ and $B$. In Fourier space these are $$(\tilde E + i \tilde B)({\ensuremath {\boldsymbol{\ell}}}) = (\tilde Q+i \tilde U)({\ensuremath {\boldsymbol{\ell}}}) e^{-2i \psi_{{\ensuremath {\boldsymbol{\ell}}}}} \label{eqn:EB}$$ [@Kamionkowski:1996zd; @Seljak:1996gy; @Kamionkowski:1996ks; @Zaldarriaga:1996xe; @Seljak:1996ti; @Cabella:2004mk; @Kamionkowski:2015yta], where $\psi_{{\ensuremath {\boldsymbol{\ell}}}}$ is the angle that ${\ensuremath {\boldsymbol{\ell}}}$ makes with ${\ensuremath {\boldsymbol{\hat \theta}}}_x$, i.e. $\tan\psi_{{\ensuremath {\boldsymbol{\ell}}}} = \ell_y/\ell_x$. The power spectra measured by Planck are then $C_\ell^{\rm EE} = \langle{|\tilde E({\ensuremath {\boldsymbol{\ell}}})|^2}\rangle/\Omega$ and $C_\ell^{\rm BB} = \langle{|\tilde B({\ensuremath {\boldsymbol{\ell}}})|^2}\rangle/\Omega$, where the average is over all ${\ensuremath {\boldsymbol{\ell}}}$ of magnitude $\ell$.[^2] The observed polarization signal $\Pi$ is typically measured in units of $\mu$K$_{\rm CMB}$, and its angular power spectra $C_\ell^{\rm EE/BB}$ have units of $\mu$K$_{\rm CMB}^2$. However, for optically thin emission, the polarization is related to the polarized emissivity $\varepsilon_\Pi$ via $$\Pi({\ensuremath {\boldsymbol{\theta}}}) = \int_0^\infty \varepsilon_\Pi(r\hat{\mathbf n}({\ensuremath {\boldsymbol{\theta}}}))\,dr, \label{eq:Pi-theta}$$ where $\hat{\mathbf n}({\ensuremath {\boldsymbol{\theta}}})$ is the 3-dimensional unit vector in the direction corresponding to angular position ${\ensuremath {\boldsymbol{\theta}}}$. The emissivity $\varepsilon_\Pi$ (and its components, $\varepsilon_Q$ and $\varepsilon_U$) have units of $\mu$K$_{\rm CMB}$pc$^{-1}$, and its 3-dimensional power spectra $P_{\varepsilon,\rm EE}({\mathbf k})$ and $P_{\varepsilon,\rm BB}({\mathbf k})$ have units of $[\varepsilon_P^2]\times\,$\[volume\], or $\mu$K$_{\rm CMB}^2$pc. For small angles or $\ell\gg 2$, the relation of 3-dimensional and 2-dimensional power spectra is usually obtained via the Limber approximation. This begins with breaking the line-of-sight integral, Eq. (\[eq:Pi-theta\]), into a series of boxes along the line of sight of width $\Delta r_i$. In each box, the emissivity can be Fourier-transformed to $\tilde\varepsilon_\Pi({\mathbf k})$: $$\tilde\varepsilon_\Pi({\mathbf k}) =\! \int_V \! \varepsilon_\Pi({\mathbf x})\,e^{-i{\mathbf k}\cdot{\mathbf x}}\,d^3{\mathbf x} \;\leftrightarrow\; \varepsilon_\Pi({\mathbf x}) = \sum_{\mathbf k} \tilde\varepsilon_\Pi({\mathbf k})\,e^{i{\mathbf k}\cdot{\mathbf x}},$$ where the Fourier wave vector ${\mathbf k}$ has (i) a transverse component ${\mathbf k}_\perp$ with a density of modes $r^2\Omega/(2\pi)^2$, and (ii) a line-of-sight component $k_\parallel = 2\pi n/\Delta r_i$ with $n\in{\mathbb Z}$. The volume of the box is $V = r^2\Omega \Delta r_i$. These transformed quantities satisfy $$\begin{aligned} \langle \tilde\varepsilon_E^\ast({\mathbf k}) \tilde\varepsilon_E({\mathbf k}') \rangle &=& (2\pi)^3 \delta^{\rm D}({\mathbf k}-{\mathbf k}') P_{\varepsilon,\rm EE}({\mathbf k}) \nonumber \\ &=& r^2\Omega\Delta r_i\,\delta^{\rm K}_{{\mathbf k},{\mathbf k}'}P_{\varepsilon,\rm EE}({\mathbf k}). \label{eq:Pow3D}\end{aligned}$$ Only the transverse ($n=0$ or $k_\parallel=0$) modes, i.e. those with ${\mathbf k}$ in the plane of the sky, survive radial integration. They relate to the projected polarization via $$\tilde\Pi({\ensuremath {\boldsymbol{\ell}}}) = \sum_i \frac1{r^2} \tilde\varepsilon_\Pi({\mathbf k} = {\ensuremath {\boldsymbol{\ell}}}/r), \label{eq:box-sum}$$ where the $1/r^2$ comes from the transformation from $d^2{\ensuremath {\boldsymbol{\theta}}}$ to $d^2{\mathbf x}_\perp$ in the Fourier integral (see Eq. (\[eqn:Fourier\])), and from Eq. (\[eq:Pow3D\]) the 2-dimensional power spectrum is $$C_\ell^{\rm EE} \approx \sum_i \frac{P_{\varepsilon,\rm EE}(k=\ell/r)}{r^2} \Delta r_i \approx \int_0^{r_{\rm max}} \frac{P_{\varepsilon,\rm EE}(k=\ell/r)}{r^2}\,dr, \label{eq:Limber}$$ where $r_{\rm max}$ is the maximum distance from which dust emission is seen. Eq. (\[eq:Limber\]) is the Limber equation, as commonly used in cosmology. The derivation contains two subtle assumptions: (i) each box can be treated as a statistically homogeneous medium; and (ii) when squaring Eq. (\[eq:box-sum\]) and taking the expected value, we can neglect correlations between different boxes $i\neq j$. In most of this paper, we will focus our attention on the ratios of the power spectra, $P_{\varepsilon,\rm EE}(k)/P_{\varepsilon,\rm BB}(k)$, or correlation coefficients between the E-mode and temperature $r = P_{\varepsilon,\rm TE}(k)/[P_{\varepsilon,\rm TT}(k)P_{\varepsilon,\rm EE}(k)]^{1/2}$. It is easily seen from Eq. (\[eq:Limber\]) that the corresponding ratio in the power spectrum, $C_\ell^{\rm BB}/C_\ell^{\rm EE}$, is a suitably weighted average of $P_{\varepsilon,\rm EE}(k)/P_{\varepsilon,\rm BB}(k)$ along the line of sight. Therefore, in attempting to explain the observed EE/BB ratio, we focus on the 3-dimensional power spectrum. When we consider the scale dependence of the polarization power spectrum, we will have to return to the full version of Eq. (\[eq:Limber\]). Magnetohydrodynamic Waves {#sec:mhd} ========================= A compressible magnetized plasma can, in the MHD limit, carry three different types of waves, linear combinations of the two transverse-vector components of the magnetic field ${\ensuremath {\mathbf{H}}}$ (since the requirement $\mathbf{\nabla} \cdot {\ensuremath {\mathbf{H}}}=0$ removes the longitudinal-vector degree of freedom) and the plasma-density degree of freedom. Here we briefly reprise the properties, relevant for this work, of these three MHD waves, which are classified into Alfvén, slow, and fast modes. We consider a magnetized plasma at rest with a homogeneous magnetic field ${\ensuremath {\mathbf{H}}}_0$ and then consider small perturbations parametrized in terms of a magnetic-field perturbation $\delta {\ensuremath {\mathbf{H}}}({\ensuremath {\mathbf{x}}},t)$ and plasma velocity ${\ensuremath {\mathbf{v}}}({\ensuremath {\mathbf{x}}},t)$. In the MHD limit, the perturbation, velocity, and background field are related (in Fourier space) by $$\omega \delta {\ensuremath {\mathbf{H}}}= -{\ensuremath {\mathbf{k}}}\times ({\ensuremath {\mathbf{v}}}\times {\ensuremath {\mathbf{H}}}_0), \label{eqn:mhd}$$ where here $\delta{\ensuremath {\mathbf{H}}}$ and ${\ensuremath {\mathbf{v}}}$ are taken to be the magnetic-field and velocity amplitudes of this particular Fourier mode. Alfvén waves ------------ The Alfvén wave has a velocity perpendicular to both ${\ensuremath {\mathbf{k}}}$ and ${\ensuremath {\mathbf{H}}}$, and it has a dispersion relation $\omega = \pm a k \cos\alpha$, where $a= H_0 (4\pi \rho)^{-1/2}$ is the Alfvén speed (and $\rho$ the plasma mass density), and $\cos\alpha = {\ensuremath {\mathbf{\hat k}}}\cdot {\ensuremath {\mathbf{\hat H}}}_0$. For this wave, $\delta {\ensuremath {\mathbf{H}}}= \pm H_0 ({\ensuremath {\mathbf{v}}}/a)$. The continuity equation, $(\partial n/\partial t) + \nabla \cdot( n {\ensuremath {\mathbf{v}}})=0$, provides a relation, $(\delta n /n_0) = {\ensuremath {\mathbf{k}}}\cdot {\ensuremath {\mathbf{v}}}/ \omega$, between the fractional density perturbation $(\delta n/n_0)$ and the velocity. Since ${\ensuremath {\mathbf{k}}}\perp {\ensuremath {\mathbf{v}}}$ in the Alfvén wave, these waves have no associated density perturbation. We thus write, $$\delta {\ensuremath {\mathbf{H}}}= -\frac{v H_0}{a} {\ensuremath {\mathbf{\hat a}}}, \label{eqn:Alfvenamplitude}$$ where ${\ensuremath {\mathbf{\hat a}}}\equiv {\ensuremath {\mathbf{\hat k}}}\times {\ensuremath {\mathbf{\hat H}}}/\sin\alpha$ is the unit vector perpendicular to ${\ensuremath {\mathbf{k}}}$ and ${\ensuremath {\mathbf{H}}}$. Slow/fast waves --------------- The slow and fast waves both have magnetic-field perturbations in a direction ${\ensuremath {\boldsymbol{\hat \theta}}}= -{\ensuremath {\mathbf{\hat k}}}\times ({\ensuremath {\mathbf{\hat k}}}\times {\ensuremath {\mathbf{\hat H}}})/\sin\alpha$ perpendicular to ${\ensuremath {\mathbf{\hat k}}}$ and ${\ensuremath {\mathbf{\hat a}}}$. The slow wave has a displacement in direction ${\ensuremath {\boldsymbol{\hat \xi}}}_s \propto \cos\alpha {\ensuremath {\mathbf{\hat H}}}+ \zeta_s \sin\alpha {\ensuremath {\mathbf{\hat k}}}_\perp$, where ${\ensuremath {\mathbf{\hat k}}}_\perp$ is a unit vector in the ${\ensuremath {\mathbf{k}}}$-${\ensuremath {\mathbf{H}}}$ plane perpendicular to ${\ensuremath {\mathbf{\hat H}}}$, and the fast-wave is in the orthogonal direction, ${\ensuremath {\boldsymbol{\hat \xi}}}_f \propto \zeta_f \cos \alpha {\ensuremath {\mathbf{\hat H}}}+ \sin\alpha {\ensuremath {\mathbf{\hat k}}}_\perp$. Here, $$\begin{aligned} \zeta_s &= &\frac{1-\sqrt{D} -\beta/2}{1+\sqrt{D} +\beta/2} \cot^2\alpha, \nonumber \\ \zeta_f &= &\frac{1-\sqrt{D} +\beta/2}{1+\sqrt{D} -\beta/2} \tan^2\alpha, \label{eqn:zetas}\end{aligned}$$ where $D = (1+\beta/2)^2 - 2\beta\cos^2\alpha$, and $\beta = P_g/P_H$ is the ratio of gas pressure to magnetic-field pressure. In the strong-field limit $\beta \to 0$, and $\beta\to\infty$ in the weak-field limit. From Eq. (\[eqn:mhd\]) it follows that for the slow wave, $$\delta {\ensuremath {\mathbf{H}}}= \frac{kv H_0}{\omega} \frac{\zeta_s \sin\alpha}{ \left( \cos^2\alpha + \zeta_s^2\sin^2\alpha \right)^{1/2}} {\ensuremath {\boldsymbol{\hat \theta}}}, \label{eqn:slowdeltaH}$$ and for the fast wave, $$\delta {\ensuremath {\mathbf{H}}}= \frac{kv H_0}{\omega} \frac{\sin\alpha}{ \left( \zeta_f^2 \cos^2\alpha + \sin^2\alpha \right)^{1/2}} {\ensuremath {\boldsymbol{\hat \theta}}}, \label{eqn:fastdeltaH}$$ where $v$ is the magnitude of the fluid velocity. For the Alfvén wave, the relationship between the magnitudes of the magnetic-field and velocity perturations is independent of the orientation of ${\ensuremath {\mathbf{k}}}$ \[cf. Eq. (\[eqn:Alfvenamplitude\])\]. The same is not true, however, for the slow/fast waves. In addition to the explicit $\alpha$ dependence in Eqs. (\[eqn:slowdeltaH\])–(\[eqn:fastdeltaH\]), there is also an $\alpha$ dependence in $\zeta_{s,f}$ and also in the dispersion relations, $$\left( \frac{ \omega}{k} \right)^2 = \frac{a^2}{2} (1 +\beta/2) \left[ 1 \pm \left( 1 - \frac{ 2 \beta \cos^2\alpha}{(1+\beta/2)^2} \right)^{1/2} \right],$$ for the fast (plus sign) and slow (minus sign) waves. The fractional density perturbation is then found from the continuity equation to be, for the slow wave, $$\frac{\delta n}{n_0} = \frac{kv}{\omega} \frac{\left( \cos^2\alpha + \zeta_s \sin^2\alpha \right)}{ \left( \cos^2\alpha + \zeta_s^2\sin^2\alpha \right)^{1/2}}, \label{eqn:slowdeltan}$$ and for the fast wave, $$\frac{\delta n}{n_0} = \frac{kv}{\omega} \frac{\left( \zeta_f\cos^2\alpha + \sin^2\alpha \right)}{ \left( \zeta_f^2 \cos^2\alpha + \sin^2\alpha \right)^{1/2}}. \label{eqn:fastdeltan}$$ The final relations then are those between the magnetic-field perturbation and the density perturbation, and the magnetic-field perturbation and the velocity perturbation. They are, for the slow wave, $$\begin{aligned} \frac{\delta n}{n_0} &=& \frac{|\delta{\ensuremath {\mathbf{H}}}|}{H_0} \left( \cos^2\alpha + \zeta_s\sin^2\alpha \right) / (\zeta_s \sin\alpha) \equiv \frac{|\delta{\ensuremath {\mathbf{H}}}|}{H_0} g_s(\alpha), \nonumber\\ \label{eqn:slowrelation}\\ \frac{|\delta{\ensuremath {\mathbf{H}}}|}{H_0} &=& \frac{k}{\omega} \frac{\zeta_s \sin\alpha}{ \left(\cos^2\alpha + \zeta_s^2 \sin^2\alpha \right)^{1/2}} |\mathbf{v}|\equiv |\mathbf{v}| h_s (\alpha) , \label{eqn:slowrelationV}\end{aligned}$$ and for the fast wave, $$\begin{aligned} \frac{\delta n}{n_0} &=& \frac{|\delta{\ensuremath {\mathbf{H}}}|}{H_0} \left( \zeta_f \cos^2\alpha + \sin^2\alpha \right) / \sin\alpha \equiv \frac{|\delta{\ensuremath {\mathbf{H}}}|}{H_0} g_f(\alpha), \nonumber\\ \label{eqn:fastrelation}\\ \frac{|\delta{\ensuremath {\mathbf{H}}}|}{H_0} &=& \frac{k}{\omega} \frac{\sin\alpha}{ \left(\zeta_f^2 \cos^2\alpha + \sin^2\alpha \right)^{1/2}} |\mathbf{v}| \equiv |\mathbf{v}| h_f(\alpha) . \label{eqn:fastrelationV}\end{aligned}$$ In the case of the Alfvén wave, as seen in Eq. (\[eqn:Alfvenamplitude\]), we have ${|\delta{\ensuremath {\mathbf{H}}}|}/{H_0} = |\mathbf{v}|/a \equiv |\mathbf{v}| h_a$. These relations allow us to determine the E- and B-mode powers under different assumptions about the power spectra for the different MHD waves. E and B modes induced by the slow, fast, and Alfvén Waves {#sec:EBmodes} ========================================================= E and B amplitudes from a single Fourier mode --------------------------------------------- Take the line of sight to be along the $z$ axis and the background field ${\ensuremath {\mathbf{H}}}_0 = H_0 (\sin\theta,0,\cos\theta)$ in the $x$-$z$ plane at an angle $\theta$ from the line of sight. Consider a perturbation of wavevector ${\ensuremath {\mathbf{k}}}= k {\ensuremath {\mathbf{\hat k}}}= k(\cos\psi,\sin\psi,0)$ in the $x$-$y$ plane of the sky (as the two-dimensional projections of other modes will experience a Limber suppression) oriented at an angle $\psi$ with respect to the $x$ axis. The angle $\alpha$ between ${\ensuremath {\mathbf{k}}}$ and ${\ensuremath {\mathbf{H}}}$ is then given by $\cos\alpha = \sin\theta \cos\psi$, as illustrated in Fig. \[fig:coords\]. We observe a two-dimensional projection of an emitting volume, and the polarized emission is assumed to have the form $$\varepsilon_P = \varepsilon_Q + i \varepsilon_U = A n H^\gamma(H_x+i H_y)^2, \label{eqn:polarization}$$ where $\gamma$ is an exponent which is equal to $-2$ if the dust alignment is independent of the magnetic-field strength, $n$ is proportional to the dust density (and has a constant background value $n_0$). Here $A<0$ is a constant; its value is taken to be negative so that the polarization is perpendicular to the magnetic field [@Chandrasekhar:1953zza]. The sign of $A$ will be significant for the temperature-polarization cross-correlation below. The polarization fluctuations are $$\delta\varepsilon_P = A n_0 H_0^{2+\gamma} \left[2 \sin\theta \,\frac{\delta H_x+ i \delta H_y}{H_0} + \gamma \sin^2\theta \,\frac{\delta H}{H_0} + \sin^2\theta \,\frac{\delta n}{n_0} \right],$$ where here $\delta H = \sin\theta \delta H_x+ \cos\theta \delta H_z$. For a given Fourier mode of wavevector ${\ensuremath {\mathbf{k}}}$ transverse to the line of sight in a box of radial width $\Delta r$, the $E$ and $B$ modes will appear in wave vector ${\ensuremath {\boldsymbol{\ell}}}= {\ensuremath {\mathbf{k}}}r$, and will have the form: $$\tilde E + i \tilde B = A n_0 H_0^{2+\gamma} \frac{\Delta r}{r^2}\,e^{-2i\psi} \left[2 \sin\theta \,\frac{\delta\tilde H_x + i \delta\tilde H_y}{H_0} + \gamma \sin^2\theta \,\frac{\delta \tilde H}{H_0} + \sin^2 \theta \,\frac{\delta\tilde n}{n_0} \right],$$ which can be decomposed into $$\begin{aligned} \tilde E &=& A n_0 H_0^{2+\gamma} \frac{\Delta r}{r^2}\, \left[ 2 \sin\theta \cos2\psi \frac{\delta \tilde H_x }{ H_0 } + 2\sin\theta \sin 2\psi \,\frac{\delta \tilde H_y}{H_0} + \gamma \sin^2\theta \cos 2\psi \,\frac{\delta \tilde H}{H_0} + \sin^2\theta \cos 2\psi \,\frac{\delta\tilde n}{n_0} \right], \nonumber \\ \label{eqn:Egen} \\ \tilde B &=& A n_0 H_0^{2+\gamma} \frac{\Delta r}{r^2}\, \left[ -2 \sin\theta \sin 2\psi \,\frac{\delta\tilde H_x }{ H_0}+ 2\sin\theta \cos 2\psi \,\frac{\delta\tilde H_y}{H_0} - \gamma \sin^2\theta \sin 2\psi \,\frac{\delta\tilde H}{H_0} - \sin^2\theta \sin 2\psi \,\frac{\delta\tilde n}{n_0} \right]. \nonumber\\ \label{eqn:Bgen}\end{aligned}$$ We now re-write the magnetic-field perturbations in terms of the two transverse-vector modes, those in the ${\ensuremath {\mathbf{\hat a}}}$ (the Alfvén wave) and ${\ensuremath {\boldsymbol{\hat \theta}}}$ (the slow and fast waves) directions: $$\begin{aligned} (\delta \tilde {\ensuremath {\mathbf{H}}})_a &=& {\ensuremath {\mathbf{\hat a}}}\delta \tilde H_a = \frac{{\ensuremath {\mathbf{\hat k}}}\times {\ensuremath {\mathbf{\hat H}}}}{\sin \alpha} \delta \tilde H_a = \frac{ (\sin\psi\cos\theta, -\cos\psi \cos\theta, -\sin\psi\sin\theta)}{\sin \alpha} \delta\tilde H_a,\nonumber \\ (\delta\tilde {\ensuremath {\mathbf{H}}})_p &=& {\ensuremath {\boldsymbol{\hat \theta}}}\delta \tilde H_p = -\frac{{\ensuremath {\mathbf{\hat k}}}\times ({\ensuremath {\mathbf{\hat k}}}\times {\ensuremath {\mathbf{\hat H}}})}{\sin\alpha} \delta \tilde H_p = \frac{ (\sin\theta \sin^2\psi,-\sin\theta \sin\psi\cos\psi,\cos\theta)}{\sin\alpha} \delta \tilde H_p,\end{aligned}$$ where $\delta \tilde H_a$ (‘a’ for Alfvén) and $\delta \tilde H_p$ (‘p’ for pseudo-Alfvén) are the magnetic-field amplitudes for the two modes. These then translate to $E$ and $B$ modes, $$\begin{aligned} \tilde E &=& A n_0H_0^{2+\gamma} \frac{\Delta r}{r^2} \left[ -\sin2\theta \frac{\sin\psi}{\sin\alpha} \frac{\delta \tilde H_a}{H_0} + \frac{ \sin^2\theta \left[ -2\sin^2\psi (1+\gamma \sin^2\alpha) + \gamma \sin^2\alpha \right]}{\sin\alpha} \frac{\delta\tilde H_p}{H_0} + \sin^2\theta \cos2\psi \frac{\delta \tilde n}{n_0} \right], \label{eqn:Eintermediate} \\ \tilde B &=& An_0 H_0^{2+\gamma} \frac{\Delta r}{r^2} \left[ -\sin2\theta \frac{\cos \psi}{\sin\alpha} \frac{\delta\tilde H_a}{H_0} - \frac{2 \sin^2\theta \sin\psi \cos\psi (1+\gamma \sin^2\alpha)}{\sin \alpha} \frac{\delta\tilde H_p}{H_0} - \sin^2\theta \sin 2\psi \frac{\delta \tilde n}{n_0} \right]. \label{eqn:Bintermediate} \end{aligned}$$ For Alfvén waves, which have no associated density perturbation, we are already done. However, the fast and slow waves both have a density perturbation. The final step is thus to re-write the $p$ and $n$ modes in terms of slow (‘s’) and fast (‘f’) modes using Eqs. (\[eqn:slowrelation\]) and (\[eqn:fastrelation\]). We then obtain $$\begin{aligned} \tilde E &=& A n_0H_0^{2+\gamma} \frac{\Delta r}{r^2} \left[ -\sin2\theta \frac{\sin\psi}{\sin\alpha} \frac{\delta\tilde H_a}{H_0} + \sum_{i=s,f} \frac{\delta \tilde H_i}{H_0} \sin^2\theta \left( \frac{ \left[ -2\sin^2\psi (1+\gamma \sin^2\alpha) + \gamma \sin^2\alpha \right]}{\sin\alpha} +g_i(\alpha) \cos2\psi \right)\right] \nonumber \\ & \equiv & A n_0 H_0^{2+\gamma} \frac{\Delta r}{r^2} \sum_{i=a,s,f} f_i^{\rm E}(\theta,\psi) \frac{\delta \tilde H_i}{H_0} = A n_0 H_0^{2+\gamma} \frac{\Delta r}{r^2} \sum_{i=a,s,f} f_i^{\rm E}(\theta,\psi) h_i(\theta,\psi) |\mathbf{v}_i|, \\ \tilde B &=& An_0 H_0^{2+\gamma} \frac{\Delta r}{r^2} \left[ -\sin2\theta \frac{\cos \psi}{\sin\alpha} \frac{\delta \tilde H_a}{H_0} - \sum_{i=s,f} \frac{\delta \tilde H_i}{H_0} \sin^2\theta \sin 2\psi \left(\frac{(1+\gamma \sin^2\alpha)}{\sin \alpha} + g_i(\alpha) \right) \right] \nonumber\\ & \equiv & A n_0 H_0^{2+\gamma}\frac{\Delta r}{r^2} \sum_{i=a,s,f} f_i^{\rm B}(\theta,\psi) \frac{\delta \tilde H_i}{H_0} = A n_0 H_0^{2+\gamma}\frac{\Delta r}{r^2} \sum_{i=a,s,f} f_i^{\rm B}(\theta,\psi) h_i(\theta,\psi) |\mathbf{v}_i|. \label{eqn:EandB}\end{aligned}$$ The intermediate lines define the angular functions $f_i^{\rm E,B}(\theta,\psi)$ which relate the polarization pattern to the magnetic field fluctuations, and the conversion into velocity fluctuations follows from Eqs. (\[eqn:slowrelationV\])–(\[eqn:fastrelationV\]). Temperature fluctuations ------------------------ The brightness temperature of the dust (synchrotron) emission is also provided, as a function of position on the sky, by Planck [@Adam:2014bub; @Ade:2014zja] (WMAP, @Page:2006hz). Since the brightness temperature of dust emission is proportional to the dust density, temperature fluctuations arise from fluctuations $\delta n$ in the dust density. The fractional intensity or temperature perturbation is thus, $$\frac{\delta \epsilon_T}{\bar\epsilon_T} = c \frac{\delta n}{n_0},$$ and projected through a box of width $\Delta r$ we have $$\tilde T({\ensuremath {\boldsymbol{\ell}}}) = c \bar\epsilon_T \frac{\Delta r}{r^2}\,\frac{\delta\tilde n({\ensuremath {\mathbf{k}}})}{n_0}.$$ We expect $c=1$ for thermal dust emission since the physical temperature of the dust grains does not depend on the gas density (it is set by radiative equilibrium). Other dust emission mechanisms, e.g. spinning dust, may depend in a complicated way on the local gas density [e.g. @Draine:1997tb; @AliHaimoud:2008dc] and hence for these we may have $c\neq 1$. Note however that our focus is on the TE cross-correlation coefficient, where $c$ cancels out. Written in terms of the wave modes, we find $$\tilde T({\ensuremath {\boldsymbol{\ell}}}) = c \bar\epsilon_T \frac{\Delta r}{r^2}\, \frac{\delta\tilde n}{n_0} = c \bar\epsilon_T \frac{\Delta r}{r^2}\, \sum_{i=s,f} g_i(\alpha) h_i(\alpha) |\mathbf{v}_i|. \label{eq:DTgh}$$ Note that the Alfvén modes do not yield any density perturbations, and hence do not contribute to $\tilde T$. ![The coordinate axes and relevant vectors are shown. The locations of $\mathbf{H}_0$ and $\mathbf{k}$ are shown by the orange and green lines, and $\mathbf{a}$ and $\hat{\theta}$ by the dashed red and blue lines, respectively. The coordinates $\alpha$ and $\varpi$ relative to $\mathbf{k}$ are also indicated. Recall that the line of sight is along the $z$ direction.\ []{data-label="fig:coords"}](AXESGRAPHK.pdf) Calculations of Power Spectra {#sec:powerspectra} ============================= We now calculate the power in E and B modes contributed by the three different types of waves. Strictly speaking, we calculate the contribution to the E- and B-mode powers at a given 3d wavenumber $k$. The observed 2d E- and B-mode powers, as a function of multipole $\ell$, are then obtained from the Limber equation which sums the contributions of wavenumbers $k=\ell/r$, from a range of distances $r$, to a given $\ell$. If, however, the EE/BB ratio is scale-independent (as we assume here and as is consistent with the measurements), then the EE/BB ratio we calculate will also be that in the observed 2d power spectrum. Similar remarks apply to the TE correlation. Parametrization of power anisotropies in the MHD waves ------------------------------------------------------ Since the background magnetic field ${\ensuremath {\mathbf{H}}}_0$ provides a preferred direction, the power spectra for the three types of MHD waves are not expected to be isotropic, but should, rather, have some $\cos\alpha$ dependence [@Shebalin:1983zz; @Goldreich:1994zz]. Here we parametrize the anisotropy as $$P_i(k,\cos\alpha) \equiv \VEV{ \left | \frac{\delta {\ensuremath {\mathbf{H}}}_i}{H_0} \right|_{{\ensuremath {\mathbf{k}}}}^2} = P_i(k) \left[h_i(\alpha) \right]^2 F_\lambda(\cos\alpha), \label{eqn:powerspectra}$$ with $$F_\lambda(\mu) =\begin{cases} (\mu^2)^\lambda, & \text{if} \quad \lambda \geq 0, \\ (1-\mu^2)^{-\lambda}, & \text{if} \quad \lambda \leq0. \end{cases} \label{eqn:flambda}$$ We work with power spectra for the magnetic-field amplitudes, but have then defined, by virtue of the $h_{s,f}(\alpha)$ in Eq. (\[eqn:powerspectra\]), the anisotropy $F_\lambda(\mu)$ relative to the velocity-perturbation amplitude. We do so to make contact with the MHD literature, wherein wave amplitudes are usually specified in terms of the velocity. With our parametrization, for $\lambda=0$ the velocity power is isotropic; for $\lambda >0$ it is weighted in modes of wavevector ${\ensuremath {\mathbf{k}}}$ parallel to ${\ensuremath {\mathbf{H}}}_0$; and for $\lambda<0$, the velocity power is weighted in modes perpendicular to ${\ensuremath {\mathbf{H}}}_0$. The EE/BB ratio --------------- Given that the EE/BB ratio seems to be roughly 2 everywhere on the sky, any MHD explanation of the EE/BB ratio must provide this ratio after averaging over all magnetic-field orientations, rather than rely on a specific orientation. There is also evidence that the angular average is warranted even along an individual line of sight: If the field direction were exactly constant along a given line of sight, then we would expect the fractional polarization for synchrotron radiation to be $\sim75\%$ [@Rybicki:1979]. Planck obtains significantly lower values (see, e.g., Fig. 22 in @Ade:2015qkp), suggesting a large dispersion in field direction even on a single line of sight. We therefore calculate the ratios $R$ of the angle-averaged E-mode and B-mode powers, induced by Alfvén, slow, and fast waves as a function of $\beta$ and the anisotropy parameter $\lambda$. The desired ratio is obtained from $$R_i(\beta,\lambda) = \frac{ \int d\Omega \left[ f_i^E(\theta,\psi) h_i(\alpha) \right]^2 F_\lambda(\cos\alpha)} { \int d\Omega \left[ f_i^B(\theta,\psi) h_i(\alpha) \right]^2 F_\lambda(\cos\alpha)}, \label{eqn:RVformula}$$ for $i=\{a,s,f\}$. Evaluation of the angular averages can be simplified by transforming to new angular coordinates $\alpha$ and $\varpi$, through $\cos\theta = \sin\alpha \cos\varpi$, $\sin\theta\sin\psi = \sin\alpha\sin\varpi$, and $\sin\theta \cos\psi = \cos\alpha$. These then are polar coordinates for the location of $\mathbf{H}_0$ about the $k$ axis, rather than the $z$ axis, as seen in Fig. \[fig:coords\]. We then integrate over $d\Omega=\sin\alpha \, d\alpha\, d\varpi$. The temperature-polarization cross-correlation ---------------------------------------------- Temperature fluctuations will arise from fluctuations in the density field, in accordance with Eq. (\[eq:DTgh\]). The Alfvén modes do not contribute to temperature fluctuations. The slow and fast modes, however, should set up a correlation between the temperature and E-mode polarization. (The TB and EB cross-correlations vanish after averaging over angles.) The relative amplitudes of the polarization and temperature fluctuations depend on a polarization fraction and the constant $c$, and so we work instead with a cross-correlation coefficient, $$r_i(\lambda) = \frac{ \int d\Omega \left[ g_i(\alpha) h_i(\alpha)\right] \left[ f_i^E(\theta,\psi) h_i(\alpha) \right] F_\lambda(\cos\alpha)} {\sqrt{ \int d\Omega \left[ g_i(\alpha) h_i(\alpha)\right]^2 F_\lambda(\cos\alpha)} \sqrt{ \int d\Omega \left[ f_i^E(\theta,\psi) h_i(\alpha)\right]^2 F_\lambda(\cos\alpha)} },$$ which corresponds to the ratio $TE/\sqrt{(TT)\, (EE)}$. Results {#sec:results} ======= The EE/BB ratio and cross-correlation coefficients are shown in Fig. \[fig:panels\] for a strong magnetic field ($\beta=0.1$), equipartition ($\beta=2$), and weak field ($\beta \gg 1$). The two observational constraints, EE/BB $\simeq2$ and TE $>0$, can be satisfied by a nearly isotropic fast mode, for a wide range of $\beta$, or by a strongly anisotropic slow mode, with $\beta \lesssim 2$. More specifically, for a fast wave with $\beta=0.1$, an isotropic spectrum ($\lambda=0$) gives EE/BB $\simeq 2$ and cross-correlation coefficient $r\simeq0.8$. For a slow wave with $\beta=0.1$, too, a strongly anisotropic spectrum with $\lambda\sim -5$ gives EE/BB $\simeq2$ and cross-correlation coefficient $r\simeq0.7$. The constraints cannot be satisfied by a pure Alfvén wave, since this incompressible mode creates no intensity fluctuation and therefore no cross correlation. ![image](MHDpanels20160601.pdf){width="7in"} All the results illustrated assume $\gamma=-2$. However, we have also examined cases in which the polarization amplitude is correlated with the magnetic field, $\gamma > -2$, as well as the inverse case, $\gamma < -2$. We find that our results for the EE/BB ratio and TE cross correlation are not strongly sensitive to the dust-alignment index in the range $-5/2 < \gamma < -3/2$. Interpretations {#sec:interpretation} =============== In this Section we try to make sense of the observations within the context of models for the ISM. We first consider MHD-turbulence models and conclude that they are unlikely to provide the whole story. We then speculate that the Planck dust-polarization data may alternatively reflect the physics driving turbulence and/or involve new physics beyond that included in the MHD-turbulence models we consider here. MHD turbulence? --------------- ### EE/BB ratio and TE correlation There are some important qualitative conclusions about MHD-turbulence models that can be inferred from the observations EE/BB $\simeq2$ and TE $>0$. (Strictly speaking, the cross-correlation coefficient we calculate here has not yet been provided by Planck. We estimate it by comparing Figs. 2 and B1 in @Adam:2014bub with Fig. D1 in @Ade:2014zja. There are uncertainties here: the cuts and assumptions that went into the latter figure are not necessarily as those that went into the first two. Even so, we infer that the cross-correlation coefficient is reasonably large and, more importantly, positive.) The models generally predict [@Cho:2002qi] that: (a) slow/Alfvén waves should have similar power spectra; (b) the slow/Alfvén should preferentially populate modes perpendicular to the magnetic field ($\lambda<0$ in our parlance); (c) the fast modes should be largely uncoupled from the slow/Alfvén modes; and (d) the fast modes should be nearly isotropic ($\lambda \simeq 0$). ![image](PowsVA20160606.pdf) ![image](PowsVS20160606) ![image](PowsVF20160606) We also need to consider the total E- and B-mode polarization powers contributed, for fixed angle-averaged velocity-perturbation power, by each of the different types of MHD waves. These are plotted in Fig. \[fig:EBpoweq\] for $\beta=0.1$ and $\beta=2$ (the results for $\beta \gg 1$ are similar to those for $\beta=2$). For $\beta\gtrsim 1$, the polarization powers contributed by all three types of waves are roughly similar. However, the polarization power in slow modes scales inversely with $\beta$ as $\beta \to 0$. Physically, this occurs because $(\omega/k)\to 0$ in this limit, indicating a vanishing restoring force. The fluid displacements, and thus density perturbations, become large. Thus, the EE/BB ratio and TE correlation will receive disproportionately large contributions from slow modes in a low-$\beta$ plasma. Looking at Fig. \[fig:panels\], along with Fig. \[fig:EBpoweq\], we see that the combination of the two constraints (EE/BB $\simeq2$ and TE $>0$) very seriously restricts the range of allowable models. There seem to be two possibilities: (1) A nearly isotropic spectrum of fast waves provides positive cross-correlation and EE/BB $\simeq2$ for any $\beta$. A combination of slow/Alfvén waves is disallowed, on the other hand for $\beta\gtrsim1$. Thus, the observations can be explained if $\beta \gtrsim1$ and Alfvén/slow waves are somehow suppressed. (2) For $\beta \ll1$, Alfvén waves can produce EE/BB $\simeq2$ if sufficiently anisotropic, but they contribute nothing to TE. Slow modes can, if sufficiently anisotropic, also contribute EE/BB $\simeq2$ and a positive TE. Given the theoretical expectation that the velocity power in slow and Alfvén waves is comparable, the slow waves will dominate at low $\beta$, and thus the anisotropy must be even greater to account for the observations. The fettle of either of these MHD-turbulence interpretations is damaged by the relative uniformity—as best can be determined—of the EE/BB ratio and TE correlation across the sky. The ISM is a complicated system that is likely to display considerable variation in the parameters $\beta$ and $\lambda$ and the relative contributions of strong/fast/Alfvén waves. While there are indeed pockets of the MHD-turbulence parameter space that can account for the observed EE/BB and TE, these predictions will not be robust if there is considerable variation of $\beta$, $\lambda$, or the mix of slow/fast/Alfvén waves within the ISM. ### Scale-dependent anisotropy? The observed power-law indexes for the $\ell$ dependences of the EE and BB power spectra agree to roughly a percent and are also very similar to those for the TT and TE power spectra [@Adam:2014bub]. As the Figures indicate, the EE/BB ratios can depend quite a bit on the anistropy parameter $\lambda$. Thus, if the power anisotropy is scale-dependent, as expected in MHD turbulence [@Goldreich:1994zz; @Cho:2002qi], then one might expect to see different power-law indexes for E modes and B modes. Some caution should be used in drawing this conclusion since a given multipole moment $\ell$ receives contributions from emission at a variety of line-of-sight distances $r$, and thus a variety of wavenumbers $k \sim \ell/r$. Still, we infer that there is no dramatic variation of the MHD power anisotropy with over the $\sim 0.1-30$ pc length scales probed by Planck. ### The wavenumber scaling {#sec:wavenumberscaling} There is also a disparity between the spectral index $\nu\simeq 2.4$ measured for the TE/EE/BB/TT power spectra, $C_\ell \propto \ell^{-\nu}$, and that, $\kappa \simeq 3.67$, in the three-dimensional power spectrum, $P(k)\propto k^{-\kappa}$ expected in MHD turbulence. The two exponents are related through the Limber equation, Eq. (\[eq:Limber\]). If the three-dimensional power spectrum is well-approximated by a single power law over the relevant distance scales, then the two-dimensional power spectrum $C_\ell$ will also be a power law and, moreover, with the same spectral index, $\nu=\kappa$. Given that the maximum distance from which we see dust emission (at least at high Galactic latitudes) is $r_{\rm max} \simeq 100-200$ pc, the range of physical length scales probed by Planck measurements over $\ell\simeq 30-600$ is roughly $L\sim 0.1-30$ pc, where $L=2 \pi/k$. An outer scale? --------------- Turbulence is expected, however, to be described by a power law only below some outer distance scale $L$, or for wavenumber $k\gtrsim k_c \sim 2\pi L^{-1}$. Suppose, for example, that the power is $P(k)=0$ for $k<k_c$ and $P(k) \propto k^{-\kappa}$ for $k>k_c$ (and with $q(r)=$ constant). In this case, we expect $C_\ell \propto \ell^{-1}$ for $\ell \ll \ell_c \equiv k_c r_{\rm max}$ and $C_\ell \propto \ell^{-\kappa}$ for $\ell \gg \ell_c$. It is conceivable that the apparent power-law index $\nu=2.42$ approximates the scaling if the $\ell=30-600$ range over which the measurements are done contains the characteristic multipole $\ell_c$ that separates the $C_\ell \propto \ell^{-1}$ low-$\ell$ behavior to the $C_\ell \propto \ell^{-\kappa}\sim \ell^{-3.67}$ behavior at higher $\ell$. If so, then the outer scale is (taking $\ell_c\simeq100$ and $r_{\rm max} \sim100$ pc) $L\sim 10$ pc, a reasonable value and not too different from the $\sim$ pc outer scale inferred from Faraday rotation and depolarization of extragalactic radio sources [@Haverkorn:2008tb]. If the $\ell=30-600$ range does indeed correspond to the outer scale of turbulence, then guidance from MHD-turbulence modeling about the power in slow/fast/Alfvén waves may be inappropriate. The observations may then have more to do with the large-scale physics—for example, stellar winds, protostellar outflows, supernovae [@Lacki:2013nda; @Padoan:2016], or Galactic spiral shocks [@Kim:2006ny]—driving the turbulence, rather than the turbulence itself. In this case, the power-law behavior in $C_\ell$ should be only an approximation, and it should be found, with improved measurement, to be shallower at lower $\ell$ and steeper at higher $\ell$. If this interpretation is correct, then extrapolations of foreground power based on measurements at $30 \lesssim \ell \lesssim 600$ to lower $\ell$ may be overestimating the low-$\ell$ CMB foregrounds. If so, this will be good news [@Kamionkowski:2015yta] for experiments, such as CLASS [@Essinger-Hileman:2014pja] and LiteBird [@Matsumura:2013aja], that go to low $\ell$ to seek this signal. Warm/neutral transition? ------------------------ Another possibility is that the ISM is not described by the conventional MHD-turbulence models. For example, it is well known that the interstellar medium is a multi-phase medium. If there is some instability that allows transitions, for example, between a warm neutral phase and a cold neutral phase then the ISM equation of state may be more complicated than that assumed in the standard MHD analysis [@Norman:1996ba; @Kritsuk:2001cm]. If so, then the normal modes of the system may not necessarily correspond to the standard slow/fast/Alfvén waves—for any value of $\beta$—but rather consist of some other linear combinations of them. Does dust trace plasma? ----------------------- The MHD approximation assumed here requires the magnetic-field lines to be tied to the plasma, and the relations \[Eqs. (\[eqn:slowrelation\]) and (\[eqn:fastrelation\])\] derived above are between the magnetic-field and plasma-density perturbations. Strictly speaking, though, the quantity $\delta n$ is the perturbation to the [*dust*]{} density. In deriving Eqs. (\[eqn:slowrelation\]) and (\[eqn:fastrelation\]), we have assumed that the dust and plasma are distributed in the same way. Although there are reasons to suspect that this assumption is largely valid, there are also indeed reasons to suspect that there may be dust-plasma relative motions of a magnitude large enough to affect our results, as we now discuss. In a turbulent ISM, one generically expects that–at least on the large scales considered in this paper–dust should be well-mixed (see, e.g., @Lazarian:2001zv). On small scales, however, the dust grains may not necessarily be well coupled to the gas. From a theoretical perspective, two major sources of coupling should be considered: collisional coupling with the atoms in the gas, and the gyromotion of charged grains in a magnetic field [@Voelk:1980; @1985prpl.conf..621D]. The product of mean atomic velocity and the collisional drag time in a (mostly) neutral medium is $$\bar v_{\rm H} t_{\rm drag} = \frac{a\rho_{\rm g}}{m_{\rm H}n_{\rm H}} = 5\,a_{-5}n_{\rm H}^{-1}\,{\rm pc}, \label{eq:bvt}$$ where we have used a grain density of $\rho_{\rm g} = 2.6$ g cm$^{-3}$, written the grain radius in units of $a_{-5} = 10^{-5}$ cm, and the hydrogen density in units of cm$^{-3}$. If magnetic fields were neglected, we would expect the dust to trace the gas for sound waves of (reduced) wavelength $\lambdabar = k^{-1}$ larger than this scale. It is easily seen that for typical ISM distances $r \sim 100$ pc, the condition of dust-gas coupling through collisions should be violated at $\ell = kr \gtrsim 20 a_{-5}^{-1}n_{\rm H}$, i.e. well within the range of interest for the Planck dust-polarization maps. On the other hand, the gyromotion of charged grains in magnetic fields restricts the motion of dust grains in directions perpendicular to the magnetic field on a length scale of $$v_{\rm A} t_{\rm L} = \frac{m_{\rm g}c}{\Phi a \sqrt{4\pi m_{\rm H}n_{\rm H}}} = 7\times 10^{-5}\, a_{-5}^2 n_{\rm H}^{-1/2}\,{\rm pc} \label{eq:val}$$ for grains with a potential $\Phi\sim 10\,$V$=0.03\,$statvolt generated by the photoelectric effect. Therefore, for Alfvén waves of (reduced) wavelength $\lambdabar = k^{-1}$ larger than this scale, we expect the dust to trace the plasma. Factors of $\beta$ and trigonometric factors may appear in the coupling to the slow and fast MHD waves, but only for extreme values would we expect the Larmor coupling to fail. A possible exception to the above argument is that gyromotion couples the dust to the magnetic field in the perpendicular direction, but not in the parallel direction. To take an extreme case, slow waves in a low-$\beta$ plasma (which have displacements mostly along the field) with $\lambdabar$ less than Eq. (\[eq:bvt\]) might primarily displace the gas, while the dust fails to participate. If the small-scale field is itself turbulent, however, grains may undergo changes in pitch angle and be forced to move with the gas [@Lazarian:2004ar]. From an empirical perspective, the similarity of the power laws for the dust-intensity and dust-polarization power spectra; the difference between the E-mode and B-mode power [@Adam:2014bub; @Ade:2015mbc; @Adam:2014gaa]; the evidence for a similar EE/BB ratio in synchrotron radiation [@Page:2006hz]; and the striking agreement of HI 21-cm and far-infrared dust maps [@Schlegel:1997yv] all suggest that the dust and plasma density are not grotesquely mismatched. If there is indeed some random component $\delta n$, not correlated with the magnetic-field perturbation, then that should drive EE/BB toward unity, given the equality of the angular averages of the $\cos2\psi$ and $\sin 2\psi$ factors that multiply $\delta n$ in Eqs. (\[eqn:Eintermediate\]) and (\[eqn:Bintermediate\]). The observations thus suggest some correlation of the dust with the magnetic field. Moreover, when considering results below, we should be looking not only for parameter combinations that provide EE/BB $\simeq2$, but perhaps also for those that provide a larger ratio. We thus proceed here under the assumption that the dust density traces the plasma density, but note that the validity of this assumption—and the consequences of its violation—warrant further investigation. Possibilities for testing the hypothesis include the frequency dependence of the dust-polarization signal (since dust segregation may depend on the grain size), cross-correlation with synchrotron polarization (which is emitted by the plasma, rather than the dust), and cross-correlation with polarized-starlight surveys. Model of Random Displacements of the Magnetized Fluid {#sec:rand} ===================================================== In the previous Section we questioned whether the Planck dust-polarization data could be explained in terms of MHD turbulence and speculated that they might have more to do with the large-scale turbulence-driving physics. Here we propose a simple phenomenological model of fluctuations of the ISM that, as we will see, can easily produce the observed EE/BB ratio and TE correlation. Instead of decomposing perturbations into slow/fast/Alfvén MHD waves, we here simply suppose that the magnetized fluid experiences a random displacement, $${\boldsymbol{\Delta}}({\ensuremath {\mathbf{x}}}) = (\Delta_1,\,\Delta_2,\,\Delta_3).$$ The continuity equation then provides the associated density perturbation, $$\frac{\delta n}{n_0} = -i \mathbf{k} \cdot \boldsymbol{\Delta} = -i k(\Delta_1 \cos\psi + \Delta_2\sin\psi),$$ and from the MHD equation, $\delta\mathbf{H} = i \mathbf{k}\times(\boldsymbol{\Delta}\times\mathbf{H}_0)$, the associated magnetic-field perturbations are, $$\begin{aligned} \frac{\delta H_x}{H_0} &=&- \tan\psi \frac{\delta H_y}{H_0} = -i k \Delta_2 \sin\theta \sin\psi \\ \frac{\delta H}{H_0} &=& i k \left( \Delta_3 \sin\theta \cos\theta \cos\psi \right. \cr &-&\left. \Delta_1 \cos^2\theta\cos\psi - \Delta_2\sin\psi \right).\end{aligned}$$ These are then inserted into Eqs. (\[eqn:Egen\])–(\[eqn:Bgen\]) to determine the E- and B-mode polarization. To calculate the power, we assume the displacement field has equal power in all three components, $\langle \Delta_i \Delta_j\rangle = \delta_{ij} F_\lambda$, where $\lambda$ now represents the anisotropy in the displacement power. ![The EE/BB ratio (top), TE cross correlation coefficient (middle), and E- and B-mode power normalized to the displacement power spectrum (bottom) are shown for the model of fluid displacements (black), as well as the individual contributions by the longitudinal (red) and transverse (blue) displacements of the MHD fluid.[]{data-label="fig:LTRand"}](RandPanels20160608.pdf){width="3.25in"} The results for the EE/BB ratio, TE cross-correlation coefficient, and individual powers are shown in Fig. \[fig:LTRand\]. This model easily explains the EE/BB $=2$ ratio and positive TE correlation with a moderately anisotropic power index of $\lambda \simeq -1$. To gain better insight into the physical mechanisms that could generate a spectrum of displacements, we decompose $\boldsymbol{\Delta}$ in the basis spanned by $\mathbf{\hat k}$, $\mathbf{\hat a}$, and $\boldsymbol{\hat \theta}$. We define longitudinal displacements as $\boldsymbol{\Delta}_\parallel = (\boldsymbol{\Delta}\cdot \mathbf{\hat k}) \mathbf{\hat k}$ and transverse displacements $\boldsymbol{\Delta}_a$ and $\boldsymbol{\Delta}_\theta$. We immediately notice that the density perturbation is entirely due to longitudinal displacements, $$\frac{\delta n}{n_0}|_{\parallel} = -i k \Delta_{\parallel}.$$ Hence, the observed, strong TE cross-correlation implies that the longitudinal modes play a significant role in the structure of the ISM on these scales. The magnetic-field fluctuations are $$\begin{aligned} \frac{\delta H_x}{H_0} &=& - \tan\psi \frac{\delta H_y}{H_0} = -i k \sin\psi \left(\Delta_\parallel \sin\alpha\sin\varpi \right. \cr &&\qquad \left. +\cos\alpha(\Delta_a \cos\varpi - \Delta_\theta \sin\varpi )\right) \\ \frac{\delta H}{H_0} &=& -i k \left( \Delta_\parallel \sin^2\alpha - \Delta_\theta \sin\alpha\cos\alpha\right).\end{aligned}$$ Using the above results, we can assess the relative contributions of longitudinal- and transverse-displacement power to the E- and B-mode power. A similar procedure as above is carried out to evaluate the EE/BB ratio, shown in Fig. \[fig:LTRand\]. The power in transverse displacements, indicated in the Figure as $\Delta_\perp$, consists of the sum of $\Delta_a$ and $\Delta_\theta$ modes. In the context of this model, the observations suggest a slightly anisotropic spectrum of longitudinal displacements. Although there is some dependence of the EE/BB ratio on $\lambda$, the dependence is relatively weak. In addition to being fairly simple, this random-displacement model is also fairly robust. There is variation in EE/BB and TE with the anisotropy parameter $\lambda$. However, the TE correlation is generically positive and the variation of EE/BB with $\lambda$ fairly slow. Clearly, this model falls far short of a theory. Still, it has strengths as a working model that may help guide a more robust astrophysical explanation for the observations. Conclusions {#sec:conclusions} =========== We have demonstrated that the EE, BB, and TE power spectra for polarized dust (and synchrotron) emission provide a new, unique, and powerful probe of the state of the magnetized ISM. We calculated the contributions to E- and B-mode power and the TE cross-correlation from the slow, fast, and Alfvén waves MHD waves and and provided results for different ratios $\beta$ of magnetic-field to gas pressures and different power anisotropies. We argued that the observations—of EE/BB/TE power and the spectral index for fluctuations—greatly reduce the available parameter space of MHD-turbulence models for the Planck dust-polarization data. We then speculated that a full explanation of the observations may involve the effects of the large-scale physics and developed a simple phenomenological model, based on random displacements of a magnetized fluid, that can account for the observations. Our work motivates a vast suite of additional investigations. First of all, we have used here only the fact that the TE cross-correlation coefficient is positive. Planck has published results for TE power, and for the TT and EE power, but those are separate analyses that use different cuts and assumptions about systematic effects. It will be valuable to measure more carefully the cross-correlation coefficient we have calculated here. Second, we have presented results for EE/BB ratios and the TE cross-correlation after averaging over all magnetic-field orientations because the observed EE/BB $\simeq2$ ratio seems to be quite generic across the sky. Still, the background-field orientation may differ from one small patch of sky to another, and so the EE/BB ratio and TE correlations should also vary. If the background field has a fixed orientation in some small patch of sky, then there should also be a local departure from statistical isotropy within that patch. There is also potentially interesting information in the $\ell$ dependence of the $C_\ell$. Is it really a power law? Or does it steepen at higher $\ell$? Are the $\ell$ dependences of the EE, BB, TE, and TT power spectra all the same? Or are there subtle variations that may reflect scale-dependent anisotropies or perhaps some other physics not accounted for here? We also suggest further investigation of the frequency dependence of dust-polarization maps and cross-correlation with synchrotron-polarization maps and starlight-polarization surveys to test the hypothesis that the dust density traces the plasma density assumed here. Finally, although we have focussed here on Planck dust-polarization maps, similar techniques can also be applied to dust-polarization data from specific molecular clouds. Fortunately, there is not only far more along these lines that can be done with existing Planck data, but also prospects for rich new data sets to build upon Planck. We thank E. Vishniac for useful discussions. MK was supported by NSF Grant No. 0244990, NASA NNX15AB18G, the John Templeton Foundation, and the Simons Foundation. RRC was supported by DOE grant DE-SC0010386. CH was supported by NASA, the U.S. Department of Energy, the David & Lucile Packard Foundation, and the Simons Foundation. 0.5in [65]{} natexlab\#1[\#1]{} Adam, R., [et al.]{} 2016, A&A, 586, A133 \[arXiv:1409.5738\] —. 2016, A&A, 586, A135 \[arXiv:1409.6728\] Ade, P. A. R., [et al.]{} 2015, A&A, 576, A104 \[arXiv:1405.0871\] —. 2015, A&A, 576, A105 \[arXiv:1405.0872\] —. 2015, A&A, 576, A107 \[arXiv:1405.0874\] —. 2015, \[arXiv:1506.06660\] —. 2016, A&A, 586, A141 \[arXiv:1505.02779\] Aghanim, N., [et al.]{} 2016 \[arXiv:1604.01029\] Ali-Haimoud, Y., Hirata, C. M., & Dickinson, C. 2009, MNRAS, 395, 1055 \[arXiv:0812.2904\] Andersson, B.-G., Lazarian, A., & Vaillancourt, J. E. 2015, , 53, 501 Armstrong, J. W., Rickett, B. J., & Spangler, S. R. 1995, , 443, 209 Brandenburg, A., & Lazarian, A. 2013, Space Sci. Rev., 178, 163 \[arXiv:1307.5496\] Cabella, P., & Kamionkowski, M. 2004, in [International School of Gravitation and Cosmology: The Polarization of the Cosmic Microwave Background Rome, Italy, September 6-11, 2003]{} Chandrasekhar, S., & Fermi, E. 1953, ApJ, 118, 113 Cho, J., & Lazarian, A. 2002, Phys. Rev. Lett., 88, 245001 \[arXiv:astro-ph/0205282\] Cho, J., Lazarian, A., & Vishniac, E. 2003, Lect. Notes Phys., 614, 56 \[arXiv:astro-ph/0205286\] Clark, S. E., Hill, J. C., Peek, J. E. G., Putman, M. E., & Babler, B. L. 2015, Phys. Rev. Lett., 115, 241302 \[arXiv:1508.07005\] Delabrouille, J., [et al.]{} 2013, A&A, 553, A96 \[arXiv:1207.3675\] Dolginov A. Z., 1972, , 18, 337 Dolginov, A. Z., & Mytrophanov, I. G. 1976, , 43, 257 Draine, B. T. 1985, in Protostars and Planets II, ed. D. C. Black & M. S. Matthews, 621–640 Draine, B. T., & Fraisse, A. A. 2009, ApJ, 696, 1, \[Erratum: ApJ757,106(2012)\] \[arXiv:0809.2094\] Draine, B. T., & Lazarian, A. 1998, ApJ, 494, L19 \[arXiv:astro-ph/9710152\] Draine, B. T., & Weingartner, J. C. 1996, ApJ, 470, 551 \[arXiv:astro-ph/9605046\] —. 1997, ApJ, 480, 633 \[arXiv:astro-ph/9611149\] Dunkley, J., [et al.]{} 2009, AIP Conf. Proc., 1141, 222 \[arXiv:0811.3915\] Elmegreen, B. G., & Scalo, J. 2004, ARAA, 42, 211 \[arXiv:astro-ph/0404451\] Essinger-Hileman, T., [et al.]{} 2014, Proc. SPIE Int. Soc. Opt. Eng., 9153, 91531I \[arXiv:1408.4788\] Finkbeiner, D. P., Langston, G. I., & Minter, A. H. 2004, ApJ, 617, 350 \[arXiv:astro-ph/0408292\] Fissel, L. M., [et al.]{} 2010, Proc. SPIE Int. Soc. Opt. Eng., 7741, 0E \[arXiv:1007.1390\] Fosalba, P., Lazarian, A., Prunet, S., & Tauber, J. A. 2002, ApJ, 564, 762 \[arXiv:astro-ph/0105023\] Goldreich, P., & Sridhar, S. 1995, ApJ, 438, 763 Goodman, A. A., Bastien, P., Menard, F., & Myers, P. C. 1990, , 359, 363 Haverkorn, M., Brown, J. C., Gaensler, B. M., & McClure-Griffiths, N. M. 2008, ApJ, 680, 362 \[arXiv:0802.2740\] Heiles, C. 1996, , 462, 316 Iroshnikov, P. S. 1964, , 7, 566 Kamionkowski, M., Kosowsky, A., & Stebbins, A. 1997, Phys. Rev. Lett., 78, 2058 \[arXiv:astro-ph/9609132\] —. 1997, Phys. Rev., D55, 7368 \[arXiv:astro-ph/9611125\] Kamionkowski, M., & Kovetz, E. D. 2014, Phys. Rev. Lett., 113, 191303 \[arXiv:1408.4125\] —. 2015, ARAA, in press \[arXiv:1510.06042\] Kataoka, A., Machida, M. N., & Tomisaka, K. 2012, ApJ, 761, 40 \[arXiv:1210.7637\] Kim, C.-G., Kim, W.-T., & Ostriker, E. C. 2006, ApJ, 649, L13 \[arXiv:astro-ph/0608161\] Koch, P. M., Tang, Y.-W., & Ho, P. T. P. 2013, ApJ, 775, 77 \[arXiv:1308.6185\] Kovetz, E. D., & Kamionkowski, M. 2015, Phys. Rev., D91, 081303 \[arXiv:1502.00625\] Kraichnan, R. H. 1965, Phys. Fluids, 8, 1385 Kritsuk, A. G., & Norman, M. L. 2002, ApJ, 569, L127 \[arXiv:astro-ph/0112437\] Lacki, B. C. 2013 \[arXiv:1308.5232\] Lazarian, A., & Yan, H. 2002, ApJ, 566, L105 \[arXiv:astro-ph/0112365\] Lazarian, A., Yan, H., & Draine, B. T. 2004, ApJ, 616, 895 \[arXiv:astro-ph/0408173\] Lithwick, Y., & Goldreich, P. 2001, ApJ, 562, 279 \[arXiv:astro-ph/0106425\] Matsumura, T., [et al.]{} 2013, \[J. Low. Temp. Phys.176,733(2014)\] \[arXiv:1311.2847\] Misawa, R., [et al.]{} 2014, Proc. SPIE Int. Soc. Opt. Eng., 9153, 91531H \[arXiv:1410.5760\] Niemack, M. D., [et al.]{} 2015 \[arXiv:1509.05392\] Norman, C. A., & Ferrara, A. 1996, ApJ, 467, 280 \[arXiv:astro-ph/9602146\] O’Dea, D. T., Clark, C. N., Contaldi, C. R., & MacTavish, C. J. 2012, MNRAS, 419, 1795 \[arXiv:1107.4612\] Padoan, P., Pan, L., Haugb[ø]{}lle, T., & Nordlund, [Å]{}. 2016, , 822, 11 \[arXiv:1509.04663\] Page, L., [et al.]{} 2007, ApJ Suppl., 170, 335 \[arXiv:astro-ph/0603450\] Pelkonen, V.-M., Juvela, M., & Padoan, P. 2007, A&A, 461, 551 \[arXiv:astro-ph/0609143\] Rybicki, G. B., & Lightman, A. P. 1979, Radiative Processes in Astrophysics, John Wiley & Sons Schekochihin, A. A., Cowley, S. C., Dorland, W., [et al.]{} 2009, ApJ Suppl., 182, 310 \[arXiv:0704.0044\] Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500, 525 \[arXiv:astro-ph/9710327\] Seljak, U. 1997, ApJ, 482, 6 \[arXiv:astro-ph/9608131\] Seljak, U., & Zaldarriaga, M. 1997, Phys. Rev. Lett., 78, 2054 \[arXiv:astro-ph/9609169\] Shebalin, J. V., Matthaeus, W. H., & Montgomery, D. 1983, J. Plasma Phys., 29, 525 Soler, J. D., Hennebelle, P., Martin, P. G., [et al.]{} 2013, ApJ, 774, 128 \[arXiv:1303.1830\] Stein, W. 1966, , 144, 318 Voelk, H. J., Jones, F. C., Morfill, G. E., & Roeser S. 1980, , 85, 316 Wilson, G. 2016, private communication Zaldarriaga, M. 2001, Phys. Rev., D64, 103001 \[arXiv:astro-ph/0106174\] Zaldarriaga, M., & Seljak, U. 1997, Phys. Rev., D55, 1830 \[arXiv:astro-ph/9609170\] [^1]: In the CMB literature this is often written $P({\ensuremath {\boldsymbol{\theta}}})$, but here we write $\Pi$ to avoid confusion with the 3D power spectrum. [^2]: The factor of $\Omega$ arises from the density of Fourier modes; the “usual” equation would read $\langle \tilde E^\ast({\ensuremath {\boldsymbol{\ell}}}) \tilde E({\ensuremath {\boldsymbol{\ell}}}') \rangle = (2\pi)^2\delta^{\rm D}({\ensuremath {\boldsymbol{\ell}}}-{\ensuremath {\boldsymbol{\ell}}}') C_\ell^{\rm EE}$, where $\delta^{\rm D}$ is the Dirac $\delta$-function. For a density of modes $\Omega/(2\pi)^2$, we have $(2\pi)^2\delta^{\rm D}({\ensuremath {\boldsymbol{\ell}}}-{\ensuremath {\boldsymbol{\ell}}}') \rightarrow \Omega\delta^{\rm K}_{{\ensuremath {\boldsymbol{\ell}}},{\ensuremath {\boldsymbol{\ell}}}'}$, where $\delta^{\rm K}$ is the Kronecker $\delta$-symbol.
{ "pile_set_name": "ArXiv" }
--- abstract: | We propose an information transmission scheme by a swarm of anonymous oblivious mobile robots on a graph. The swarm of robots travel from a sender vertex to a receiver vertex to transmit a symbol generated at the sender. The codeword for a symbol is a pair of an initial configuration at the sender and a set of terminal configurations at the receiver. The set of such codewords forms a code. We analyze the performance of the proposed scheme in terms of its code size and transmission delay. We first demonstrate that a lower bound of the transmission delay depends on the size of the swarm, and the code size is upper bounded by an exponent of the size of the swarm. We then give two algorithms for a swarm of a fixed size. The first algorithm realizes a near optimal code size with a large transmission delay. The second algorithm realizes an optimal transmission delay with a smaller code size. We then consider information transmission by swarms of different sizes and present upper bounds of the expected swarm size by the two algorithms. We also present lower bounds by Shannon’s lemma and noiseless coding theorem. [**Keyword**]{}: Mobile robots, information transmission, coding theory. author: - 'Yukiko Yamauchi[^1]' - 'Masafumi Yamashita[^2]' bibliography: - 'papers.bib' title: Coding theory for noiseless channels realized by anonymous oblivious mobile robots --- Introduction ============ Memory is indispensable to a computer system to demonstrate its computation ability. Its importance does not diminish in a distributed system. We thus tend to guess that a distributed system lacking memory can solve no problems but trivial ones. A *swarm* of anonymous oblivious mobile robots is a distributed system whose components called robots move on a continuous space or a graph. It is typically characterized by the lack of identifiers and common and local memories. Contrary to the quess, in spite of the absence of memory, it has been shown to have rich ability to solve a variety of problems (e.g., [@BMPT11; @CFPS12; @CDPIM10; @DFSY15; @DLPRT12; @DFSV18; @DP07; @DPV10; @FIPS13; @FPSW08; @FYOKY15; @LYKY18; @PPV15; @SY99; @YS10; @YUKY17; @YUKY16; @YY14]). In order to solve those problems, the robots indeed need to “remember” key information to solve a problem such as the current search direction in exploration [@BMPT11; @DLPRT12; @DYKY18; @FIPS13], the currently agreed common coordinate system in pattern formation [@FYOKY15; @SY99; @YS10; @YY14; @YUKY16; @YUKY17], and the current phase in forming a sequence of patterns [@DFSY15]. A main idea to make up the lack of memory is to use “external memory” composed of the locations of all robots, i.e., its (global) configuration, and the information that the robots need to remember is embedded in the current configuration. However for the robots to keep external memory stable is by no means easy, and interesting tricks have been developed to realize the idea based on concepts such as the smallest enclosing circle [@DPV10; @FPSW08; @FYOKY15; @LYKY18; @SY99; @YS10; @YY14], the rotation group [@YUKY17; @YUKY16] and the Fermat point [@CFPS12]. Proposing a suitable external memory is thus considered to be a key in the design of algorithm for a swarm of anonymous oblivious mobile robots. This paper investigates an information transmission problem on a graph, which asks an algorithm to transmit information from a sender to a receiver by using a swarm of anonymous oblivious mobile robots moving on the graph, and analyzes the bounds on the amount of information that the swarm of robots can carry in its external memory and the transmission delay. Let $G = (V,E)$ be a connected undirected (possibly infinite) graph with two distinguished vertices $u_S$ and $u_R$ called the *sender* and the *receiver*, respectively. We want to transmit a symbol $s$ in $S = \{ s_1, s_2, \ldots , s_{\alpha} \}$ from $u_S$ to $u_R$ by using a swarm of anonymous oblivious mobile robots on $G$, where the probability that $s = s_i$ is $p_i$. Sometimes we assume that the distance between $u_S$ and $u_R$ is sufficiently larger than $\alpha$ and/or the number $k$ of the robots to reduce possible disturbance caused by boundary conditions in analyses. Each vertex of $G$ can accommodate at most one robot of the swarm. Thus the *configuration* of the swarm is represented by a subset $U$ of $V$. Any configuration of the swarm must be connected; ${\cal C} = \{ U \mid G[U] \text{ is connected} \}$ denotes the set of all connected configurations, where $G[U]$ is the subgraph of $G$ induced by $U$. When the current configuration is $C$, a robot $r$ at $u \in C$ can move to one of its neighbors $v \not\in C$ if the next configuration $(C \setminus \{u\}) \cup \{v\}$ is also connected. At most one robot can move at one time step. The *behavior* of the swarm of robots is determined by a deterministic algorithm, which specifies the next configuration by choosing a robot to move and its destination. The input of the algorithm is the current configuration. Thus the behavior of the swarm is always deterministic. A configuration $C$ is said to be *initial* (resp. *terminal*) if $u_S \in C$ (resp. $u_R \in C$). The sets ${\cal C}_I = \{ C \in {\cal C} \mid u_S \in C \}$ and ${\cal C}_T = \{ C \in {\cal C} \mid u_R \in C \}$ denote the sets of initial and terminal configurations, respectively. Let ${\cal A} = \{ A_k \mid k = 1, 2, \ldots \}$ be a set of algorithms, where $A_k$ is an algorithm for the swarm of $k$ robots. For an initial configuration $C_0 \in {\cal C}_I$, where $|C| = k$, the behavior $C_0, C_1, \ldots$ of the swarm of $k$ robots under $A_k$ may eventually reach $C_d \in {\cal C}_T$ for the first time. In this case we define that $\tau(C_0) = C_d$ and the transmission delay is $d$. Otherwise, $\tau(C_0) = \bot$. Suppose that there are $\alpha$ initial configurations $C_0^1, C_0^2, \ldots , C_0^{\alpha}$ such that $\tau(C_0^i) \not= \bot$ for $i = 1, 2, \ldots , \alpha$ and $\tau(C_0^i) \not= \tau(C_0^j)$ for all $1 \leq i < j \leq \alpha$. Assuming that both $u_S$ and $u_R$ are accessible to the list of pairs $(C_0^i, \tau(C_i^0))$ for $i = 1, 2, \ldots , \alpha$, we can transmit a symbol $s_i \in S$ as follows: 1. The sender initializes the swarm of $k_i$ robots by moving the robots to the vertices in $C_0^i$ and starts algorithm $A_{k_i}$, where $k_i = |C_0^i|$. 2. The receiver recognizes $\tau(C_0^i)$ when a robot reaches $u_R$, and knows that the sender transmitted $s_i$. Here we assume that the sender has a bag of infinitely many robots and can initialize the swarm of any number of robots, and the receiver recognizes $\tau(C_0^i)$ as soon as a robot reaches $u_R$. In this paper, we fix this information transmission scheme and analyze its performance. Given a set of algorithms $\cal A$, the performance of the scheme mainly depends on the selection of an initial configuration $C_0^i$ for each $s_i$. A simple approach is to choose configurations with the same size $k$. That is, $k_i = k$ for all $i = 1, 2, \ldots , \alpha$. We first investigate this approach. Let $\mu_{A_k}$ be the maximum number such that there are $\mu_{A_k}$ initial configurations $C_0^1, C_0^2, \ldots , C_0^{\mu_{A_k}}$ with size $k$ satisfying $\tau(C_0^i) \not= \tau(C_0^j)$ for all $1 \leq i < j \leq \mu_{A_k}$. Then we are interested in upper and lower bounds on $\mu_{A_k}$ and $\mu_k$, where $\mu_k = \max_{A_k} \mu_{A_k}$ denotes the maximum amount of information that the swarm of $k$ robots can carry in external memory under this information transmission scheme. However, assigning all symbols to initial configurations with the same size is not always a good approach from the view of transmission delay, since the value of $k_0$ that satisfies $\mu_{k_0} \geq \alpha$ can be large and, as we will show later, the transmission delay is roughly greater than $dist(u_S,u_R) k_0$ where $dist(u_S,u_R)$ is the distance between $u_S$ and $u_R$ (i.e., the number of edges in the shortest path connecting them). Let $k_i$ be the size of $C_0^i$ assigned to $s_i$ for $i = 1, 2, \ldots , \alpha$. Then we are interested in upper and lower bounds on the average size $K^* = \sum_{i=1}^{\alpha} p_i k_i$, since there is a chance that $K^* < k_0$ holds for some $\cal A$. The reduction of $K^*$ contributes to the reduction of energy, as well as the reduction of transmission delay, whose lower bound is roughly estimated by $dist(u_S,u_R) K^*$. However, its upper bound still heavily depends on $\cal A$. We mainly investigate information transmission on *8-grids* $G_8(m,n) = ({\mathbb Z}_{m,n}^2, N_8)$, where ${\mathbb Z}_{m,n}^2 = \{(i,j) \in {\mathbb Z} \times {\mathbb Z} \mid 0 \leq i \leq m-1, 0 \leq j \leq n-1 \}$ and $N_8 =\{ ((i,j),(i',j') \in {\mathbb Z}_{m,n}^2 \times {\mathbb Z}_{m,n}^2 : (|i - i'| \leq 1) \wedge (|j - j'|) \leq 1) \wedge ((i,j) \not= (i',j')) \}$. [**Our contributions. **]{} Our contributions can be summarized as follows: 1. We present exponential upper bounds of $\mu_k$ on general graphs and 8-grids. 2. Focusing on “narrow” channel realized by $G_8(m,2)$, we present two algorithms that promises exponential code size to a fixed swarm size $k$. The first algorithm promises a near optimal code size $2^{k-14}$, but has a large transmission delay. The second algorithm promises a near optimal transmission delay, but has a smaller code size $2^{\lfloor k/2 \rfloor}$. 3. We step into the second approach with these two algorithms and show upper bounds of the expected swarm size. 4. We finally analyze lower bounds of the expected swarm size based on the Shannon’s lemma and noiseless coding theory. [**Related works. **]{} Study of information transmission by a binary (or $q$-ary) code is a main stream in computer science known as information theory (or coding theory). It is a fully developed research area, and there are many standard textbooks (e.g., [@CT06]). One can notice the similarity between the first approach and the equal-length coding, and the second one and the variable-length coding. Our results are also related with Shannon’s source coding theorem, which relates the optimal average code length with the entropy of the information source. Information transmission on a graph by a swarm of anonymous oblivious mobile robots has not been investigated to the best of our knowledge although, including a swarm of mobile robots, distributed systems consisting of mobile entities have been extensively investigated in the last couple of decades. Two survey books [@FPS12; @FPS19] include their models and algorithms to solve typical problems such as gathering and convergence [@CFPS12; @PPV15; @SY99], pattern formation [@DPV10; @FPSW08; @FYOKY15; @SY99; @YS10; @YUKY16; @YY14], scattering and covering [@CDPIM10; @DP07; @TKGT18; @IKY14], flocking and marching [@AFSY08; @CP07; @GP04; @YXCD08], and searching and exploration [@BMPT11; @DLPRT12; @DYKY18; @FIPS13]. A main theme in these studies is to solve the problems without using memory, partly motivated by a challenge to prejudice that those systems cannot solve non-trivial problems and partly because memoryless algorithms are usually strong against transient failures [@D74]. For example, exploration of a finite square grid by a metamorphic robot of size 5 is proposed [@DYKY18]. The metamorphic robot consists of $5$ anonymous oblivious mobile robots (called *modules* in the literature) and remembers the current search direction (i.e., right, left, up or down) in its configuration during the exploration. In the existing pattern formation algorithms for anonymous oblivious mobile robots, in order to control the move order among the anonymous robots, they agree on a common coordinate system and remember it in external memory [@DPV10; @FPSW08; @FYOKY15; @SY99; @YS10; @YUKY16]. In each of the oblivious algorithms one can find a trick to maintain external memory. [**Organization of the paper. **]{} We define the swarm of anonymous oblivious robots on a graph and formalize the noiseless communication channel realized by the swarm with an algorithm in Section \[sec:preliminary\]. Section \[sec:eqsizecode\] first presents an upper bound of $\mu_{G_8(m,n),k}$. Then, we present two algorithms for information transmission by a swarm of fixed size. We consider the second approach in Section \[sec:varsizecode\] with the two algorithms and present lower and upper bounds of the expected swarm size. We conclude this paper with Section \[sec:concl\] which also includes future directions and open problems. Preliminaries {#sec:preliminary} ============= Swarm of anonymous oblivious mobile robots ------------------------------------------ Let $G = (V, E)$ be a simple connected undirected graph. For $u \in V$, $N_G(u)$ denotes the set of neighbors of $u$ in $G$, and for a subset $U \subseteq V$, $G[U]$ denotes the subgraph of $G$ induced by $U$. For two vertices $u,v \in V$, $dist_G(u,v)$ denotes the distance between $u$ and $v$. We define two infinite regular graphs. Let ${\mathbb Z}$ be the set of integers. The *$4$-grid* is an infinite graph $G_4 = ({\mathbb Z}^2, N_4)$, where $N_4 = \{((i,j), (i', j')) \in {\mathbb Z}^2 \times {\mathbb Z}^2 \mid ((i=i') \wedge (|j-j'|=1)) \vee ((|i-i'|=1) \wedge (j=j'))\}$ and the *$8$-grid* is an infinite graph $G_8 = ({\mathbb Z}^2, N_8)$, where $N_8 = \{((i,j),(i'.j')) \in {\mathbb Z}^2 \times {\mathbb Z}^2 \mid (|i-i'|\leq 1) \wedge (|j-j'|\leq 1) \wedge ((i,j) \not= (i',j'))\}$. We consider a swarm $R = \{r_1, r_2, \ldots, r_k\}$ of anonymous oblivious mobile robots on $G$. We use the indices just for description since a robot is anonymous. A robot is *oblivious*, which means that it does not have memory to remember information obtained in the past. A vertex of $G$ can accommodate at most one robot at each time step, and the set of vertices $C \subseteq V$ that accommodate robots is called the *configuration* of $R$. We say that a configuration $C$ is *connected* if $G[C]$ is connected. A robot $r$ at $u \in V$ can move to a vertex $v \in N_G(u) \setminus C$ through edge $(u,v)$ if the new configuration $C(u;v) = (C \setminus \{u\}) \cup \{v\}$ obtained by this movement of $r$ is connected. In what follows, we always assume that a configuration means a connected one. By ${\cal C}$ we denote the set of (connected) configurations $C$. Each robot $r \in R$ knows $G$ and is aware of the vertex $u$ it resides. It repeats a Look-Compute-Move cycle once initialized. In each Look-Compute-Move cycle, it first observes the positions of other robots on $G$ in Look phase. The visibility range is finite, but is large enough to observe all the robots, so that $r_i$ recognizes the current configuration $C$. In Compute phase, given $C$ and $u$ as inputs, a common deterministic algorithm $A$ (on $r$) computes its next position $v$ in such a way that $C(u;v)$ is connected, where $v = u$ is possible and it means to stay at $u$. In this paper, we assume that $A$ outputs $v = u$ on all robots $r$ but one.[^3] Finally, it moves to $v$ through $(u,v)$ in Move phase. Observe that an oblivious robot which does not have memory (except the input buffer of $A$ for $C$ and $u$, and the work space for $A$) can execute $A$. We consider discrete time $0, 1, 2, \cdots$. Given an algorithm $A$ and an initial configuration $C_0 \in {\cal C}_k$, all robots are initialized at time $0$, and synchronously execute a Look-Compute-Move cycle in each time step. Then $A$ chooses a single robot $r$ at a vertex $u \in C_0$ and moves it to one of its neighbors $v$ to yield a configuration $C_1 = C_0(u;v)$ at time 1. The swarm of robots repeats this process and yields configurations $C_2, C_3, \ldots$. We call the evolution of configurations, i.e., ${\mathcal B} = C_0, C_1, C_2, \ldots$, the *behavior* of algorithm $A$ from an initial configuration $C_0$. Since $A$ is deterministic, ${\mathcal B}$ from $C_0$ under $A$ is uniquely determined. Graph as a noiseless communication channel ------------------------------------------ We investigate information transmission from the *sender* $u_S \in V$ to the *receiver* $u_R \in V$ by a swarm of robots, i.e., we regard $G$ as a communication channel. The sender $u_S$ is a memoryless information source that generates a symbol in $S = \{s_1, s_2, \ldots, s_{\alpha}\}$, where $s_i$ is generated with probability ${\mathrm{P}}(s_i) = p_i$ for each $i=1,2, \ldots, \alpha$. Thus $\sum_{i=1}^{\alpha} p_i = 1$. Let ${\cal C}_I$ be the set of configurations that contain $u_S$ and ${\cal C}_T$ be the set of configurations that contains $u_R$. With each symbol $s_i$, we associate a configuration $\gamma_I(s_i) \in {\cal C}_I$ and a set of configurations $\gamma_T(s_i) \subseteq {\cal C}_T$, and regard the pair $(\gamma_I(s_i), \gamma_T(s_i))$ as the *codeword* of $s_i$. Here $\gamma_I(s_i) \neq \gamma_I(s_j)$ and $\gamma_T(s_i) \cap \gamma_T(s_j) = \emptyset$, for any $i \neq j$. To send $s_i$ from $u_S$, as explained in Section 1, $R$ is initialized with $\gamma_I(s_i)$ at $u_S$. When a configuration $C_d$ of $R$ in $\gamma_T(s_i)$ is reached, $u_R \in C_d$ receives $s_i$. We call the pair $\gamma=\langle \gamma_I, \gamma_T \rangle$ *code* of $S$, and $|\gamma|$, which is the size of $\gamma$, is the size $\alpha$ of $S$. Now we consider a coding scheme with an algorithm $A$. Let ${\mathcal B} = C_0, C_1, \ldots$ be a behavior of $R$ under $A$. We define a function $\tau_A: {\cal C}_I \to {\cal C}_T \cup \{\bot\}$ as follows: If ${\mathcal B}$ eventually reaches a configuration $C_d \in {\cal C}_T$ for the first time, then $\tau_A(C_0) = C_d$. We call $d$ the *transmission delay*. Otherwise, if ${\mathcal B}$ does not reach a configuration in ${\mathcal C}_T$ forever, then $\tau_A(C_0) = \bot$. We say algorithm $A$ *realizes* code $\gamma$ if $\tau_A(\gamma_I(s_i)) \in \gamma_T(s_i)$ for each $i=1,2, \ldots, |\gamma|$. Let $\mu_A = |\{\tau_A(C) \mid C \in {\mathcal C}_I\} \setminus \{\bot\}|$. \[obs:1\] Let $A$ be any algorithm. 1. There is a code $\gamma$ with $|\gamma| = \mu_A$ which is realizable by $A$. 2. There is no code $\gamma$ with $|\gamma| = \mu_A+1$ which is realizable by $A$. As for $1$, we construct a code $\gamma^A = \langle \gamma_I^A, \gamma_T^A \rangle$ with $|\gamma^A| = \mu_A$ realizable by $A$ as follows: Let $S_{\mu_A} = \{ s_1, s_2, \ldots, s_{\mu_A}\}$. By definition, there are $\mu_A$ configurations $C_0^1, C_0^2, \ldots , C_0^{\mu_A}$ in ${\mathcal C}_I$ such that $\tau_A(C_0^i) \not= \bot$ for all $i = 1, 2, \ldots , \mu_A$ and $\tau_A(C_0^i) \not= \tau_A(C_0^j)$ for all $1 \leq i < j \leq \mu_A$. We then define $\gamma_I^A(s_i) = C_0^i$ and $\gamma_T^A(s_i) = \{\tau_A(C_0^i)\}$ for $i = 1, 2, \ldots , \mu_A$. It is easy to observe that $\gamma^A$ is indeed a code of size $\mu_A$. As for $2$, suppose that there exists a code $\gamma = \langle \gamma_I, \gamma_T \rangle$ with $|\gamma|=\mu_A +1$ to derive a contradiction. Since $|\gamma|=\mu_A +1 > \mu_A$, there are two symbols $s_i$ and $s_j(\neq s_i)$ such that $\tau_A(\gamma_I(s_i)) = \tau_A (\gamma_I(s_j))$, a contradiction. The proof of Theorem \[obs:1\] gives a construction method to construct code $\gamma^A$ from algorithm $A$. The transmission delay of $\gamma^A$ depends on the choice of $\gamma_I^A(s_i) \in {\mathcal C}_I$ for each $s_i \in S_{\mu_A}$; any $C$ such that $\tau_A(C) = \tau_A(\gamma_I^A(s_i))$ can be chosen instead. Let $d_A(C)$ be the transmission delay of the behavior of $R$ starting from $C$ under $A$. To reduce the transmission delay of $\gamma^A$, it is natural to choose an initial configuration $C$ that minimizes $d_A(C)$, and we thus assume that such a $C$ is chosen in the construction of $\gamma^A$. Then the transmission delay $d_{\gamma^A}$ of $\gamma^A$ is ${\displaystyle \max_{s_i \in S_{\mu_A}} d_A(\gamma_I^A(s_i)) }$. We are interested in the maximization of $\mu_A$ and the minimization of $d_A$. Let ${\displaystyle \mu_{N,k} = \max_A \mu_A }$ and ${\displaystyle d_{N,k} = \min_A d_A }$, where ${\displaystyle d_A = \max_{C: \tau_A(C) \neq \bot} d_A(C) }$ and $N=\langle G, u_S, u_R \rangle$. General graphs -------------- Since $\mu_{N,1} = 1$, we assume $k \geq 2$ in what follows. We also assume $dist_G(u_S,u_R) \geq k$, i.e., ${\mathcal C}_I \cap {\mathcal C}_T = \emptyset$. Indeed we are interested in the case where $dist_G(u_S, u_R) \gg k$. \[O201\] For any $N=\langle G, u_S, u_R \rangle$, $\mu_{N,k} \leq |{\mathcal C}_I|$. \[T201\] Let $N=\langle G, u_S, u_R \rangle$ with $dist_G(u_S, u_R) \geq k$. If $G$ has a cut vertex whose removal disconnects $u_S$ and $u_R$, then $\mu_{N,k} = 0$, that is, $\tau_A(C_0) = \bot$ for any algorithm $A$ and any configuration $C_0 \in {\mathcal C}_I$. To derive a contradiction, we assume that there is an algorithm $A$ and a configuration $C_0 \in {\mathcal C}_T$ such that the behavior ${\mathcal B} = C_0, C_1, \ldots$ of $R$ under $A$ eventually reaches a configuration $C_d \in {\mathcal C}_T$. Let $v \in V$ be a cut vertex of $G$ whose removal from $G$ disconnects $u_S$ and $u_R$. Thus $G[V \setminus \{v\}]$ consists of more than one connected components. Let $G_1 = (V_1 \setminus \{v\}, E_1)$ be the one that contains $u_S$. We first assume $C_0 \subseteq V_1$. Let $t$ be the smallest index such that $C_t \subseteq V_1$ and $C_{t+1} \not\subseteq V_1$. Then there is a vertex $w \in V \setminus V_1$ such that $w \in C_{t+1}$ and $(v,w) \in E$, and the robot at $v$ at $t$ moves to $w$. It is a contradiction since $v \not\in C_{t+1}$ and hence $C_{t+1}$ is not connected. Next assume that $C_0 \not\subseteq V_1$. Since $dist_G(u_S,u_R) \geq k$, there is a time instant $t$ such that the robot $r$ at $v$ moves for the first time. Then $C_{t+1}$ is not connected since $v \not\in C_{t+1}$, and a contradiction is derived. \[T202\] Let $N = \langle G_4, u_S, u_R \rangle$ with $dist_{G_4}(u_S, u_R) \geq k$. Then $\mu_{N,k} = 0$, that is, for any algorithm $A$ and for any configuration $C_0 \in {\mathcal C}_I$, $\tau_A(C_0) = \bot$. To derive a contradiction, we assume that there is an algorithm $A$ and a configuration $C_0 \in {\mathcal C}_I$ such that the behavior $B= C_0, C_1, \ldots $ of $R$ under $A$ eventually reaches a configuration $C_d \in {\mathcal C}_T$. Since $dist_G(u_S, u_R) \geq k$, $d \geq 1$ and $|i_R - i_S| + |j_R - j_S| \geq k$, where $u_S = (i_S, j_S)$ and $u_R = (i_R, j_R)$. We assume without loss of generality that $i_R > i_S$ and $j_R \geq j_S$. For configuration $C_t$, let $x_{\max}(t) = \max \{i \mid (i,j) \in C_t\}$ and $y_{\max} = \max \{j \mid (i,j) \in C_t \}$. If $x_{\max}(0) \geq i_R$, then $j_S \leq y_{\max} < j_R$. Thus either $i_S \leq x_{\max}(0) < i_R$ or $j_S \leq y_{\max}(0) < j_R$ holds. We assume without loss of generality that $i_S \leq x_{\max} < i_R$. Let $t$ be the smallest index such that $x_{\max}(0) + 1 = x_{\max}(t)$ holds. That is, $x_{\max}(t-1) = x_{\max}(0)$. Suppose that the robot $r$ at vertex $u \in C_{t-1}$ moves. Then $C_t$ is not connected, since $r$ is the only robot with $x$-coordinate $x_{max}(0)+1$ in $C_t$ and $u \not\in C_t$. \[theorem:computable\] Let $G$ be a finite graph and $N = \langle G, u_S, u_R \rangle$. Then, $\mu_{N,k}$ and $d_{N,k}$ are computable. We consider the transition diagram (i.e., directed graph) $X = ({\mathcal C} \cup \{ v_S, v_R \},F)$ of the swarm on $G$, where $(C,C') \in F$ if and only if 1) there are $u \in C$ and $v \in C'$ such that $C' = C(u;v)$, 2) $C = v_S$ and $C' \in {\mathcal C}_I$, or 3) $C \in {\mathcal C}_T$ and $C' = v_R$. Then $\mu_{N,k}$ is the size of the maxflow of $X$ from $v_S$ to $v_R$. As for $d_{N,k}$, there are only a finite number of different sets $\Pi$ of $\mu_{N,k}$ disjoint paths connecting $v_S$ and $v_R$. By comparing the length of the longest path in each $\Pi$, we can calculate $d_{N,k}$. By Theorem \[theorem:computable\], we can design an optimal algorithm $A$ to achieve $\mu_{N,k}$ such that $A$ makes $R$ follow one of the $\mu_{N,k}$ disjoint paths connecting $v_S$ and $v_R$. Although an optimal $A$ exists, for a general $N$, it is in generally hard to explicitly describe $A$ and to estimate $\mu_{N,k}$ or $d_{N,k}$. However, we have the following rather obvious upper bound of $d_{N,k}$ when the distance between $u_S$ and $u_R$ is large. \[T204\] Let $G$ be a finite graph and $N = \langle G, u_S, u_R \rangle$, where $dist_G(u_S, u_R) \geq 2k-1$ and $C_0 \cap C_d = \emptyset$ for any $C_0 \in {\cal C}_I$ and $C_d \in {\cal C}_T$. Then $d_{N,k} \geq k(dist_G(u_S, u_R) - 2(k-1))$. For any algorithm $A$, let ${\mathcal B} = C_0, C_1, C_2, \ldots$ be the behavior from any initial configuration $C_0 \in {\mathcal C}_I$ under $A$. We assume that ${\mathcal B}$ eventually reaches a configuration $C_d \in {\mathcal C}_T$. For each robot $r_i \in R$, let $x_i$ and $y_i$ be the vertices that it resides at time 0 and $d$, respectively. Thus $x_i \in C_0$ and $y_i \in C_d$. Obviously $dist_G(u_S, x_i) \leq k-1$ and $dist_G(u_R, y_i) \leq k-1$. Hence each robot moves at least $dist_G(u_S, u_R) - 2(k-1)$ until time $d$, which implies that $d_A(C_0) \geq k (dist_G(u_S, u_R) - 2(k-1))$. The general policy to construct a code with a small transmission delay is to use a swarm with a small number of robots. Finite 8-grids {#sec:eqsizecode} ============== We investigate $8$-grids in the rest of this paper. Let ${\mathbb Z}_{m,n}^2 = \{(i,j) \in {\mathbb Z} \times {\mathbb Z} \mid 0 \leq i \leq m-1, 0 \leq j \leq n-1\}$ and $G_8(m,n) = G_8[{\mathbb Z}_{m,n}^2]$. Clearly, $\mu_{\langle G_8(m,1), u_S, u_R \rangle, k} = 0$ and $d_{\langle G_8(m,1), u_S, u_R \rangle, k} = \infty$ for any $m > k \geq 2$. We illustrate a swarm in a 2D square grid instead of $G_8(m,n)$. Each cell of the square grid is associated with the underlying $x$-$y$ coordinate system and each vertex $(i,j) \in G_8(m, n)$ corresponds to cell $(i,j)$. Hence $N_{G_8(m,n)}((i,j))$ corresponds to the eight adjacent cells, i.e., $(i+1,j), (i+1,j+1),(i,j+1),(i-1,j+1),(i-1, j),(i-1,j-1),(i,j-1),(i+1,j-1)$. We call the cells $\{(i, j) \mid i=0,1,2,\ldots, m-1\}$ the *$j$th row* and the cells $\{(i, j) \mid j=0,1,2,\ldots, n-1\}$ the *$i$th column*. For the sake of simplicity, we assume $u_S = (0, 0)$ and $u_R = (m-1, 0)$. Upper bound of $\mu_{G_8(m,n),k}$ --------------------------------- We derive an upper bound of $\mu_{(G_8(m,n),k)}$ by estimating the size $|{\mathcal C}_I|$ of initial configurations by Observation \[O201\]. Since $u_S$ and $u_R$ are fixed when $G_8(m,n)$ is given, instead of $N = \langle G_8(m,n), u_S, u_R \rangle$, we omit $u_S$ and $u_R$ to denote $\mu_{G_8(m,n),k}$. \[L301\] For any $u_S$ and $u_R$, the number of initial configurations of $G_8$ is at most $2^{6(k-1)}$. Let $C \in {\mathcal C}_I$ be any initial configuration. Consider the depth first traversal of an arbitrary spanning tree of $G[C]$ rooted at $u_S$. The track is represented by a sequence of the eight types of movements, $(+1, 0)$, $(+1, +1)$, $(0, +1)$, $(-1, +1)$, $(-1, 0)$, $(-1, -1)$, $(0, -1)$, and $(+1, -1)$. Thus any configuration $C \in {\mathcal C}_I$ is represented by a sequence of movements whose length is $2(k-1)$. Hence $|{\mathcal C}_I| < 8^{2(k-1)} = 2^{6(k-1)}$. \[C301\] For any $m, n \geq 1$, $\mu_{G_8(m,n),k} \leq 2^{6(k-1)}$. The number of initial configurations in $G_8(m,n)$ is obviously not greater than that of $N = \langle G_8, u_S, u_R \rangle$. When $n = 2$, we can obtain a better bound of $\mu_{G_8(m,2),k}$. A trivial upper bound of $\mu_{G_8(m,2),k}$ is $3^k$, since $|{\mathcal C}_I| \leq 3^k$ and each column of any initial configuration in ${\mathcal C}_I$ must contain at least one robot to keep the connectivity. \[L302\] If $m \geq k$, $\mu_{G_8(m,2)} < (1 + \sqrt{2})^k$. Recall that $u_S = (0,0)$ and $u_R = (m-1,0)$. Let ${\mathcal C}_I$ be the set of initial configurations of the swarm of $k$ robots on $G_8(m,2)$. There are two types of configurations in ${\mathcal C}_I$: 1. The set of configurations $C \in {\mathcal C}_I$ such that $(0,1) \not\in C$. 2. The set of configurations $C \in {\mathcal C}_I$ such that $(0,1) \in C$. Note that $(0,0) \in C$ for any $C \in {\mathcal C}_I$ by definition. Let $n_k$ be the number of configurations such that the $0$th column contains at least one robot. Such a configuration may not contain $(0,0)$. We have the following recurrence formula: $$n_{k} = 2 n_{k-1} + n_{k-2},$$ with $n_{1} = 2$ and $n_{2} = 5$. We can calculate the general term $n_k$ using a standard method. Since the two roots of equation $x^2 - 2x - 1 = 0$ are $1 \pm \sqrt{2}$, we have two equations $$\begin{aligned} n_{k} - (1-\sqrt{2})n_{k-1} &=& (1+\sqrt{2})^{k-2}(3 + 2\sqrt{2}), {\rm~~and} \\ n_{k} - (1+\sqrt{2})n_{k-1} &=& (1-\sqrt{2})^{k-2}(3 - 2\sqrt{2}). \end{aligned}$$ Thus $$n_k = \frac{3+2\sqrt{2}}{2\sqrt{2}} (1+\sqrt{2})^{k-1} - \frac{3-2\sqrt{2}}{2\sqrt{2}} (1-\sqrt{2})^{k-1}.$$ Since $|{\mathcal C}_I| = n_{k-1} + n_{k-2}$, $$\begin{aligned} |{\mathcal C}_I| &=& \frac{3+2\sqrt{2}}{2\sqrt{2}} \left((1+\sqrt{2})^{k-2} + (1+\sqrt{2})^{k-3} \right) - \frac{3-2\sqrt{2}}{2\sqrt{2}} \left((1-\sqrt{2})^{k-2} + (1-\sqrt{2})^{k-3} \right) \\ &=& \sqrt{2} \left(1 + \frac{3}{2\sqrt{2}} \right)(1+\sqrt{2})^{k-2} + \sqrt{2} \left(1 + \frac{3}{2\sqrt{2}} \right) (1-\sqrt{2})^{k-2} \\ &<& \sqrt{2}(1+\sqrt{2})^{k-1} + \sqrt{2}(1+\sqrt{2}) \\ &<& (1+\sqrt{2})^{k-1}(1+\sqrt{2}) \\ &=& (1+\sqrt{2})^{k}. \\\end{aligned}$$ Lower bound of $\mu_{G_8,k}$ ---------------------------- We present an algorithm $ALG_1$ on $G_8(m,2)$ by which we can construct a code $\gamma$ with size $|\gamma| = 2^{k-14}$. Thus $2^{k-14} \leq \mu_{G_8(m,2),k} \leq \mu_{G_8(m,n),k} \leq \mu_{\langle G_8, u_S, u_R \rangle,k}$ hold for any $n \geq 2$. Referring to Lemma \[L302\], $ALG_1$ produces a code with a near optimal code size $2^{k-14}$. However, the transmission delay of the code is not so good. We hence propose another algorithm $ALG_2$ by which we can produce a code with an optimal transmission delay, in the next subsection, although its code size is small and $2^{\lfloor k/2 \rfloor}$. Algorithm $ALG_1$ divides the swarm into a codeword and a controller. Roughly, the codeword is external memory to remember the codeword of the symbol transmitted, and the controller, which is placed following the codeword, is used to keep the external memory stable. The controller consists of $14$ robots, which is further divided into a buffer consisting of four robots, a temporal memory consisting of five robots, a signal consisting of one robot, and a counter consisting of four robots. In the canonical state, the configuration $C$ contains exactly one robot at each column as illustrated in Figure \[figure:G8-m-2-control\]. Algorithm $ALG_1$ shifts $C$ right by [*two*]{} (not one) columns. Suppose that the leftmost cell of $C$ is in the column $i_s$. Then the rightmost cell is in the column ($i_s + k - 1$). The leftmost $(k - 14)$ robots in $C$ represents the symbol it is transmitting. We can thus transmit one of $2^{k-14}$ symbols or, in other words, this part can be regarded as memory of $(k-14)$ bits, by associating the robot in row 0 (resp. row 1) with 0 (resp. 1). The configuration in Figure \[figure:G8-m-2-control\] is thus transmitting a binary sequence 011010100110. We regard the swarm as a sequence of arrays whose element can take $0$ or $1$; the codeword array $code[0..k-15]$, followed by the buffer array $bf[0..4]$, the temporal memory array $cp[0..4]$, the signal array $sign[0]$, and the counter array $cnt[0..3]$ (Figure \[figure:G8-m-2-control\]). ![Illustration of Algorithm $ALG_1$ for $G_8(m,2)$. It illustrates a configuration in the canonical state. The array top is an array representation of the configuration bottom.[]{data-label="figure:G8-m-2-control"}](A1-structure.pdf){width="8cm"} \ \ \ \ \ \ \ \ \ \ \ Starting from a configuration in the canonical state, $ALG_1$ changes the bits in $code[0..k-1]$ one by one, and shifts the swarm right by two columns when the configuration returns to the canonical state. Here it is worth noting that a bit may be changed more than once. First, $cp[0..4]$ copies $code[0..4]$ and two robots in $code[0..1]$ move to make a “wave”. There are at most two columns containing two robots during the shift procedure. When there are two such columns, the right one is called the *head* (column of the wave) and the left one the *tail* (column of the wave). The shift procedure shifts a bit from left to right, and the wave indicates the bits that $ALG_1$ is currently shifting. However, when the wave is created, the bits maintained in the head and the tail are lost. Array $cp$ is used to save these bits. When the shift of the codeword finishes,the wave is in the controller, and the two robots in the head and the tail are simply sent to the right end of the configuration to reset the controller to all zero. The core of the shift procedure is to shift a bit in codeword right by two columns. Let $head$ (resp. $tail$) denote the $x$-coordinate of the head (resp. the tail). The shift is indeed carried out at the tail. The part of the configuration to the left of the tail always remembers a prefix $H = b_1 b_2 \ldots b_h$ of the bit sequence $B = b_1 b_2 \ldots b_{k-14}$ transmitting. To extend $H$ by appending $b_{h+1}$, $ALG_1$ moves the correct robot in the tail to right so that the correct bit $b_{h+1}$ is created in the tail. As a matter of fact, the tail has already moved to right, though. The shift procedure is a repetition of the following phases: 1. The controller updates $cp[0..4]$ to $b_{h+1} b_{h+2} \ldots b_{h+5}$. 2. The correct robot in the tail moves to the next column to create bit $b_{h+1}$, by referring to $cp[0]$ to decide the correct robot. This move of the robot shifts the tail right by one column. To implement the above procedure, the update and the wave shift phases have to be synchronized. Although the second phase finishes in one step, the first phase needs a number of steps to complete, and the update phase must be finished before the shift of tail starts. Algorithm $ALG_1$ keeps inequality $head-tail \leq 2$ during the procedure. The end of phase 2 is indicated by $tail = head-1$, and roughly until the following phase 1 finishes, a robot in the head does not move and hence equation $tail = head-1$ holds. When the update of $cp[0..4]$ in phase 1 finishes with cooperation of $cnt$, a robot in the head moves to the next column. When it moves, equation $tail = head-2$ holds, which however is not a sign to finish phase 1 (and to start phase 2), since $cnt$ must be reset before starting phase 2. The completion of the reset is signaled by $sign$. It is set once $tail = head-1$ holds. It is reset, when $tail = head-2$ holds and $cnt$ is reset to all 0. The update of $cp$ is carried out in the following way: Since $cp$ stores $b_{h+1} b_{h+2} \ldots b_{h+5}$, $ALG_1$ updates it to $b_{h+2} b_{h+3} \ldots b_{h+6}$. For $i = 1, 2, 3$, in this order, $cp[i]$ is copied to $cp[i-1]$, and finally it stores $b_{h+6}$ to $cp[4]$, where $b_{h+6}$ is stored to the right of the head and hence it is possible. The synchronization of these five updates are done with the cooperation of $cnt$. Here $cnt[0..3]$ works as a unary counter, and it is incremented whenever an update finishes (or the update is not necessary because it has already had the correct bit value). Once all bits in $cnt$ are set to 1, as explained, the head starts shifting. Finally, the buffer $bf[0..3]$ is used when $ALG_1$ is going to shift two bits $b_{k-15}$ and $b_{k-14}$ which were stored in the last two columns. These column are now the tail and the head, and the bits are stored in $cp[0..1]$. The buffer is used to store $b_{k-15}$ and $b_{k-14}$. It is also necessary to separate $code$ and $cp$, so that the wave will not directly shift into $cp$ and the bits will not be lost. In the following description of Algorithm $ALG_1$, $cnt$ increments as $0000 \to 1000 \to 1100 \to 1110 \to 1111$, and the counter value is $c$ when the number of $1$’s in $cnt[0..3]$ is $c$. [**Algorithm $ALG_1$**]{} 1. When the configuration is canonical, first store $b_{1} b_{2} \ldots b_{5}$ to $cp$, then move the robots in the first and the second column to the third and the fifth column to initialize the wave where $head = 3$ and $tail = 1$, i.e., $head - tail = 2$ holds after the moves. 2. When there are no columns in $code$ that contain two robots, but there are head and tail of the wave, move the two robots to right and append to the right end of the configuration, which move reset the controller all 0, and the configuration returns to the canonical state. 3. When the wave is in $code$ and $sign=0$ (Note that $cnt = 0000$ always holds in this case.): 1. If $head-tail = 2$, referring to $cp[0]$, the correct robot in the tail moves to an empty cell of the $(tail+1)$st column. 2. If $head-tail = 1$, set $sign$ to $1$. 4. When the wave is in $code$ and $sign = 1$: 1. If $cnt = c < 4$, copy $cp[c+1]$ to $cp[c]$ if they are different. Then increment $cnt$. 2. If $cnt = c = 4$, copy $code[head+1]$ to $cp[4]$ if they are different. (As a result, $cp[4] = code[head+1]$ holds.) If $head-tail = 1$, one of the two robots in the head moves to an empty cell of column $head+1$. If $head-tail = 2$, first reset $cnt$ to 0, and then reset $sign$ to 0 (when $cnt = 0$). \[L303\] There is a code $\gamma$ based on algorithm $ALG_1$ whose size is $|\gamma| = 2^{k-14}$. The transmission delay $d_{ALG_1}$ is at most $(10k-123.5)(m - k)$. The fact $|\gamma| = 2^{k-14}$ is obvious from the definition of algorithm $ALG_1$. As for the transmission delay, one cycle of transitions shifts a configuration in the canonical state right by two columns. During the movement, the wave moves from the first two bits of the codeword to the end of the counter. Thus the total number of shifts of the wave is $2k$. For each bit of the codeword, each bit of the counter moves twice, the signal twice, and each bit of the copy part at most once. Hence, the total number of robot moves in the controller is $(k-14)(5+2+8) = 18k-252$. In the beginning of the shift, each bit of the copy part moves at most once. Hence the total number of robot moves to move the codeword by one column is at most $$\frac{2k + (18k-252) + 5}{2} = 10k-123.5.$$ The distance from the last bit of the counter to $u_R$ is at most $(m - k)$. Hence $d_{ALG_1} \leq (10k-123.5)(m-k)$. An obvious consequence of Lemma \[L303\] is that $2^{k-14} \leq \mu_{G_8(m,2),k} \leq \mu_{G_8(m,n)}$ for all $n \geq 2$. The transmission delay of the code by $ALG_1$ is roughly $10k(m-k)$ when $m (\gg k)$ is large, which is not so fast compared with its lower bound $k(m-2(k-1))$ of Theorem \[T204\]. This fact motivates the study of a faster algorithm in the next subsection. Algorithm $ALG_2$ for faster information transmission ----------------------------------------------------- It is easy to design a code with a small transmission delay if we do not need to care the code size. In this subsection, we design a code based on an algorithm $ALG_2$ whose code size is $2^{\lfloor k/2 \rfloor}$. However its transmission delay is near optimal. For simplicity, we assume that $k$ satisfies $\lfloor k/2 \rfloor = \ell$ for some positive integer $\ell \geq 2$. Algorithm $ALG_2$ divides the swarm into a codeword and a copy, each of which consists of $\ell$ robots. In the same way as $ALG_1$, the codeword represents $\ell$ bits and the copy follows the codeword. When $k$ is odd, a single robot follows the copy, and it does not represent any bit. In the canonical state, each column contains exactly one robot as illustrated in Figure \[fig:shift-by-1\]. Algorithm $ALG_2$ shifts the canonical state by one column. Suppose that the leftmost cell of the canonical state is in the column $i$. Then the rightmost cell is in the column $(i+k-1)$. The left $\ell$ robots represent the symbol it is transmitting. We can thus transmit one of $2^{\ell}$ symbols, or in other words, these $\ell$ robots can be regarded as memory of $\ell$ bits $B =b_1 b_2 \ldots b_{\ell}$. We now regard the robots in the left $\ell$ columns as an array $code[0..\ell-1]$ and those in the right $\ell$ columns as an array $copy[0..\ell-1]$. When $k$ is odd, the last robot is put in the $0$th row of column $i+k-1$. The canonical state in Figure \[fig:shift-by-1\] represents a binary sequence $010$. Starting from a configuration where the swarm is in the canonical state, $ALG_2$ shifts the swarm right by one column when the configuration returns to the canonical state. In the canonical state, $ALG_2$ first shifts the codeword by generating a wave; the leftmost robot moves to the empty cell of its right column. From now on, we regard the column with two robots as the wave.[^4] Once the wave is initialized, the wave moves to right by repeating the following procedure; when the wave is in the $j$th column from the left end of the swarm, one of the two robots in the wave moves to the empty cell of the $(j+1)$st column so that the $j$th column represents $b_{j}$, which is stored at $copy[j-1]$. After the movement, the wave is in the $(j+1)$st column and the left $j$ columns of the swarm represent $b_1 b_2 \ldots b_j$. When the wave reaches the copy, it repeats the same procedure until it reaches the right end of the swarm. Specifically, when the wave reaches the column of $copy[0]$ (i.e., $\ell$th column from the left end of the swarm), $copy[\ell-1]$ specifies which of the two robots in the wave move to which cell because $\ell \geq 2$. Now, $B$ is fully stored in the left $\ell$ columns of the swarm and the wave proceeds to right with copying these columns. When $k$ is odd, the last robots moves to the $0$th row of the right end of the swarm. Then, the swarm retains the canonical state. ![Codeword for $010$ in $ALG_2$. []{data-label="fig:structureA2"}](A2-structure.pdf){width="3cm"} ![Illustration of a right shift by one column realized by $ALG_2$. A robot represented by a black circle is stationary, while a one represented by a white circle moves and yields a new configuration.[]{data-label="fig:shift-by-1"}](A2.pdf){width="14cm"} \[L331\] There is a code $\gamma$ based on algorithm $ALG_2$ whose size is $|\gamma|=2^{\lfloor k/2 \rfloor}$. The transmission delay $d_{ALG_2}$ is $k(m-k)$. The fact that $|\gamma|=2^{\lfloor k/2 \rfloor}$ is obvious from the definition of algorithm $ALG_2$. As for the transmission delay, one cycle of transmission shifts the canonical state right by one column. During the movement, the wave moves from the left end of the swarm to the right end with making one robot move in each column. Thus the total number of movement in a cycle is $k$. Consequently, $d_{ALG_2} = k(m-k)$. By comparing two transmission delays in Lemmas\[L303\] and \[L331\], when $m (\gg k)$ is large, $ALG_2$ is 10 times as fast as $ALG_1$. Variable swarm size code {#sec:varsizecode} ======================== The information transmission scheme in Introduction allow us to assign configurations of different sizes to different symbols. Suppose that, in a code $\gamma$, a configuration with size $k_i$ is assigned to a symbol $s_i$ whose occurrence probability is $s_i$. Let $K_{\gamma} = \sum_{i=1}^{\alpha} k_i p_i$ be the average swarm size of $\gamma$ and define $K^* = \min_{\gamma} K_{\gamma}$, which is the optimal average swarm size. This section analyzes bounds of $K^*$. Lower bounds of $K^*$ --------------------- We derive a lower bound of $K^*$ on $G_8(m,n)$. We can assume without loss of generality that $p_i \leq p_{i+1}$ for $i = 1, 2, \ldots , \alpha - 1$ and $k_i \leq k_{i+1}$ for $i = 1, 2, \ldots , \alpha - 1$, since, if $k_i > k_{i+1}$ holds, we could reduce $K^*$ by exchanging the assignment to $s_i$ and $s_{i+1}$. The size of any code realized by a swarm of $k$ robots on $G_8(m,n)$ is bounded from above by $2^{6(k-1)}$, i.e., $\mu_{G_8(m,n)} \leq 2^{6(k-1)}$ by Lemma \[L301\]. A lower bound of $K^*$ can be achieved by a code constructed under the assumption that $2^{6(k-1)}$ configurations of size $k$ are assignable to symbols for each $k$. Let $H(S) = - \sum_{i = 1}^{\alpha} p_i \log p_i$ be the entropy of $S$. \[T401\] Suppose that $m > k$ and $n\geq 2$. Then for any code for $S$ on $G_8(m,n)$, $$K ^* > \frac{1}{6} H(S) + 1 - \frac{1}{6} \log \lceil \log \frac{63 \alpha+1}{6} \rceil.$$ By assumption $k_i \leq k_{i+1}$ holds for $i = 1, 2, \ldots , \alpha - 1$. We first calculate $k_{\alpha}$. By assumption $k_{\alpha} = \max_i k_i$ holds. Then $\sum_{k = 1}^{k_{\alpha}-1} 2^{6(k-1)} < \alpha \leq \sum_{k = 1}^{k_{\alpha}} 2^{6(k-1)}$. By a simple calculation, we have $k_{\alpha} = \left\lceil \log \frac{63 \alpha+1}{6} \right\rceil$. Let $q_i = 2^{-6(k_i - 1)}/k_{\alpha}$. Then, $q_i > 0$ for all $i = 1, 2,\ldots, \alpha$ and $$\sum_{i = 1}^{\alpha} q_i = \sum_{k = 1}^{k_{\alpha}} \frac{2^{-6(k-1)}}{k_{\alpha}} 2^{6(k-1)} \leq 1.$$ By Shannon’s lemma, we have $$\begin{aligned} H(S) = - \sum_{i=1}^{\alpha} p_i \log p_i &<& - \sum_{i=1}^{\alpha} p_i \log q_i \\ &=& - \sum_{i=1}^{\alpha} p_i \log \frac{2^{-6(k_i-1)}}{k_{\alpha}} \\ &=& \sum_{i=1}^{\alpha} p_i \left(6(k_i-1) + \log k_{\alpha} \right) \\ &=& 6 \sum_{i=1}^{\alpha} p_i k_i -6 + \log k_{\alpha} \\ &=& 6 K^* -6 + \log k_{\alpha}. \end{aligned}$$ Thus $$\begin{aligned} K^* &>& \frac{1}{6} H(S) +1 - \frac{1}{6} \log k_{\alpha} \\ &>& \frac{1}{6} H(S) + 1 - \frac{1}{6} \log \left\lceil \log \frac{63 \alpha+1}{6} \right\rceil.\end{aligned}$$ Note that $\log \log \alpha > \frac{1}{6} \log \left\lceil \log \frac{63 \alpha+1}{6} \right\rceil - 1$, we have $$K^* > \frac{1}{6}(H(S) - \log \log \alpha).$$ A bound of $K^*$ for any code on $G_8(m,2)$ can be derived by using Lemma \[L302\], by a similar discussion. \[C401\] Suppose that $m > k$. Then for any code for $S$ on $G_8(m,2)$, $$K^* > 0.78 H(S) - 0.79 \log \left( 1 + 0.79 \log \alpha \right).$$ Since $1 + 0.79 \log \alpha < \log \alpha$ when $\alpha \geq 2^5$, $K^* > 0.78 H(S) - 0.79 \log \log \alpha$, which may be approximated by $\frac{4}{5}(H(S) - \log \log \alpha)$ when $\alpha \geq 2^5$. Note that regardless of $\alpha$, $H(S)$ can be 0, when $p_1 = 1$, which means that these lower bounds can be negative and become meaningless. On the other hand, when $p_i = 1/\alpha$ for all $i = 1, 2, \ldots , \alpha$, $H(S)$ take the maximum value $\log \alpha$. Thus $K^*$ on $TG_8(m,2)$ is roughly bounded from below by $0.78 \log \alpha$, when $\log \log \alpha$ is negligible. Upper bound of $K^*$ based on $ALG_1$ ------------------------------------- To derive an upper bound of $K^*$, we construct a code $\gamma_{ALG_1}$ for $S$ based on $ALG_1$. Although there are $2^{k-14}$ size $k$ configurations assignable to symbols for any $k > 14$, there are no size $k$ configurations assignable to symbols for $k \leq 14$. For each $k \leq 14$ however, we can easily construct a code with code size 1 by using a trivial locomotion algorithm illustrated in Figure \[figure:loco\]. ![Locomotion of a small number of robots on $G_8(m,2)$[]{data-label="figure:loco"}](locomotion.pdf){width="11cm"} Thus we use the configurations defined by $ALG_1$ as well as one configuration defined above for each $k \leq 14$. Let $X$ be the set of all configurations assignable. Code $\gamma_{ALG_1}$ is defined as follows: We assign a configuration of size $k_i$ in $X$ to a symbol $s_i$ in such a way that $k_i \leq k_{i+1}$ holds for $i = 1, 2, \ldots , \alpha - 1$. By definition, $k_{\alpha} \leq \alpha$. Let $K_{ALG_1}$ denote the average code size of $\gamma_{ALG_1}$. \[T402\] When $m > \alpha$, $K_{ALG_1} < H(S) + 15$ holds. Consider the “tower” of codewords of $\gamma_{ALG_1}$. Each of the first $14$ levels contains a single codeword, and the $\ell$th layer contains $2^{\ell-14}$ codewords for $\ell \geq 15$. Shannon’s noiseless coding theorem guarantees the existence of a prefix-free code whose average code length is smaller than $H(S) + 1$. Let $\ell_i$ be the length of the codeword of $s_i$ in such a prefix-free code. By the definition of $\gamma_{ALG_1}$, $k_i - 14 \leq \ell_i$. Thus $$K_{ALG_1} \leq \sum_{i=1}^{\alpha} p_i (\ell_i + 14) < H(S)+15.$$ Thus the performance of $\gamma_{ALG_1}$ in terms of the average code size is sufficiently good on $G_8(m,2)$, since, very roughly, $K_{\gamma}$ of any code $\gamma$ on $G_8(m,2)$ is bounded from below by $\frac{4}{5}(H(S) - \log \log \alpha)$. Code $\gamma_{ALG_1}$ can be used not only on $G_8(m,2)$ but also on $G_8(m,n)$ for $n > 2$. On $G_8(m,n)$, its performance may not be good, however, since our lower bound of $K^*$ on $G_8(m,n)$ is $\frac{1}{6}(H(S) - \log \log \alpha)$, and there is still a large gap from $K_{ALG_1}$. Average transmission delay -------------------------- We next analyze the average transmission delay. For any code $\gamma$ on $G_8(m,2)$, let $D_{\gamma} = \sum_{i=1}^{\alpha} d_i p_i$. We are interested in the optimal average transmission delay $D^* = \min_{\gamma} D_{\gamma}$. First we derive a lower bound of $D^*$. Consider any code $\gamma$ on $G_8(m,n)$, where $m > 2 \alpha$. We may assume without loss of generality that $\alpha \geq k_i$ for any $i = 1, 2, \ldots , \alpha$. By Theorem \[T204\], $d_i \geq k_i (m - 2k_i + 1)$. Since $\alpha \geq k_i$, $$D_{\gamma} \geq (m+1) K_{\gamma} - 2 \sum_{i = 1}^{\alpha} k_i^2 p_i \geq (m + 1 - 2\alpha)K_{\gamma}.$$ Since $K_{\gamma} \geq K^*$, we have: \[O401\] $D^* \geq (m + 1 - 2 \alpha) K^*$. To derive an upper bound of $D^*$, Let $D_{ALG_1}$ be the average delay of $\gamma_{ALG_1}$. By Lemma \[L303\], $d_i \leq (10k_i-123.5)(m-k_i)$. Since $\alpha \geq k_i$ holds for all $i = 1, 2, \ldots , \alpha$, we have: \[O402\] $D_{ALG_1} \leq 10 m K_{ALG_1}$. Thus $\gamma_{ALG_1}$ may not be a good choice from the view of the average transmission delay, even considering the fact that $K_{ALG_1} \leq H(S) + 15$. The average number of steps necessary to move one column is approximated from above by $10 H(S)$ when $m \gg \alpha$. Since algorithm $ALG_2$ has a smaller transmission delay, we analyze a code $\gamma_{ALG_2}$ based on $ALG_2$. Since $ALG_2$ is defined only for $k \geq 4$, like $\gamma_{ALG_1}$, consider set $X$ that contains standard configurations defined in $ALG_2$ for all $k \geq 4$, as well as a single configuration of size $i$ for each of $i = 1, 2, 3$. Like $\gamma_{ALG_1}$, $\gamma_{ALG_2}$ is defined as follows: We assign a size $k_i$ configuration in $X$ to a symbol $s_i$ in such a way that $k_i \leq k_{i+1}$ holds for $i = 1, 2, \ldots , \alpha - 1$. Let $K_{ALG_2}$ denote the average code size of $\gamma_{ALG_2}$. \[T403\] Suppose that $m > \alpha > 1$. Then $K_{ALG_2} < 2H(S)$. Recall that $p_i \geq p_{i+1}$ for $i = 1, 2, \ldots , \alpha - 1$. Then $p_i \leq 1/i$, and hence $-\log p_i \geq \log i$. Consider the “tower” of codewords of $\gamma_{ALG_2}$. Each of the first $3$ levels contains a single codeword, and the $\ell$th layer contains $2^{\lfloor \frac{k}{2} \rfloor}$ codewords for $k \geq 4$. Code $\gamma_{ALG_2}$ packs the symbols in its order $s_1, s_2, \ldots, s_{\alpha}$ to codewords from the first level to the higher level in its order. We then divide the tower into three small towers, $T_1$, $T_2$, and $T_3$, where $T_1$ consists of the first 3 levels, $T_2$ consists of even levels greater than 3, and $T_3$ consists of odd levels greater than 4. Hence, $h$th level of $T_2$ (and $T_3$ also) corresponds to a binary tree in the sense that it contains $2^{h-1}$ codewords. First consider a codeword for $s_i$ in $T_2$. Let $h_i$ be the level that contains a configuration assigned to $s_i$ in $T_2$. The total number of symbols assigned to the level lower than $(h_i-1)$ in $T_1$, $T_2$, and $T_3$ is smaller than $i$. Thus we have $3 + 2(1+2+\cdots + 2^{h_i-1}) < i$, which implies that $h_i < \log i - 1$. Since the size of $k_i$ is $2h_i + 2$, $k_i < 2 \log i$. We then consider a codeword $s_i$ in $T_3$. Let $h_i$ be the level of $s_i$ in $T_3$. In the same way, $3 + 2(1+2+\cdots + 2^{m_i-1}) + 2^{m_i} < i$, which implies $h_i < \log i - \log 3$. Since the size of $k_i$ is $2h_i + 3$, $k_i < 2 \log i + 3 - 2 \log 3 < 2 \log i$. Thus, we have $$\begin{aligned} \sum_{i=1}^{\alpha} p_i k_i &<& \sum_{i=1}^{\alpha} 2 p_i \log i \\ &=& 2 \sum_{i=1}^{\alpha} p_i \log i \\ &\leq& 2 \sum_{i=1}^{\alpha} p_i (- \log p_i) \\ &=& 2 H(S). \end{aligned}$$ Thus the upper bound of $K_{ALG_2}$ obtained by the theorem is roughly twice as much as that of $K_{ALG_1}$. Let $D_{ALG_2}$ be the average transmission delay of $\gamma_{ALG_2}$. By Lemma \[L331\], $d_i = k_i (m - k_i)$. Thus \[O402\] $D_{ALG_2} \leq m K_{ALG_2}$. Since $K_{ALG_2} < 2 H(S)$, we have $\leq D^* \leq 2m H(S)$. The average number of steps necessary to move one column is approximated from above by $2 H(S)$ when $m \gg \alpha$. Thus the variable swarm size code by $ALG_2$ is 5 times as fast as that by $ALG_1$. Conclusion {#sec:concl} ========== We proposed an information transmission scheme by a swarm of anonymous oblivious mobile robots on a graph. We mainly analyzed the performance of our scheme in terms of code size and transmission delay in the 8-gird and proposed two algorithms one achieves exponential code size with large transmission delay and the other achieves optimal transmission delay with small code size. We finally extended these algorithms for variable swarm size codes. There are many open problems related to the proposed scheme. First, we could not find any algorithm with optimal code size and optimal transmission delay. Second, the gap between the upper bound and the lower bound of the (expected) swarm size of the fixed swarm size code and of the variable swarm size code needs to be closed. One approach is a more sophisticated technique to upper bound the number of initial configurations and terminating behaviors. Third, parallel movement of robots might speed up the transmission and makes algorithms simpler. One of the most important future directions is robustness. We put our basis on Shannon’s noiseless coding theorem, and the next step is to consider faulty robots. We believe Shannon’s noisy channel coding theorem help the investigation. Another direction is to investigate local algorithms that restricts the visibility of the robots to a constant distance. [^1]: Corresponding author. Faculty of Information Science and Electrical Engineering, Kyushu University, Japan. E-mail: `[email protected]` [^2]: Faculty of Information Science and Electrical Engineering, Kyushu University, Japan. E-mail: `[email protected]` [^3]: Since all robots know $G$ and $C$, $A$ on all robots can choose the same vertex $u \in C$ to move the robot at $u$ to a different vertex $v \in N_G(u) \setminus C$. Algorithm $A$ may output $v = u$ on all robots $r$. [^4]: The wave in $ALG_2$ consists of just one column and this is different from the wave in $ALG_1$, that consists of two columns for synchronization.
{ "pile_set_name": "ArXiv" }
--- author: - 'C. Argiroffi' - 'G. Micela' - 'A. Maggio' bibliography: - 'argiroffi.bib' title: 'Simbol-X capability of detecting the non-thermal emission of stellar flares' --- Introduction ============ Stellar flares are phenomena where the magnetic field rapidly releases large amount of energy. It is supposed that: 1) magnetic reconnections generate non-maxwellian population of fast particles; 2) these particles hit and heat the chromospheric plasma, evaporating it, and producing non-thermal emission (hard X-rays and $\gamma$-rays); 3) the evaporated chromospheric plasma fills coronal structures and generates thermal emission (mostly in the UV and soft X-ray bands). In solar data a power-law spectrum, compatible to the bremsstrahlung emission of fast particles hitting a thick target, is observed [i.e. @HudsonRyan1995]. This hard power-law emission precedes the soft X-ray and UV thermal emission. Information on stellar emission in the hard X-ray band is very poor, since most X-ray observations are dedicated only to the soft band. Recently @OstenDrake2007 presented the detection of non-thermal emission in the hard X-rays during a stellar flare. The study of the flare non-thermal emission allows to study: 1) the depicted scenario for stellar flares, and hence whether and how the thermal emission is caused by the non-thermal particles; 2) the flare energy balance, since only a small amount of released energy is lost by thermal radiation. We investigate the Simbol-X capability of detecting non-thermal emission associated with stellar flares, and distinguishing it from a hot thermal one. Results about this issue are important also for other astrophysical subjects where non-thermal emission is likely present (galaxy clusters, SNR). Method ====== We adopt the following procedure: 1) we simulate MPD and CZT spectra adopting as model the superposition of a thermal and a non-thermal component; 2) we assume as null hypothesis that the simulated spectra can be described by thermal components only; hence we fit simultaneously MPD and CZT spectra adopting a 2-$T$ optically thin plasma as model; 3) we assume that the null hypothesis of only thermal emission is rejected, and hence non-thermal emission is detected and recognized, when the [*EM*]{} of the hottest component is significantly greater than zero and its temperature is larger than 300MK. To produce simulated MPD and CZT spectra we consider typical flares of nearby active stars. We are not interested in time resolved spectroscopy, hence we consider all the parameters as time-averaged over the flare duration. In all cases we consider a flare duration of 20ks. For each star we produce different simulated spectra: we keep frozen the thermal component, and vary the normalization of the non-thermal component. With this choice for a fixed thermal emission component we are able to check the amount of non-thermal emission needed to detect it. We use the [mekal]{} XSPEC model to simulate the thermal component (Table \[tab:par\] contains the parameters of the thermal emission for the selected stellar targets). We simulate the non-thermal emission using the broken power law XSPEC model ([bknpower]{}). We assume: a photon index of 3 for $E<E_{0}$, in order to have a negligible contribution of the non-thermal component at low energy; a break point $E_{0}=10$keV; a photon index of -2 for $E>E_{0}$. Figure \[fig:mod\] shows one of the explored models. [lcccc]{} Name & $D$ & [*EM*]{} & $T$ & $\log L_{XT}^{a}$\ & & & &\ & (pc) & ($10^{50}\,{\rm cm^{-3}}$) & (MK) & (erg/s)\ Prox Cen & 1.3 & 5 & 20 & 27.5\ AD Leo & 4.7 & 10 & 20 & 27.7\ EV Lac & 5.0 & $2\times10^2$ & 20 & 28.9\ AB Dor & 14.9 & $2\times10^2$ & 20 & 29.0\ HR 1099 & 29.0 & $1\times10^3$ & 20 & 29.6\ PMS stars & 150 & $5\times10^4$ & 20 & 31.3\ \ Results and conclusions ======================= Inspecting all the simulated flares we find that: non-thermal emission is detected and recognized with Simbol-X observations whether the CZT detector collects more than $\sim20$ photons in the $20-80$keV band. For each simulated spectrum we compute the thermal and non-thermal luminosity. We compare the results obtained with those derived from solar flares by @IsolaFavata2007 [see Fig. 5]. We find that for some of the considered flares the non-thermal emission can be detected with Simbol-X if thermal vs. non-thermal solar flare relation is valid also at high luminosities.
{ "pile_set_name": "ArXiv" }
--- abstract: 'In the framework of a holographic QCD approach we study an influence of matters on the deconfinement temperature, $T_c$. We first consider quark flavor number ($N_f$) dependence of $T_c$. We observe that $T_c$ decreases with $N_f$, which is consistent with a lattice QCD result. We also delve into how the quark number density $\rho_q$ affects the value of $T_c$. We find that $T_c$ drops with increasing $\rho_q$. In both cases, we confirm that the contributions from quarks are suppressed by $1/N_c$, as it should be, compared to the ones from a gravitational action (pure Yang-Mills).' author: - Youngman Kim - 'Bum-Hoon Lee' - Siyoung Nam - Chanyong Park - 'Sang-Jin Sin' title: Deconfinement phase transition in holographic QCD with matter --- Introduction ============ To understand QCD phase structure at finite temperature and/or density has been an fascinating theme in hadron physics. Some of studies so far are basically based on phenomenological models or effective field theories of QCD such as chiral perturbation theory (ChPT) and Nambu-Jona-Lasinio (NJL) model. Lattice QCD has been a powerful tool for QCD phase diagram at finite temperature and, recently, at finite density. Recently interesting developments in AdS/CFT [@adscft] to study strongly interacting system such as QCD, which goes under the name of AdS/QCD, have been made. Confinement is realized with an infrared (IR) cut off $z_{IR}$ in AdS space [@polchinski], and flavors are introduced by adding extra probe branes [@karch]. More phenomenological approaches were also suggested to construct a holographic model dual to QCD, for example,  [@EKSS; @PR; @Brodsky]. The deconfinement temperature is estimated in a Hawking-Page type transition analysis in cutoff AdS space in  [@Herzog:2006ra]. In this analysis, the contribution from mesons (quarks) are not considered, since they are suppressed by $1/N_c$ compared to the gravitational part, and consequently, $T_c$ may correspond to that of pure Yang-Mills. The quark (baryon) chemical potential is introduced through AdS/CFT to study physics of dense matter [@dAdSQCD]. In the present work, we consider contributions from mesons, motivated largely by lattice QCD results, to see flavor $N_f$ and quark number density dependence of $T_c$. As it is well know from many lattice QCD results so far, the details of the transition strongly depends on the number of quark flavors. The transition temperature with no dynamical quarks is roughly $270~{\rm MeV}$, for instance see [@Karsch], while with dynamical quarks it is around $170~{\rm MeV}$ [@KLP; @Aoki]. In addition, $T_c$ will decreases with increasing $N_f$ [@KLP]. Apart from the temperature axis in the QCD phase diagram, lattice QCD is now investigating the QCD equation of state at nonzero chemical potential [@FK; @AEH]. In [@FK], it is shown that the transition temperature decreases with the baryon chemical potential up to  $1~{\rm GeV}$. Finally, we remark that $T_c$ in the present work denotes the critical temperature of the deconfinement (first order) phase transition, while at low density there is a cross over as observed in lattice QCD  [@FK; @Aoki; @Bernard]. We attribute this difference to the large $N_c$ nature of the Hawking-Page type transition. A Hawking-Page type transition with quark flavors ================================================= A study [@Herzog:2006ra] based on a Hawking-Page type transition in the AdS/QCD models calculated the deconfinement temperature. In the analysis, the contribution from mesons are not considered, since they are suppressed by $1/N_c$ compared to the gravitational part: the gravitational coupling scales as $\kappa \approx g_s \approx 1/N_c$, and the contribution from the mesons scales only as $N_c$. Here we briefly summarize the analysis of  [@Herzog:2006ra] done in the hard wall model [@EKSS; @PR]. In the hard wall model, the AdS space is compactified such that $z_0<z<z_{IR}$, where $z_0\rightarrow 0$. The value of $z_{IR}$ is fixed by the rho-meson mass ($m_\rho$) at zero temperature: $m_\rho (\simeq 770~{\rm MeV})\simeq3\pi/(4z_{IR})$$\rightarrow$ $1/z_{IR}\simeq 320~{\rm MeV}$[@EKSS; @PR]. The Euclidean gravitational action given by $$\label{action1} S_{grav} ~=~ -\frac{1}{2\kappa^2} \int d^5x \sqrt{g}\left(\textrm{R}+\frac{12}{L^2}\right)$$ where $\kappa^2 = 8\pi G_5$ and $L$ is the length scale of the $AdS_5$, there are two relevant solutions for the equations of motion derived from the above action.The one is cut-off thermal AdS(tAdS) with the line element $$ds^2=\frac{L^2}{z^2}\left(d\tau^2+dz^2+d\vec{x}^2_3\right),$$ where the radial coordinate runs from the boundary of tAdS space $z=0$ to the cut-off $z_{IR}$, which corresponds to an infrared cut-off in energies proportional to $1/z_{IR}$ from the point of view of the boundary dual theory. The other solution is cut-off AdS black hole(AdSBH) with the line element $$ds^2=\frac{L^2}{z^2}\left(f(z)d\tau^2+\frac{dz^2}{f(z)}+d\vec{x}^2_3\right)$$ where $f(z)=1-(z/z_h)^4$ and $z_h$ is the horizon of the black hole. Note that there will be no cut-off in this space for $z_{IR}<z_h$. The Hawking temperature of the black hole solution is $T = 1/(\pi z_h)$ which is given by regularizing the metric near the horizon. In the tAdS case, the periodicity in the Euclidean time-direction is fixed by comparing two geometries at an UV cut-off $\epsilon$ where the periodicity of the time-direction in both cases is locally the same. Then, the time periodicity of tAdS is given by $$\beta = \pi z_h \sqrt{f(\epsilon)}.$$ Now we calculate the action density $V$, which is defined by the action divided by the common volume factor of $R^3$. The regularized action density of the tAdS is given by $$V_1(\epsilon) = \frac{4L^3}{\kappa^2} \int^{\beta '}_{0} d\tau \int^{z_{IR}}_{\epsilon}\frac{dz}{z^5}\, ,$$ and that of the AdSBH is given by $$V_2(\epsilon) = \frac{4L^3}{\kappa^2}\int^{\pi z_h}_{0} d\tau \int^{\bar{z}}_{\epsilon}\frac{dz}{z^5}$$ where $\bar{z} = min(z_{IR},z_h)$. Then, the difference of the regularized actions is given by $$\Delta V_g = \lim_{\epsilon\rightarrow 0}\left[V_2(\epsilon) -V_1(\epsilon)\right]= \left\{\begin{array}{ll} \frac{L^3 \pi z_h}{\kappa^2} \frac{1}{2z_h^4} & z_{IR} < z_h\\ \\ \frac{L^3 \pi z_h}{\kappa^2} \left( \frac{1}{z_{IR}^4} - \frac{1}{2z_h^4}\right) & z_{IR} > z_h.\end{array} \right.$$ This is the result of [@Herzog:2006ra] in the hard wall model. When $\Delta V_g$ is positive(negative), tAdS (the black hole) is stable. Thus, at $\Delta V_g=0$ there exists a Hawking-Page transition. In the first case $z_{IR} < z_h$, there is no Hawking-Page transition and the thermal AdS is always stable. In the second case $z_{IR} > z_h$, the Hawking-Page transition occurs at $$\label{tempads} T_0 = 2^{1/4}/(\pi z_{IR})$$ and at low temperature $T < T_0$ (at high temperature $T > T_0$) the thermal AdS (the AdS black hole) geometry becomes a dominant background. Now, we include the bulk matter into the theory. The action for mesons is given by $$\label{action2} S_{matter} ~=~ M_5 \int d^5x \sqrt{g}~ \textrm{Tr} \left[\frac{1}{2}|D_\mu \Phi|^2 + \frac{1}{2} M_{\Phi}^2|\Phi^2| +\frac{1}{4}\left(F_{L}^2 +F_{R}^2\right)\right]$$ where $L^2 M_{\Phi}^2=-3$, $D_\mu \Phi=\partial_\mu \Phi +i A_{L\mu} \Phi -i\Phi A_{R\mu}$, $A_{L,R} = A_{L,R}^a t^a$ and $F_{\mu\nu}=\partial_\mu A_\nu - \partial_\nu A_\mu +i[A_\mu,A_\nu]$. The vector fields and the axial vector fields are defined by $V=(A_L +A_R)/\sqrt{2}$ and $A=(A_L -A_R)/\sqrt{2}$ respectively. For later convenience, we present some relations here: $$\frac{1}{\kappa^2}=\frac{1}{8\pi G_5}, ~~~~ \frac{1}{G_5}=\frac{32N^2_c}{\pi L^3}, ~~ ~~ \textrm{and}~~~~M_5=\frac{N_c}{12\pi^2 L},$$ where the second relation comes from Ref. [@Csaki:2006ji]. In this section, we study the quark flavor number dependence of the critical temperature of the Hawking-Page type transition at zero density. Here we assume that there is no condensate in the vector and axial-vector channel, and so we turn off the vector gauge fields temporarily until next section. On thermal AdS background, the solution of the equation of motion for a scalar field is given by $$v(z) = a z + b z^3 \, ,$$ where $v(z)=<\Phi>$. According to the AdS/CFT correspondence, the coefficient $a$ is a mass of the boundary quark and $b$ corresponds to chiral condensate $<\bar qq>$. When considering the AdS black hole background, the boundary theory becomes a finite temperature field theory. The equation of motion of the scalar field is $$\left[\partial_z^2 -\frac{4-f}{zf}\partial_z +\frac{3}{z^2f}\right]v(z) =0$$ and the solution of the equation is given by [@GY; @Kim:2006ut] $$\label{sinbh} v(z) = a \cdot _2F_1\left(\frac{1}{4}, \frac{1}{4}, \frac{1}{2}, \frac{z^4}{z_h^4}\right) z + \tilde b \cdot _2F_1\left(\frac{3}{4}, \frac{3}{4}, \frac{3}{2}, \frac{z^4}{z_h^4} \right) z^3 .$$ For simplicity, we take the chiral limit where the quark mass is zero, $a=0$. We assume that chiral symmetry restoration and the deconfinement take place at the same temperature, and so the chiral condensate is zero in the deconfined phase. Since AdS black hole background corresponds to a deconfined phase of the boundary theory, we have $\tilde b=0$. At low temperature described by the thermal AdS background, the boundary theory corresponds to a confined phase, so $b\neq 0$. Therefore, the scalar field solutions in two backgrounds are given by $v_{tAdS}\simeq b z^3$ and $v_{BH}\simeq 0$. Now we calculate the matter contribution to the Hawking-Page transition. For the thermal AdS, the scalar field contribution to the action is $$V_{1m} = \frac{L^3\pi z_h}{2}\cdot 3M_5 N_f b^2z^2_{IR}.$$ In the black hole background, since $v_{BH}\simeq 0$ there is no contribution of the scalar field, i.e. $V_{2m}=0$. Therefore, the values of $\Delta V_m$ of the matter part for $z_{IR} < z_h$ and $z_{IR} > z_h$ are the same : $$\Delta V_m = -\frac{L^3 \pi z_h}{\kappa^2}\cdot \frac{b_t^2 L^2 N_f z_{IR}^2}{32N_c}.$$ Note that the parameter $b$ is given by $$a \simeq 0 ~~\textrm{and}~~ b \simeq \frac{\xi}{L z_{IR}^3}$$ where $\xi \simeq4$ [@PR]. Finally, the total difference of the action in each background becomes $$\Delta V = \left\{\begin{array}{ll}\frac{L^3 \pi z_h}{2\kappa^2}\left[ \frac{1}{z_h^4} - \left(\frac{N_f}{N_c}\right)\frac{\xi^2}{16 z_{IR}^4}\right] & z_{IR} < z_h\\ \\ \frac{L^3 \pi z_h}{2\kappa^2} \left[ \frac{2}{z_{IR}^4} - \frac{1}{z_h^4} - \left(\frac{N_f}{N_c}\right)\frac{\xi^2}{16 z_{IR}^4}\right] & z_{IR} > z_h. \end{array} \right.$$ If we define the critical temperature of the pure $AdS_5$ gravity (pure Yang-Mills) as $T_0 =\frac{1}{\pi z_h} =\frac{2^{1/4}}{\pi z_{IR}}$ following (\[tempads\]) the critical temperatures modified by mesons is given by $$T = \left\{\begin{array}{ll} {T}_0 \left(\frac{\xi^2}{32} \frac{N_f}{N_c}\right)^{\frac{1}{4}} & z_{IR} <z_h\\ \\ {T}_0 \left(1 -\frac{\xi^2}{32}\frac{N_f}{N_c}\right)^{\frac{1}{4}} & z_{IR}>z_h.\end{array}\right.$$ In the first case $z_{IR} <z_h$, a Hawking-Page type transition occurs, which seems unphysical, since it implies that there is a phase transition at low temperature. Unfortunately, we have no clear understanding on this phase transition. If we include, however, the back-reaction of the matter field on the background and work on a deformed AdS metric, then the unphysical phase transition may disappear. In the second case $z_{IR}>z_h$, at high (low) temperature the AdS black hole (the thermal AdS) is again stable, but due to the matter field, the critical temperature is smaller than that of the pure AdS gravity theory, which is consistent with an observation made in [@KLP]. A Hawking-Page type transitions at finite density ================================================= In this section, we consider the quark number density dependence on the deconfinement temperature. In QCD, quark chemical potential introduced as $\mu_q\bar\psi\gamma_0\psi$, and so according to an AdS/CFT dictionary we need to introduce a bulk U(1) field in AdS$_5$ whose boundary value is $\mu_q$. To this end, we follow [@DH] and generalize the symmetry to U$(N_f)\times $U$(N_f)$. The equation of motion for the time component of the U(1) vector field is given by $$\partial_z\left[\frac{1}{z}\partial_z V_{\tau}(z)\right] = 0\, ,$$ and a solution of the equation of motion is given by $$V_{\tau} = c_1 + c_2 z^2\, .$$ Since the factors $g^{\tau\tau}g^{zz}$ in the equation of motion for tAdS and AdSBH backgrounds are same, we conclude that the equations of motion of both backgrounds are same and the corresponding forms of the solutions are also same. That is, we can use the same form of the solution $V_{\tau}$ in both background tAdS and AdSBH. According to the AdS/CFT correspondence, the coefficient of the non-normalizable term, $c_1$, is proportional to coupling with the dual operator of the boundary theory. Since the time component of the U(1) vector field is dual to the quark number current, $c_1$ must correspond to the quark chemical potential. Meanwhile, the coefficient of the normalizable term, $c_2$, corresponds to the expectation value of the dual operator so that $c_2$ is interpreted as the quark number density, $c_2=12\pi^2\rho_q/N_c$ [@DH]. Following the same procedure used in the previous section, we arrive at $$V_{v1} = \pi z_h M_5 N_f L^5 c_2^2 z_{IR}^2$$ for the tAdS and $$V_{v2}=\left \{\begin{array}{ll} \pi z_h M_5 N_f L^5 c_2^2 z_{h}^2 & z_h<z_{IR}\\ \\\pi z_h M_5 N_f L^5 c_2^2 z_{IR}^2 & z_h>z_{IR} \end{array} \right.$$ for AdSBH. From these results, the differences of the action reads $$\Delta V_{v}=\left\{\begin{array}{ll} -\pi z_h M_5 N_f L^5 c_2^2(z^2_{IR}- z_{h}^2) & z_h<z_{IR}\\ \\0 & z_h>z_{IR} \end{array} \right.$$ The final result for $z_h<z_{IR}$ is $$\label{hp2} \Delta V = \frac{L^3 \pi z_h}{\kappa^2}\left[ \frac{1}{z_{IR}^4} -\frac{1}{2z_h^4} -\frac{L^4 N_f c^2_2}{48N_c}\left(z^2_{IR}- z_{h}^2\right)\right]$$ When $\Delta V < 0$, the dominant geometry is a AdS black hole and the boundary theory corresponding to this geometry is in a confined phase. As in Ref. [@Herzog:2006ra], in the pure gravity theory, the critical temperature is given by $T_0 = \frac{2^{1/4}}{\pi z_{IR}}$. So when considering the quark density, using the relation $T = 1/(\pi z_h)$, the critical temperature $T_c$ is determined by solving the following equation $$\label{eTc} 0~=~ T^4 - \frac{2}{\pi^4} \left( \frac{1}{z_{IR}^4} -\frac{L^4 N_f }{48N_c} c^2_2 \left( z^2_{IR} - \frac{1}{\pi^2 T^2} \right) \right)\Bigg{|}_{T =T_c} .$$ In (\[eTc\]), the last term $z^2_{IR}- z_{h}^2$, which is from quark number density, is positive, and so the critical temperature at finite density is always lower than that of the pure gravity theory. Now, we consider a case where $z_h\ll z_{IR}$ to see the quark number dependence of $T_c$ clearly. We note here that typically $z_h^2/z_{IR}^2\approx 0.6$. In this case, the critical temperature is given by $$\label{Tc} T_c (\rho_q) = \frac{2^{1/4}}{\pi} \left( \frac{1}{z_{IR}^4} -\frac{L^4 N_f z^2_{IR}}{48N_c} c^2_2 \right)^{1/4},$$ which is a consistent result with the lattice QCD data. Note here again that $c_2\sim \rho_q$ [@DH]. In Fig. \[TcF\], we plot Eq. (\[eTc\]), together with the lattice result [@FK; @FKtalk], where $R\equiv T_c(\rho_q)/T_0$ and $\bar\rho_q =\rho_q z_{IR}^3$. Our result in Fig. \[TcF\] is consistent with lattice QCD results at low density, [*i.e*]{}, see [@FK]. Finally, we combine the results obtained in the previous and in this sections. The final form of $\Delta V$ is given by $$\Delta V = \left\{ \begin{array}{ll} \frac{L^3 \pi z_h}{2\kappa^2}\left[ \frac{1}{z_h^4} - \left(\frac{N_f}{N_c}\right) \frac{\xi^2}{16 z_{IR}^4}\right] & z_{IR} < z_h\\ \\ \frac{L^3 \pi z_h}{2\kappa^2} \left[ \frac{2}{z_{IR}^4} - \frac{1}{z_h^4}- \left(\frac{N_f}{N_c}\right) \frac{\xi^2}{16 z_{IR}^4} -\frac{L^4 N_f c^2_2}{48N_c}\left(z^2_{IR}- z_{h}^2\right) \right] & z_{IR} > z_h. \end{array} \right.$$ Discussion ========== We have studied the effects of matters on the deconfinement transition in the context of a Hawking-Page type analysis. We observed, as it should be, that the corrections from mesons to the deconfinement temperature are suppressed by $1/N_c$. We found that $T_c$ decreases with the number of quark flavor $N_f$. At finite density, we obtained the density dependence of $T_c$, and it shows a similar behavior calculated in lattice QCD. We remark here that, as it is well known, in the presence of the dynamical quarks the Polyakov loop is no longer a good order parameter to describe the deconfinement transition [@BU]. In lattice QCD, however, it has been established that the Polyakov loop could be served as an approximate order parameter for the deconfinement transition. While, one of the key quantities that connects the Hawking-Page transition to the deconfinement is the Polyakov loop. In the present work, we assume that the the gravity description of confinement/deconfinement through the Hawking-Page transition is still viable with dynamical quarks. It may be interesting to see how the expectation value of the Polyakov loop behaves in the gravity side, when the dynamical quarks in the fundamental representation are present. The work of Bum-Hoon Lee, Siyoung Nam and Chanyong Park was supported by the Science Research Center Program of the Korea Science and Engineering Foundation through the Center for Quantum Spacetime(CQUeST) of Sogang University with grant number R11 - 2005 - 021. [99]{} J. M. Maldacena, “The large N limit of superconformal field theories and supergravity,” Adv. Theor. Math. Phys.  [**2**]{}, 231 (1998) \[Int. J. Theor. Phys.  [**38**]{}, 1113 (1999)\] \[arXiv:hep-th/9711200\];\ S. S. Gubser, I. R. Klebanov and A. M. Polyakov, “Gauge theory correlators from non-critical string theory,” Phys. Lett.  B [**428**]{}, 105 (1998) \[arXiv:hep-th/9802109\]. J. Polchinski and M. J. Strassler, “Hard scattering and gauge/string duality,” Phys. Rev. Lett.  [**88**]{}, 031601 (2002) \[arXiv:hep-th/0109174\]. A. Karch and E. Katz, “Adding flavor to AdS/CFT,” JHEP [**0206**]{}, 043 (2002) \[arXiv:hep-th/0205236\]; J. Erlich, E. Katz, D. T. Son and M. A. Stephanov, “QCD and a holographic model of hadrons,” Phys. Rev. Lett.  [**95**]{}, 261602 (2005) \[arXiv:hep-ph/0501128\]. L. Da Rold and A. Pomarol, “Chiral symmetry breaking from five dimensional spaces,” Nucl. Phys.  B [**721**]{}, 79 (2005) \[arXiv:hep-ph/0501218\]. S. J. Brodsky and G. F. de Teramond, “Hadronic spectra and light-front wavefunctions in holographic QCD,” Phys. Rev. Lett.  [**96**]{}, 201601 (2006) \[arXiv:hep-ph/0602252\]. E. Witten, “Anti-de Sitter space and holography,” Adv. Theor. Math. Phys.  [**2**]{}, 253 (1998) \[arXiv:hep-th/9802150\]. C. P. Herzog, “A holographic prediction of the deconfinement temperature,” Phys. Rev. Lett.  [**98**]{}, 091601 (2007) \[arXiv:hep-th/0608151\]; K.-Y. Kim, S.-J. Sin and I. Zahed, “Dense hadronic matter in holographic QCD,” hep-th/0608046; N. Horigome and Y. Tanii, JHEP [**0701**]{}, 072 (2007)\[hep-th/0608198\]; S. Nakamura, Y. Seo, S.-J. Sin and K.P. Yogendran, “A New Phase at Finite Quark Density from AdS/CFT,” hep-th/0611021; S. Kobayashi, D. Mateos, S. Matsuura, R. C. Myers and R. M. Thomson, JHEP [**0702**]{}, 016 (2007) \[hep-th/0611099\]. F. Karsch, Nucl. Phys. Proc. Suppl. [**83**]{}, 14 (2000) \[arXiv:hep-lat/9909006\]. F. Karsch, E. Laermann and A. Peikert, Nucl. Phys. [**B605**]{}, 579 (2001) \[arXiv: hep-lat/0012023\]. Y. Aoki, et al, Nature [**443**]{}, 675 (2006). Z. Fodor and S.D. Katz, JHEP [**0203**]{}, 014 (2002) \[arXiv:hep-lat/0106002\]. C.R. Allton, et al, Phys. Rev. [**D68**]{}, 014507 (2003)\[arXiv:hep-lat/0305007\]. C. Bernard, et al, Phys. Rev. [**D 71**]{}, 034504 (2005) \[hep-lat/0405029\]; K. Ghoroku and M. Yahiro, Phys. Rev. [**D73**]{}, 125010 (2006) \[hep-ph/0512289\]. Y. Kim, S. J. Sin, K. H. Jo and H. K. Lee, “Vector Susceptibility and Chiral Phase Transition in AdS/QCD Models,” arXiv:hep-ph/0609008. S. K. Domokos and J. A. Harvey, “Baryon number-induced Chern-Simons couplings of vector and axial-vector mesons in holographic QCD,” arXiv:0704.1604 \[hep-ph\]. C. Csaki and M. Reece, “Toward a systematic holographic QCD: A braneless approach,” arXiv:hep-ph/0608266. T. Banks and A. Ukawa, Nucl. Phys. [**B225**]{} 145 (1983). F. Csikor, G.I. Egri, Z. Fodor and S.D. Katz, “Lattice QCD at non-vanishing density: phase diagram, equation of state,” Contributed to Workshop on Strong and Electroweak Matter (SEWM 2002), Heidelberg, Germany, 2-5 Oct 2002, e-Print: hep-lat/0301027.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Allowing the energy of a gravitational field to serve partially as its own source allows gravitating bodies to exhibit stronger fields, as if they were more massive. Depending on degree of compaction of the body, the field could be one to five times larger than the newtonian field. This is a comfortable range of increase in field strength and may prove to be of convenience in the study of velocity curves of spirals, of velocity dispersions in clusters of galaxies and in interpreting the Tully-Fisher or Faber-Jackson relations in galaxies or systems of galaxies. The revised gravitation admits of superposition principle but only approximately in systems whose components are widely separated. The revised dynamics admits of the equivalence principle in that, the effective force acting on a test particle is derived from a potential, and could be eliminated in a freely falling frame of reference.' author: - 'Y. Sobouti' date: title: 'Revised dynamics or dark matter in galactic and extra galactic scales? ' --- > [The only justification for our concepts and system of concepts is that they serve to represent the complex of our experiences. Beyond this they have no legitimacy. I am convinced that the philosophers have had a harmful effect upon the progress of scientific thinking in removing certain fundamental concepts from the domain of empiricism, where they are under our control, to the intangible heights of the a priori. Albert Einstein, 1922.]{} Introduction ============ [*The problem of dark matter surrounding spiral galaxies*]{}, wrote Van Albada et al. (1985), [*is one of the most enigmatic questions in present day astrophysics. A number of years of intensive research have brought little or no clarification.*]{} Earlier suggestions to resolve mass discrepancy were on the conservative side. Bahcall and Cassertano (1985) spoke of [*missing mass*]{} as synonymous to [*missing light*]{}. Tremaine and Lee (1987)with [*dark matter*]{} meant the matter whose existence was inferred only through its gravitational effects. Critics were vocal. Outstanding among them Milgrom (1983 a,b,c), the architect of [*MOND*]{}, spoke of the [*dark side of the dark matter hypothesis*]{} and [*of its arbitrariness in that one invokes the dark matter in the correct amount and spatial distribution needed to explain the mass discrepancy in each and every case for itself*]{} (1987). In the course of the past two decades acurate observations of velocity curves of spiral galaxies have become available.See, e. g., Begeman et al. (1991), Sanders (1996), Sanders and Verheijen (1998), Mc Gauph and de Blok (1998). Credible velocity dispersions of galactic systems have emerged. See, for example, Tullly et al. (1996) for data on Ursa Majoris cluster. On the theoretical side many ingenious and bold candidates for dark matter have been offered. A list compiled by Ostriker and Steinhardt (2003) includes items such as cold collisionless dark matter, strongly self interacting dark matter, warm dark matter, repulsive dark matter, fuzzy dark matter, self annihilating dark matter, decaying dark matter, etc. In spite of all these developments, however, there is no clue as how to detect the dark matter and get to the physics that it obeys. The issue does not seem less enigmatic than what von Albada and his colleagues described in 1985. Arguments pro and con are the same as a quarter of century ago. Proponents of dark matter assume the validity of Newtonian dynamics at all distance scales and look for one or another form of hypothetical matter to provide the missing gravity. Skeptics, on the other hand, see no logic in resorting to a concept that solves the riddle of dynamics but remains unamenable to further validation by any other known physical means. A historical reminder might be timely. In the closing decade of the 19th. century, physicists had agreed that light was an electromagnetic wave. And out of the experience with waves in other contexts had required a medium of propagation for it, the ether. Yet ether evaded all attempt of detection no matter how ingeniously the detection devices were designed. On the other hand the notion of medium of propagation seemed so obvious to everyone, that some of the brightest minds of the time invoked theories that would have kept the ether, but in a concealed form. The famous transformations of Lorentz were developed primarily to explain the null results of the experiments of Fiseau, Michelson, Michelson & Morely, etc., rather than as a foundation for the special relativity that emerged later. It was up to Einstein to think and later to speak out that [*our concepts have no legitimacy beyond what the complex of our experiences bestows upon them*]{}. If experiments do not reveal the ether it could be dispensed with. If experiments show that the speed of light is the same in all reference frames, that could be a fact to build the new physics upon it. In these early years of the 21st. Century, it would not be unwise to take Einstein’s advice seriously. Accept the validity of newtonian dynamics in solar and similar systems, where it has been tested, but look for alternatives in larger galactic and extragalactic scales, where it has not been verified. Many researchers have actually adopted such a point of view before. See Sanders and Mc Gauph (2002) for a review of Milgrom’s [*MOND*]{} and earlier attempts. Here, we adopt a variational approach to the problem. We amend the classical action by adding a term to it that eventually makes provision for the field energy to serve partially as its own source. The procedure is apt to a general relativistic generalization, in preparation for a future presentation. Variational formulation ======================= The system to be considered is a collection of point masses $m_i$ at positions ${\bf x}_i$, $i=1,\cdots,N$; plus the field $\phi({\bf x})$ at point ${\bf x}$. Both newtonian equation of motion and equation of gravitational field are derivable from the following action integral. $$\begin{aligned} \al\al I=\int L[\textbf{x}_i,\phi(\textbf{x})]~ dt,\\ \al\al L[\textbf{x}_i,\phi(\textbf{x})]=\sum m_i\left[\frac{1}{2}\dot{\bf x}_i^2+\phi\left({\bf x}_i\right)\right]-\frac{1}{8\pi G} \int\left|\nabla\phi({\textbf{x}}\right|^2d^3x.\nonumber\end{aligned}$$ At point ${\bf x}$ the gravitational energy density is $(8\pi G)^{-1}|\nabla\phi(\x)|^2$. By Einstein, this energy has an equivalent mass. We propose to entertain the possibility of using this mass or a fraction $\alpha$ of it as the source of extra gravitation on the particles. The effect on $m_i$ at position ${\bf x}_i$ will be $\alpha Gm_i(8\pi G c^2)^{-1}|\nabla\phi(\x)|^2|\x-\x_i|^{-1}$. The contribution to the action integral from all space and all particles will be $$\begin{aligned} \al\al I_{int}=\int L_{int}[\textbf{x}_i,\phi(\textbf{x})]~ dt,\nonumber\\ \al\al L_{int}[\x_i, \phi(\textbf{x})]=\frac{1}{8\pi}\frac{\alpha}{c^2}\sum_i m_i \int\frac{|\nabla\phi({\bf x}|^2}{|{\bf \textbf{x}}-{\bf \textbf{x}}_i|}d^3 x.\end{aligned}$$ The dimensionless constant $\alpha$ is of the order of unity. It is introduced for possible later adjustments. Note that $Gm_i/c^2$ is of the order of the Schwarschild radius of $m_i$. Before proceeding further, let us emphasize that the argument given in composing $I_{int}$ is for mnemonic purposes only. The formal statement of the assumption is the following: The classical action integral from which Newton’s laws are inferred is inadequate in galactic and extragalactic scales. We propose to amend it by adding $I_{int}$ to the classical action. The addition is a scalar, quadratic in field gradients, and contains the particle coordinates and masses. It will alter both the field equation and the equations of motion. The field equation for $\phi(\x)$ is obtained by requiring the functional derivative of the total action, $I+I_{int}$, with respect to $\phi(\x)$to vanish. Thus, $$\begin{aligned} \frac{\delta (I+I_{int})}{\delta\phi(\textbf{x})}&=& {\bf\nabla}. \left[{\bf \nabla}\phi(\textbf{x})\left\{1-\frac{\alpha G}{c^2}\sum \frac{m_i}{|\textbf{x}-\textbf{x}_i|}\right\}\right]\nonumber\\&+& 4\pi G\sum m_i\delta(\textbf{x}-\textbf{x}_i)=0.\end{aligned}$$ By writing $\delta(\textbf{x}-\textbf{x}_i)=\frac{1}{4\pi}\nabla^2{|\textbf{x}-\textbf{x}_i|^{-1}}$ one immediately integrates Eq. (3) into $$\begin{aligned} \nabla\phi\left\{1-\frac{\alpha G}{c^2}\sum_i\f{m_i}{|\textbf{x}-\textbf{x}_i|}\right\}=-G\nabla\sum \frac{m_i}{|\textbf{x}-\textbf{x}_i|}.\end{aligned}$$ Denoting the newtonian potential by $$\begin{aligned} \al\al \phi_N(\textbf{x})=-G\sum\f{m_i}{|\textbf{x}-\textbf{x}_i|},\end{aligned}$$ Eq.(4) integrates into $$\begin{aligned} \phi(\x)&=&\f{c^2}{\alpha}\ln\left[1+\alpha\f{\phi_N (\bf x)}{c^2}\right]\nonumber\\&=& \phi_N\left[1-\f{1}{2}\alpha\f{\phi_N}{c^2}+\f{1}{3}\alpha^2\f{\phi^2_N}{C^4}+\cdots\right].\end{aligned}$$ Equation of motion for $m_i$ is obtained likewise, $$\begin{aligned} \al\al\frac{\delta (I+I_{int})}{\delta \textbf{x}_i}=\nonumber\\ \al\al\hspace{0.5cm} m_i\left[\ddot{ \textbf{x}}_i-\nabla_i\phi(\textbf{x}_i)-\f{1}{8\pi}\f{\alpha }{c^2}\nabla_i\int\f{|\nabla\phi(\textbf{x})|^2}{|\textbf{x}-\textbf{x}_i|}d^3 x\right]=0.\end{aligned}$$ Newtonian limits in all cases are obtained by letting $\alpha=0$. To elucidate the significance of the proposed amendment to the classical action and the way it relates to mass discrepancy, several simple examples are worked out in section 3. Applications ============ Gravitational field of spheres ------------------------------- Consider a sphere of uniform density $\rho$, radius $R$, and total mass $M$: From Eqs. (5) and (6) the outside solution is $$\begin{aligned} \al\al \phi(r) = \f{GM}{s}\ln\f{r-s}{r}=-\f{GM}{r}\left(1+\f{1}{2}\f{s}{r}+\cdots\right), \hspace{0.3cm} r\geq R \cr \al\al \f{d\phi}{dr} = \f{GM}{r(r-s)}, \hspace{.3cm}r\geq R,\end{aligned}$$ where $s=\alpha GM/c^2$ is of the order of Schwarschild radius of $M$. For the inside, noting that $\phi_N(r)=-\f{1}{2}\f{G M}{R^3}\left(3R^2-r^2\right)$, one obtains $$\begin{aligned} \al\al \phi(r)=\f{c^2}{\alpha}\ln\left[1-\f{3}{2}\f{s}{R}+\f{1}{2}\f{s}{R}\f{r^2}{R^2}\right], \hspace{0.3cm}r\leq R,\cr \al \al \f{d\phi}{dr}=\f{GM}{R^2}\f{r}{R}\left[1-\f{3}{2}\f{s}{R}+\f{1}{2}\f{s}{R}\f{r^2}{R^2}\right]^{-1}, \hspace{0.3cm} r\leq R.\end{aligned}$$ Motion of test satellites ------------------------- Equation of motion is obtained by substituting Eqs. (8) & (9) in Eq. (7) and reducing it. The result is $$\begin{aligned} \al\al m\ddot\mathbf{r}=-\f{G m M}{r^2}\left[1+u\left(\f{s}{R}\right)+\f{s}{r-s}\right]\f{{\bf r}}{r},\end{aligned}$$ where $u(s/R)$ is derived, numerically calculated, and plotted in the Appendix. As the central body contracts and $R$ reduces from an initially large value to its allowed minimum $\f{3}{2}s; u(s/R)$ increases from zero to 4. The first term on the right hand side of Eq. (10) is recognized as the classical expression. The second term, $u(s/R)$, is the pivotal one and as mentioned it could be four times as large as the classical term. In newtonian parlance this amounts to saying that a compact object may, gravitationally, present itself up to five times more massive than the same object in a diffuse state. This is what we propose as a partial answer to mass discrepancy!\ The third term in Eq. (10) arises from deviations of the force law from $r^{-2}$. It is infinitesimal compared with the previous two terms, though it may observably perturb elliptical orbits in secular time scales. N-body problem -------------- Consider a collection of spheres each of mass $M_i$, radius $R_i$, and positions $\x_i$. Assume the system members are well separated such that $|\x_i-\x_j|=r_{ij}\gg R_i$ for all $i \neq j$. It is demonstrated in the Appendix, that on account of this assumption, the integral appearing in Eq. (7) splits into a sum given by Eq. (A. 7). Thus, one finds $$\begin{aligned} \al\al \left. \ddot \x_i-G\sum M_j^{\rm eff}\f{(\x_i-\x_j)}{|\x_i-\x_j|^3}\right]=0.\end{aligned}$$ where $M_j^{\rm eff}=M_j(1+u_j), u_j=u(s_j/R_j)$. Actually a term of the order $G\sum_jM_js_j|r_{ij}-s_j|^{-1}$, is neglected in Eq (11), a) because of the extra small factor $s_j$ and b) because of its steeper, $r^{-3}$, decrease with distance compared with $r^{-2}$ dependence of the retained terms. Albeit the approximations, Eq. (11) is identical with the newtonian many body equation of motion, except that instead of the classical masses, the effective masses, $M_j^{\rm eff}=M_j(1+u_j)$, play the role. Depending on the degree of compactness of the constituent bodies, the effective masses could, in the case of uniform densities, be 1 to 5 times larger than the classical ones. Having noted this, it is easy to write the first integrals of Eq. (11): $$\begin{aligned} \al\al {\rm Momentum}: {\bf P}=\sum M_i^{\rm eff}\dot\x_i= const.\\ \al\al {\rm Angular~ momentum}: {\bf L}=\sum M_i^{\rm eff}\x_i\times\dot\x_i= const.\\ \al \al {\rm Energy}: E=T+V= {\rm const.,}~~~\\ \al\al\hspace{1.37cm} T=\f{1}{2}\sum_iM_i^{\rm eff}\dot\x_i^2,~~~ V=-\f{1}{2}G\sum_{i,j}M_i^{\rm eff} M_j^{\rm eff}/r_{ij}\nonumber\\ \al\al{\rm Virial~~ theorem}: 2\bar T+\bar V=0,\end{aligned}$$ where ’bars’ on $T$ and $V$ indicate averages over system members or over time periods.\ Increased effective mass in Eq. (11) and the virial theorem may prove useful in explaining the missing mass or missing gravity issues, in the velocity curves of galaxies and/or velocity dispersions in clusters of galaxies. Concluding remarks ================== > [*... it is contrary to the mode of thinking in science to conceive of some thing that can act itself, but which cannot be acted upon. Albert Einstein, 1922.*]{} The quotation is from Einstein’s argument to set the stage for the rejection of absolute space-time continuum of special relativity in the presence of matter, and his admiration of Mach’s reservation to ascribe an absolute meaning to [*inertia*]{} irrespective of matter elsewhere. In an attempt to amend newtonian dynamics, we have followed Einstein’s advice and tried to alter the laws of gravity and of motion simultaneously. The interaction lagrangian of Eq. (2) contains both the particle coordinates, $\x_i$, and the field gradient, $\nabla\phi$. Noteworthy in $L_{int}$ is the field energy that eventually appears as the increased effective mass, $M^{{\rm eff}}=M(1+u)$ in Eqs. (10-15). We recall that the main issues in the analysis of the velocity curves of spirals are: a) insufficiency of the observed and/or estimated stellar masses to provide the required dynamical effect, and b) the flatness of the velocity curves beyond what the distribution of observed masses would permit. Similarly, in clusters of galaxies, again the observed masses are much smaller than the virial ones, inferred from the velocity dispersion in the cluster. The increased effective mass discussed above could provide a partial answer in both cases of luminous mass deficiency. The flatness of the rotation curves is a matter of the distribution of effective masses. For example, if there are enough low luminosity collapsed faint objects in the outskirts of spiral arms the velocity curves could be flattened enough. The field in Eq. (6), unlike the Newtonian one, is not proportional to the mass of the constituent members and does not admit of superposition principle. Even in the case of approximate superposition in systems with widely separated components, the field of each component is a highly nonlinear function of its mass. This nonlinearity could prove useful in discussions of Tully-Fisher relation in spirals, an empirical relation between the brightness of the galaxy and the asymptotic orbital speed at the flat end of the velocity curve. The Tully-Fisher relation, with the assumption of a further empirical mass to light ratio for the galaxy, translates into a power law between the asymptotic speed and the total mass, $v_{asymp} \propto M^\beta$, $2.5 <\beta<3.5$. In the present dynamics, $\beta$ will sensitivity depend on the compaction factor, the fraction, and the distribution of the compact members of the galaxy. Same considerations and reservations holds for the Faber-Jackson relation, another empirical relation between the velocity dispersion in cluster of galaxies and the luminosity of the cluster. In section 3.3, conservation laws were derived for the case of approximate superposition. One can, however, do better. The total lagrangian, $L+L_{{\rm int}}$, is invariant under time translations, coordinate translations, and coordinate rotations. Therefore, seven first integrals of motion, corresponding to energy, momenta, and angular momenta, should exist. This point of view will be presented elsewhere. Birkhof’s theorem states that in newtonian and relativistic regimes the gravitational field of a spherically symmetric system is independent from the internal structure of the system. This is not the case in the present theory as it is highlighted by the presence of $u(s/R)$ in Eq. (10). This may open the possibility of studying the history of contraction or collapse of an object by logging the external gravitational field of the object as it evolves. From Eq.(7), the effective force on a test particle is derived from a potential. It could be eliminated in a freely falling frame of reference, meaning that the dynamics developed here admits of equivalence principle. This is hardly surprising. For, in the action integrals of Eqs. (1) and (2) one single mass $m_i$ is used to compose the kinetic and the gravitational potential energies of the system. Finally a criticism: a) The mass equivalent of energy is a relativistic concept. Yet, a gravitational effect was attributed to it through the law of gravity of Newton. b) The effective mass attains its full significance when the source object shrinks into sizes of the order of Schwarzschild radius. Logically, a covariant general relativistic approach should be adopted. Presently, we are looking for a field equation and a geodesic one in which some sort of mutual interactions between distant points, similar to one in $L_{{\rm int}}$ are taken into account. As in Eq. (2), there will be a free parameter $\alpha$ in the formulation. We will require the emerging dynamics to reduce a) to the general relativistic one in the limit $\alpha \rightarrow 0$; b) to the present dynamics in the limit of weak fields but $\alpha \neq0$, and c) to the newtonian one in the limit of zero $\alpha$ and weak field. This will be presented elsewhere. Reduction of $L_{int}$, Equation(2) ==================================== We have already seen the formal role of the interaction lagrangian, $L_{int}$ in altering the equations of motion and of the field. Its numerical values are needed in the study of orbits, Eq.(7); virial theorem, Eq. (15), etc. Here, we calculate it a) for a single uniform spherical mass, and b) for a collection of unform spheres.\ a)For a spherically symmetric field the non newtonian integral appearing in Eq. (7) is $$\begin{aligned} l_{int}(r)=\f{1}{8\pi}\f{\alpha}{c^2}\int\f{|d\phi(r')/dr'|^2}{|{\bf r}-{\bf r}'|}d^3r'.\end{aligned}$$ In an expansion of $ [\bf r- \bf r' ]^{-1}$ in Legendre polynomials, only the $ l=0$ term will have non vanishing contribution; for $d\phi/dr'$ has no directional dependence. We carry out angular integrations and split the radial range of the integration into three intervals $0<r'<R$, $R<r'<r$ and $r<r'<\infty$. Thus, $$\begin{aligned} l_{int}&=&\f{\alpha}{2c^2}\left[\f{1}{r}\int_0^R\left|\f{d\phi(r')}{dr'}\right|^2r'^2dr'\right. +\f{1}{r}\int_R^r\left|\f{d\phi(r')}{dr'}\right|^2r'^2dr' \nonumber\\&+& \left.\int _r^\infty\left|\f{d\phi(r')}{dr'}\right|^2r'dr'\right].\end{aligned}$$ The expression for the field gradient in the first integral is the interior solution of Eq. (9) and in the other two are the outer solution of Eq. (8). The remaining mathematical manipulations are elementary, though lengthy. One eventually arrives at $$\begin{aligned} l_{int}=\f{GM}{r}u\left(\f{s}{R}\right), \hspace{.4cm}s=\f{\alpha GM}{c^2},\end{aligned}$$ where $$\begin{aligned} u(y)=-\f{1}{2}+\f{3}{y}\left[1-\sqrt{\frac{(2-3y)}{y}} \arctan \sqrt{\frac{y}{(2-3y)}}~\right]\end{aligned}$$ The parameter $s/R$ indicates the degree of compaction of the central body. As the body contracts from an initially dispersed state, $(s/R)\approx 0$, to its maximum allowable compact state, $(s/R=2/3)$, $u(s/R)$ increases from zero to 4. A plot of $u(s/R)$ is given in figure below. In newtonian language, the effect on the motion of an orbiting satellite is to make the central body look $(1+u)$ times more massive than what it does in newtonian regime.\ b)Multi-component systems: In Eq. (A.1) The main contribution to the integral comes from the immediate vicinity of the central body. As one approaches the gravitating body from outside, $\left|d\phi(r')/dr'\right|^2$ grows roughly as $r'^{-4}$ up to the surface of the body and fades out to zero at its center. Assume a system consisting of two well separated spheres $(M_1, R_1)$ at $\x_1$ and $(M_2, R_2)$ at $\x_2$ with $|\x_1-\x_2|\gg R_1 \& R_2$. From Eqs. (5) and (6), we read $$\begin{aligned} |\nabla\phi(\x')|^2&=&\left[1-\f{\alpha G}{c^2}\left(\f{M_1}{|\x'-\x_1|}+\f{M_2}{|\x'-\x_2|}\right)\right]^{-2}\nonumber\\ &\times& G^2\left[\f{M_1^2}{|\x'-\x_1|^4}+\f{M_2^2}{|\x'-\x_2|^4}\right.\nonumber\\ &&\hspace{0.7cm}+\left. 2M_1M_2\f{(\x'-\x_1). (\x'-\x_2)}{|\x'-\x_1|^3|\x'-\x_2|^3}\right].\end{aligned}$$ Again, the main contributions come from the vicinities of the two spheres. In the immediate vicinity of sphere 1, however, $G^2M_1^2|\x'-\x_1|^{-4}$ is large. While, with the assumption of large separation of the two spheres, $G^2M_2^2|\x'-\x_2|^{-4}$ is insignificant. Vice versa for the vicinity of sphere 2. The expression containing $(\x'-\x_1).(\x'-\x_2)$ practically adds up to zero upon integration over $\x'$, because of its directional dependencies. We conclude that\ \ \ $$\begin{aligned} l_{int}(x)\approx \f{GM_1u_1}{|\x-\x_1|}+\f{GM_2u_2}{|\x-\x_2|},~~ u_i=u\left(\f{s_i}{R_i}\right).\end{aligned}$$ Similarly, for a system of many, but well separated, components one obtains $$\begin{aligned} l_{int}(x)\approx G\sum_i\f{M_i u_i}{|\x-\x_i|}, \hskip 2em u_i(s_i/R_i).\end{aligned}$$ Albada, von, T. S., Bahcall, J. N., Begman, K., and Sancisi, R.: 1985, Astrophys. J., [**295**]{}, 305 Bahcall, J. N., Cassertano, S.: 1985, Astrophys. J. Lett., 293, L7 Begeman, K. G., Broeils, A. H., and Sanders, R. H.: 1991, MNRAS, [**249**]{}, 523 Einstein, A., 1922, The Meaning of relativity, Princeton Univ. Press,3rd. paperback,1972 Mc Gaugh, S. S., and de Blok, W. J. G.: 1998 a,b, Astrophys. J., [**499**]{}, 41 (a), 66 (b) Milgrom, M.: 1983 a,b,c, Astrophys. J., [**270**]{}, 365 (a), 375 (b), and 384 (c) Milgrom, M.: 1987, in [*ark matter in the universe*]{}, p. 231, eds.: Bahcall,J. Piran,T. and Weinberg, S., Wold Scientific Publ. Co., Singapore Ostriker, J. M. and Steinhardt, P., 2003, arXiv:astro-ph/0306402 Sanders, R. H.: 1996, Astrophys. J., [ **473**]{}, 117 Sanders, R. H., and Verheijen, M. A. W.: 1998, Astrophys. J., [**503**]{}, 97 Sanders, R. H., and Mc Gaugh, S. S.: 2002, arXiv:astro-ph/0204521 v1 Tremaine, S., Lee, H. M.: 1987, in [*Dark matter in the universe*]{}, p.,410, eds. Bahcall, J., Piran, T., and Weinberg, S., World Scientific Pub. Co., Singapore Tully, R. B., Verheijen, M. A. W., Pierce, M. J., Huang, J. S., and Wainscoat, R.: 1996, Astron. J., [ **112**]{}, 2471
{ "pile_set_name": "ArXiv" }
--- abstract: 'There is a remarkable well-known connection between the $\operatorname{G}(4,n)$ cluster algebra and $n$-particle amplitudes in $\mathcal{N}=4$ SYM theory. For $n \ge 8$ two long-standing open questions have been to find a mathematically natural way to identify a finite list of amplitude symbol letters from among the infinitely many cluster variables, and to find an explanation for certain algebraic functions, such as the square roots of four-mass-box type, that are expected to appear in symbols but are not cluster variables. In this letter we use the notion of “stringy canonical forms” to construct polytopal realizations of certain compactifications of (the positive part of) the configuration space $\operatorname{Conf}_n(\mathbb{P}^{k-1}) \cong$ $\operatorname{G}(k,n)/T$ that are manifestly finite for all $k$ and $n$. Some facets of these polytopes are naturally associated to cluster variables, while others are naturally associated to algebraic functions constructed from Lusztig’s canonical basis. For $(k,n) = (4,8)$ the latter include precisely the expected square roots, revealing them to be related to certain “overpositive" functions of the kinematical invariants.' author: - 'Nima Arkani-Hamed' - Thomas Lam - Marcus Spradlin title: 'Non-perturbative geometries for planar $\mathcal{N}=4$ SYM amplitudes' --- I. Introduction =============== Scattering amplitudes are boundary observables in flat space, depending only on the kinematical data specifying the helicities and momenta of the scattering particles. It is thus natural to ask whether there is some question that can be posed in kinematic space whose answer yields the amplitudes directly, without referring to auxiliary notions such as unitary evolution in the bulk of spacetime or a string worldsheet. The challenge appears daunting since there are no obvious physical notions of locality or time associated with, say, the space of $n$ null momenta relevant for the scattering of massless particles. We must instead cast out more adventurously, looking for new sorts of mathematical structures in this naively barren space, with the power to generate all of the richness and complexity needed for scattering amplitudes compatible with locality and unitarity. The past several years have seen significant inroads in this program, associated with deep new combinatorial and geometric structures in kinematic space connecting various aspects of amplitudes to mathematical notions of total positivity, cluster algebras and motives in startling new ways. A prototype of the kind of description we seek is provided for the all-loop integrand in planar $\mathcal{N}=4$ super-Yang-Mills (pSYM) theory by the amplituhedron [@Arkani-Hamed:2013jha], which can be understood purely in the kinematical momentum-twistor space: the (super)integrand is the unique canonical form [@Arkani-Hamed:2017tmz] with logarithmic singularities on (and only on) all boundaries of the amplituhedron. Thus the integrand is fully determined by some geometry in kinematic space (the amplituhedron) and a question asked of that geometry (the determination of its canonical form). The amplituhedron provides a geometric origin for all of the singularities of the integrand as a rational function. But the simplicity gained in dealing with rational functions comes at a significant cost: the amplituhedron is inexorably tied to perturbation theory. Indeed, there is a different geometry for every loop order. For the full amplitude we should instead expect some geometric origin for the much more intricate pattern of branch cuts, which are present non-perturbatively. Of course the question of determining the geometry of branch cuts from first principles, for instance from an analysis of Landau equations, was an infamously difficult one in the S-matrix program in the 1960s. But there is hope for planar theories of massless particles, like pSYM theory, where it is known that for any fixed particle number, the number of branch points associated with solutions to the Landau equations is finite [@Prlina:2018ukf]. We also expect that perturbation theory for the planar theory has a finite radius of convergence. Given this encouragement, there is a natural candidate for the non-perturbative geometry we seek. The kinematic data is provided by $n$ momentum-twistor four-vectors [@Hodges:2009hk] $Z_1^I, \ldots, Z_n^I$, and the action of the conformal group SL(4) tells us we can associate this with a point in the Grassmannian $\operatorname{G}(4,n)$, a $4(n{-}4)$-dimensional space. Restricting for simplicity to the case of MHV amplitudes, the amplituhedron asks for the external data to lie in the [*positive*]{} Grassmannian $\operatorname{G}_+(4,n)$ [@ArkaniHamed:2012nw], so the positivity of external data should clearly be an important ingredient. Even more concretely, there is apparently a fascinating connection between the positivity of kinematic data and the Landau equations—in all examples studied to date, the Landau equations admit [*no*]{} solutions when the external data is taken to be in $\operatorname{G}_+(4,n)$! This is a highly nontrivial fact, implying that amplitudes have no branch points inside the positive domain, again suggesting that this region should play a starring role in defining the non-perturbative geometry relevant to amplitudes. MHV amplitudes enjoy an additional little group symmetry under which $Z_i \to t_i Z_i$, and as is familiar, they are therefore functions of cross-ratios, depending only on $3(n{-}5)$ rather than $4(n{-}4)$ variables. Thus an obvious guess for the non-perturbative geometry is “$\operatorname{G}_+(4,n)$ modulo the the little group torus", denoted $\operatorname{G}_+(4,n)/T$. Indeed, quite apart from this more recent motivation involving positivity, it has long been appreciated [@Golden:2013xva] that there is a deep connection between MHV amplitudes and the configuration space $\operatorname{Conf}_n(\mathbb{P}^3) \cong \operatorname{G}(4,n)/T$. All evidence available to date from explicit multi-loop computations [@Goncharov:2010jf; @Caron-Huot:2019bsq; @Dixon:2016nkn] supports the hypothesis that the symbol alphabets of $n=6,7$ particle amplitudes are the cluster variables of the $\operatorname{G}(4,n)$ Grassmannian cluster algebra. This has long been a source of inspiration for determining the non-perturbative geometry of pSYM theory, but for $n \ge 8$ there are a couple of open questions indicating that the cluster algebra itself is not the end of the story. First, as we have stressed, we expect that the number of branch points for all pSYM amplitudes, and hence the number of symbol letters, should be finite for any $n$. However, the $\operatorname{G}(k,n)$ cluster algebra has infinitely many cluster variables if $(k{-}2)(n{-}k{-}2) > 3$, including the cases $k = 4$, $n \ge 8$ of interest to amplitudes. Is there a mathematically natural way to extract some “finite subset” of variables relevant to scattering amplitudes? Second, cluster variables for $\operatorname{G}(k,n)$ are always polynomials in minors of the $Z$ matrix. But beginning with $n=8$, even for MHV amplitudes that should be polylogarithmic, we expect (specifically, starting at three loops [@Prlina:2017azl; @Prlina:2017tvx]) symbol letters that are algebraic functions of these minors. The most familiar and famous of these is the square root associated with the four-mass box integral [@Hodges1977; @tHooft:1978jhc], but a variety of algebraic letters appear in various contexts (see for example [@Chicherin:2017dob; @Gehrmann:2018yef; @Chicherin:2018yne; @Heller:2019gkq; @Zhang:2019vnm], and [@Besier:2019kco; @Bourjaily:2019igt] for some tools for dealing with them). How can we see them arise in a mathematically natural way? One might think that defining $\operatorname{G}_+(k,n)/T$ is completely straightforward, and indeed there is no subtlety associated with thinking of what the [*interior*]{} of this space. But in the S-matrix program it is precisely the boundaries of kinematic space that are of particular interest, since these are where amplitudes can have singularities. So from the perspective of determining a non-perturbative geometry for amplitudes, it is imperative to understand the [*boundary*]{} structure of $\operatorname{G}_+(k,n)/T$, and this involves making a choice of compactification. In this letter we propose natural compactifications of $\operatorname{G}_+(k,n)/T$ that address both of the above questions, providing us (when $k=4$) with candidate non-perturbative geometries for the amplitudes of pSYM theory. We provide explicit polytopal realizations of these compactifications using the “stringy canonical forms" of [@AHL]. This construction has a number of connections to other ideas, such hypersimplex decompositions and tropical Grassmannians, and we defer a systematic exposition to a longer companion paper [@toappear]. Our purpose in this letter is instead to summarize some of the essential ideas and results relevant to pSYM theory. In Sec. II we introduce certain compactifications of $\operatorname{G}_+(k,n)/T$ that manifestly have a finite number of facets for any $k, n$. In cases when the corresponding cluster algebra is finite each facet is naturally associated with a cluster variable, but in infinite cases there are additional facets. In Sec. III we describe a natural way of associating cluster algebraic functions, that generalize the notion of cluster variables, to such facets. In Sec. IV we study the case $(k,n) = (4,8)$ in detail. We find that the “extra” facets are associated with the famous square root associated to the four-mass box. Ancillary files contain data pertaining to several of the polytopes we study. II. Kinematic Space Polytopes ============================= Scattering amplitudes of $n$ particles in pSYM theory are functions on $\operatorname{Conf}_n(\mathbb{P}^3)$, the configuration space of $n$ points in $\mathbb{P}^3$. (Strictly speaking this is true only for MHV amplitudes; non-MHV amplitudes are most naturally thought of as differential forms [@Arkani-Hamed:2017vfh] on a $(\mathbb{C}^*)^{n-1}$ bundle over $\operatorname{Conf}_n(\mathbb{P}^3)$.) There is a birational isomorphism [@Golden:2013xva] $\operatorname{Conf}_n(\mathbb{P}^{k-1}) \simeq \operatorname{G}(k,n)/(\mathbb{C}^*)^{n-1}$ because generic points in this configuration space can be represented by maximal rank $k \times n$ matrices $Z$, with the columns representing the homogeneous coordinates of $n$ points in $\mathbb{P}^{k-1}$, modulo SL($k$) and modulo independent rescaling of each column. The (open) *positive domain* of $\operatorname{Conf}_n(\mathbb{P}^{k-1})$ (with respect to the ordering $1,\ldots,n$) is defined as the positive Grassmannian $\operatorname{G}_+(k,n)$ modulo the torus action $T = \mathbb{R}_+^n$ that rescales columns. The problem before us is that of understanding the boundary structure of this domain under a suitable compactification. There exist many inequivalent compactifications, with the choice appropriate for any particular application determined by, indeed one should say defined by, the class of functions under consideration. Ultimately it is the scattering amplitudes themselves that dictate the compactification of $\operatorname{G}_+(4,n)/T$ that is relevant to pSYM theory. The case $k=2$ is well-known to both mathematicians and physicists: $\operatorname{G}_+(2,n)/T$ is the moduli space of $n$ ordered points on the real line, and its Deligne-Mumford compactification [@DM] has long been known [@Koba:1969kh] to underlie the structure of open string amplitudes. The most natural generalization to $k>2$ is the positive Chow quotient of the Grassmannian [@toappear], which we will refer to as the *totally nonnegative configuration space* $\overline{\operatorname{G}_+(k,n)/T}$. Geometrically, it is the closure of $\operatorname{G}_+(k,n)/T$ inside the Chow quotient of the Grassmannian $\operatorname{G}(k,n)/\!/T$ [@Kapranov]. In [@AHL] it has recently been shown that Koba-Nielsen-like string worldsheet integrals can be used to construct polytopal realizations of various positive spaces, including $\overline{\operatorname{G}_+(k,n)/T}$, generalizing the well-known realization of the compactification of $\operatorname{G}_+(2,n)/T$ as the $A_{n-3}$ associahedron. The space $\overline{\operatorname{G}_+(k,n)/T}$ comes with a stratification that will be discussed elsewhere [@toappear], with strata labeled by several combinatorially interesting data including positroid decompositions of the $k,n$ hypersimplex, faces of the tropical positive Grassmannian, positive tropical Plücker vectors, etc. In this note we content ourselves with presenting the simplest data about its polytopal realization, which we denote by $\mathcal{C}(k,n)$. Here we briefly review the key steps of [@AHL]. A *positive parameterization* is a $d=(k{-}1)(n{-}k{-}1)$-parameter family of $k \times n$ matrices $Z(x_1,\ldots,x_d)$ that covers all of $\operatorname{G}_+(k,n)/T$ as the parameters range over the positive orthant $(x_1,\ldots,x_d) \in \mathbb{R}^d_+$. In the following it will be important that we always use a *cluster parameterization*, which means that the parameters are Fock-Goncharov coordinates in the initial cluster of the $\operatorname{G}(k,n)$ cluster algebra (for which we use the conventions of Fig. 2 of [@CDFL]). For example, $$\begin{aligned} Z = \begin{pmatrix} 0 & 1 & 1 & 1 & 1 \\ -1 & 0 & 1 & 1 + x_1 & 1 + x_1 + x_1 x_2 \end{pmatrix}\end{aligned}$$ is a cluster parameterization of $\operatorname{G}_+(2,5)/T$ because $$\begin{aligned} \frac{\langle 12 \rangle \langle 34 \rangle}{ \langle 14 \rangle \langle 23 \rangle} = x_1\,, \qquad \frac{\langle 13 \rangle \langle 45 \rangle}{ \langle 15 \rangle \langle 34 \rangle} = x_2\,.\end{aligned}$$ We use $\langle i_1 i_2 \ldots i_k \rangle = \det(Z_{i_1} \cdots Z_{i_k})$ to denote Plücker coordinates on $\operatorname{G}(k,n)$, where $Z_i$ is the $i$th column of $Z$. The canonical form [@Arkani-Hamed:2017tmz] on $\operatorname{G}_+(k,n)/T$ is $\Omega = \bigwedge d\log x_i$. If $P$ is a homogeneous polynomial in Plücker coordinates we let $\operatorname{Newt}(P)$ denote the Newton polytope of $P$ in $\mathbb{R}^d$ respect to the variables $(x_1,\ldots,x_d)$. The string integral construction associates to any sufficiently large (defined in [@AHL]) collection $\mathcal{P} = \{P_1, \ldots, P_\ell\}$ of such polynomials a polytopal realization of a compactification of $\operatorname{G}_+(k,n)/T$ obtained by taking the convex hull of the Minkowski sum of the corresponding $\operatorname{Newt}(P_i)$. We define $\mathcal{C}(k,n)$ to be the polytope obtained by taking $\mathcal{P}$ to be the set of all $\binom{n}{k}$ Plücker coordinates. If $k=2$ this gives the familiar realization [@Arkani-Hamed:2017mur] of the $A_{n-3}$ associahedron. However for $k>2$, $\mathcal{C}(k,n)$ is different than the corresponding cluster polytope [@FZY], and is in particular a manifestly finite polytope for any $k$ and $n$, even when $(k{-}2)(n{-}k{-}2) > 3$ in which case the cluster algebra is infinite and it is not clear that there even exists a cluster polytope. Using [@Minksum; @Normaliz] we have computed the vertices, found the bounding hyperplanes, and analyzed the polyhedral combinatorics of $\mathcal{C}(k,n)$ for various $(k,n)$. Here for brevity we summarize just the $f$-vectors $$\begin{aligned} \begin{split} (3,6):&\quad (1, 48, 98, 66, 16, 1)\,, \\ (4,7):&\quad (1, 693, 2163, 2583, 1463, 392, 42, 1)\,, \\ (3,8):&\quad (1, 13612, 57768, 100852, 93104, 48544, \\ & \qquad \qquad \quad \ \, \, \quad \qquad 14088, 2072, 120, 1)\,, \\ (4,8):&\quad (1, 90608, 444930, 922314, 1047200, 706042, \\ & \qquad \qquad\ \ \ \ \quad \quad 285948, 66740, 7984, 360, 1)\,. \end{split} \label{eq:gknmodt}\end{aligned}$$ The first two of these have appeared as the duals of the fans associated to the tropical positive Grassmannian $\operatorname{Trop}^+ \operatorname{G}(k,n)$ [@SpeyerWilliams], and some applications to physics for all four have recently been discussed in [@Cachazo:2019apa; @Drummond:2019qjk; @Cachazo:2019ble]. Note that these polytopes are neither simple nor simplicial. Except for the special case $n=6$, the $\mathcal{C}(4,n)$ polytopes are not invariant under the parity transformation $Z_i \mapsto *(Z_{i-1}Z_i Z_{i+1})$ that is a symmetry of MHV amplitudes. There are several ways of constructing parity-invariant polytopes, for example by taking a larger set $\mathcal{P}$ that includes the parity conjugates of all Plücker coordinates, or by taking a smaller set $\mathcal{P}$ of Plücker coordinates that is closed under parity. We find that the second option gives a particularly interesting polytope we call $\mathcal{C}^\dagger(k,n)$, obtained by taking $\mathcal{P}$ to be the subset of Plücker coordinates having the form $\langle i\,i{+}1\,j\,j{+}1\rangle$ or $\langle i\,j{-}1\,j\,j{+}1\rangle$. This is the largest subset of Plücker coordinates that is closed under parity, and for $n=7,8$ gives parity-invariant polytopes with $f$-vectors $$\begin{aligned} \begin{split} (4,7):&\quad (1, 595, 1918, 2373, 1393, 385, 42, 1)\,,\\ (4,8):&\quad (1, 49000, 249306, 536960, 635176, 447284, \\ &\qquad\qquad\qquad\,\,\, \ 189564, 46312, 5782, 274, 1)\,. \end{split} \label{eq:unnamed}\end{aligned}$$ Interestingly the number 595 appeared in [@SpeyerWilliams], where it was noted to be the number of facets of (the dual of) $\mathcal{C}(4,7)$ that are simplicial. However this seems to be a coincidence: we find that 50356 (not 49000) facets of (the dual of) $\mathcal{C}(4,8)$ are simplicial. The virtue of using a cluster parameterization in the string integral construction is that it ties the geometry of the resulting polytope to the combinatorics of cluster algebras. Specifically, we find that the (outward) normal rays to all facets of $\mathcal{C}(k,n)$ are generated by ${\bf g}$-vectors [@FZ4] of the $\operatorname{G}(k,n)$ cluster algebra in all of the cases listed above where the latter is finite. We remind the reader that the ${\bf g}$-vectors associated to the cluster variables in any one cluster generate a cone, the cones associated to different clusters are non-overlapping, and the union of all cones (called the “cluster fan”) covers all of $\mathbb{R}^d$ in finite cases. One of the salient features of infinite cluster algebras is the last of these is no longer true: there are directions in $\mathbb{R}^d$ that are outside the cluster fan, and so are not associated to any cluster. On the other hand the cones associated to the outward pointing normal rays to any polytope (the “normal fan”) manifestly cover all of space. In infinite cases some normal rays to our polytopes point in directions outside the cluster fan, and it is interesting to identify their cluster-algebraic significance. Specifically, in the infinite case $(4,8)$ we find that the normal rays to 356 of the facets of $\mathcal{C}(4,8)$ lie along ${\bf g}$-vectors of the $\operatorname{G}(4,8)$ cluster algebra, consistent with the results reported in [@Drummond:2019qjk; @Southampton; @HP] for the fan associated to $\operatorname{Trop}^+ \operatorname{G}(4,8)$. However, the remaining 4 normal rays are not even inside the cluster fan, according to the criterion given in Theorem 3.25 of [@CDFL]. The normal rays to the 274 facets of $\mathcal{C}^\dagger(4,8)$ are a proper subset of those of the facets of $\mathcal{C}(4,8)$; 272 of them lie along ${\bf g}$-vectors and 2 of them are outside the cluster fan. The cluster variables associated to the 272 ${\bf g}$-vector facets of $\mathcal{C}^\dagger(4,8)$ include the 108 symbol letters of the two-loop MHV amplitude [@CaronHuot:2011ky] and the 64 additional rational letters of the two-loop NMHV amplitude [@Zhang:2019vnm]. It would be interesting to see if this 272-letter alphabet exhausts the rational letters of (at least the MHV) 8-particle amplitudes to all loop order. (All evidence available to date suggests that the corresponding statement is true for the 9- and 42-letter symbol alphabets for the cases $n=6, 7$ respectively [@Caron-Huot:2019bsq; @Dixon:2016nkn].) In the following two sections we turn our attention to the remaining 2 (4) normal rays of $\mathcal{C}^\dagger(4,8)$ ($\mathcal{C}(4,8)$), which lie outside the cluster fan and so are not naturally associated to any cluster variables. We will see that the canonical basis element associated to the first integer point along each of these rays is overpositive, and we conjecture that the basis elements associated to points further along the rays are encapsulated in quadratic generating functions with positive roots. In the case of $\mathcal{C}^\dagger(4,8)$, these turn out to be precisely the roots associated to the four-mass box integral [@Hodges1977; @tHooft:1978jhc]. III. Cluster Canonical Bases ============================ Cluster variables in a rank $d$ cluster algebra are naturally associated with various lattice points in $\mathbb{Z}^d$, defined with respect to some initial seed cluster. One rather intuitive one is the notion of a “denominator vector". The Laurent phenomenon tells us that every cluster variable can be expressed as a ratio polynomial/monomial in the initial seed cluster variables, and the denominator vector of a cluster variable is the exponent vector of the monomial appearing in this expression. A more canonical object is the ${\bf g}$-vector we have already alluded to; one can think of the [**g**]{}-vector as the denominator vector with respect to an ordering defined by the $B$-matrix of the initial cluster. More precisely, one can define a partial ordering on the space of $n$-dimensional vectors [**g**]{} by saying that ${\bf g}' \preceq {\bf g}$ iff ${\bf g}' -{\bf g}$ is in the cone spanned by the columns of the initial $B$-matrix. If we expand any cluster variable $x$ as a sum of Laurent monomials in initial cluster variables, the [**g**]{}-vector of $x$ is defined to be that of the term whose [**g**]{}-vector is smallest with respect to $\preceq$. As we have mentioned, in finite-type cluster algebras, the [**g**]{}-vectors associated to the cluster variables of any cluster define a cone, and these cones are remarkably non-overlapping and cover all of $\mathbb{R}^d$. This is related to another beautiful fact in finite type: the collection of monomials associated with all the clusters provides a basis for the entire cluster algebra. This is not obvious. The Laurent phenomenon guarantees that any product of cluster variables can be represented as a Laurent polynomial in terms of some initial seed cluster variables, but the claim is that any product of cluster variables from arbitrarily distant clusters can be written as a polynomial made of sums of monomials of variables in the same cluster. Beyond finite type, while the cluster cones are still non-overlapping, they do not cover all of space. Related to this, cluster monomials no longer provide a basis for the full cluster algebra, as there are directions in [**g**]{}-vector space that can not be spanned by products of cluster variables. There is a large literature on the construction of various “canonical bases" of cluster algebras in infinite type. For $\operatorname{G}(k,n)$ every integer point in [**g**]{}-vector space can be assigned [@CDFL] (partly conjecturally) a polynomial in cluster variables, Lusztig’s canonical basis element [@Lusztig]. When the integer point lies inside a [**g**]{}-vector cone of some cluster, the corresponding function is simply the obvious associated monomial in cluster variables, but when the integer point is on a ray not pointing in the direction of a cluster variable, the corresponding basis element is not a cluster variable, but something else. We will encounter precisely this situation in our study of $\mathcal{C}(k,n)$, and explicitly for the $(k,n)=(4,8)$ case of interest for amplitudes. The facets of $\mathcal{C}(4,8)$ have some normal rays that point in non-cluster directions, and it is natural to ask for the generating function for the canonical basis elements associated to all points along each such ray. This will motivate an association between the poles of this generating function and symbol letters, revealing a remarkable connection between $\mathcal{C}(4,8)$, the canonical basis, and the quadratic equation associated with the four-mass box symbol letters. A Rank-2 Example ---------------- Before jumping into the intricacies of $\operatorname{G}(4,8)$ let us warm up by illustrating many of the salient points with the simplest example of an infinite-type cluster algebra. Largely following [@sherman2004positivity], consider the rank-2 cluster algebra with initial exchange matrix $$\begin{aligned} B = \begin{pmatrix} 0 & -2 \\ 2 & 0 \end{pmatrix}\end{aligned}$$ whose cluster variables $x_n$ are determined in terms of those of the initial cluster $(x_1, x_2)$ by $$\begin{aligned} \label{eq:mutation} x_{n+1} = \frac{1 + x_n^2}{x_{n-1}}\,.\end{aligned}$$ For $a,b \in \mathbb{Z}$, the ${\bf g}$-vector of a monomial $x_1^a x_2^b$ in the initial variables is defined to be $(a,b)$. As mentioned above, for general cluster variables, the [**g**]{}-vector is given by the exponent vector in the Laurent expansion which is smallest by the partial ordering specified by $B$. For example, two cluster variables are $$\begin{aligned} x_4 = \frac{1}{x_2} + \frac{1}{x_1^2x_2} + 2 \frac{x_2}{x_1^2} + \frac{x_2^3}{x_1^2}\end{aligned}$$ and $$\begin{aligned} x_0 = \frac{x_1^2}{x_2} + \frac{1}{x_2}\,,\end{aligned}$$ where in each case the terms are written in increasing order with respect to $\preceq$. We see that the ${\bf g}$-vector of $x_4$ is $(0,-1)$. In general, it is easy to work out that the [**g**]{}-vector associated to $x_n$ is $(n{-}4,3{-}n)$ for $n \geq 3$ and $(2{-}n,n{-}1)$ for $n \leq 2$. The union of ${\bf g}$-vector cones *almost* covers all of $\mathbb{R}^2$, but they accumulate along a single missing ray generated by $(1,-1)$. It is thus clear that cluster monomials don’t provide a basis for the full cluster algebra. For instance, the smallest monomial in the product $x_1 x_4$ has ${\bf g}$-vector equal to $(1,-1)$, and so can’t be written as a sum of cluster monomials. What is needed to complete a basis for the cluster algebra? It is easy to give an elementary answer to this question in this very simple example. We begin by noting that the mutation relation (\[eq:mutation\]) may be recast into the form of a recurrence relation $$\begin{aligned} x_{n+1} = A x_n - x_{n-1} ~ \mbox{with} ~ A = \frac{x_1}{x_2} + \frac{1}{x_1 x_2} + \frac{x_2}{x_1}\,. \label{eq:chebyshev}\end{aligned}$$ In fact $x_1,x_2$ can be replaced by any $x_n,x_{n+1}$ in the expression for $A$; it is an invariant across all clusters. We can use this generating function to give an explicit expression for all cluster variables. If we define the generating function $X(t) = \sum_{k \ge 0} x_k t^k$, then the recurrence relation implies that $(1 - A t + t^2) X(t) = x_0 + t (x_1 - A x_0)$, and hence $$\begin{aligned} X(t) = \frac{x_0(1 - A t) + x_1 t}{1 - A t + t^2}\,.\end{aligned}$$ By factoring the quadratic equation in terms of its roots $R_{\pm}$, we can also express $x_n$ as $$\begin{aligned} x_n = \frac{x_0 (R_-^{n+1} - R_+^{n+1}) + (x_1 - A x_0)(R_-^n - R_+^n)}{R_- - R_+}\end{aligned}$$ where $$\begin{aligned} R_\pm = \frac{A \pm \sqrt{A^2 - 4}}{2}\,.\end{aligned}$$ Here we see the first occurrence of an interesting quadratic equation and its associated roots in the generating function $X(t)$ for the cluster variables themselves. The variable $A$ has a deeper significance. Note that while e.g. $x_1 x_4$ can’t be expressed as a sum of cluster monomials, we can write $x_1 x_4 = A + x_2 x_3$, suggesting that $A$ should be considered an element of the basis. Its [**g**]{}-vector is $(1,-1)$, which lies along the missing ray, so it is a basis element that is not a cluster variable. What basis elements should we associate with the other integer points $(p,-p)$ along this ray? Most naively, in analogy with the case of cluster variables, we might think that these should just be the powers $A^p$. But the variable $A$ has a very interesting and peculiar property that suggests this is the wrong answer. Cluster variables are “critically positive", in the sense that they approach zero on some boundaries of the positive part of the cluster variety. But this is not true of $A$! Note that $A = r + 1/r + 1/(x_1 x_2)$ where $r=x_1/x_2$, and thus $A \geq 2$. $A$ is thus “overpositive", and can’t reach zero on any boundary of the positive part. Relatedly, we can’t understand the positivity of $A - 2$ in the way that one familiarly understands positivity of polynomials in cluster variables. Usually, an expression can be determined to be positive simply by expanding in the cluster variables of an initial seed and seeing that all terms in the Laurent expansion are positive. But that is not the case for $A-2$; this expression is positive despite the appearance of a negative sign in its Laurent expansion simply because $A^2 - 4 > (r+1/r)^2 - 4 = (r - 1/r)^2 > 0$. Motivated by this observation, we say that an element $x$ of the algebra is *positive* if it is positive-valued when $x_1, x_2 > 0$ and that a positive $x$ is *Laurent positive* (with respect to the initial cluster) if it is a linear combination of initial cluster monomials with positive coefficients; otherwise $x$ is *nontrivially positive*. Finally we say that $x$ is *overpositive* if there exists a positive $y$ such that $x/y$ is non-constant and $\min(x/y)>0$. Note that $$\begin{aligned} A^2 = \frac{2}{x_1^2} + \frac{2}{x_2^2} + \frac{1}{x_1^2 x_2^2} + \frac{x_1^2}{x_2^2} + \frac{x_2^2}{x_1} + 2\end{aligned}$$ so we see that $A^2 - z$ is Laurent positive only for $z \le 2$ (this is another way to see that $A$ is overpositive). It is still positive for $2 < z \le 4$ since we have already seen that $A \ge 2$, and an important consequence of this nontrivial positivity is that $R_\pm$ are both real and positive-valued in the positive part of the cluster varitey. Finally, it is easy to see that $A^2 - z$ is not positive for $z > 4$. Moving to higher powers, it is natural to ask for some basis $T_p(A)$ of polynomials in $A$, beginning with $T_0(A) = 1$, $T_1(A) = A$ and $T_2(A) = A^2 - 2$, such that $T_p(A)$ is maximally Laurent positive (that means, with no further subtractions possible). Using $A = r + 1/r + 1/(x_1 x_2)$, such polynomials can be determined by requiring $G_p(z + 1/z) = z^p + 1/z^p$. The generating function for basis elements of this form along the missing ray, $g(t) = \sum_{p \ge 0} G_p(A) t^p$, is given by $g(t) = (1 - t^2)/(1 - A t + t^2)$. Note the second appearance of the the same quadratic polynomial in $A$ we saw earlier, this time not in computing cluster variables, but more fundamentally as generating the basis elements for this “critically Laurent positive" basis along the missing $(1,-1)$ ray. It is also natural to consider another basis $F_p$ with generating function $f(t) = (1 - A t + t^2)^{-1}= \sum_{p \ge 0} F_p(A) t^p$, which is expected to be the analogue of Lusztig’s canonical basis for this rank 2 cluster algebra. We will not attempt to explain the deep significance of this basis here; for now, we simply wish to emphasize that the nontriviality of the generating function for both the bases we have highlighted is associated with the surprising overpositivity of the non-cluster variable $A$ associated with the first integer point on the non-cluster ray. We also emphasize that while the generating functions $f(t)$ and $g(t)$ clearly differ, they have the same poles in $t$; we expect this will be true of any suitably reasonable basis. A Definition of Cluster Algebraic Functions ------------------------------------------- Let ${\mathcal A}$ be a cluster algebra of rank $d$, let ${\bf X} \simeq \mathbb{Z}^d$ be a lattice that parameterizes bases of ${\mathcal A}$, and let $\mathcal{B}({\bf g})$ be a basis element associated to the lattice point ${\bf g} \in {\bf X}$. For a ray in ${\bf X}$ with integer points $1 \cdot {\bf g}, 2 \cdot {\bf g}, \ldots$ (note that we always take ${\bf g}$ to be the first integer point along its ray) we define the generating function $$\begin{aligned} \label{eq:genfuncdef} f_{\bf g}(t) = \sum_{k \ge 0} \mathcal{B}(k {\bf g}) t^k\,.\end{aligned}$$ In general $f_{\bf g}(t)$ depends on the choice of basis, but if ${\bf g}$ lies in the cluster fan of ${\mathcal A}$ then $\mathcal{B}(k {\bf g}) = \mathcal{B}({\bf g})^k$, where $\mathcal{B}({\bf g})$ is the cluster monomial associated to ${\bf g}$, and hence $$\begin{aligned} \label{eq:simplepole} f_{\bf g}(t) = \frac{1}{1 - t \mathcal{B}({\bf g})}\,.\end{aligned}$$ Motivated by (\[eq:simplepole\]), in cases where $f_{\bf g}(t)$ is a rational function of $t$ we denote the roots of $1/f_{\bf g}(1/t)$ by $R({\bf g})$. We conjecture that these are always positive on ${\mathcal A}_{> 0}$, the positive part of ${\mathcal A}$. We call $R({\bf g})$ the *cluster algebraic function* associated to the ray generated by ${\bf g}$. If ${\bf g} \in {\bf X}$ is a ${\bf g}$-vector of the cluster algebra ${\mathcal A}$ then $R({\bf g})$ contains a unique element, the cluster variable associated to ${\bf g}$, but if ${\bf g}$ lies outside the cluster fan of ${\mathcal A}$ then $R({\bf g})$ is a finite collection of algebraic functions of cluster variables. Further aspects of such functions will be discussed in [@toappear]. IV. $\operatorname{G}(4,8)$ and the Four-Mass Box ================================================= Here we apply the proposal just introduced to the polytopes constructed in section II. As reported there, we find that the normal rays to 356 facets of $\mathcal{C}(4,8)$ are generated by ${\bf g}$-vectors of the $\operatorname{G}(4,8)$ cluster algebra, and therefore are naturally associated to 356 of its cluster variables. The other 4 facets have normal rays generated by $$\begin{aligned} \begin{split} {\bf g}_1 &= (-1, 1, 0, 1, 0, -1, 0, -1, 1)\,, \\ {\bf g}_2 &= (0, -1, 0, -1, 0, 1, 0, 1, 0)\,, \\ {\bf g}_3 &= (-1, -1, 1, -1, 2, 0, 1, 0, -1)\,, \\ {\bf g}_4 &= (1, 0, -1, 0, -2, 1, -1, 1, 1)\,. \end{split} \label{eq:fourmissing}\end{aligned}$$ The first two of these are also normal rays to $\mathcal{C}^\dagger(4,8)$. We adopt Corollary 7.3 of [@CDFL] as a (conjectural) way to assign a canonical basis element to any lattice point in $\mathbb{R}^d$, regardless of whether it is in the cluster fan. In the notation defined in that paper, the semistandard Young tableaux $T_1, \ldots, T_4$ associated to (\[eq:fourmissing\]) are respectively $$\begin{aligned} \begin{ytableau} 1 & 3 \\ 2 & 5 \\ 4 & 7 \\ 6 & 8 \end{ytableau}\,, \qquad \begin{ytableau} 1 & 2 \\ 3 & 4 \\ 5 & 6 \\ 7 & 8 \end{ytableau}\,, \qquad \begin{ytableau} 1 & 1 & 2 & 4 \\ 2 & 3 & 3 & 6 \\ 4 & 5 & 5 & 7 \\ 6 & 7 & 8 & 8 \end{ytableau}\,, \qquad \begin{ytableau} 1 & 1 & 2 & 3 \\ 2 & 4 & 4 & 5 \\ 3 & 6 & 6 & 7 \\ 5 & 7 & 8 & 8 \end{ytableau}\,.\end{aligned}$$ Let us first consider $T_1$, which is the same as the $T_4$ considered in Example 8.1 of [@CDFL], where the variables associated to the first two points along the ray generated by ${\bf g}_1$ were computed as $$\begin{aligned} \begin{split} \mathcal{B}(1 \cdot {\bf g}_1) &= \operatorname{ch}(T_1) = A\,,\\ \mathcal{B}(2 \cdot {\bf g}_1) &= \operatorname{ch}(T_1 \cup T_1) = A^2 - B \label{eq:firsttwo} \end{split}\end{aligned}$$ in terms of the quantities $$\begin{aligned} \begin{split} A &= \langle 1256 \rangle \langle 3478\rangle - \langle 1278 \rangle \langle 3456 \rangle - \langle 1234 \rangle \langle 5678 \rangle\,, \\ B &= \langle 1234 \rangle \langle 3456 \rangle \langle 5678 \rangle \langle 1278 \rangle\,. \end{split} \label{eq:abdef}\end{aligned}$$ We have computed the next variable along this ray, $$\begin{aligned} \mathcal{B}(3 \cdot {\bf g}_1) = \operatorname{ch}(T_1 \cup T_1 \cup T_1) = A^3 - 2 A B\,. \label{eq:third}\end{aligned}$$ Based on (\[eq:firsttwo\]) and (\[eq:third\]) we conjecture that $$\begin{aligned} \label{eq:genfuncconjecture} f_{{\bf g}_1}(t) = \sum_{k \ge 0} \operatorname{ch}(T_1^{\cup k}) t^k = \frac{1}{1 - t A + t^2 B}\,.\end{aligned}$$ Note that this encapsulates an infinite number of predictions about the behavior of the canonical basis along the direction ${\bf g}_1$. According to the proposal outlined in the previous section, the variables associated to this ray are therefore $$\begin{aligned} R({\bf g}_1) = \frac{A \pm \sqrt{A^2 - 4 B}}{2}\,.\end{aligned}$$ Exactly as in the rank two toy example we studied in the previous section, using the representation of $A, B$ in terms of initial cluster variables it is straightforward to check that $A$ is overpositive and $\Delta \equiv A^2 - 4 B$ is positive, even though $A^2 - r B$ is Laurent positive only for $r \le 2$ and not positive if $r > 4$. This can be seen by noting that expanding in terms of initial cluster variables, we have $A = x + y + \cdots$, where $x,y$ are monomials such that $B = x y$. This shows that $A^2 - 4B$ is positive and that $A^2 - 2 B$ is Laurent positive. Furthermore, it can be checked that the exponent vectors of $x,y$ are separated from the exponent vectors of all the other monomials in $A$ by a hyperplane. This means that we can scale the cluster variables in such a way that $x,y$ dominate by arbitrarily large factors relative to the other monomials in $A$, and shows that we can make $A^2 - r B$ negative for $r>4$, so $r=4$ is the critical case. Even more exciting is the fact that $\sqrt{\Delta}$ is precisely one of the two cyclic incarnations of the square root that appears in the four-mass box integral [@Hodges1977; @tHooft:1978jhc]; its cyclic partner comes from the ray generated by ${\bf g}_2$. These are known to be the only square roots that appear in the symbol alphabets of the one-loop N${}^2$MHV [@Britto:2004nc] and two-loop NMHV amplitudes [@Zhang:2019vnm]; they are also expected to appear in MHV amplitudes at three loops and beyond [@Prlina:2017tvx]. Finally we turn to ${\bf g}_3$ and its cyclic partner ${\bf g}_4$. It is straightforward to compute the variable $G = \mathcal{B}(1 \cdot {\bf g}_3) = \operatorname{ch}(T_3)$ associated to the lattice point ${\bf g}_3$. Remarkably we find that $G$, which has torus weight 2 in each of the eight $Z_i$, is related to $A$ by a braid element [@FraserBraid] of the $\operatorname{G}(4,8)$ cluster modular group. Under the same braid transformation we find that $B \mapsto B'$ where $$\begin{aligned} B' = B \,\langle 2345 \rangle \langle 4567 \rangle \langle 1678 \rangle \langle 1238 \rangle\,.\end{aligned}$$ Therefore $\Delta$ maps to $\Delta' = G^2 - 4 B'$, which again is nontrivially positive. While the presence of the additional root $\sqrt{\Delta'}$ (and its cyclic partner) associated to $\mathcal{C}(4,8)$ has a simple and beautiful mathematical origin, it is unclear whether these roots play any role in physics; for example, whether they correspond to the Landau singularities. This may be another sign (beyond the consideration of parity discussed above) that $\mathcal{C}^\dagger(4,8)$ (which lacks these extra roots) is more relevant to the physics of MHV amplitudes than $\mathcal{C}(4,8)$. Outlook ======= Our work raises a number of related questions. On the mathematical side, it would be interesting to verify the conjecture (\[eq:genfuncconjecture\]) and to explore the corresponding generating functions at higher $n$ and $k$. More generally, it would be interesting to understand under what circumstances the generating function defined in (\[eq:genfuncdef\]) is rational (for example, is this true in $\operatorname{G}_+(k,n)$ for all ${\bf g}$?) and to prove our conjecture that in such cases the roots of its denominator are always positive. On the physics side it would be interesting to determine if the 272 cluster variables associated to $\mathcal{C}^\dagger(4,8)$, together with the algebraic letters of four-mass-box type, indeed constitute the all-loop symbol alphabet of the 8-particle MHV amplitude. Evidence in support of this suggestion could be provided either by explicit computation or by using them as a bootstrap ansatz. Also, we have so far only discussed the crudest relation between polytopes and cluster algebras, according to which symbol letters of the latter are related to facets of the former. For finite algebras this connection was first observed in [@Golden:2013xva] but in recent years much finer connections have been explored, for example the observation [@Drummond:2017ssj] that two cluster variables can appear next to each other in a symbol only if the corresponding facets intersect. From (\[eq:unnamed\]) we see that $\mathcal{C}^\dagger(4,7)$ has 385 codimension-2 faces, suggesting that only 385 distinct pairs of adjacent symbol entries (all of the ones tabulated in [@Drummond:2017ssj] except for those described in equation (12b)) appear in 7-particle MHV amplitudes. This is consistent with all evidence available to date [@Southampton]. It would be interesting to explore this type of finer structure for $n>7$, and especially to understand the “cluster adjacency” properties of algebraic symbol letters. We leave the most ambitious question for last. Supposing that an appropriate non-perturbative geometry for pSYM theory is indeed provided by a construction of the type that we have described, what is the non-perturbative question we should ask of this space, whose answer gives a scattering amplitude? Acknowledgements ================ It is a pleasure to thank J. Drummond, J. Foster, [" O]{}. G[ü]{}rdo[ğ]{}an, S. He, C. Kalousios and A. Volovich for stimulating conversations, and C. Fraser and J.-R. Li for helpful correspondence. This work was supported in part by DOE grants DE-SC0009988 (NAH) and DE-SC0010010 Task A (MS). [99]{} N. Arkani-Hamed and J. Trnka, JHEP [**1410**]{}, 030 (2014) \[arXiv:1312.2007 \[hep-th\]\]. N. Arkani-Hamed, Y. Bai and T. Lam, JHEP [**1711**]{}, 039 (2017) \[arXiv:1703.04541 \[hep-th\]\]. I. Prlina, M. Spradlin and S. Stanojevic, Phys. Rev. Lett.  [**121**]{}, no. 8, 081601 (2018) \[arXiv:1805.11617 \[hep-th\]\]. A. Hodges, JHEP [**1305**]{}, 135 (2013) \[arXiv:0905.1473 \[hep-th\]\]. N. Arkani-Hamed, J. L. Bourjaily, F. Cachazo, A. B. Goncharov, A. Postnikov and J. Trnka, arXiv:1212.5605 \[hep-th\]. J. Golden, A. B. Goncharov, M. Spradlin, C. Vergu and A. Volovich, JHEP [**1401**]{}, 091 (2014) \[arXiv:1305.1617 \[hep-th\]\]. S. Caron-Huot, L. J. Dixon, F. Dulat, M. Von Hippel, A. J. McLeod and G. Papathanasiou, JHEP [**1909**]{}, 061 (2019) \[arXiv:1906.07116 \[hep-th\]\]. A. B. Goncharov, M. Spradlin, C. Vergu and A. Volovich, Phys. Rev. Lett.  [**105**]{}, 151605 (2010) \[arXiv:1006.5703 \[hep-th\]\]. L. J. Dixon, J. Drummond, T. Harrington, A. J. McLeod, G. Papathanasiou and M. Spradlin, JHEP [**1702**]{}, 137 (2017) \[arXiv:1612.08976 \[hep-th\]\]. I. Prlina, M. Spradlin, J. Stankowicz, S. Stanojevic and A. Volovich, JHEP [**1805**]{}, 159 (2018) \[arXiv:1711.11507 \[hep-th\]\]. I. Prlina, M. Spradlin, J. Stankowicz and S. Stanojevic, JHEP [**1804**]{}, 049 (2018) \[arXiv:1712.08049 \[hep-th\]\]. A. Hodges, Twistor Newsletter 5, 1977; reprinted in *Advances in twistor theory*, eds. L.P. Hugston and R. S. Ward (Pitman, 1979). G. ’t Hooft and M. J. G. Veltman, Nucl. Phys. B [**153**]{}, 365 (1979). D. Chicherin, J. Henn and V. Mitev, JHEP [**1805**]{}, 164 (2018) \[arXiv:1712.09610 \[hep-th\]\]. T. Gehrmann, J. M. Henn and N. A. Lo Presti, JHEP [**1810**]{}, 103 (2018) \[arXiv:1807.09812 \[hep-ph\]\]. D. Chicherin, T. Gehrmann, J. M. Henn, P. Wasser, Y. Zhang and S. Zoia, Phys. Rev. Lett.  [**122**]{}, no. 12, 121602 (2019) \[arXiv:1812.11057 \[hep-th\]\]. M. Heller, A. von Manteuffel and R. M. Schabinger, arXiv:1907.00491 \[hep-th\]. C. Zhang, Z. Li and S. He, arXiv:1911.01290 \[hep-th\]. M. Besier, P. Wasser and S. Weinzierl, arXiv:1910.13251 \[cs.MS\]. J. L. Bourjaily, A. J. McLeod, C. Vergu, M. Volk, M. Von Hippel and M. Wilhelm, arXiv:1910.14224 \[hep-th\]. N. Arkani-Hamed, S. He and T. Lam, “Stringy Canonical Forms,” to appear. N. Arkani-Hamed, M. Spradlin and T. Lam, to appear. N. Arkani-Hamed, H. Thomas and J. Trnka, JHEP [**1801**]{}, 016 (2018) \[arXiv:1704.05069 \[hep-th\]\]. P. Deligne and D. Mumford, [*Publications Math[é]{}matiques de l’Institut des Hautes [É]{}tudes Scientifiques*]{} [**36**]{}, no. 1, 75 (1969). Z. Koba and H. B. Nielsen, Nucl. Phys. B [**12**]{}, 517 (1969). M. Kapranov, Adv. Sov. Math. [**16**]{}, no. 2, 29 (1993), arXiv:alg-geom/9210002. W. Chang, B. Duan, C. Fraser and J.-R. Li, arXiv:1907.13575 \[math.RT\]. N. Arkani-Hamed, Y. Bai, S. He and G. Yan, JHEP [**1805**]{}, 096 (2018) \[arXiv:1711.09102 \[hep-th\]\]. S. Fomin and A. Zelevinsky, Annals Math. [**158**]{}, no. 3, 977 (2003) \[arXiv:hep-th/0111053\]. W. Bruns, B. Ichim, T. Römer, R. Sieg and C. Söger, Normaliz: Algorithms for rational cones and affine monoids, available at <http://normaliz.uos.de> C. Weibel, Minksum: A Reverse-search algorithm for Minkowski sums, available at <https://sites.google.com/site/christopheweibel/research/minksum> D. Speyer and L. Williams, J. Algebr. Comb. [**22**]{}, no. 2, 189 (2005), arXiv:math/0312297. J. Drummond, J. Foster, [" O]{}. G[ü]{}rdo[ğ]{}an and C. Kalousios, arXiv:1907.01053 \[hep-th\]. F. Cachazo and J. M. Rojas, arXiv:1906.05979 \[hep-th\]. F. Cachazo, B. Umbert and Y. Zhang, arXiv:1911.02594 \[hep-th\]. S. Fomin and A. Zelevinsky, Compos. Math. [**143**]{}, no. 1, 112 (2007), arXiv:math/0602259. J. Drummond, J. Foster, [" O]{}. G[ü]{}rdo[ğ]{}an and C. Kalousios, “Algebraic singularities of scattering amplitudes from tropical geometry,” to appear. N. Henke and G. Papathanasiou, “How tropical are seven- and eight-particle amplitudes?,” to appear. S. Caron-Huot, JHEP [**1112**]{}, 066 (2011) \[arXiv:1105.5606 \[hep-th\]\]. G. Lusztig, J. Am. Math. Soc.  [**3**]{}, no. 2, 447 (1990). P. Sherman and A. Zelevinsky, Moscow Math. J. [**4**]{}, no. 4, 947 (2004), arXiv:math/0307082. R. Britto, F. Cachazo and B. Feng, Nucl. Phys. B [**725**]{}, 275 (2005) \[hep-th/0412103\]. C. Fraser, arXiv:1702.00385 \[math.CO\]. J. Drummond, J. Foster and [" O]{}. G[ü]{}rdo[ğ]{}an, Phys. Rev. Lett.  [**120**]{}, no. 16, 161601 (2018) \[arXiv:1710.10953 \[hep-th\]\].
{ "pile_set_name": "ArXiv" }
--- abstract: 'Dichotomous noise appears in a wide variety of physical and mathematical models. It has escaped attention that the standard results for the long time properties cannot be applied when unstable fixed points are crossed in the asymptotic regime. We show how calculations have to be modified to deal with these cases and present as a first application full analytic results for hypersensitive transport.' author: - 'I. Bena and C. Van den Broeck' - 'R. Kawai' - Katja Lindenberg title: Nonlinear Response With Dichotomous Noise --- While the Wiener process together with its “time derivative," the Gaussian white noise, is certainly the method of choice to describe Brownian motion, the motion induced by another fundamental stochastic process, namely, the dichotomous Markov process (see, e.g. [@kampen; @fulinsky94]), has its own virtues and interest. Systems driven by dichotomous noise can often be described analytically, including the Gaussian white noise case as a specific limit, and allow the analytic investigation of the effects of the finite correlation time of the noise, notably in noise induced transitions, noise induced phase transitions, stochastic resonance, and ratchets. Dichotomous noise is often a good representation of the actual physical situation, e.g., thermal transitions between two configurations or states, and can easily be implemented as an external noise, with the additional advantage that the support of this noise is finite. A widely studied generic stochastic equation that describes the temporal evolution of a single scalar variable $x(t)$ is $$\dot{x}=\xi(t)v(x)+F, \label{1}$$ where the dot stands for the time derivative, $\xi(t)$ is a symmetric dichotomous Markov process that takes on the values $\pm 1$ with transition rate $k$, $v(x)$ is a given velocity profile, and $F$ is a constant external force. One can of course embellish this description in a variety of ways such as, e.g., by allowing for a state- and/or time-dependent external force, but here we adhere to this simple form. Existing results include the steady state distribution [@klyatskin77; @horsthemke] and first passage time moments, see e.g. [@MFPT]. When Eq. (\[1\]) is defined on an interval with periodic boundary conditions, one can evaluate the stationary probability flux and from it the average asymptotic drift velocity or diffusion coefficient, a problem that has recently received a great deal of attention in the context of Josephson junctions and Brownian motors [@stokesdrift]. Although these results are often claimed to be completely general, our study shows that this is [*not*]{} the case. Indeed, to our knowledge, with the exception of [@bala01], all the existing results are limited to motion that asymptotically [*does not cross unstable fixed points of the dynamics*]{}. The main purpose of our work is to point out where the existing results break down, and to present the procedure to obtain a fully general solution for the asymptotic average velocity, including, as a first direct application, the problem of hypersensitive response [@ginzburg]. ![ The net potentials $\mp V(x) -Fx$. []{data-label="fig1"}](fig1.eps){width="2.8in"} Consider, then, the prototypical stochastic differential equation (\[1\]). We take the velocity profile $v(x)$ to be periodic, $v(x)=v(x+L)$, with zero average, $\int_{0}^{L} v(x)\,dx =0$. A schematic representation of the two configurations assumed by the “net potentials" $\mp V(x) -Fx$ associated with right hand side (with $v(x)\equiv -dV(x)/dx$) is shown in Fig. \[fig1\]. The fixed points of the dynamics are the points at which $\pm v(x)+F$ vanish, i.e., the local extrema of the net potentials. The stochastic dynamics (\[1\]) can equivalently be described by the master equation for the probability densities $P_{+}(x,t)$ and $P_{-}(x,t)$ for being at $x$ at time $t$ if $\xi=+1$ and $-1$, respectively, with $x \in [0,L]$ and periodic boundary conditions: $$\begin{aligned} \frac{\partial P_{\pm}(x,t)}{\partial t}&=&\mp\frac{\partial}{\partial x} \left\{\left[v(x)\pm F\right]P_{\pm}(x,t)\right\} \nonumber\\ &&-k\left[P_{\pm}(x,t)- P_{\mp}(x,t)\right]. \label{4}\end{aligned}$$ To find analytic expressions for the asymptotic steady state (i.e., time-independent) probabilities and mean velocity $\left<\dot{x}\right>$, it is more convenient to introduce the sum and difference probability densities $P(x,t)=P_{+}(x,t)+P_{-}(x,t)$ and $p(x,t)=P_{+}(x,t)-P_{-}(x,t)$. Summation of the equations in (\[4\]) then immediately leads to the conclusion that in the steady state the probability flux $J$ associated with $P(x)$, namely, $J=FP(x)+v(x)p(x)$, is a constant whose value is to be determined. It also leads to a direct relation between the mean velocity in the stationary state and the flux, $\langle \dot{x} \rangle = \int_0^L \left\{ \left[ v(x)+F\right] P_+(x) + \left[ -v(x)+F\right]P_-(x)\right\} dx = LJ$. Again in the steady state, subtraction of the equations in (\[4\]) and the constant flux condition leads to the following first-order differential equation for $p(x)$: $$\left[F^2-v^2(x)\right] \frac{dp(x)}{dx}-2\left[v(x)\,v'(x)-kF\right]p(x)+Jv'(x)=0, \label{15}$$ where $v'(x)=dv(x)/dx$. The solution to this equation, together with the constant flux condition and the normalization condition for $P(x)$, $\int_0^LP(x)dx=1$, can be used to determine $P(x)$, the flux $J$ and the mean velocity $\left< \dot{x}\right>$. The crux of the problem now resides in finding the solution to Eq. (\[15\]). This solution is straightforward when $\left[F^2-v^2(x)\right]$ [*has no zeroes*]{}, that is, when the net potentials have no extrema within the interval $(0,L)$. In this case the standard method of variation of parameters leads to the familiar solution $$p(x)=-\frac{J}{\left|F^2-v^2(x)\right|}\left[ CG(x,0)+ H(x,0;x)\right], \label{17}$$ where $C$ is a constant of integration that arises from the general solution to the homogeneous part of Eq. (\[15\]), the second contribution is the particular solution, and we have defined the functions $$\begin{aligned} H(z,y;x) &=&\int_y^z\mbox{sgn}[F^2-v^2(x')]v'(x') G(x,x')dx', \nonumber\\ G(z,y)&\equiv& \exp\left[-2kF\int_y^z \frac{dx}{F^2-v^2(x)}\right]. \label{new}\end{aligned}$$ The usual procedure to determine $C$ is to require periodicity of $p(x)$, recalling that $v(x)$ is periodic. One finally obtains $p(x)=[J-FP(x)]/Jv(x)$, and $$\begin{aligned} P(x)&&= \frac{J}{F} \left\{1-\frac{v(x)}{\left[1-G(0,L)\right] [F^2-v^2(x)]}\right. \nonumber\\ &&\left. \times \int_x^{x+L}dx' v'(x')G(x,x') \right\}. \label{21}\end{aligned}$$ The normalization of $P(x)$ determines the value of the flux $J$ and leads to the following result for the mean velocity at the steady state: $$\begin{aligned} \langle \dot{x}\rangle &=& F \left\{1-\frac{1}{L\left[ 1-G(0,L)\right]} \int_0^L dx \frac{v(x)}{F^2-v^2(x)} \right. \nonumber\\ &&\left. \times \int_x^{x+L} dx' v'(x') G(x,x') \right\}^{-1}.\end{aligned}$$ The standard results shown above are applicable not only in the absence of fixed points, but also when the asymptotic behavior is governed by [*stable*]{} fixed points. In this latter case the dynamics settles into an alternating motion between these points, so that they delimit the interval in which the steady state probability is non-zero [@horsthemke]. The associated normalizable divergences at the fixed points represent regions where the probability density for finding the system is high. The situation is entirely different, both physically and mathematically, when the system can cross [*unstable*]{} fixed points within the interval $(0,L)$ in the long time limit. A simple illustration is provided by the example $v(x)=\sin\,x$. In the absence of an external force, the dynamics is restricted to an interval $[k\pi,\,(k+1)\pi]\,\,\,(k\,\,\,{\mbox {integer}})$. Even though the application of an external forcing $|F|<1$ cannot induce “escape" from this interval in either of the separate dynamics $\dot{x}=\sin\, x + F$ and $\dot{x}=-\sin\, x + F$, running solutions with finite average velocity appear when the dynamics switches back and forth between the two [@ginzburg] (see Fig. \[fig1\]). The explicit calculation of this velocity is one of our main goals. Clearly, the solution (\[21\]) is no longer correct because it contains non-integrable singularities (see below) at the unstable fixed points where the probability of finding the system is expected to be [*low*]{}, not high. For simplicity we restrict our presentation to velocity profiles $v(x)$ that are continuously decreasing functions of $x$ on $[0,\,L/2]$ and symmetric about $L/2$, $\,v(x+L/2)=-v(x)$. This implies that $P(x+L/2)=P(x)$ and $p(x+L/2)=-p(x)$, so that we can limit our analysis to the interval $[0,\,L/2]$. In this “minimal scenario", the equation $F^2\,-\,v^2(x)=0$ has only two solutions in the interval $[0,\,L/2]$, namely, $x_1$, corresponding to an unstable fixed point in the $\xi=-1$ dynamics $[F=v(x_1)]$, and $x_2$, a stable fixed point in the $\xi=+1$ dynamics $[F=-v(x_2)]$, with $x_2>x_1$. The steady state results leading to Eq. (\[15\]) still apply, but the solution to Eq. (\[15\]) is more delicate than the “blind" integration that yields Eq. (\[21\]). Indeed, the coefficient of the first derivative is zero at the fixed points, which now lie entirely within the support of the probability distribution. Thus, the equation becomes singular. The method of variation of parameters for an equation of the type (\[15\]) leads to a solution which is a sum of the general solution of the homogeneous equation and a particular solution of the inhomogeneous equation, as in Eq. (\[17\]). The subtlety lies in the determination of the constant of integration $C$, which in the previous case was fixed simply by imposing periodic boundary conditions. In the vicinity of the stable fixed point $x_2$ this straightforward procedure leads to the dependence $P(x)\sim {|x-x_2|^{{k}/{|v'(x_2)|}-1}}$, which is continuous when ${k}/{|v'(x_2)|} >1$ and divergent but integrable for ${k}/{|v'(x_2)|}\leq 1$. This result causes no mathematical difficulty and is consistent with the physical intuition that probability near a stable fixed point should indeed build up, especially when the switching rate is low. At the unstable fixed point $x_1$, however, this procedure leads to an apparent non-integrable divergence, $P(x) \sim {|x-x_1|^{-{k}/{|v'(x_1)|}-1}}$, which is clearly unphysical and mathematically improper in view of the requirement of normalization. The fallacy lies in the assumption that a single constant $C$ is valid throughout the region $(0,L)$. In fact, the solution (\[17\]) is valid in the separate intervals $[0,x_1)$, $(x_1,x_2)$, and $(x_2,L/2]$, [*but not necessarily with the same constant of integration*]{} in all of them. Indeed, there is [*exactly one*]{} choice of this constant valid for both $[0,x_1)$ and $(x_1,x_2)$ such that the divergence at $x_1$ is removed, and another choice valid in the interval $(x_2,L/2]$ that ensures required continuity and periodicity. In other words, even though the general solution of the homogeneous equation always has a divergence, there exists exactly one solution to the full inhomogeneous equation (\[15\]) that has no divergence and is actually completely smooth at $x_1$. This solution is given by Eq. (\[17\]) with the choice $C=- H(x_1,0;0)$ in both intervals $[0,x_1)$ and $(x_1,x_2)$ [@elsewhere]. This choice insures that the coefficient of the divergent term vanishes at $x=x_1$. We conclude that for ${ x \in [0,x_2)}$, $$P(x)=\frac{J}{F}\left[1+\frac{v(x)}{\left|F^2-v^2(x)\right|} H(x,x_1;x)\right]. \label{28}$$ Note that $P(x)$ is now continuous at $x_1$, and that $\lim_{x \searrow x_1}P(x)=\lim_{x \nearrow x_1}P(x)={J}F^{-1}\left\{1-[2(k/|v'(x_1)|+1)]^{-1}\right\}$ is indeed finite. For $x\in (x_2,L/2]$, the result (\[17\]) for $p(x)$ applies again, but now the constant $C$ is determined by imposing the continuity of $p(x)$ at $x=L/2$. One finds: $$\begin{aligned} P(x)&=&\frac{J}{F}\left\{1+\frac{v(x)} {\left|F^2-v^2(x)\right|} \left[ H(x,L/2;x) \right. \right. \nonumber\\ && \left. \left. +G(0,L/2)H(x_1,0;x)\right]\right\}. \label{32}\end{aligned}$$ At the stable fixed point $x_2$, $P(x)$ has the behavior described earlier, i.e., it is continuous for $k/|v_2'(x_2)|>1$ and divergent but integrable for $k/|v'(x_2)| \leq 1$. The values of the flux $J$ and of the average velocity follow from these results by imposing the normalization of $P(x)$: $$\begin{aligned} \langle \dot{x}\rangle &=& F\left\{ 1+\frac{2}{L} \int_0^{x_2} dx \frac{v(x)}{\left|F^2-v^2(x)\right|} H(x,x_1;x)\right. \nonumber\\ &&\left. + \frac{2}{L}\int_{x_2}^{L/2} dx \frac{v(x)}{\left|F^2-v^2(x)\right|} \left[ H(x,L/2;x) \right. \right. \nonumber\\ && \left. \left. +G(0,L/2)H(x,0;x)\right]\right\}^{-1}. \label{eq34}\end{aligned}$$ This is our main new result. Note that the above procedure can be repeated straightforwardly but tediously for more complicated cases involving several stable and unstable fixed points. ![ Probability density $P(x)$ vs $x/L$ for the parameter values $f=0.5$, $\Gamma=0.4$, and $\alpha=1.0$. Histogram: simulation results. Curve: exact theory. []{data-label="fig2"}](fig2.eps){width="2.8in"} To illustrate our findings with explicit results, we turn to a particular case of a piecewise linear profile: $$v(x)=\left\{ \begin{array}{ll} v_0, &\mbox{for}\,\,\, x\,\in [0,\,L/2\,-\,2\,l) \\ v_0\,\left(\displaystyle \frac{L}{2\,l}\,-\,1\,-\, \displaystyle \frac{x}{l}\right), & \mbox{for}\,\,\,x\,\in\,[L/2\,-\,2\,l,\,L/2) \\ -v(x\,-L/2)\,\,, &\mbox{for}\,\,\, x\, \in [L/2,\,L) \end{array} \right. \label{eq35}$$ with $l\leq L/4$ and, of course, the periodicity condition $v(x+L)=v(x)$. It is convenient to introduce the following dimensionless variables: $$f=F/v_0 , \qquad \alpha=lk/v_0 , \qquad \Gamma=4l/L. \label{36}$$ ![ Mean velocity as a function of the applied force for various values of $\alpha$. Note the hypersensitive response. []{data-label="fig3"}](fig3.eps){width="2.8in"} In this case the function $H(x,0;0)/v_0\equiv T(x)$ becomes $T(x)=0$ for $x\in [0,L/2-2l)$ and, for $x\in [L/2-2l, L/2)$, $$\begin{aligned} T(x)&=& \left|\frac{1+f}{1-f}\right|^{\alpha}\exp\left[-\frac{4\alpha f(1-\Gamma)} {(1-f^2)\Gamma}\right] \nonumber\\ &\times& \int_1^{L/2l-1-x/l}\mbox{sgn}(f^2-s^2) \left|\frac{f-s}{f+s}\right|^{\alpha} ds. \label{eq38}\end{aligned}$$ Explicit [*exact*]{} results for the probability densities and for the resultant average velocity for all values of $f$ are available and will be detailed elsewhere [@elsewhere]. Here we exhibit only some of these results for the new case $0<f<1$. Figure \[fig2\] depicts a typical probability density $P(x)$ vs $x/L$ that clearly shows the agreement between the exact theoretical results and simulations. For the average velocity we find $$\begin{aligned} \langle \dot{x} \rangle &=& F\left\{1-\frac{\Gamma T(x_1)}{4\alpha f} \left[\exp\left(\frac{4\alpha f(1-\Gamma)}{\Gamma(1-f^2)}\right) -1\right] \right. \nonumber\\ &&\hspace{-1cm}+\,\frac{\Gamma(1-f)^{\alpha}}{2(1+f)^{\alpha}} \exp\left(\frac{4\alpha f(1-\Gamma)}{\Gamma(1-f^2)}\right)\times \nonumber\\ &&\hspace{-1cm}\times\left\{ \int_{-1}^{1}dt\, \frac{t\,\left[T[L/2-l(t+1)]-T(x_1) \right]} {\left|t-f\right|^{1+\alpha}\left|t+f\right|^{1-\alpha}}\right. \nonumber\\ &&\hspace{-1cm}+ \left[T(x_1) \left(1+ \displaystyle\left(\frac{1+f}{1-f}\right)^{2\alpha} \exp\left(-\frac{4\alpha f(1-\Gamma)}{\Gamma(1-f^2)}\right) \right)\right.\nonumber\\ &&\left.\left.\left.\hspace{-1cm}-T(L/2)\displaystyle\right] \int_{-1}^{-f}dt\, \frac{t} {\left|t-f\right|^{1+\alpha}\left|t+f\right|^{1-\alpha}} \right\}\right\}^{-1 }. \label{velocity}\end{aligned}$$ The above integrals can be evaluated explicitly for specific values of $\alpha$, in particular, for $\alpha=1/2$, $1$, and $2$ [@elsewhere]. The analytic and simulation results for the variation of the average velocity with $f$ (again with full agreement) are shown in Fig. \[fig3\]. In the limit of slow switching rate, that is, in the adiabatic regime $\alpha \rightarrow 0$, one recovers the region of hypersensitive response discussed in [@ginzburg], namely, $\langle\dot{x}\rangle\approx 2v_0\alpha/\Gamma$. The physics of this result is explained as follows: The stable and unstable fixed points of the dynamics, which coincide in the absence of forcing, are shifted apart by an amount of order of $ l F /v_0$ for a small force. Upon each switch of the dichotomous process and for sufficiently slow switching rate, the particle will glide down to the next stable fixed point, crossing the location of the unstable fixed point of the alternate dynamics, cf. Fig. \[fig1\]. As the average time between switches is $k^{-1}$, and the distance covered is $L/2$, the mean velocity is just $\langle \dot{x} \rangle = Lk/2$, independently of $F$ and $v_0$. The typical time $\tau$ for a particle to escape the region around the fixed point is determined by the relation $(lF/v_0)\exp(v_0 \tau/l) \approx l$. The crucial observation for hypersensitive response is that the necessary condition, $\tau << k^{-1}$ or $\alpha << {(-lnf)}^{-1}$, can be satisfied by very small forces for moderately small $\alpha$ because of the logarithmic dependence on $f$, see Fig. \[fig3\]. In this figure one also observes the region of “normal" (i.e., linear) response for higher forcing or frequency and, more relevant to our preoccupation here, the strongly nonlinear dependence at very low forcing. In fact in the limit $f \rightarrow 0$, one finds from Eq. (\[velocity\]) that $\langle\dot{x}\rangle/v_0\approx [2\alpha \Gamma (ln f)^2]^{-1}$. In other words, hypersensitive response is very pronounced in this region, with the velocity picking up with an infinite derivative at $f=0$. In conclusion, the procedure presented here has resolved all technical problems related to steady state dichotomous dynamics, making possible the analytic description of cases involving the crossing of fixed points in the asymptotic regime. This work was partially supported by the National Science Foundation under grant Nos. PHY-9970699 and DMS-0079478. [100]{} N. G. Van Kampen, [*Stochastic Processes in Physics and Chemistry*]{}, North-Holland, Amsterdam, 1992. A. Fulinski, Phys. Rev. E [**50**]{}, 2668 (1994). Klyatskin, V. I., Radiophys. Quantum Electron. [**20**]{}, 382 (1977). W. Horsthemke and R. Lefever, [*Noise Induced Transitions*]{} (Springer-Verlag, Berlin, 1984). J. M. Sancho, Phys. Rev. A [**31**]{}, 3523 (1985); J. Masoliver, K. Lindenberg, and B. J. West, Phys. Rev. A [**33**]{}, 2177 (1986); [*ibid*]{} [**34**]{}, 1481 (1986); [*ibid*]{} [**34**]{}, 2351 (1986); M. A. Rodriguez and L. Pesquera, Phys. Rev. A [**34**]{}, 4532 (1986); C. R. Doering, Phys. Rev. A [**35**]{}, 3166 (1987); V. Balakrishnan, C. Van den Broeck, and P. Hänggi, Phys. Rev. A [**38**]{}, 4213 (1988); U. Behn and K. Schiele, Z. Phys. B [**77**]{}, 485 (1989); M. Kuś, E. Wajnryb and K. Wódkiewicz, Phys. Rev. A [**43**]{}, 4167 (1991); J. Olarrea, J. M. R. Parrondo, and F. J. de la Rubia, J. Stat. Phys. [**79**]{}, 669 (1995); [*ibid.*]{} [**79**]{}, 683 (1995); C. Doering, W. Horsthemke, and J. Riordan, Phys. Rev. Lett. [**72**]{}, 2984 (1994). M. M. Millonas and D. R. Chialvo, Phys. Rev. E [**53**]{}, 2239 (1996); V. Berdichevsky and M. Gitterman, Phys. Rev. E [**56**]{}, 6340 (1997); S. H. Park, S. Kim, and C. S. Ryu, Physics Lett. A [**225**]{}, 245 (1997); I. Bena, M. Copelli, and C. Van den Broeck, J. Stat. Phys. [**101**]{}, 415 (2000). V. Balakrishnan, C. Van den Broeck, and I. Bena, Stochastics and Dynamics [**1**]{}, 537 (2001). V. Berdichevsky and M. Gitterman, Physica A [**249**]{}, 88 (1998); S. L. Ginzburg and M. A. Pustovoit, Phys. Lett. A [**291**]{}, 77 (2001); [*ibid*]{}, Phys. Rev. Lett. [**80**]{}, 4840 (1998); [*ibid*]{}, J. Exp. Th. Phys. [**89**]{}, 80 (1999); [*ibid*]{}, Eur. Phys. J. B [**19**]{}, 101 (2001). I. Bena et al., in preparation.
{ "pile_set_name": "ArXiv" }
--- abstract: | A remarkably small scatter in the N/O ratios for the HII regions in low-metallicity blue compact galaxies (BCG) has been found recently. It lead to the conclusion that N and O are both produced by massive stars. Conversely, the N/Si ratios in damped Ly$\alpha$ absorbers (DLA) show a large scatter. This result provides support for the time-delay model of nitrogen production in intermediate-mass stars, nitrogen production in massive stars being not required. The goal of this study is to test whether these observational data are compatible with each other and with the existing ideas on the chemical evolution of galaxies. We find that it is possible to reconcile the constancy of N/O ratios in low-metallicity BCGs and the scatter of N/Si ratios in DLAs under the following three assumptions: [*1*]{}) a significant part of nitrogen is produced by intermediate-mass stars, [*2*]{}) star formation in BCGs and DLAs occurs in bursts separated by quiescent periods, [*3*]{}) the previous star formation events are responsible for measured heavy element abundances of HII regions in BCGs. Since the reality of low N/Si ratios in DLAs is not beyond question, the possibility that the nitrogen in low-metallicity BCGs has been produced by massive stars cannot be rejected without additional observations. author: - 'L.S. Pilyugin' date: 'Received ; accepted ' subtitle: 'Blue compact galaxies versus damped Ly$\alpha$ absorbers' title: 'On the origin of nitrogen in low-metallicity galaxies:' --- Introduction ============ The theoretical stellar yields of nitrogen are rather uncertain. It has been found (Renzini & Voli 1981; Marigo et al 1996,1998; van den Hoek & Groenewegen 1997) that nitrogen production takes place in the asymptotic giant branch phase of the evolution of intermediate-mass stars, but the mass range of the nitrogen-producing stars and the predicted amount of freshly produced nitrogen depend on poorly known parameters. The theoretical yields of nitrogen by massive stars are essentially unknown. Therefore, an “empirical” approach (analysis of the relation of N/O with O/H) has been widely used to study of the origin of nitrogen (Edmunds & Pagel 1978; Lequeux et al 1979; Matteucci & Tosi 1985; Garnett 1990; Pilyugin 1992, 1993; Vila-Costas & Edmunds 1993; Marconi et al 1994; among others). Pagel (1985) called attention to the large scatter in N/O at fixed O/H in low-metallicity dwarf galaxies. Given the time delay between the injection of nitrogen by intermediate-mass stars and that of oxygen by shorter lived massive stars (the time – delay hypothesis: Edmunds & Pagel 1978) and the hypothesis of self-enrichment of star formation regions (Kunth & Sargent 1986), models for the chemical evolution of dwarf galaxies reproducing the observed scatter of N/O have been constructed (Garnett 1990, Pilyugin 1992, Marconi et al 1994). A significant part (if not all) of nitrogen was assumed in these models to be produced by intermediate-mass stars. The HII regions with high N/O abundance ratios have been considered to be in early stages of star formation (before self-enrichment in oxygen occured), reflecting the chemical composition of the whole galaxy. The HII regions in advanced stages of the star formation bursts would have small N/O ratios due to self-enrichment in oxygen. These HII regions would then reflect the local chemical composition, and the scatter in their N/O ratios would be caused by different degrees of temporal decrease of N/O ratio within the HII regions. During the interburst period the nitrogen is ejected by the intermediate-mass stars, and the matter ejected by massive and intermediate-mass stars is supposed to be well mixed throughout the whole galaxy. Some scatter in global N/O ratios among dwarf galaxies can be caused by enriched galactic winds (Pilyugin 1993, 1994; Marconi et al 1994). In this framework, the chemical composition of low-metallicity dwarf galaxies was expected to be characterised by high N/O ratios with a relatively small dispersion, and the large observed dispersion in the N/O ratios is supposed to be due to a temporary N/O decrease inside HII regions in which the chemical composition is determined spectroscopically. In contrast with previous measurements, Thuan et al (1995) and Izotov & Thuan (1999) have found a remarkably small scatter in the N/O ratios of the HII regions in low-metallicity (with 12+logO/H $\leq$ 7.6) BCGs. Izotov & Thuan concluded that these galaxies are presently undergoing their first burst of star formation, and that nitrogen in these galaxies is produced by massive stars only. Since the intermediate and low-mass stars certainly do not make an appreciable contribution to oxygen production, the N/O ratio corresponding to the ejecta of massive stars and observed in low-metallicity BCGs is a lower limit for the global N/O ratios. This conclusion is in conflict with the fact that the nitrogen to $\alpha$-element abundance ratios measured in some DLAs are well below than the typlical value observed in low-metallicity BCGs (Lu et al 1998, and references therein). \[Since oxygen abundance measurements are not available for DLAs, $[N/S]$ or $[N/Si]$ ratios are considered instead of $[N/O]$. This is justified by the fact that there is no reason to believe that the relative abundances of O, S and Si which are all produced in Type II supernovae are different from solar in DLAs.\] However, as suggested by Izotov & Thuan (1999), the nitrogen to $\alpha$-element abundance ratios in DLAs can be significantly underestimated if the absorption lines originate in the HII instead of the HI gas. Nevertheless, the possibility of a truly low nitrogen abundances in some DLAs cannot be excluded without additional observations, and these DLAs with low nitrogen to $\alpha$-element abundance ratios do not confirm the idea that the N/O ratio in low-metallicity BCGs is a lower limit for the global N/O ratio in galaxies. The best way to find the lower limit of N/O ratio and hence the amount of nitrogen produced by massive stars would be to determine the N/O ratios in galactic halo stars. Unfortunately, at the present state their N/O ratios cannot be determined with a precision better than a factor 2 or 3. The only firm conclusion is that nitrogen has a strong primary component (Carbon et al 1987). The goal of this study is to test whether the constancy of the N/O ratios in low-metallicity BCGs and and the scatter of the N/Si ratios in DLAs are compatible with each other and with the existing ideas on the chemical evolution of galaxies. Possible interpretation of nitrogen abundances in BCGs and DLAs =============================================================== Here we will demonstrate that the constancy of the N/O ratios in low-metallicity BCGs and the scatter of the N/Si ratios in DLAs can be reconciled under the following three assumptions: [*1*]{}) a significant part of nitrogen is produced by intermediate-mass stars, [*2*]{}) star formation in BCGs and DLAs occurs in bursts separated by quiescent periods, [*3*]{}) the previous star formation events are responsible for heavy element abundances observed in the HII regions of the BCGs. As it was discussed in Introduction, there are no crucial arguments in favor of or against assumption [*1*]{}. Assumption [*2*]{} - that star formation in BCGs and DLAs occurs in bursts separated by quiescent periods - is commonly accepted in the case of BCGs after it was initially suggested by Searle & Sargent (1972). In order to reproduce the observed properties of low-metallicity BCGs, only a few (in some cases only one or two) star formation bursts during their life are required (Tosi 1994). Papaderos et al (1998) have found that the spectrophotometric properties of SBS 0335-052 can be accounted for by a stellar population not older than $\sim$ 100 Myr. The possibility of an underlying old (10 Gyr) stellar population with mass not exceeding $\sim$ 10 times that of young stellar population mass hawever cannot be definitely ruled out on the basis of the spectrophotometric properties. In other words, the time interval between possible previous and current star formation events can be as large as $\sim$ 10 Gyr. In the case of DLAs, assumption [*2*]{} seems to be also acceptable, despite the fact that DLAs do not constitute an homogeneous class of galaxies but belong to the wide variety of morphological types of galaxies (Le Brun et al 1997). Star formation in a given region in galaxies of any morphological type seems to be episodic. The DLA observations sample the general interstellar medium at random times along random lines of sight and may or may not see a region where a star formation event occured in a recent past. Assumption [*3*]{} - that previous star formation events are responsible for the observed heavy element abundances in BCGs - is equivalent to say that the element abundances of HII regions in BCGs are not yet polluted by the stars of the present star formation event, and that their abundances reflect the average N/O in the galaxy, which results from cumulative previous star formation. Martin (1996) has found that the current event of star formation in the most metal-poor known blue compact galaxy I Zw 18 started 15-27 Myr ago. The duration of current star formation burst in another extremely metal-poor blue compact galaxy SBS 0335-052 (Papaderos et al 1998) is also in excess of the lifetime of the most massive stars. Therefore, a selection effect in favor of observations of young HII regions in which the massive stars had not yet have time to explode as supernovae cannot be only reason why the HII regions are not observed as self-enriched. Massive stars in the current star formation burst have often had time to synthesize heavy elements and to eject them via stellar winds and supernova explosions into the surrounding interstellar gas. Kunth & Sargent (1986) suggest that the heavy elements produced in this way initially mix into H II region only, i.e. the giant H II regions are self-enriched. However, it is possible that the nucleosynthetic products of massive stars are in high stages of ionization and do not make appreciable contribution to the element abundance as derived from optical spectra (Kobulnicky & Skillman 1997, Kobulnicky 1999). Indeed, the oxygen abundance in SBS 0335-052 has been measured within the region of 3.6 kpc (Izotov et al 1997). There is a supershell of radius $\sim$ 380 pc. There is no difference in oxygen abundances inside and outside the supershell as it should be expected since $\sim$ 1500 supernovae are required to produce this supershell (Izotov et al 1997). Other star-forming galaxies, which are chemically homogeneous despite the presence of multiple massive star clusters, are reported by Kobulnicky & Skillman (1998). This can be considered as evidence that the nucleosynthetic products of massive stars in giant HII regions are hidden from optical spectroscopic searches because they are predominantly found in a hot, highly – ionized superbubble. It should be noted however that some fraction of supernova ejecta can mix with dense clouds changing their chemical composition. If such cloud survives and produces a subgroup of stars shortly, the star formation region will have sub-generations of stars with different chemical composition. This seems to be the case in the Orion star formation region (Cunha & Lambert 1994, Pilyugin & Edmunds 1996). With assumptions [*1*]{} and [*2*]{} the behaviour of the N/O ratio is described by models of type suggested by Garnett (1990), Pilyugin (1992), Marconi et al (1994). Figgure \[figure:8566f1\] illustrates the time behaviour of N/O in the interstellar medium of a galaxy in the case when all the nitrogen is produced by intermediate-mass stars (see also Fig.7 in Pilyugin 1992). Since only low-metallicity galaxies are considered here, the adopted total astration level in the models is small (less than 0.05), and the nitrogen yield is assumed to be independent on metallicity. The nitrogen yields by stars in the mass interval 3$\div$4M$_{\odot}$ were taken from Renzini and Voli (1981). As can be seen on Fig.\[figure:8566f1\], the N/O ratio increases for about 1 Gyr after the star formation burst and then remains constant. This is due to the assumption that nitrogen is not produced by stars with masses less than $\sim$ 3$M_{\odot}$ (Renzini and Voli 1981). If stars with masses less than $\sim$ 3$M_{\odot}$ also make some contribution to the nitrogen production (Marigo et al 1996, van den Hoek and Groenewegen 1997), this does not change appeciably the picture because the amount of nitrogen ejected decreases strongly with decreasing stellar mass (Renzini and Voli 1981, Marigo et al 1996, van den Hoek and Groenewegen 1997). Therefore, low-metallicity systems with a large time interval (more than $\sim$ 1 Gyr) between successive star formation events will have close values of N/O ratios before the current star formation event (Fig.\[figure:8566f1\]). With this behaviour of the N/O ratio and taking into account assumption [*3*]{}, the nitrogen abundances in BCGs and DLAs can be interpreted in the following way. The low-metallicity BCGs are systems with a small amount of old (with age $>$ 1 Gyr) underlying stellar population over which the current star formation burst is superposed; only the stars from the previous star formation event(s) are responsible for the observed chemical composition in the giant HII regions in these galaxies. The DLAs with low nitrogen to $\alpha$-element ratios correspond to systems probed less than 1 Gyr after the last local star formation event, but after a time sufficient for disappearance of the superbubble and mixing of the freshly produced heavy elements in the interstellar medium. Conversely, the DLAs with nitrogen to $\alpha$-element ratios close to that in low-metallicity BCGs correspond to systems in which the time interval after last star formation event is sufficiently large for intermediate-mass stars to have substantially enhanced the nitrogen to $\alpha$-element abundance ratios. Discussion and conclusions ========================== We have shown that the observed nitrogen abundances in both BCGs and DLAs can be reproduced if a significant part of nitrogen is produced by intermediate-mass stars. If one assumes instead that the nitrogen abundances measured in low-metallicity BCGs is produced by massive stars, the observational data for BCGs and DLAs are incompatible with each other. Can this be considered as a crucial argument in favor of dominant production of nitrogen by intermediate-mass stars? Since the low nitrogen to $\alpha$-element abundance ratios obtained in some DLAs is not beyond question, the possibility that nitrogen in low-metallicity BCGs is produced by massive stars cannot be excluded. The crucial argument in favor of dominant production of nitrogen by intermediate-mass stars would be the reliable proof of the existence of systems with low N/O ratios. Solid determinations of nitrogen abundances in damped Ly$\alpha$ absorbers can clarify this matter. The Zn/S abundance ratios in DLAs with low measured nitrogen to $\alpha$-element abundance ratios can also tell us something about the reality of low N/O ratios in these objects. The measured value of $[Zn/S]$ in DLA associated with QSO 0100-130 is close to solar ratio (Prochaska & Wolfe 1998, Lu et al 1998), indicating that Type I supernovae have already contributed to the zink abundance. The measured $[N/S]$ in this DLA is close to the $[N/O]$ ratio in low-metallicity BCGs. If values of Zn/S in DLAs with low N/O ratios are close to the solar ratio, this will be a strong argument in favor of an underestimation of the nitrogen to $\alpha$-element abundance ratios. The time scale for nitrogen enrichment is shorter than the time scale for iron enrichment, and a system with high Fe/O ratio should not have a low N/O ratio. If values of $[Zn/S]$ in DLAs with measured low N/O ratios are close to $[Fe/O]$ ratios in the galactic halo stars, this can be considered as an argument in favor of genuinely low nitrogen to $\alpha$-element abundance ratios in these DLAs. Of course, the best way to find the amount of nitrogen produced by massive stars would be an undisputable determination of N/O ratios in galactic halo stars. In summary: If a significant part of nitrogen is produced by intermediate-mass stars, it is possible to reconcile the observational data for BCGs and DLAs. If nitrogen is mainly produced by massive stars, the observational data for BCGs and DLAs are incompatible with each other. Since the low nitrogen to $\alpha$-element abundance ratios obtained in some DLAs are not beyond question, the possibility that nitrogen measured in low-metallicity BCGs is produced by massive stars cannot however be completely excluded. I would like to thank Drs. N.G.Guseva and Y.I.Izotov for fruitful discussions. I thank the referee, Prof. J.Lequeux, for helpful comments and suggestions which resulted in a better presentation of the work. This work was partly supported through INTAS grant 97-0033. This study has been done using the NASA’s Astrophysical Data Service. Carbon D.F., Barbuy B., Kraft R.P., Friel E.D., Suntzeff N.B., 1987, PASP 99, 335 Cunha K., Lambert D.L., 1994, ApJ 426, 170 Edmunds M.G., Pagel B.E.J., 1978, MNRAS 185, 77p Garnett D.R., 1990, ApJ 363, 142 Izotov Y.I., Lipovetsky V.A., Chaffee F.H., Foltz C.B., Guseva N.G., Kniazev A.Y., 1997, ApJ 476, 698 Izotov Y.I., Thuan T.X., 1999, ApJ, 511, 639 Kobulnicky H.A., 1999, astro-ph 9901260 Kobulnicky H.A., Skillman E.D., 1997, ApJ 489, 636 Kobulnicky H.A., Skillman E.D., 1998, ApJ 497, 601 Kunth D., Sargent W.L.W., 1986, ApJ 300, 496 Le Brun V., Bergeron J., Boisse P., Deharveng J.M., 1997, A&A 321, 733 Lequeux J., Peimbert M., Rayo J.F., Serrano A., Torres-Peimbert S., 1979, A&A 80, 1979 Lu L., Sargent W.L.W., Barlow T.A., 1998, ApJ 115, 55 Marconi G., Matteucci F., Tosi M., 1994, MNRAS 270, 35 Marigo P., Bressan A., Chiosi C., 1996, A&A 313, 545 Marigo P., Bressan A., Chiosi C., 1998, A&A 331, 564 Martin C.L., 1996, ApJ 465, 680 Matteucci F., Tosi M., 1985, MNRAS 217, 391 Pagel B.E.J., 1985, in Production and Distribution of CNO Elements, Eds. Danziger I.J., Matteucci F., Kjar K. (Garching: Europen Southern Observatory), p. 155 Papaderos P., Izotov Y.I., Fricke K.J., Thuan T.X., Guseva N.G., 1998, A&A 338, 43 Pilyugin L.S., 1992, A&A 260, 58 Pilyugin L.S., 1993, A&A 277, 42 Pilyugin L.S., 1994, A&A 287, 387 Pilyugin L.S., Edmunds M.G., 1996, A&A 313, 792 Prochaska J.X., Wolfe A.M., 1998, astro-ph/9810381 Renzini A., Voli M., 1981, A&A 94, 175 Searle L., Sargent W.L.W., 1972, ApJ 173, 25 Thuan T.X., Izotov Y.I., Lipovetsky V.A., 1995, ApJ 445, 108 Tosi M., 1994, in: Dwarf Galaxies, Eds. Meylan G., Prugniel P., (Garching: Europen Southern Observatory), p. 465 van den Hoek L.B., Groenewegen M.A.T., 1997, A&AS 123, 305 Vila-Costas M.B., Edmunds M.G., 1993, MNRAS 265, 199
{ "pile_set_name": "ArXiv" }
--- abstract: 'The International Linear Collider (ILC) has recently proven its technical maturity with the publication of a Technical Design Report, and there is a strong interest in Japan to host such a machine. We summarize key aspects of the Beyond the Standard Model physics case for the ILC in this contribution to the US High Energy Physics strategy process. On top of the strong guaranteed physics case in the detailed exploration of the recently discovered Higgs boson, the top quark and electroweak precision measurements, the ILC will offer unique opportunities which are complementary to the LHC program of the next decade. Many of these opportunities have connections to the Cosmic and Intensity Frontiers, which we comment on in detail. We illustrate the general picture with examples of how our world could turn out to be and what the ILC would contribute in these cases, with an emphasis on value-added beyond the LHC. These comprise examples from Supersymmetry including light Higgsinos, a comprehensive bottom-up coverage of NLSP-LSP combinations for slepton, squark, chargino and neutralino NLSP, a $\ttau$-coannihilation dark matter scenario and bilinear R-parity violation as explanation for neutrino masses and mixing, as well as generic WIMP searches and Little Higgs models as non-SUSY examples.' author: - | [*Howard Baer$^1$, Mikael Berggren$^2$, Jenny List$^2$, Mihoko M. Nojiri$^{3,4}$,*]{}\ [*Maxim Perelstein$^5$, Aaron Pierce$^6$, Werner Porod$^7$, Tomohiko Tanabe$^8$*]{}\ $^1$University of Oklahoma, Norman, OK 73019, USA\ $^2$DESY, Notkestra[ß]{}e 85, 22607 Hamburg, Germany\ $^3$Theory Center, KEK, Tsukuba, Ibaraki 305-0801, Japan\ $^4$Kavli IPMU, The University of Tokyo, Kashiwa, Chiba 277-8583, Japan\ $^5$Cornell University, Ithaca, NY 14850, USA\ $^6$University of Michigan, Ann Arbor, MI 48109, USA\ $^7$University of Würzburg, 97074 Würzburg, Germany\ $^8$ICEPP, The University of Tokyo, Bunkyo-ku, Tokyo 113-0033, Japan\ title: | Physics Case for the ILC Project:\ Perspective from Beyond the Standard Model --- Introduction {#sec:intro} ============ Connection to Cosmic Frontier {#sec:cosmo} ============================= Connection to Intensity Frontier {#sec:intensity} ================================ ILC Stories {#sec:stories} =========== While there is a compelling case that new structure is needed in the laws of physics at the $100-1000$ GeV scale, exactly what that structure might be is one of the great mysteries of our time. Theoretical guidance allows one to extend the Standard Model to include new ideas such as supersymmetry, dark matter, grand unification, see-saw neutrinos, extra dimensions, string theory, and many others. But whether, or how, any of these ideas would manifest themselves at the weak scale is open to speculation; the truth will only be gleaned by a program of detailed experimental tests. The ILC– with its uniquely clean and flexible experimental environment – is prepared for almost any possibility. In this white paper, we focus on scenarios in which new particles appear within kinematic reach of the ILC. Here, we present seven such scenarios – each a particular story of how nature might be constructed at the weak scale, and what role the ILC could play in revealing the new physics. Coming to terms with electroweak naturalness {#subsec:natsusy} -------------------------------------------- At the ILC, SUSY [*is*]{} simplified {#subsec:simplified} ------------------------------------ LHC and ILC complementarity: SUSY is complex! {#subsec:stc} --------------------------------------------- LHC and ILC complementarity: Electroweakinos {#subsec:ewkinos} --------------------------------------------- Bilinear R-Parity Violation: Neutrino Physics at Colliders {#subsec:bRPV} ----------------------------------------------------------- Production of dark matter in the laboratory? {#subsec:WIMPs} --------------------------------------------- Little Higgs: an alternative to SUSY {#subsec:littlehiggs} ------------------------------------ Summary {#sec:summary} =======
{ "pile_set_name": "ArXiv" }
--- abstract: | The *Web Geometry Laboratory* ([[*WGL*]{}]{}) project’s goal is to build an adaptive and collaborative blended-learning Web-environment for geometry. In its current version (1.0) the WGL is already a collaborative blended-learning Web-environment integrating a dynamic geometry system (DGS) and having some adaptive features. All the base features needed to implement the adaptive module and to allow the integration of a geometry automated theorem prover (GATP) are also already implemented. The actual testing of the WGL platform by high-school teachers is underway and a field-test with high-school students is being prepared. The adaptive module and the GATP integration will be the next steps of this project. author: - Pedro Quaresma - Vanda Santos - 'Seifeddine Bouallegue[^1]' title: 'The Web Geometry Laboratory Project[^2]' --- Introduction {#sec:introduction} ============ The use of intelligent computational tools in a learning environment can greatly enhance its dynamic, adaptive and collaborative features. It could also extend the learning environment from the classroom to outside of the fixed walls of the school. To build an adaptive and collaborative blended-learning environment for geometry, we claim that we should integrate dynamic geometry systems (DGSs), geometry automated theorem provers (GATPs) and repositories of geometric problems (RGPs) in a Web system capable of individualised access and asynchronous and synchronous interactions. A system with that level of integration will allow building an environment where each student can have a broad experimental learning platform, but with a strong formal support. In the next paragraphs we will briefly explain what do we mean by each of these features and how the Web Geometry Laboratory (WGL) system cope, or will cope, with that. #### A blended-learning environment is a mixing of different learning environments, combining traditional face-to-face classroom (synchronous) methods with more modern computer-mediated (asynchronous) activities. A Web-environment is appropriate for both situations (see Figure \[fig:globalnet\]). #### An adaptive environment is an environment that is able to adapt its behaviour to individual users based on information acquired about its user(s) and its environment and also, an important feature in a learning environment, to adapt the learning path to the different users needs. In the [[*WGL*]{}]{} project this will be realised through the registration of the geometric information of the different actions made by the users and through the analysis of those interactions [@Quaresma2012b]. #### A collaborative environment is an environment that allows the knowledge to emerge and appear through the interaction between its users. In WGL this is allowed by the integration of a DGS and by the users/groups/constructions relationships. Using a DGS, the constructions are made from free objects and constructed objects using a finite set of property preserving manipulations. These property preserving manipulations allow the development of “visual proofs”, these are not formal proofs. The integration in the WGL of a GATP will give its users the possibility to reason about a given DGS construction, this is an actual formal proof, eventually in a readable format. They can be also used to test the soundness of the constructions made by a DGS [@Janicic2010; @Janicic2007]. As said above to have an adaptive and collaborative blended-learning environment for geometry we should integrate intelligent geometric tools in a Web system capable of asynchronous and synchronous interactions. This integration is still to be done, there are already many excellent DGSs [@wikipedia2013], some of them have some sort of integration with GATPs, others with RGP [@Janicic2010; @Quaresma06d]. Some attempts to integrate these tools in a learning management system (LMS) have already been done, but, as far as we know, all these integrations are only partial integrations. A learning environment where all these tools are integrated and can be used in a fruitful fashion does not exist yet [@Santos2013]. The Web Geometry Laboratory Framework {#sec:CollaborativeEnvironmentForGeometry} ===================================== [r]{}[4cm]{} ![image](WGLglobalNet.pdf){width="4cm"} A class session using [[*WGL*]{}]{} is understood as a Web laboratory where all the students (eventually in small groups) and the professor will have a computer running [[*WGL*]{}]{} clients. Also needed is a [[*WGL*]{}]{} server, e.g. in a school Web-server (see Figure \[fig:globalnet\]). The [[*WGL*]{}]{} server is the place where all the information is kept: the login information; the group definition; the geometric constructions of each user; the users activity registry; etc. In the [[*WGL*]{}]{} server is also kept the DGS applet and the GATP will also execute there. Each client will have an instance of the DGS applet, using the server to all the needed information exchange. After installing a [[*WGL*]{}]{} server the administrator of the system should define all the teachers that will be using the system. The teachers will be privileged users in the sense that they will be capable of define other users, their students. In the beginning of each school year the teachers will define all his/her students as regular users of the [[*WGL*]{}]{}. The teacher may also define groups of users (students), these groups can be define at any given time, e.g. for a specific class, and it will be within this groups that the collaboration between its members will be possible. The definition of the groups and the membership relation between groups and its members will be the responsibility of the teachers that could create groups, delete groups and/or modify the membership relation at any given time. Each user will have a “space” in the server where he/she can keep all the geometric construction that he/she produces. Each user will have full control over this personal scrapbook, having the possibility of saving, modifying and deleting each and every construction he/she produces using the DGS applet. To allow the collaborative work a permissions system was implemented. This system is similar to the “traditional Unix permissions” system. The users will own the geometric construction defining the [**r**]{}eading, [**w**]{}riting and [**v**]{}isibility permissions (rwv) per geometric construction. The users to groups and the constructions to groups relationships can be established in such a way that the collaborative working, group-wise, is possible. By default, the teacher will own all the groups he/she had created granting him/her, in this way, access to all the constructions made by the students. The default setting will be [rwvr-v—]{}, meaning that the creator (owner) will have all the permissions, other users belonging to his/her groups will have “read” access and all the others users will have none. At any given moment he/she can download (read) the construction into the DGS, modify it and, eventually, upload the modified version into the database. The collaborative module of [[*WGL*]{}]{} distinguishes students having the lock over the group construction from those without the lock. The students with the lock will have a full-fledged DGS applet, and they will be working with the group construction (see Figure \[fig:withlock\]). The students without the lock will have also the two DGS applets, but the construction in the “group shared construction” one is a synchronised version of the one being developed by the student with the lock, and a full-fledged version that can be used to develop his/her own efforts. A text-chat will be available to exchange information between group members. The teacher could always participate in this efforts having for that purpose an interface where he/she can follow the students and groups activities. The [[*WGL*]{}]{} collaborative features are thought mostly for a blended-learning setting, that is, a classroom/laboratory where the computer-mediated activities are combined with a face-to-face classroom interaction. Nevertheless given the fact that the [[*WGL*]{}]{} is a Web application the collaborative work can extend itself to the outside of the classroom and be used to develop collaborative work at home, e.g. solving a given homework. In this setting the only drawback it will be a slow connection to the [[*WGL*]{}]{} server. We estimate that a normal bandwidth ($\geq 20$Mb) will be enough. The [[*WGL*]{}]{} as a Web client/server application; the database (to keep: constructions; users information; constructions; permissions; user’s logs); the DGS applet; the GATP and the synchronous and asynchronous interaction are all implemented using free cross-platform software, namely PHP, Javascript, Java, AJAX, JQuery and MySQL, and also Web-standards like XHTML, CSS style-sheets and XML. The WGL is a internationalised tool (i18n/l10n) with already translations for Portuguese and Serbian, apart from default support for English. All this will allow to build a collaborative learning environment where the capabilities of tools such as the DGS and the GATP can be used in a more rich setting that in an isolated environment where (eventually) every students could have a computer with a DGS but where the communication between them would be non-existent. The exchange of text, oral and geometric information between members of a group will enrich the learning environment. Learning environments supported by computer are seen as an important means for distance education. The DGS are also important in classroom environments, as a much enhanced substitute for the ruler and compass physical instruments, allowing the development of experiments, stimulating learning by experience. There are several DGS available, such as: [*GeoGebra*]{}, [*Cinderella*]{}, [*Geometric Supposer*]{}, [*GeometerSketchpad*]{}, [*CaR*]{}, [*Cabri*]{}, [*GCLC*]{} but none of then defines a Web learning environment with adaptive and collaborative features [@Santos2013]. The program [ *Tabul[æ]{}*]{} is a DGS with Web access and with collaborative features. This system is close to [[*WGL*]{}]{}, the permissions system and the fact that the DGS is not “hardwired” to the system but it is an external tool incorporated into the system, are features that distinguish positively [[*WGL*]{}]{} from [*Tabul[æ]{}*]{}. The adaptive features, the connection to the GATP and the internationalisation/localisation are also features missing in [ *Tabul[æ]{}*]{} [@Santos2013]. Conclusions and Future Work {#sec:ConclusionsFutureWork} =========================== When we consider a computer system for an educational setting in geometry, we feel that a collaborative, adaptive blended-learning environment with DGS and GATP integration is the most interesting solution. That leads to a Web system capable of being used in the classroom but also outside the classroom, with collaborative and adaptive features and with a DGS and GATPs integrated. The [[*WGL*]{}]{} system is a work-on-progress system. It is a client/server modular system incorporating a DGS, some adaptive features, i.e., the individualised scrapbook where all the users can keep their own constructions and with a collaborative module. Given the fact that it is a client/server system the incorporation of a GATP (on the server) it will not be difficult. One of the authors has already experience on that type of integration [@Janicic2007; @Quaresma2011; @Quaresma06d]. A first case study, involving two high-schools (in the North and Center of Portugal) three classes, two teachers and 44 students and focusing in the use of [[*WGL*]{}]{} in a classroom, is already being prepared and it will be implemented in the spring term of 2013. The next task will be the adaptive module, the logging of all the steps made by students and teacher and the construction of student’s profiles on top of that. The last task will be the integration of the GATP in the [[*WGL*]{}]{}. We hope that at the end the [[*WGL*]{}]{} can became an excellent learning environment for geometry. A prototype of the [[*WGL*]{}]{} system is available at <http://hilbert.mat.uc.pt/WebGeometryLab/>. You can enter as “anonymous/anonymous”, a student-level user, or as “cicm2013/cicm”, a teacher-level user. [1]{} Predrag Janičić, Julien Narboux, and Pedro Quaresma. The [A]{}rea [M]{}ethod: a recapitulation. , 48(4):489–532, 2012. Predrag Janičić and Pedro Quaresma. Automatic verification of regular constructions in dynamic geometry systems. In Francisco Botana and Tomás Recio, editors, [*ADG2006*]{}, volume 4869 of [*LNAI*]{}, pages 39–51. Springer, 2007. Pedro Quaresma. housands of [G]{}eometric problems for geometric [T]{}heorem [P]{}rovers ([TGTP]{}). In Pascal Schreck, Julien Narboux, and J[ü]{}rgen Richter-Gebert, editors, [*ADG2010*]{}, volume 6877 of [ *LNAI*]{}, pages 169–181. Springer, 2011. Pedro Quaresma and Yannis Haralambous. Geometry Constructions Recognition by the Use of Semantic Graphs. In [*Atas da XVIII Confer[ê]{}ncia Portuguesa de Reconhecimento de Padr[õ]{}es, [R]{}ec[P]{}ad 2012*]{}, Tipografia Damasceno, Coimbra, 2012. Pedro Quaresma and Predrag Janičić. Integrating dynamic geometry software, deduction systems, and theorem repositories. In Jonathan [M. Borwein]{} and William [M. Farmer]{}, editors, [ *MKM2006*]{}, volume 4108 of [*LNAI*]{}, pages 280–294. Springer, 2006. Vanda Santos and Pedro Quaresma. Collaborative aspects of the [WGL]{} project. , 2013. (to appear). Wikipedia. List of interactive geometry software. <http://en.wikipedia.org/wiki/List_of_interactive_geometry_software>, April 2013. [^1]: IAESTE traineeship PT/2012/71. [^2]: The final publication is available at http://link.springer.com.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We describe how substantial domain-independent language-processing systems for French and Spanish were quickly developed by manually adapting an existing English-language system, the SRI Core Language Engine. We explain the adaptation process in detail, and argue that it provides a fairly general recipe for converting a grammar-based system for English into a corresponding one for a Romance language.' author: - | Manny Rayner and David Carter\ SRI International\ Suite 23, Millers Yard\ Cambridge CB2 1RQ\ United Kingdom\ [{manny, dmc}@cam.sri.com]{} Pierrette Bouillon\ ISSCO, University of Geneva\ 54, route des Acacias\ 1227 Geneva\ Switzerland\ [[email protected]]{} title: Adapting the Core Language Engine to French and Spanish --- Introduction ============ In this paper, we will describe how substantial domain-independent language-processing systems for French and Spanish were quickly developed by manually adapting an existing English system, the SRI Core Language Engine. The resulting systems have been integrated as components in the Spoken Language Translator (SLT; [@SLT-HLT; @SLT-report]). The English to French version of SLT [@RaynerBouillon:95] is of a standard comparable to the original English to Swedish version. The syntactic rule-set for French covers nearly all the basic constructions of the language, including the following: declarative, interrogative and imperative clauses; formation of YN and WH-questions using inversion, complex inversion and “est-ce que”; clitic pronouns; adverbial modification; negation; nominal and verbal PPs; complements to “être” and “il y a”; relative clauses, including those with “dont”; partitives, including use of “en”; passives; pre- and post-nominal adjectival modification, including comparative and superlative; code expressions; sentential complements and embedded questions; complex determiners; numerical expressions; date and time expressions; conjunction of most major constituents; and a wide variety of verb types, including modals and reflexives. There is a good treatment of inflectional morphology which includes all major paradigms. The coverage of the Spanish grammar is comparable in scope, though slightly less extensive. The French and Spanish versions of the CLE are both “reversible”, and can be used for either analysis or generation. We will describe the adaptation process in detail, and argue that it provides a fairly general recipe for converting a grammar-based system for English into a corresponding one for a Romance language. Due to space limitations, and since it is rather the better of the two, we will concentrate on the French version. Examples will be taken from the Air Travel Planning (ATIS) domain used in the current SLT prototype. The rest of the paper is organized as follows. Section \[Section:CLE\] gives an overview of the CLE, focussing on the aspects relevant to this paper. Section \[Section:Morphology\] describes the French morphology rules. Sections \[Section:French\] and \[Section:Spanish\] describe the French and Spanish grammars. Section \[Section:Conclusions\] concludes. Overview of the Core Language Engine {#Section:CLE} ==================================== The CLE is a general language-processing system, which has been developed by SRI International in a series of projects starting in 1986. The original system was for English only. A Swedish version [@CLE §14.2] was developed in a collaboration with the Swedish Institute of Computer Science; the French and Spanish versions described here were developed in collaborations with ISSCO and the University of Seville respectively. The CLE is extensively described elsewhere [@CLE; @CLARE-report; @SLT-report], so this section will only give the minimum background necessary to understand the remainder of the paper. The basic functionality offered by the CLE is two-way translation between surface form and a representation in terms of a logic-based formalism called QLF [@AlshawiCrouch:92]. The modules comprising a version of the CLE for a given language can be divided into three groups, which we refer to as “code”, “rules” and “preferences”. The “code” modules constitute the language-independent compilers and interpreters that make up the basic processing engine; the other two types of module between them constitute a declarative description of the language. The “rules” contain domain-independent lexico-grammatical information for the language in question; they encode a relationship between surface strings and QLF representations. Thus for any given surface string, the rules define a set of possible QLF representations of that string. Conversely, given a well-formed QLF representation, the rules can be used to produce a set of possible surface-form realisations of the QLF. The code modules support compilation of the rules into forms that allow fast processing in both directions: surface-form $\rightarrow$ QLF (analysis) and QLF $\rightarrow$ surface-form (generation). The relationship between surface form and QLF is in general many-to-many. “Preference” modules contain data in the form of statistically learned distributional facts, based on analysis of domain corpora [@AlshawiCarter:94; @RaynerBouillon:95]). Using this extra information, the system can distinguish between plausible and implausible applications of the rules with a fairly high degree of accuracy. In particular, the preference information makes it possible in a given application domain to select the intended readings of ambiguous utterances. We will not consider the preference modules further here, as the statistical training procedures are completely language-independent and invisible to the developer. The non-trivial problems involved in adapting the CLE to a new language arise in connection with the “rule” modules, which we will now consider in more detail. These fall into the following main categories: Lexicon: : The largest single set of rules is that which comprises the lexicon. This is in fact divided up into three subsets: a function-word lexicon; a set of macros specifying generic lexical entries for common types of content word (e.g. “count noun”, “intransitive verb”, etc.); and a content word lexicon which defines lexical entries for other words in terms of the generic macros. Morphology: : This set of rules defines the inflectional morphology of the language, allowing analysis of words in terms of stems and inflections. The morphology rules are described further in Section \[Section:Morphology\]. Syntax: : Syntax rules are written in a unification-based feature-grammar formalism. The style of the CLE grammar is loosely based on GPSG; detailed descriptions of the CLE grammar for English are available in [@CLE-grammar-chapter] and [@SLT-grammar-chapter]. Semantics: : The CLE grammar is “sign-based” [@CLE-semantics-chapter]; each syntax rule is coupled with one or more semantic counterpart, which defines the piece of QLF form produced by that rule. QLF representations are built up compositionally using unification only. Reference resolution and scoping: : Further sets of rules can be used to convert QLF representations into representations in full first-order logic. This phase of processing is required, for example, when using the CLE for database query applications [@AlshawiCrouch:92; @Rayner:93] As noted in [@CLE §14.2.2] the effort involved in adapting a set of rule modules to a new language depends on how directly they refer to surface form; unsurprisingly, modules defining surface phenomena are the ones which require most work. When adapting the system to French and Spanish, the problems arose almost exclusively in connection with morphology and syntax rules. Other parts of the English system were adapted with little effort. In particular, the semantic rule-sets for English could be used for the new languages with only minimal changes. The following two sections describe in detail the issues pertaining to the morphology and syntax rule-sets respectively. Morphology and spelling {#Section:Morphology} ======================= In order to handle the more complex inflectional morphology of Romance and other European languages, a morphological processor based on feature-augmented two-level morphology was developed [@Carter:95]. This allows the complex spelling changes occurring in these languages to be handled quickly in both analysis and generation. Compilation of the full sets of two-level rules describing spelling changes and of production rules describing legal affix combinations takes of the order of a minute, allowing changes to the rules to be debugged relatively easily. Further flexibility is gained by not requiring the lexicon to be present at compile time (contrast [@KaplanKay:94]); thus the lexicon can be incremented and tested without any recompilation being required. Two-level spelling rules were also used to describe the inter-word effects that are particularly common in French. The total number of rules required to describe inflectional morphology was around 75 for French and 50 for Spanish (inter-word rules being responsible for much of the difference). We concentrate here on the French phenomena, which are more complex. Intra-word spelling changes --------------------------- Intra-word spelling changes for French present several problems not encountered in English inflectional morphology. Some of these are technical in nature, and easily dealt with. In particular, French exhibits many multiple letter changes, e.g. “chameau+e” $\rightarrow$ [*chamelle*]{}, “peign+rai” $\rightarrow$ [*peindrai*]{}. For reasons explained in [@Carter:95], these must be handled by a separate rule for each letter that changes, rather than one for the whole changed substring. Also, some changes can be optional. For example, the “y” in verbs such as “payer” can remain the same or change to “i” before silent “e”: “pay+e” $\rightarrow$ either [*paye*]{} or [*paie*]{}. This phenomenon is rare or absent in English, but is handled easily by making the relevant spelling rule optional. Less trivial problems, however, arise from the fact that spelling changes in French generally cannot be predicted from the surface form of the word alone. This means the application of the rules must be controlled; we do this by specifying feature constraints, which must match between the rule and all morphemes it applies to. The following extended example describes our treatment of one of the most challenging cases. Nouns, adjectives and verbs ending in “-et” or “-el” can either double the “t” or “l” before a silent “e” or change the prefinal “e” to “è”: “cadet+e” $\rightarrow$ [*cadette*]{}, but “complet+e” $\rightarrow$ [*complète*]{}. The application of the spelling rules is therefore controlled by means of a feature [spelling\_type]{}, with value [double]{} in the first case and [change\_e\_è]{} in the second. This situation is further complicated by two facts. Firstly, the surface “èl” or “èt” of the verbs is ambiguous between a deep “el” or “et”, and “él” or “ét”. For example, we have [*achète*]{} $\leftarrow$ “achet+e”, but [*affrète*]{} $\leftarrow$ “affrét+e”. For this reason, we introduce a third value for [spelling\_type]{}: [change\_é\_è]{}. “Affrét” has thus the feature [spelling\_type=change\_é\_è]{}, “achet” [spelling\_type=change\_e\_è]{} and “appel” [spelling\_type=double]{}. Secondly, the “e” that begins future and conditional endings sometimes affects preceding letters as if it were silent, and sometimes as if it were not. For example, “appel+erai” $\rightarrow$ [*appellerai*]{}, doubling the “l” just as in “appel+e” $\rightarrow$ [*appelle*]{}, where the final “e” actually is silent. However, “céd+erai” $\rightarrow$ [*céderai*]{}, not \*[*cèderai*]{} as would be expected from the silent-e behaviour “céd+e” $\rightarrow$ [*cède*]{}. To make this distinction, we use a feature [muet]{} (“silent”) for specifying if the “e” in the suffix is silent, as “e” ([muet=y]{}), not silent, as “ez” ([muet=n]{}) or the “e” of the future or conditional tenses, for example “erai/erais” ([muet=fut\_cond\_e]{}). Then, we restrict the rule for doubling the consonant with the features [spelling\_type=double, muet=y$\vee$fut\_cond\_e]{}, and the one for “é” $\rightarrow$ “è” with the features [spelling\_type=change\_é\_è,muet=y]{}. Inter-word spelling changes --------------------------- In English, inter-word spelling changes occur only in the alternation between “a” and “an” before consonant and vowel sounds respectively. In French, such changes are far more widespread and can be complex. However, they can be handled by judiciously specifying contexts in two-level rules and, in a few cases, by postulating non-obvious underlying lexical items. Some important cases are: - The “e” in the function words “de”, “je”, “le”, “me”, “ne”, “que”, “se” and “te” is elided before (most) words starting in a vowel sound, except when the function word follows a hyphen: “le homme” $\rightarrow$ [*l’homme*]{}, “je ai” $\rightarrow$ [*j’ai*]{}, but “puis-je avoir” does not elide, so the elision rule specifies that the hyphen be absent from the context. “Ce” also elides when used as a pronoun (“ce est” $\rightarrow$ [*c’est*]{}, but when used as a determiner it takes the form “cet” before a vowel: [*cet homme*]{}. We therefore take the underlying form of the determiner to be “cet”, which [*loses*]{} its “t” when followed by a consonant-initial word (“cet soir” $\rightarrow$ [*ce soir*]{}). Numerals do not allow elision either: “le onze” does not become \*[*l’onze*]{}. We therefore treat the lexical form as being “\#onze”, where “\#” acts as an underlying consonant but is realised as a null. (Syntax plays a role here too: “le un” $\rightarrow$ [*l’un*]{} when is a determiner, but not when it is a numeral. Thus lexically we have “un” as determiner and “\#un” as numeral). - The very common preposition/article combinations “de”/“à” and “le”/“les”: “de le” $\rightarrow$ [*du*]{}, “à les” $\rightarrow$ [*aux*]{}, etc. These contractions span constituent boundaries (we view [*du vol*]{} as being syntactically \[PP de \[NP le vol\]\]) so need to be treated as spelling effects. Also, vowel elision takes precedence: “de le homme” $\rightarrow$ [*de l’homme*]{}, not \*[*du homme*]{}. - Hyphens between verbs and clitic pronouns are treated as lexical items in our grammar. They are realised as [*-t-*]{} when preceded by “a” or “e” and followed by “e”, “i” or “o”: “va - il” $\rightarrow$ [*va-t-il*]{}, but “vont - ils” $\rightarrow$ [*vont-ils*]{}. Hyphens joining nouns or names are treated as different lexical items not subject to this change: “les vols Atlanta - Indianapolis” does not involve introduction of “t”. French syntax {#Section:French} ============= When comparing the French and English grammars, there are two types of objects of immediate interest: [*syntax rules*]{} and [*features*]{}. Looking first at the rules themselves, about 80% of the French syntax rules are either identical with or very similar to the English counterparts from which they have been adapted. Of the remainder, some rules (e.gthose for date, time and number expressions) are different, but essentially too trivial to be worth describing in detail. Similar considerations apply to features. We will concentrate our exposition on the rules and features which are both significantly different, and possess non-trivial internal structure. Examining the grammar, we find that there are three large interesting groups of rules and features, describing three separate complexes of linguistic phenomena: question-formation, clitic pronouns and agreement. As we have argued previously [@RaynerBouillon:95], all of these are rigid and well-defined types of construction which occur in all genres of written and spoken French. It is thus both desirable and reasonable to attempt to encode them in terms of feature-based rules, rather than (for instance) expecting to derive them as statistical regularities in large corpora. In Sections \[Section:Clitics\], \[Section:Questions\] and \[Section:Agreement\], we describe how we handle these key problems. Question-formation {#Section:Questions} ------------------ We start this section by briefly reviewing the way in which question-formation is handled in the English CLE grammar. There are two main dimensions of classification: questions can be either WH- or Y-N; and they can use either the inverted or the uninverted word-order. Y-N questions must use the inverted word-order, but both word-orders are permissible for WH-questions. The phrase-structure rules analyse an inverted WH-question as constituting a fronted WH+ element followed by an inverted clause containing a gap element. The feature [inv]{} distinguishes inverted from uninverted clauses. The following examples illustrate the top-level structure of Y-N, unmoved WH- and moved WH-questions respectively. > \[Does he love Mary\]$_{S:[inv=y]}$ > \[Who loves Mary\]$_{S:[inv=n]}$ > \[\[Whom\]$_{NP}$ \[does he love \[\]$_{NP}$\]$_{S:[inv=y]}$\] The French rules for question formation are structurally fairly similar to the English ones. However, there are several crucial differences which mean that the constructions in the two languages often differ widely at the level of surface form. Two phenomena in particular stand out. Firstly, English only permits subject-verb inversion when the verb is an auxiliary, or a form of “have” or “be”; in contrast, French potentially allows subject-verb inversion with any verb. For this reason, English question-formation using auxiliary “do” lacks a corresponding construction in French. Secondly, French permits two other common question-formation constructions in addition to subject-verb inversion: prefacing the declarative version of the clause with the question particle “est-ce que”, and “complex inversion”, i.e. fronting the subject and inserting a dummy pronoun after the inverted verb. In certain circumstances, primarily if the subject is the pronoun “ça”, it is also possible to form a non-subject WH-question out of a fronted WH+ phrase followed by an uninverted clause containing an appropriate gap. We refer to this last possibility as “pseudo-inversion”. If the subject is a pronoun, only inversion and the “est-ce que” construction are allowed; if it is [*not*]{} a pronoun, only the “est-ce que” construction and complex inversion are valid. In addition, a subject pronoun following an inverted verb needs to be linked to it by a hyphen, which can be realised as a “-t-” (cf. Section \[Section:Morphology\]). Figure \[Figure:FrenchQ\] presents examples illustrating the main French question constructions. > [*Y-N, inversion:*]{}\ > Aime-t-il Marie?\ > \ > [*Y-N, “est-ce que”:*]{}\ > Est-ce que Jean aime Marie?\ > \ > [*Y-N, complex inversion:*]{}\ > Jean aime-t-il Marie?\ > \ > [*WH, subject question, no inversion:*]{}\ > Quel homme aime Marie?\ > \ > [*WH, inversion:*]{}\ > Quelle femme aime-t-il?\ > \ > [*WH, “est-ce que”:*]{}\ > Quelle femme est-ce que Jean aime?\ > \ > [*WH, complex inversion:*]{}\ > Quelle femme Jean aime-t-il?\ > \ > [*WH, pseudo-inversion:*]{}\ > Combien ça coûte? Modification of the English syntax rules to capture the basic requirements so far is quite simple. In our grammar, we added three extra rules to cover the “est-ce que”, complex-inversion and pseudo-inversion constructions: the second of these rules combines the complex-inverted verb with the following dummy pronoun to form a verb, in essence treating the dummy pronoun as a kind of verbal affix. A further rule deals with the hyphen linking an inverted verb with a following subject. With regard to the feature-set, the critical change involves the [inv]{} feature. In English, as we saw, this feature had two possible values, [y]{} and [n]{}. In French, the corresponding feature has five values: [inverted]{}, [uninverted]{}, [est\_ce\_que]{}, [complex]{} and [pseudo]{}, distinguishing clauses formed using the different question-formation constructions. (It is important to note, though, that the semantic representation of the clause is the same irrespective of its inversion-type). To enforce the restrictions concerning combinations of inversion-type and subject form, we also added a new clausal feature which distinguished clauses in which the subject is a pronoun. The attractive aspect of this treatment is that the remaining English rules used for question-formation can be retained more or less unchanged. In particular, the English semantic rules can still be used, and produce QLF representations with similar form. It would almost be true to claim that the above constituted our entire treatment of French question-formation. In practice, we have found it desirable to add a few more features to the grammar in order to block infelicitous combinations of the inversion rules with certain commonly occurring lexical items. It is possible that the effect of these features could be achieved equally well by statistical modelling or other means, but we describe them here for completeness: Restrictions on use of “est-ce que”: : Question-formation with “est-ce que” is strongly dispreferred when the main verb is a clause-final occurrence of “être”, or existential “avoir” (as in “il y a”). For example: > ?Quand est-ce que le prochain vol est?\ > ?Combien de vols est-ce qu’il y a? We enforce this by adding a suitable feature to the verb category. Fronting of “heavy” NPs: : Most languages prefer not to front “heavy” NPs, and this dispreference is particularly strong in French. We have consequently added an NP feature called [heavy]{}, which has the value [y]{} on NPs containing PP and VP post-modifiers. Thus for example generation of > Quels vols en partance de Dallas y a-t-il? is blocked, but the preferable > Quels vols y a-t-il en partance de Dallas? is permitted. Inverted subject NPs: : Occurrence of some pronouns (in particular “cela”, and “ça”) is strongly dispreferred in inverted subject position. A binary feature enforces this as a rule, for example blocking > Combien coûte ça pour aller à Boston? but instead permitting > Combien ça coûte pour aller à Boston? Clitics {#Section:Clitics} ------- The most difficult technical problems in adapting an English grammar to a Romance language are undoubtedly caused by clitic pronouns. In contrast to English, certain proform complements of verbs do not appear in their normal positions; instead, they occur adjacent to the main verb, and possibly joined to it by a hyphen. The position of the clitics in relation to the verb (pre- or post-verbal) is determined by the mood of the verb, and whether or not the verb is negated. If two or more clitics are affixed to the verb, their internal order is determined by their surface forms. Several attempts to account for the above and other data have previously been described in the literature e.g. [@Grimshaw:82; @BesGardent:89; @Estival:90; @MillerSag:95]; we have in particular been influenced by the last of these,. Although the underlying framework is very different from the HPSG formalism used by Miller and Sag, our basic idea is the same: to treat “clitic movement” by a mechanism similar to the one used to handle WH movement. More specifically, we introduce two sets of new rules. The first set handles the “surface” clitics. They define the structure of the verb/clitic complex, which we, like Estival, regard as a constituent of category V composed of a main verb and a “clitic-list”. A second set of “gap” rules defines empty constituents of category NP or PP, occurring at the notional “deep” positions occupied by the clitics. Thus, for example, on our account the constituent structure of “Est-ce que vous le voulez?” will be > \[Est-ce que \[vous$_{NP}$ \[le voulez\]$_V$ \[\]$_{NP}$\]$_S$\]$_S$ where the “gap” NP category represents the notional direct object of “voulez”, realised at surface level by the pre-verbal clitic “le”. To make this work, we add an extra feature, [clitics]{}, to all categories which can participate in clitic movement: in our grammar, these are V, VP, S, NP and PP. The [clitics]{} feature is used to link the cliticised V constituent and its associated clitic gap or gaps. We have found it convenient to define the value of the [clitics]{} feature to be a bundle of five separate sub-features, one for each of the five possible clitic-positions in French. Thus for instance the second-position clitics “le”, “la” and “les” are related to object-position clitic gaps through the second sub-feature of [clitics]{}; the fourth-position “y” clitic is related to its matching PP gap through the fourth sub-feature; and so on. The linking relation between a clitic-gap and its associated clitic is formally exactly the same as that obtaining between a WH-gap and its associated antecedent, and can if desired be conceptualized as a type of coindexing. The [clitics]{} feature-bundle is threaded through the grammar rule which defines the structure of the list of clitics associated with a cliticised verb, and enforces the constraints on ordering of surface clitics. These constraints are encoded in the lexical entry for each clitic. This basic framework is fairly straight-forward, though a number of additional features need to be added in order to capture the syntactic facts. We summarize the main points: Position of surface clitics: : Clitics occur post-verbally in positive imperative clauses, otherwise pre-verbally. The clitic-list constituent consequently needs to share suitable features with the verb it combines with. Surface form of clitics: : The first- and second-person singular clitics are realised differently depending on whether they occur pre- or post-verbally: for example “Vous me réservez un vol” versus “Réservez-moi un vol”. Moreover, “me” and “te” are first-position clitics (e.g. “Vous me les donnez”), while “moi” and “toi” are third-position (“Donnez-les-moi”). This alternation is achieved simply by having separate lexical entries for each form. The entries have different syntactic features, but a common semantic representation. Special problems with the “en” clitic: : The most abstruse problems occur in connection with the “en” clitic, and are motivated by sentences like > Combien en avez-vous? Here, our framework seems to dictate a constituent structure including three gaps, viz: > \[Combien \[\[en avez\]$_V$ \[vous$_{NP}$ \[\[\]$_V$ \[\[\]$_{NP}$ \[\]$_{PP}$\]$_{NP}$\]\]$_S$\]$_S$\]$_S$ in which the V gap links to “avez”, the NP gap to “combien”, and the PP gap to “en”. The specific difficulty here is that the “en” PP gap ends up as an NP modifier (it modifies the NP gap). Normally, however, PP modifiers of NPs cannot be gaps, and the above type of construction is the only exception we have found. Rather than relax the very common [NP $\rightarrow$ NP PP]{} rule to permit a gap PP daughter, we introduce a second rule of this type which specifically combines certain NPs, including suitable gaps resulting from WH-movement, and an “en” clitic gap. A feature, [takespartative]{}, picks out the NPs which can participate as left daughters in this rule. Agreement {#Section:Agreement} --------- Although grammatical agreement is a linguistic phenomenon that plays a considerably larger role in French than in English, the adjustments needed to the lexicon and syntax rules are usually obvious. For instance, a feature has to be added to the both daughters of the rule for pre-nominal adjectival modification, to enforce agreement in number and gender. In nearly all cases, this same procedure is used. A feature called [agr]{} is added to the relevant categories, whose value is a bundle representing the category’s person, number and gender, and the [agr]{} feature is shared between the categories which are required to agree. There are however some instances where agreement is less trivial. For example, the subject and nominal predicate complement of “être” may occasionally fail to agree in gender, e.g. > La gare est le plus grand bâtiment de la ville. However, if the predicate complement is a pronoun (“lequel”, “celui-ci”, “quel”[^1]...) agreement in both gender and number is obligatory: thus for instance > Quel/\*quelle/\*quels est le premier vol. It would be most unpleasant to duplicate the syntax rules, with separate versions for the pronominal and non-pronominal cases. Instead, we add a second agreement feature ([compagr]{}) to the NP category, which is constrained to have the same value as [agr]{} on pronominal NPs; subject/predicate agreement can then use the [compagr]{} feature on the predicate, getting the desired behaviour. Similar considerations apply to the rule allowing modification of a NP by a “de” PP. In general, there is no requirement on agreement between the head NP and the NP daughter of the PP. However, for certain pronominal NP (“lequel”, “l’un”, “chacun”) gender agreement is obligatory, e.g. > lequel/\*laquelle de ces vols\ > laquelle/\*lequel de ces dates This is dealt with correspondingly, by addition of a new agreement feature specific to the [NP $\rightarrow$ NP PP]{} rule. Spanish syntax {#Section:Spanish} ============== This section briefly describes the interesting features of the Spanish syntactic rule-set. In general, the Spanish rules were distinctly simpler than the French ones. With a few exceptions noted below (in particular, prodrop), the current Spanish syntax rules are essentially a slightly modified subset of the French ones. Despite this, they give very adequate coverage of the ATIS domain, the only in which they have so far been seriously tested. In a little more detail: Question-formation: : The Spanish rules for question-formation are similar to, but less elaborate than the French ones. Subject-verb inversion is allowed with any subject; there is no restriction that it be pronominal. There are no constructions corresponding to “est-ce que” or complex inversion. When the inverted subject is a pronoun, it does not require a preceding hyphen linking it to the verb. Clitics: : The Spanish clitic system is also considerably simpler than the French one. There are fewer clitics; in particular, there are no clitics corresponding to the French “y” and “en”, which as we saw in Section \[Section:Clitics\] above gave rise to many of the difficult problems in French. Postverbal clitics are affixed directly to the verb, rather than being joined by hyphens. Since CLE morphotax rules have a uniform format [@CLE §3.9], this only involved moving the relevant syntax rules to the morphology rule file. Phrasal rules: : The rules for Spanish numbers, dates and times are substantially different from the French ones, and those for dates in particular needed to be rewritten more or less from scratch. The issues involved are however straight-forward. Also, the form of the Spanish superlative adjective is slightly different: the postnominal superlative adjective has no extra article, e.g. “le vol le \[plus cher\] versus “la plaza \[menos cara\]”. The necessary adjustments are again simple. Relative clauses: : A less trivial difference involves relative clauses. In Spanish, the main verb of the relative clause must be in the subjunctive mood if it modifies an argument of a verb in the imperative mood. Thus for example > Which is the first flight that serves a meal?\ > $\rightarrow$ Cuál es el primer vuelo que sirve una comida? (“sirve” = present indicative), but > Show me flights that serve a meal!\ > $\rightarrow$ Enséñeme los vuelos que sirva una comida (“sirva” = present subjunctive). Handling this alternation correctly involves trailing an extra feature through many grammar rules, so as to link the main verb in the relative clause to the main verb in the clause immediately above it. Prodrop: : The second substantial change required when adapting the French grammar to Spanish was necessitated by the prodrop rule: Spanish, unlike French, permits and indeed encourages omission of the subject when it is a pronoun. Perhaps surprisingly, prodrop in fact only resulted in a few divergences between the Spanish and French grammars. A new syntax rule of the form [S $\rightarrow$ VP]{} was added (it is in fact a slightly modified version of the French imperative-formation rule). The associated semantic rule fills in a representation of the omitted clausal subject from the main verb; to make this possible, the semantic entries for inflected verbs are all modified to contain an extra feature encoding the possible prodrop subject. The details are straight-forward. Conclusions {#Section:Conclusions} =========== The preceding sections describe in essence all the changes we needed to make in order to adapt a substantial English language processing system to French and Spanish. Due to space limitations, we have been obliged to present some of the details in a more compressed form than we would ideally have wished, but nothing important has been omitted. Creation of a good initial French version required about five person-months of effort; after this, the Spanish version took only about two person-months. We do not believe that we were greatly aided by any special features of the Core Language Engine, other than the fact that it is a well-engineered piece of software based on sound linguistic ideas. Our overall conclusion is that an English-language system conforming to these basic design principles should in general be fairly easy to port to Romance languages. Acknowledgements {#acknowledgements .unnumbered} ================ The work described here was supported by SRI International, Suissetra, and Telia Research AB, Sweden. We would like to acknowledge the assistance provided by Gabriela Fernandez of the University of Seville in developing the Spanish version of the system, and thank Sabine Lehmann, David Milward and Steve Pulman for helpful comments. Agnäs, M-S., Alshawi, H., Bretan, I., Carter, D.M., Ceder, K., Collins, M., Crouch, R., Digalakis, V., Ekholm, B., Gambäck, B., Kaja, J., Karlgren, J., Lyberg, B., Price, P., Pulman, S., Rayner, M., Samuelsson, C. and Svensson, T. 1994. Spoken Language Translator: First Year Report. SRI technical report CRC-043[^2] Alshawi, H. (ed.) 1992. The Core Language Engine. MIT Press. Alshawi, H. Carter, D., Crouch, R., Pulman, S., Rayner, M. and Smith, A. 1992. CLARE: A Contextual Reasoning and Cooperative Response Framework for the Core Language Engine SRI technical report CRC-028. Alshawi, H., and Carter, D. 1994. Training and Scaling Preference Functions for Disambiguation. Computational Linguistics, 20:4. Alshawi, H. and Crouch, R. 1992. “Monotonic Semantic Interpretation”. Proceedings of 30th ACL. Bès, G. and Gardent, C. 1989. French Order without Order. Proceedings of 4th European ACL. Carter, D. 1995. Rapid Development of Morphological Descriptions for Full Language Processing Systems. Proceedings of 7th European ACL. Also SRI Technical Report CRC-047 Estival, D. Generating French with a Reversible Unification Grammar. 1990. Proceedings of 13th COLING. Grimshaw, J. 1982. On the Lexical Representation of Romance Reflexives. In J. Bresnan (ed.), The Mental Representation of Grammatical Relations. MIT Press. Kaplan, R., and Kay, M. 1994. Regular Models of Phonological Rule Systems. Computational Linguistics, 20:3, 331–378. Miller, P. and Sag, I. 1995. French Clitic Movement Without Clitics or Movement. CSLI Technical Report. Pulman, S. 1992. Unification-Based Syntactic Analysis. In [@CLE] Rayner, M. 1993. Abductive Equivalential Translation and its Application to Natural Language Database Interfacing. PhD thesis, Royal Institute of Technology/Stockholm University. Also SRI Technical Report CRC-052 Rayner, M. 1994. English linguistic coverage. In [@SLT-report] Rayner, M., Alshawi, H., Bretan, I., Carter, D.M., Digalakis, V., Gambäck, B., Kaja, J., Karlgren, J., Lyberg, B., Price, P., Pulman, S. and Samuelsson, C. 1993. A Speech to Speech Translation System Built From Standard Components. Proc. 1st ARPA workshop on Human Language Technology. Also SRI Technical Report CRC-031. Rayner, M. and Bouillon, P. 1995. Hybrid Transfer in an English-French Spoken Language Translator. Proceedings of IA ’95, Montpellier, France. Also SRI Technical Report CRC-056. van Eijck, J. and Moore, R. 1992. Semantic Rules for English. In [@CLE] [^1]: Most French grammars regard “quel” as an adjective, but for semantic reasons we have found it more convenient to treat it as a pronoun in this type of construction and as a determiner in expression like “quel vol”. [^2]: All SRI Cambridge technical reports are available through WWW from [http://www.cam.sri.com]{}
{ "pile_set_name": "ArXiv" }
--- abstract: 'We show that the n-dimensional MICZ-Kepler system arises from symplectic reduction of a simple mechanical system on the cone over the rotation group $SO(n)$. As a corollary we derive an elementary formula for its general solution. The punch-line of our computation is that the additional MICZ-Kepler $|\phi|^2/r^2$ type potential term is the rotational part of the cone’s kinetic energy.' address: | Dept. of Mathematics\ University of California, Santa Cruz\ Santa Cruz CA author: - Richard Montgomery date: 'November 15, 2011 (Preliminary Version)' title: 'MICZ-Kepler = dynamics on the cone over the rotation group' --- Introduction ============ The classical mechanical formulation of a Hydrogen atom is identical to the Kepler problem for a planet moving in the gravitational field of a massive Sun. The MICZ \[McIntosh-Cisneros-Zwanziger\] -Kepler system is an integrable extension of the Kepler problem in which the charged proton at the origin of the Hydrogen atom is simultaneously a Dirac monopole. The original references are [@Zwanziger; @Cisneros]. See [@Meng1; @Meng2] for history, references, and the n-dimensional generalization. Our purpose is to show that the MICZ-Kepler system is the reduction of a natural mechanical system (no magnetic fields) on the cone over the rotation group. See theorem 1 below and eq (\[H1\]). We then use this realization to write down an explicit Lie-theoretic formula (eq \[solution\]) for the system’s general solution. Meng [@Meng2] formulated the MICZ-Kepler system as a system on what he called a ‘Sternberg phase space’ -and what we will call the ‘adjoint bundle phase space’ - associated to a principal $SO(n-1)$ bundle over $\R^n \setminus \{0\}$. This phase space arises as the symplectic reduction of the cotangent bundles of the same principal bundle. See . It follows that there is a Hamiltonian system whose configuration space is Meng’s principal bundle and whose reduction yields the MICZ-Kepler systems. Our object is to find this system, and then use it. The heart of the computation (noticing that the Hamiltonian eq(\[H1\]) becomes (\[H2\]) in adjoint bundle variables) is the observation that the angular part of the Kepler kinetic energy plus the quadratic color charge term of the MICZ-Keple Hamiltonian equal the kinetic energy for the bi-invariant metric on the full rotation group. (For the quantum-mechanical analogue of this computation compare the algebra around eq (\[H2\]) to the algebra at the end of section 2 of [@Zwanziger]. ) MICZ-Kepler =========== The MICZ -Kepler system of equations can be written as a mixed 1st-2nd order system for a curve $(q(t), \phi (t))$ in a vector bundle over $\R^n \setminus \{0\}$. Here $q(t)$ denotes the curve in $\R^n$ and $\phi(t)$ lies in the moving fiber over this curve. This fiber is the Lie algebra ${\mathfrak so(n-1)}$ of $SO(n-1)$. The bundle is called the adjoint bundle, and is an associated vector bundle to a certain principal bundle $SO(n-1) \to Q \to \R^n \setminus \{0\}$, called the Dirac monopole and described in the next section. So, relative to a local trivialization of the bundle, $\phi(t)$ takes values in the Lie algebra $so(n-1)$. The principal bundle is endowed with a canonical connection whose curvature is $F$, and which induces a covariant derivative $D$ on the adjoint bundle. Then the MICZ-Kepler system is $$\ddot q = - \frac{q}{r^3} + \frac{ |\phi |^2 q} {r^4} + \phi \cdot F( \dot q, \cdot)$$ $$D \phi / dt = 0.$$ Some clarifications are in order. The curvature is a two-form with values in the adjoint bundle. Consequently $F(\dot q, \cdot)$ is a one-form with values in the adjoint bundle. The Killing form endows the adjoint bundle with a natural fiber inner product, so that $\phi \cdot F(, \dot q, \cdot)$ is a one-form on $\R^n$. We turn this one-form into a vector using the standard flat metric. In coordinates then: $ \phi \cdot F( \dot q, \cdot) ^j = \phi_a F^a _{ij} (q) \dot q ^i $. In a local trivialization the second equation reads $D \phi/dt = d \phi /dt + [ A _i \dot q^i , \phi]$ where $A$ is the connection one-form, an ${\mathfrak so(n-1)}$-valued one-form , in this trivialization. The system has two basic conserved quantities: its energy $$\label{E1} H = \frac {1}{2}( | \dot q | ^2 + \frac{| \phi| ^2 }{r^2}) - \frac{1}{r},$$ and the square norm of the adjoint variable, or $so(n-1)$- $$\text{Casimir} = \| \phi (t) \|^2.$$ There are other conserved quantities, an angular momentum and a Runge-Lenz (or Laplace) vector, but we will not need them here. The adjoint bundle canonically fibers into adjoint-orbit fibers which are preserved by parallel transport. Consequently the fiber variable $\phi (t)$ stays on whichever adjoint orbit fiber it begins on at time $0$. As a particular case, the zero-orbit is preserved and corresponds to setting $\phi = 0$ in the equations, which reduces them to the standard equations of the Kepler problem. Meng [@Meng2] takes $\phi$ to lie in an adjoint orbit of a particular type which he calls “magnetic”. We allow any $\phi$, hence any adjoint orbit. The n-dimensional Dirac Monopole ================================ We describe the bundle-with-connection needed to define the MICZ-Kepler equations of the previous section. We take our description from Meng, who called it the Dirac monopole, since that is what it is when $n = 3$. Consider the space $C_0$ of all orthogonal frames $f = (f_1, f_2, \ldots, f_n)$ on $\R^n$ normalized so that their lengths are all equal: $|f_i| = |f_j|$ all $i,j$. The map $$f \mapsto f_n , \qquad C_0 \to \R^n \setminus \{0\},$$ gives $C_0$ the structure of a principal $SO(n-1)$ bundle over $\R^n \setminus \{0\}$. Write $r =|f_n|$. Then, we have an $SO(n)$-equivariant diffeomorphism $$\R^+ \times SO(n) \to C_0 ; (r, g) \mapsto (r ge_1, rg e_2, \ldots r ge_n) = (f_1, \ldots f_n),$$ where $e_1, \dots, e_n$ is the standard basis of $\R^n$. From this perspective, the bundle projection becomes $(r, g) \mapsto r g e_n$. Put spherical coordinates on $\R^n$ so that $\R^n \setminus \{0\} \cong \R^+ \times S^{n-1}$. Recall that $S^{n-1} = SO(n)/SO(n-1)$ where the $SO(n-1) \subset SO(n)$ is the stabilizer of $e_n$. The restriction of the principal bundle $C_0 \to \R^n \setminus \{0\}$ to $S^{n-1} \subset \R^n \setminus \{0\}$ defines the homogeneous principal bundle $\pi_{S^{n-1}} : SO(n) \to S^{n-1}$. The Killing form on $so(n)$ endows $SO(n)$ with a bi-invariant metric $d^2 s_{SO(n)}$, and relative to this metric, the orthogonal complement to the fibers of the projection $\pi_{S^{n-1}}$ define an $SO(n)$-equivariant connection for the homogeneous principal bundle. Extend this connection trivially in the radial ($r$) direction to arrive at the “Dirac connection” of $C_0 \to \R^n \setminus \{0 \}$. We refer to the bundle $C_0$ endowed with this connection as ‘the monopole’. For coordinate expressions for the connection see . Adjoint and co-adjoint bundles. =============================== Suppose that $G \to Q \to S$ is a principal $G$ bundle. Then the adjoint bundle $Ad(Q) : = Q \times_G {\mathfrak g} \to S$ is the vector bundle associated to $Q$ via the adjoint action of $G$ on ${\mathfrak g}$. This means we divide the product $Q \times {\mathfrak g}$ by the $G$-induced equivalence relation $(q, \xi) \sim (qg, Ad_{g ^{-1}} \xi)$. Write equivalence classes $[q, \xi]$. The fiber of the adjoint bundle is ${\mathfrak g}$. The co-adjoint bundle is defined similarly, using the co-adjoint action. If ${\mathfrak g}$ is endowed with a bi-invariant inner product such as the Killing form on ${\mathfrak g} = so(n-1)$, then we get $G$-equivariant isomorphisms ${\mathfrak g} \cong {\mathfrak g}^*$ and hence we can identify the adjoint bundle with the co-adjoint bundle. This identification sends adjoint orbit fiber to the corresponding co-adjoint orbit fiber. We make this identification throughout the paper. Thus the variable $\phi(t)$ in the MICZ-Kepler system is a section of the adjoint bundle $Ad(C_0)$ along the curve $q(t) \in \R^n \setminus \{0\}$ A connection on $Q$ induces on any associated vector bundle, and so on the adjoint bundle. We can describe parallel transport in ${\mathfrak g} (Q)$ along a curve $c(t) \in S$ as follows. Pick a point $\phi (0) = [q(0), \xi] \in {\mathfrak g} (Q)$. Consider the horizontal lift $q(t) \in Q$ of $c(t)$. Then $\phi(t) = [q(t), \xi] \in {\mathfrak g} (Q) $ is the parallel transport of $\phi(0)$ along $c(t)$. The 2nd MICZ-Kepler equation states that $\phi(t)$ is covariantly constant, and hence of the form just described. Metric cones and submersions ============================ Let $r = |f_n| \to 0$ in the construction of $C_0 \cong \R^+ \times SO(n)$ so that the whole of $SO(n)$ is crunched to a point defined by $r=0$. We we have formed the cone $C = Cone(SO(n)) \supset C_0$ over the full rotation group $SO(n)$. The bundle projection $C_0 \to \R^n \setminus \{0\}$ extends to the cone point $r=0$, sending it to the origin. We now put a canonical metric on the cone $C$. Recall that the cone over a manifold $(X, d^2 s_X)$ is given by the metric $dr^2 + r^2 d^2 s_X$ on $\R^+ \times X$ where $r \in \R^+$. As $r \to 0$ the metric factor involving $X$ shrinks to $0$ so it makes sense to identify $\{0\} \times X$ to a single point, which is the cone point. In this we get a metric on $Cone(X) = [0, \infty) \times X/ \{0 \} \times X$ which is Riemannian away from the cone point. We form the metric cone $C = Cone(SO(n))$ by applying this cone construction to $X = SO(n)$ endowed with a Killing induced bi-invariant metric $d^2 s _SO(n)$. Such a metric is well-defined up to scale. That scale will be fixed by insisting that the bundle projection $C_0 \to \R^n \setminus \{0\}$ is a submersion. Recall the notion of a submersion $\pi: Y \to S$ between manifolds $Y, S$. Suppose that $\pi$ is a submersion. Take any point $y \in Y$ and consider the orthogonal complement at $y$ to the fiber $\pi ^{-1} (s)$ through $y$. Here $s = \pi(y)$. Following the principal bundle language as above, call this orthogonal complement ${\mathcal H}_y \subset T_y Y$ the ‘horizontal space’ at $y$. The differential $d \pi_y$ of $\pi$, restricted to the horizontal space, is necessarily a linear bijection onto the tangent space to $s$. Now the horizontal space inherits an inner product from $Y$. If this restricted differential is an isometry between inner product spaces for all $y$, then $\pi$ is said to be a submersion. Consider the standard metric $d^2 s_{S^{n-1}}$ on the unit sphere $S^{n-1}$. The Killing scale for $SO(n)$ is now fixed by insisting that the bundle projection $SO(n) \to S^{n-1}$ be a submersion. This scaling is the one for which the standard basis elements $e_i \wedge e_j$ of $so(n)$ have length $1$. Here $e_i \wedge e_j$ is the skew-symmetric operator sending $e_i$ to $e_j$ and $e_j$ to $-e_i$, $i \ne j$. Let $\theta_{ij}$ be the dual basis, viewed as left-invariant one-forms. Then $d^2 s _{SO(n) } = \Sigma (\theta_{i j})^2$so that $$\begin{aligned} \label{metricC} d^2 s _{C } &=&dr ^2 + r^2 \Sigma (\theta_{i j})^2 \end{aligned}$$ We verify that $SO(n) \to S^{n-1}$ and $C_0 \to \R^n \setminus \{0\}$ are submersions. The connection form for the Dirac monopole is the $so(n-1)$-valued one-form $$\label{monopole} A = \Sigma_{i < j < n} \theta_{ij} e_i \wedge e_j , \qquad ( \text{ on } SO(n) \text{ or } C_0 ).$$ (Note we must use [*left*]{} invariant one forms, rather than right-invariant formsas can be seen by the fact that the connection is not $G$-invariant, but rather $G$-equivariant with $G$ acting on the Lie lagebra by the adjoint action.) The horizontal distribution is defined by $\theta_{ij} =0, i , j \ne n$. The vertical distribution is defined by $\theta_{i n} = 0, i = 1, 2, \ldots , n-1$, together with $dr = 0$ in the $C_0$ case. Thus the Killing metric on $SO(n)$ splits orthogonally relative to the horizontal-vertical splitting $$\begin{aligned} \label{metricS1} d^2 s _{SO(n) } &=& \Sigma (\theta_{i n})^2 + \Sigma_{ i < j < n} (\theta_{ij})^2 \\ & = & \pi^* d^2 s _{S^{n-1}} + d^2 s_{fiber}\end{aligned}$$ The fact that $exp (t e_i\wedge e_n)$ has period $2 \pi$ shows that the metric scalings are correct for the submersion: a 2$\pi$ periodic horizontal geodesic in $SO(n)$ maps to a great circle of circumference $2 \pi$ in $S^{n-1}$. Now, written out in spherical coordinates the metric on $\R^n$ is $ds^2 _{\R^n} = dr^2 + r^2 ds^2_{S^{n-1}}$, which is to say that metrically speaking $\R^n = Cone(S^{n-1})$. The corresponding radial-horizontal spherical-vertical splitting of the metric on $C$ is $$\begin{aligned} \label{metricS1} d^2 s _{C } &=& (dr^2 + r^2 \Sigma (\theta_{i n})^2 )+ r^2 \Sigma_{ i < j < n} (\theta_{ij})^2 \\ & = & \pi^* ds^2 _{\R^n} + d^2 s_{fiber}\end{aligned}$$ which shows the projection $C \to \R^n$ becomes a submersion away from the cone point. Main result: MICZ-Kepler from a mechanical system on the cone. ============================================================== A natural mechanical system consists of a configuration space $Q$, endowed with a metric $ds^2_Q$ and a potential function $V: Q \to \R$. This data defines a Hamiltonian on $T^*Q$ whose Hamiltonian $H$ is kinetic plus potential: $H = K + V$, where the kinetic energy $K$ is induced by the metric. In standard canonical coordinates $(q, p) = (q_i, p^i)$ for $T^*Q$ we have $K(q,p) = \frac{1}{2} g^{ij} (q) p_i p_i p_j$ if $ds^2 _Q = g_{ij} dq^i dq^j$. Take the canonical metric cone $C = Cone(SO(n))$ described in the previous section so as to get a metric on $C_0 = Cone(SO(n) \setminus \{0\}) \to \R^n \setminus \{0 \}$. Take potential function $V = -\frac{1}{r} :C_0 \to \R$ where $r: Cone(SO(n)) \to [0, \infty)$ denote the cone’s radial coordinate. Then the resulting natural mechanical system (Hamiltonian (\[H1\]) below) on $T^* C_0$ is invariant under the lifted action of $SO(n)$, and so is also invariant by $SO(n-1)$. The symplectic reduction of this system by $SO(n-1)$ at any particular $\mu \in so(n-1)^* \cong so(n-1)$ yields the generalized MICZ-Kepler system associated to the adjoint orbit through $\mu$. We see from the expression (\[metricC\]) for the metric on the cone that the Hamiltonian of this theorem is $$\label{H1} H = \frac{1}{2} (p_r ^2 + \frac{1}{r^2} \Sigma \xi_{ij}^2 )- \frac{1}{r}$$ where $r,p_r$ are canonical coordinates on $T^* \R^+$ and the $\xi_{ij}$ are the Lie-Poisson coordinates – linear coordinates on $so(n)^*$ - induced by the choice of basis $e_i \wedge e_j$ for $so(n)$. Any solution $(q(t), \phi(t))$ to the generalized MICZ-Kepler system can be constructed as follows. Fix $\xi \in so(n), w \in S^{n-1}$. Fix a solution $r(t)$ to the 1-dimensional Kepler problem: $\ddot r = - V_{\mu} ^{\prime} (r)$ where the effective potential is $V_{\mu} (r)= -\frac{1}{r} + \frac{\mu^2}{ r^2}$ with $\mu^2 = | \xi |^2$, and a solution $u(t)$ to the ODE $\dot u = \frac{1}{r(t)^2}$. Then $$\label{solution} q (t) = r(t) exp(u (t) \xi) w$$ The adjoint bundle variable $\phi(t)$ is obtained by parallel translating an initial adjoint vector $\phi(0) = [f, A_f (\xi)]$ along $q(t)$, where $f \in C_0$ is any element projecting to $q(0)$ and $A$ is the connection one-form (\[monopole\]). [**Special Cases.**]{} [**1. Kepler.**]{} Take $\xi$ horizontal over the initial $q(0)$, so that $A(\xi) = 0$ and $\phi(t) = 0$. For simplicity, take the initial $q(0)$ in the direction $w = e_n$. Then horizontality implies $\xi = \Sigma v_i e_i \wedge e_n = {\vec v} \wedge e_n$ is an infinitesimal rotation in the ${\vec v}, e_n$ plane. The solution $q(t)$ then lies in this plane. Set $\theta(t) = u(t) | \xi| $. Then $(r(t), \theta(t))$ is a solution to Kepler’s equations expressed in polar coordinates. [**2. Magnetic cone.**]{} Meng takes his $\phi (0)$ to be of “magnetic type’, which means, relative to a local trivialization, that $\phi(0) ^2 = - \mu Id.$ In other words, up to scale $\phi(t)$ is an almost complex structure on $\R^n$, compatible to the standard complex structure. Thus $\xi$ satisfies $\xi ^2 = -\mu ^2 Id$. Set $J = \xi/\mu$ so that $J$ is an honest almost complex structure. Then $exp(u \xi ) = cos( \mu u) I + \sin (\mu u) J $. The solution $q(t)$ lies on the two-plane spanned by $q(0)$ and $J q(0)$. Indeed, it is another Keplerian conic on that plane, as Meng showed in [@Meng2]. [**3. Generic.**]{} Take $\xi$ generic, meaning that it has $[n/2]$ distinct nonzero eigenvalues $\pm i \omega_j$ linearly independent over the rationals. We can, by conjugating by a rotation, put $\xi$ into the normal form $\Sigma \omega_j e_{2j -1} \wedge e_{2j}$. Then $\theta \mapsto exp(\theta \xi)$ is a dense curve on a standard maximal torus in $SO(n)$. For negative energy the corresponding one dimensional Kepler motion is periodic with period $T$ , and without collision (since $\xi \ne 0$). We can arrange that $2 \pi /T$ is rationally independent of the $\omega_j$. Then the corresponding solution curve $q(t)$ forms a dense winding on a kind of “annular projection” $r_{min} \le r \le r_{max}$ to $\R^n$ of a torus of dimension $[n/2] + 1$. Proof of theorem 1. =================== We apply the general theory of reduction of cotangent bundles of a principal bundle. We first describe that general theory. See [@Montgomery1], particularly pp. 160-163, or the earlier references [@Montgomery1; @Weinstein; @Sternberg] for perhaps more leisurely descriptions. Let $G \to Q \to S$ be a principal $G$-bundle. $G$ acts on $T^*Q$ with $G$-equivariant momentum map $J: T^* Q \to {\mathfrak g}^*$. The quotient $(T^*Q)/G$ is naturally a Poisson manifold whose symplectic leaves are the symplectic reduced spaces $J^{-1} ({\mathcal O}_ \mu)/G = J^{-1} (\mu)/G_{\mu}$ where ${\mathcal O}_{\mu}= G \cdot \mu$ denotes the coadjoint orbit through $\mu \in {\mathfrak g}^*$. The general theory proceeds by using a connection on $Q$ to define a symplectic isomorphism with these reduced spaces. Differentiate the sequence of maps $\xymatrix{ G \ar[r] & Q \ar[r]^{\pi} & S}$ at fixed $q \in Q$ to obtain the sequence of linear maps $\xymatrix{ {\mathfrak g} \ar[r]^{\sigma_q } & TQ \ar[r]^{d \pi_q} & T_s S}$ where $s = \pi(q)$. The sequence is exact: the image of $\sigma_q$ equals the kernel of $d\pi_q$. (This common image is called the vertical space at $q$.) Letting $q$ vary parametrically we obtain the ‘Atiyah sequence’ (described in [@Atiyah]) $$\label{AtiyahSeq} \xymatrix{ Q \times {\mathfrak g} \ar[r]^{\sigma} & TQ \ar[r]^{d \pi} & \pi_S ^* TS}$$ which is an exact sequence of $G$-equivariant vector bundles over $Q$. Dualizing the first map of (\[AtiyahSeq\]), and composing with the projection yields the momentum map : $$J: \xymatrix{ T^*Q \ar[r]^{\sigma^{*}} & Q \times {\mathfrak g}^{*} \ar[r]& {\mathfrak g}^*}.$$ A connection $A$ for $Q \to S$ induces a $G$-invariant splitting of (\[AtiyahSeq\]) and hence a $G$-equivariant isomorphism: $$TQ \cong \pi_S^* TS \oplus (Q \times {\mathfrak g}).$$ Dualizing yields the $G$-equivariant isomorphism: $$T ^*Q \cong \pi_S^* T^* S \oplus (Q \times {\mathfrak g}^*). \label{dualIso}$$ Now $G$ acts on the bundle $Q \times {\mathfrak g}$ by $g(q, \xi) = (qg, Ad_{g^-1} \xi)$ as per the equivalence relation used to define the adjoint bundle and thus the action of $G$ on $Q \times {\mathfrak g}^*$ is the one used to define the co-adjoint bundle. Forming the quotient by $G$ we thus get the bundle isomorphism $$\Psi_A: (T^*Q)/G \cong T^*S \oplus Ad^* (Q)$$ over $S$ which is our desired identification. We refer to the right hand side of this isomorphism as being “on the Adjoint bundle side” in what follows. From our factorization of $J$ we see that under the isomorphism $\Psi_A$ the reduced spaces $J^{-1} (\mu)/G_{\mu} = J^{-1} ({\mathcal O}_ \mu)/G$ become the submanifolds $T^*S \oplus ({\mathcal O}_{\mu}) (Q)$, which are the Adjoint bundle phase spaces and the Sternberg phase spaces of ([@Meng2]). If ${\mathfrak g}$ is endowed with a bi-invariant Killing form as above, then the co-adjoint orbit bundle $({\mathcal O}_{\mu}) (Q)$ is identified with a corresponding adjoint orbit. We need more detail regarding the isomorphism $\Psi_A$ and the Poisson brackets on the universal phase spaces in order to pull-back the cone Hamiltonian (\[H1\]) and compute equations of motion. The connection defines horizontal lift operators $h_q: T_s S \to T_q Q$, $q \in Q$, which are linear operators whose image is the horizontal space ${\mathcal H}_q = ker(A(q)) \subset T_q Q$ of the splitting of $TQ$. The dual of $h_q$ is $h_q ^*: T^* _q Q \to T^* _s S $ and is one factor of the isomorphism (\[dualIso\]). Write $[q, P]$ for the equivalence class in $T^*Q/G$ of $(q, P) \in T^*Q$. Then $$\Psi_A ([q,P] ) = (\pi(q), h_q^* P) \oplus [q, J(q,P)] \in T^*S \oplus Ad^* (Q).$$ We describe $\Psi_A$ in coordinates. Let $(x^i , g) \in \R^d \times G$ be coordinates on $Q$ induced by a local trivialization $Q_U \cong U \times G$ of $Q \to S$, together with coordinates on $U \subset S$. Then, over $U$ we have $T^*Q \cong T^* U \times T^* G = T^*U \times G \times {\mathfrak g}^*$ with coordinates $(x^i, p_i, g, \xi_a)$ where the coordinates $\xi_a$ are Lie-Poisson linear coordinates on ${\frak g}^*$. relative to a basis $e_a$ of ${\frak g}$. Thus $(T^* Q)/G \cong T^*U \times {\mathfrak g}^*$ with coordinates $x_i, p_i, \xi_a$. On the other hand, the same data $(x^i, g)$ and basis $e_a$ yield coordinates $x_i, \pi_i, \xi_a$ for $T^* S \oplus Ad^* (Q)$. Relative to these two sets of coordinates the map $\Psi_A$ is the minimal coupling procedure $\Psi_A (x, p_i, \xi) = (x^i , p_i - \xi_a A^a _i (x), \xi_a) = (x^i , \pi_i, \xi_a)$ where $A(x) = \Sigma e_a A^a _i (x) dx$ is the connection one-form relative to the local trivialization and coordinates. The brackets on the Adjoint bundle side are $\{x_i , \pi_j \} = \delta ^i _ j, \{ \xi_a \xi_b \} = -c^d _{a b} \xi _d$ and $\{ \pi_i, \pi_j \} = -\xi_a F ^a_{ij}$, $\{\pi_i, \xi_a \} = D_i \xi ^a = [A_i, \xi]_a$. Here $F^a _{ij}$ is the expression for the curvature of $A$ in this local trivialization. (Compare eqs (12.2) of [@Montgomery1] to (3.2) [@Meng2]. Note that in the triple of displayed equations immediately following (12.2) of [@Montgomery1] most terms should have a capital $P$ immediately in front of them.) These agree with the brackets found in Meng for the case of $so(n-1)$. Now we recompute the Hamiltonian (\[H1\]) on the Adjoint bundle side using $\Psi_A$. The dual of the metric splitting (\[metricS1\]) yields $$\begin{aligned} \label{K1} K_C & = &\frac{1}{2} [p_r ^2 + \frac{1}{r^2}(\Sigma_{i =1} ^n \xi_{in} ^2 + \frac{1}{2} \Sigma_{ i < j < n} \xi_{ij} ^2)] \\ & = & \frac{1}{2} [ p_r ^2 + \frac{1}{r^2} (h^*K_{S^n} + |\phi|^2) ]\end{aligned}$$ where $h^* : T^* SO(n) \to T^* S^{n-1}$ is the connection induced dual of the horizontal lift. The fiber term $\Sigma_{ i < j < n} \xi_{ij} ^2$ corresponds, on the Adjoint bundle side, to the Casimir function $|\phi \|^2$ of $SO(n-1)$, viewed as a function on the adjoint bundle. So the Hamiltonian, viewed on adjoint bundle side, reads $$H = \frac{1}{2} (p_r ^2 + \frac{1}{r^2} K_{S^n} + \frac{1}{r^2} |\phi |^2) - \frac{1}{r}$$ The sum of the first two terms $ \frac{1}{2} (p_r ^2 + \frac{1}{r^2} K_{S^n})$ is the usual kinetic energy $\frac{1}{2} \Sigma_{i =1} ^n \pi_i ^2 $ on $\R^n$ written in spherical variables. Thus $$H = \frac{1}{2} (\Sigma \pi_i ^2 + \frac{1}{r^2} |\phi |^2) - \frac{1}{r} \label{H2}$$ which is the MICZ-Kepler Hamiltonian. Proof of the corollary. ======================= We compute the equations of motion on $C_0$, using the expression (\[H1\]) for the Hamiltonian and the fact that $\Omega = \|\xi \|^2$ is a Casimir. Set $\mu^2 = \Omega$ and $V_{\mu} (r) = -\frac{1}{r} + \frac{\mu^2}{r^2}$ Then Hamilton’s equations on $T^* C_0 = T^* \R^+ \times T^* SO(n) = \R^+ \times \R \times SO(n) \times so(n)$ are: $$\dot r = \{ r, H \} = p_r$$ $$\dot p_r = \{p_r , H \} = - V_{\mu^2} ^{\prime} (r)$$ $$\dot g = g \frac{ \partial H}{\partial \xi} = g \frac{1}{r^2} \xi$$ $$\dot \xi_a = \{\xi_a , H \} =\frac{1}{2r^2} \{ \xi_a, \Omega\} = 0$$ The first pair of equations decouple from the second pair, and assert that $(r,p_r)$ evolves as per the one-dimensional radial Kepler equation with effective potential $V_{\mu}$. The last equation asserts that $\xi \in so(n)$ is constant. The equation for $g$ asserts that $g(t) = g_0 exp(u(t) \xi)$ where $du/dt = 1/r^2$. Indeed, the solution to $\dot g = g \xi$ through $g = Id$ is the one-parameter subgroup $exp(t \xi)$ and this flow is generated by the Hamiltonian $\frac{1}{2} \Omega$. We have scaled the Hamiltonian on $SO(n)$ by the (time-dependent) factor $1/r^2$ and used left-invariance. To rewrite $g_0 exp (u(t) \xi) = exp(u(t) \tilde \xi)g_0$ we can set $\tilde \xi = g_0 \xi g_0 ^{-1}$. We now have the solution the Kepler equation on the cone: $(r(t), g(t))$ with $g(t) = exp(u(t) \xi) g_0$. Recall the bundle projection is $(r, g) \mapsto r ge_n$ and use that any unit vector $w$ can be written $g_0 e_n$ to the expression for $q(t)$ in the corollary. QED Other groups ============ The tricks used here apply to any Lie group $G$ in place of $SO(n)$ provided that $G$ is endowed with an faithful orthogonal representation on $\R^n$ which is transitive on the unit sphere $S^{n-1}$. We get the theorem that the Kepler problem on $Cone(G)$ is equivalent to the ‘MICZ-Kepler-$G$ ’ problem whose ‘color variables’ lie in an adjoint orbit bundle for $G$ over $\R^n \setminus \{0\}$. The standard families of such groups are the unitary groups $U(n)$ on $\R^{2n} = \C^n$ and $Sp(n; \H)$ on $\H^n = \R^{4n}$. [20]{} M. Atiyah, [*Complex Analytic Connections in Fibre Bundles*]{}, Trans. AMS, [**85**]{}, 181-207, (1957); or Collected works, vol. 1 , pp. 97-102, Clarendon Press - Oxford, (1987). G. Meng, [*MICZ-Kepler Problem in all Dimensions*]{}, J. Math. Phys. 48, 032105 (2007) (2007), arXiv:0507028 \[math-ph\]. G. Meng, [*The Poisson Realization of so(2,2k+2) on Magnetic Leaves*]{}, (2012), arXiv :1211:5992 \[math-ph\] H. McIntosh, A. Cisneros, Degeneracy in the presence of a magnetic monopole, J. Math. Phys. 11 (1970), 896-916. R. Montgomery, “A tour of subriemannian geometries , their geodesics, and applications". \[monograph\], Mathematical Surveys and Monographs, vol. 91, American Math. Society, Providence, Rhode Island, 2002. R. Montgomery , [*Canonical Formulations of a Particle in a Yang-Mills Field*]{}, Lett. Math. Phys. [**8**]{}, 59-67, (1984). S. Sternberg, [*On minimal coupling and the symplectic mechanics of a classical particle in the presence of a Yang-Mills field*]{}, Proc Nat Acad Sci, [**74**]{}, 5253-5254, (1977). A. Weinstein, [*A universal phase space for a particle in a Yang-Mills field*]{}, Lett. Math. Phys., [**2**]{}, 417-420, (1978). D. Zwanziger, Exactly soluble nonrelativistic model of particles with both electric and magnetic charges, Physical Review [**176**]{}, no. 5, (1968), 1480-1488.
{ "pile_set_name": "ArXiv" }
--- abstract: | A [*pure Dirac’s method*]{} of Yang-Mills expressed as a constrained $BF$-like theory is performed. In this paper we study an action principle composed by the coupling of two topological $BF$-like theories, which at the Lagrangian level reproduces Yang-Mills equations. By a pure Dirac’s method we mean that we consider all the variables that occur in the Lagrangian density as dynamical variables and not only those ones that involve temporal derivatives. The analysis in the complete phase space enable us to calculate the extended Hamiltonian, the extended action, the constraint algebra, the gauge transformations and then we carry out the counting of degrees of freedom. We show that the constrained $BF$-like theory correspond at classical level to Yang-Mills theory. From the results obtained, we discuss briefly the quantization of the theory. In addition we compare our results with alternatives models that have been reported in the literature.\ author: - Alberto Escalante - 'J. Berra' title: ' A pure Dirac’s method for Yang-Mills expressed as a constrained BF-like theory' --- INTRODUCTION ============= Nowadays, the study of topological field theories is a topic of great interest in physics. The importance for studying those theories lies in a closed relation with physical theories as for instance, Yang-Mills \[YM\] and General Relativity [@carlo; @tomas]. Topological field theories are characterized by being devoid of local physical degrees of freedom[^1] , they are background independent and diffeomorphisms covariant [@3a; @4a]. Relevant examples of topological field theories are the so called $BF$ theories. $BF$ theories were introduced as generalizations of three dimensional Chern-Simons actions and in a certain sense this is the simplest possible gauge theory. It can be defined on spacetimes of any dimension. It is background free, meaning that to formulate it we do not need a pre-existing metric or any other such geometrical structure on spacetime [@5a; @6a]. At classical level, the theory has no local degrees of freedom; all the interesting observables are global in nature and this seems to remain true upon quantization [@baez]. Thus $BF$ theory serves as a simple starting point for the studies of background free theories. In particular, general relativity in 3-dimensions is a special case of $BF$ theory, while general relativity in 4-dimensions can be viewed as a $BF$ theory with extra constraints [@8aa]. Furthermore, we are able to find in the literature several examples where $BF$ theories come to be relevant models for instance, in alternative formulations of gravity such as the MacDowell-Mansouri approach [@7a]. MacDowell-Mansouri formulation of gravity consists in breaking down the internal symmetry group of a $BF$-theory from $SO$(5) to $SO$(4), to obtain Palatini’s action plus a sum of the second Chern and Euler topological invariants. Because these terms have trivial local variations that do not contribute classically to the dynamics, one thus obtain essentially general relativity. On the other hand, within the framework of \[YM\] theories, we can find some cases where $BF$ theories have been relevant, an example of this is Martellini’s model [@8a]. This model consists in to express \[YM\] theory as a $BF$-like theory in order to expose its relation with topological $BF$ theory. Thus, the first-order formulation (BF-YM) is equivalent on shell to the usual second-order formulation (YM). In fact, after a Wick rotation both formulations of the theory possess the same perturbative quantum properties; the Feynman rules, the structure of one loop divergent diagrams and renormalization has been studied founding that there exists an equivalence of the $uv$-behaviour for both approaches [@Martellini; @1]. The main advantage of this formulation lies in the possibility to express some observables like the color magnetic operator in the continuum and express it as the dual ’t Hooft observable employing the abelian projection gauge [@Cattaneo]. Nevertheless, the canonical quantization perspective is more subtle under Wick rotations, because it shows that in order to make the theory euclidean, we must to complexify the Poisson algebra [@Martellini; @2]. Recently other approaches use a well known duality between Maxwell-Chern Simons theory and a self dual massive model, this description has been extended to topological massive gauge theories providing a topological mechanism to generate mass for the bosonic $p$-tensor fields in any spacetime dimension [@Bertrand; @Bertrand1].\ On the other side, we are able to find that Lisi in [@Lisi] and Smolin in [@Smolin] have worked with $BF$-like theories written as an extension of Plebanski’s action, and in those works they got a consistent dynamics for any group $G$ containing the local Lorentz group, and then by using a simple mechanism which breaks down the symmetry, they obtain as resulting dynamics to \[YM\] coupled to general relativity plus corrections.\ At the light of these facts, in this paper we analyze a $BF$-like action yielding \[YM\] equations of motion. We study the principal symmetries of that action by means of a detailed canonical analysis using a pure Dirac’s method, and the quantization procedure is discussed. With the terminology *a pure Dirac’s method* we mean that we shall consider in the Hamiltonian framework all the fields that define our theory are dynamical ones. Of course, our approach differs from the standard Dirac’s analysis, because the standard analysis is developed on a smaller phase space by considering as dynamical variables only those variables with time derivative occurring explicitly in the Lagrangian. In this paper, our approach present clear advantages in respect to the standard one, namely; by working on the full phase space, we will able to know the full structure of the constrains and their algebra, the equations of motion obtained from the extended action and the full structure of the gauge transformations as well. The approach used in the present paper, has been performed to diffeomorphism covariant field theories [@4aa; @4bb], showing results that are not obtained by means of a standard Dirac’s analysis. In those works were reported the full structure of the constraints on the full phase space for the Second-Chern class and the latter for general relativity in the $G \rightarrow 0$ limit, being $G$ the gravitational coupling constant. The correct identification of the constraints is a very important step because are used to carry out the counting of the physical degrees of freedom and they let us to know the gauge transformations if there exist first class constraints. On the other hand, the constraints are the guideline to make the best progress for the quantization of the theory, therefore it is mandatory to know their full structure [@4a]. It is worthwhile to mention, that the constraints obtained by performing a pure Dirac’s formalism, the algebra among them is closed, and is not necessary to fix by hand the constraints as in the case of Plebanski theory [@Peldan; @ww], because the method itself provides us the required structure. One example of ambiguities found by developing the hamiltonian analysis on a reduced phase space, is presented in three dimensional tetrad gravity, in despite of the existence of several articles performing the hamiltonian analysis, in some papers it is written that the gauge symmetry is Poincare symmetry [@Witten], in others that is Lorentz symmetry plus diffeomorphisms [@Carlip], or that there exist various ways to define the constraints leading to different gauge transformations. We think that the complete Hamiltonian method ( a pure Dirac’s formalism) is the best tool for solving those problems.\ Finally we show that the action analyzed in this paper is the coupling of topological theories namely; the $BF$-like action studied here can be split in two terms lacking of physical degrees of freedom, the complete action, however, does has physical degrees of freedom, the \[YM\] degrees of freedom.\ The paper is organized as follows: In section II, we show that \[YM\] action differ from the $BF$-like action studied here, because it is expressed as the coupling of two topological terms just as General Relativity in Plebanski’s formulation [@ww]. By a pure Dirac’s method of the $BF$-like theory, one realize that the couplet terms incorporate reducibility conditions in the constraints, thus, the topological invariance of the full action will be broken emerging degrees of freedom. In Section III, a complete canonical analysis analysis of Martellini’s model is performed, this exercise has not reported in the literature, then we compare the results of this section with those obtained in previous sections. We finish with some remarks about our results. \[c2\] A pure Dirac’s method for \[YM\] theory expressed as a constrained BF-like theory ================================================================================= The action of our interest is given by $$S[A,B]= \int_M \ast B^{I} \wedge B^{I} -2 B^{I}\wedge \ast F^{I}, \label{bf-ym1}$$ where $F^{I}$ corresponds to the curvature of the connection one-form $A^{I}$ valued on the algebra of $SU(N)$. In this manner, the action (\[bf-ym1\]) takes the form $$S[B,A]=\int_{M}\frac{1}{4}B_{\mu\nu}^{I}B^{\mu\nu I}-\frac{1}{2}B^{\mu\nu I}\left(\partial_{\mu}A^I_{\nu}-\partial_{\nu}A_{\mu}^{I}+f^{IJK}A_{\mu}A_{\nu}\right), \label{bf-ym}$$ the equations of motion obtained from (\[bf-ym\]) are given by $$B_{\mu\nu}^{I}=F_{\mu\nu}^{I},\; \;\;\;\;\;\;\; D_{\mu}B^{\mu\nu I}=0,$$ which correspond to \[YM\] equations of motion. By substituting the former equations of motion in the action we recover the \[YM\] action $$S[A]=-\int_{M}\frac{1}{4}F_{\alpha\beta}^{I}F^{\alpha\beta }_ Id^{4}x, \label{eq.4}$$ where $F_{\alpha\beta}^{I}=\partial_{\mu}A^I_{\nu}-\partial_{\nu}A_{\mu}^{I}+f^{I}{_{JK}}A^J_{\mu}A^K_{\nu}$, is the curvature tensor valued on a Lie algebra. It is important to remark, that the role of the dynamical variables occurring in the actions (\[bf-ym\]) and (\[eq.4\]) is quite different. For the former, $A_\mu$ and $B_{\mu \alpha}$ both are determined by the dynamics. For the later, $A_\mu$ is determined by the dynamics and $F_{\mu \alpha}$ is not a dynamical variable anymore, it is a label.\ In order to procedure with our analysis, we are able to observe that if we split the action (\[bf-ym\]) in two parts, say $$\label{S1} S_{1}[B]=\int_{M}\frac{1}{4}B_{\mu\nu}^{I}B^{\mu\nu }_I,$$ and $$\label{S2} S_{2}[A,B]=\int_{M}\frac{1}{2}B_ I^{\mu\nu }\left(\partial_{\mu}A_{\nu}^{I}-\partial_{\nu}A_{\mu}^{I}+f^I_{{JK}}A_{\mu}^{J}A_{\nu}^{K}\right),$$ we obtain two topological field theories. In fact, we can see immediately that $S_{1}[B]$ is topological since does not have dynamical variables occurring in the action. On the other hand, we will show below that the action (\[S2\]) is a topological one as well. May be for the lector this part is not relevant, however, we need to remember that topological field theories are characterized by being devoid of local degrees of freedom. That is, the theories are susceptible only to global degrees of freedom associated with non-trivial topologies of the manifold in which they are defined and topologies of the gauge bundle [@21a; @22a]. Thus, it is mandatory to perform the canonical analysis of the action (\[S2\]) because there is a gauge group.\ So, by performing the 3+1 decomposition in (\[S2\]) we obtain $$S_{2}[A,B]=\int\int_{\Sigma}\left[B^{0iI}(\dot{A}_{i}^{I}-\partial_{i}A_{0}^{I}+f^{IJK}A_{0}^{J}A_{i}^{K})+\frac{1}{2}B^{ijI}F_{ij}^{I}\right], \label{49n}$$ where $F_{ij}^{I}=\partial_{i}A_{j}^{I}-\partial_{j}A_{i}^{I}+f^{I}{_{JK}}A_{i}^{J}A_{j}^{K}$.\ Dirac’s method calls for the definition of the momenta $(\Pi^{\alpha I},\Pi^{\alpha\beta I})$ canonically conjugate to the dynamical variables $(A_{\alpha}^{I},B_{\alpha\beta}^{I})$ $$\Pi^{\alpha I}=\frac{\delta \mathcal{L}}{\delta \dot{A}_{\alpha}^{I}}, \;\;\;\;\; \Pi^{\alpha\beta I}=\frac{\delta\mathcal{L}}{\delta \dot{B}_{\alpha\beta}^{I}},$$ On the other hand, the matrix elements of the Hessian $$\frac{\partial^{2}\mathcal{L}}{\partial(\partial_{\mu}A_{\alpha}^{I})\partial(\partial_{\mu}A_{\rho}^{J})},\;\;\; \frac{\partial^{2}\mathcal{L}}{\partial(\partial_{\mu}B_{\alpha\beta}^{I})\partial(\partial_{\mu}A_{\rho}^{J})},\;\;\; \frac{\partial^{2}\mathcal{L}}{\partial(\partial_{\mu}B_{\alpha\beta}^{I})\partial(\partial_{\mu}B_{\rho\gamma}^{J})},$$ are identically zero, thus the rank of the Hessian is zero. Therefore, we expect $10(N^{2}-1)$ primary constraints. From the definition of the momenta we identify the following 10 primary constraints $$\begin{split} \phi^{0I}:\Pi^{0I}&\approx 0,\\ \phi^{iI}:\Pi^{iI}-B^{0iI}&\approx 0,\\ \phi^{0i}_I:\Pi^{0i}_I&\approx 0,\\ \phi^{ij}_I:\Pi^{ij}_I&\approx 0, \end{split}$$ The canonical Hamiltonian density for the system has the following form $$\begin{split} \mathcal{H}_{c}&=\dot{A}_{\mu}^{I}\Pi^{\mu }_I+\dot{B}_{0i}^{I}\Pi^{0i}_I+\dot{B}_{ij}^{I}\Pi^{ij}_I-\mathcal{L}\\ &=-A_{0}^{I}D_{i}\Pi^{0i}_I-\frac{1}{2}B^{ij}_IF_{ij}^{I}, \end{split}$$ Thus, by taking in to account the primary constraints, we can identify the primary Hamiltonian given by $$H_{P}=H_{c}+\int d^{3}x\left[\lambda_{0}^{I}\phi^{0}_I+\lambda_{i}^{I}\phi^{i}_I+\lambda_{0i}^{I}\phi^{0i}_I+\lambda_{ij}^{I}\phi^{ij}_I\right],$$ where $\lambda_{0}^{I},\lambda_{i}^{I},\lambda_{0i}^{I},\lambda_{ij}^{I}$ are Lagrange multipliers enforcing the constraints. The fundamental Poisson brackets for our theory are given by $$\begin{split} \{A_{\alpha}^{I}(x),\Pi^{\mu I}(y)\}&=\delta^{\mu}_{\alpha}\delta^{IJ}\delta^{3}(x-y),\\ \{B_{\alpha\beta}^{I}(x),\Pi^{\mu\nu I}(y)\}&=\frac{1}{2}\left(\delta_{\alpha}^{\mu}\delta_{\beta}^{\nu}-\delta_{\beta}^{\mu}\delta_{\alpha}^{\nu}\right)\delta^{3}(x-y). \end{split}$$ In this manner, by using the fundamental Poisson brackets for our theory we find that the following $10(N^{2}-1)\times 10(N^{2}-1)$ matrix whose entries are the Poisson brackets among the primary constraints $$\begin{aligned} \{ \phi^{0}_I(x),\phi^{0}_J(y) \}&=&0, \qquad \{ \phi^{0}_I(x),\phi^{i}_J(y) \} = 0, \nonumber \\ \{ \phi^{0}_I(x),\phi^{0i}_J(y) \}&=&0, \qquad \{ \phi^{0}_I(x),\phi^{ij}_J(y) \} = 0, \nonumber \\ \{ \phi^{i}_I(x),\phi^{j}_J(y) \}&=&0, \qquad \{ \phi^{i}_I(x),\phi^{0j}_J(y) \} = \frac{1}{2}\delta{i}_{j}\delta_{IJ}\delta^{3}(x-y), \nonumber \\ \{ \phi^{i}_I(x),\phi^{jk}_J(y) \}&=&0, \qquad \{ \phi^{ij}_I(x),\phi^{kl}_J(y) \} = 0, \nonumber \end{aligned}$$ has rank $6(N^{2}-1)$ and $(4(N^{2}-1))$ null vectors. This means that we expect $4(N^{2}-1)$ secondary constraints $$\begin{split} \dot{\phi^{0}_I}&=\{\phi^{0}_I,H_{P}\}\approx 0 \;\; \Rightarrow \;\; \psi_{I}:=D_{i}\Pi^{i}_I\approx 0,\\ \dot{\phi^{ij}_I}&=\{\phi^{ij}_I,H_{P}\}\approx 0 \;\; \Rightarrow \;\; \psi_{ij}^I:=\frac{1}{2}F_{ij}^I\approx 0, \end{split}$$ and the rank allows us to fix the following Lagrange multipliers $$\begin{split} \dot{\phi^{0iI}}&=\{\phi^{0iI},H_{P}\}\approx 0 \;\; \Rightarrow\;\; \lambda_{i}^{I}=0,\\ \dot{\phi^{ijI}}&=\{\phi^{ijI},H_{P}\}\approx 0 \;\; \Rightarrow\;\; \lambda_{0i}^{I}=2D_{j}B^{ij}_I+2f{_{I}}^{JK}A_{0}^{J}\Pi^{iK}. \end{split}$$ For this theory there are not, third constraints. In this manner, with all the constraints at hand, we need to identify those that are first and second class kind. For this purpose, we can observe that the $14(N^{2}-1)\times 14(N^{2}-1)$ matrix whose entries are the Poisson brackets among the primary and secondary constraints given by $$\begin{aligned} \{ \phi^{0P}(x),\phi^{0I}(y) \}&=&0, \qquad \{ \phi^{0P}(x),\phi^{iI}(y) \} = 0, \nonumber \\ \{ \phi^{0P}(x),\phi^{0iI}(y) \}&=&0, \qquad \{ \phi^{0P}(x),\phi^{ijI}(y) \} = 0, \nonumber \\ \{ \phi^{lP}(x),\phi^{iI}(y) \}&=&0, \qquad \{ \phi^{lP}(x),\phi^{0iI}(y) \} = \frac{1}{2}\delta^{i}_{l}\delta^{PI}\delta^{3}(x-y), \nonumber \\ \{ \phi^{lP}(x),\phi^{ijI}(y) \}&=&0, \qquad \{ \phi^{lmP}(x),\phi^{ijI}(y) \} = 0, \nonumber \\ \{ \phi^{0P}(x),\psi^{I}(y) \}&=&0, \qquad \{ \phi^{lP}(x),\psi^{I}(y) \} = f^{PIK}\Pi^{lK}\delta^{3}(x-y), \nonumber \\ \{ \phi^{0lP}(x),\psi^{I}(y) \}&=&0, \qquad \{ \phi^{lmP}(x),\psi^{I}(y) \} = 0, \nonumber \\ \{ \psi^{P}(x),\psi^{I}(y) \}&=&f^{PIK}D_{i}\Pi^{iK}, \qquad \{ \psi^{lmP}(x),\psi^{I}(y) \} = 0, \nonumber \\ \{ \phi^{0P}(x),\psi^{ijI}(y) \}&=&0, \qquad \{ \phi^{lP}(x),\psi^{ijI}(y) \} =\frac{1}{2}\left(-\delta^{l}_{j}\delta^{PI}\partial_{i}+\delta^{l}_{i}\delta^{PI}\partial_{j}+f^{PIK}(\delta^{l}_{i}A_{j}^{K}-\delta^{l}_{j}A_{i}^{K})\right)\delta^{3}(x-y), \nonumber \\ \{ \phi^{lmP}(x),\psi^{ijI}(y) \}&=&0, \qquad \{ \psi^{lmP}(x),\psi^{ijI}(y) \} = 0, \nonumber \end{aligned}$$ has rank=$6(N^{2}-1)$ and $8(N^{2}-1)$ null-vectors. From the null-vectors it is possible to identify the following $8(N^{2}-1)$ first class constraints $$\begin{split} \gamma^{0}_I&=\Pi^{0}_I\approx 0,\\ \gamma^{ij}_I&=\Pi^{ij}_I\approx 0,\\ \gamma^{I}&=D_{i}\Pi^{iI}+2f^{I}{_{JK}}B_{0i}^{J}\Pi^{0iK},\\ \gamma_{ij}^I&=\frac{1}{2}F_{ij}^{I}+\frac{1}{2}[D_{i}\Pi{^0}{_{j}}^{I}-D_{j}\Pi{^0}{_{i}}^{I}]. \end{split} \label{65v}$$ On the other side, the rank allows us to identify the next $6(N^{2}-1)$ second class constraints $$\begin{split} \chi^{0i}_I&=\Pi^{0i}_I\approx 0,\\ \chi^{i}_I&=\Pi^{i}_I-B^{0i}_I\approx 0. \end{split}$$ However, we can observe in (\[65v\]) that the fourth constraint can be written as $$\Upsilon_{i}^I\equiv \eta_{i}{^{jk}}\gamma_{jk}^{I}=\frac{1}{2}\eta_{i}{^{jk}}F^{I}_{jk}+2\eta_{i}{^{jk}}D_{j}\Pi{^0}{_{k}}^{I},$$ thus, $D_{i}\Upsilon^{iI}-\eta_{k}{^{ij}}f^{I}{_{JK}}F_{ij}^{J}\Pi^{0kK} =0$ because of Bianchi’s identity $\eta^{ijk}D_{i}F^{I}_{jk}=0$. In this way, the former relation represents a reducibility condition. This means that the number of independent first class constraints corresponds to $[8-1](N^2-1)=7(N^2-1)$. Therefore, the counting of degrees of freedom can be carry out as follows: There are $20(N^{2}-1)$ canonical variables, $7(N^2-1)$ independent first class constraints and $6(N^{2}-1)$ independent second class constraints. Therefore, the system expressed by the action principle (\[bf-ym\]) is devoid of physical degrees of freedom and corresponds to be a topological theory. In this manner, the separated actions (\[S1\]) and (\[S2\]) are topological field theories.\ Now we will perform a pure Dirac’s analysis for the coupled action (\[bf-ym\]) and we will show that will be not topological anymore. For our aims, we perform the $3+1$ decomposition of (\[bf-ym\]) obtaining $$S[A,B]=\int\int_{\Sigma}\left[\frac{1}{2}B_{0i}^{I}B^{0iI}+\frac{1}{4}B_{ij}^{I}B^{ijI}-B^{0iI}(\dot{A}_{i}^{I}-\partial_{i}A_{0}^{I}+f^{ijk}A_{0}^{J}A_{i}^{K})-\frac{1}{2}B^{ijI}F_{ij}^{I}\right]d^{3}xdt,$$ thus, to perform a pure Dirac’s method we need the definition of the momenta $(\Pi^{\alpha I},\Pi^{\alpha\beta I})$ canonically conjugate to $(A_{\alpha}^{I},B_{\alpha\beta}^{I})$ $$\label{momenta bf} \Pi^{\alpha I}=\frac{\delta \mathcal{L}}{\delta \dot{A}_{\alpha}^{I}}, \;\;\;\;\; \Pi^{\alpha\beta I}=\frac{\delta\mathcal{L}}{\delta \dot{B}_{\alpha\beta}^{I}},$$ on the other hand, the matrix elements of the Hessian $$\frac{\partial^{2}\mathcal{L}}{\partial \dot{A}_{\alpha}^{I}\partial\dot{A}_{\rho}^{J}},\;\;\; \frac{\partial^{2}\mathcal{L}}{\partial\dot{B}_{\alpha\beta}^{I}\partial\dot{A}_{\rho}^{J}},\;\;\; \frac{\partial^{2}\mathcal{L}}{\partial\dot{B}_{\alpha\beta}^{I}\partial\dot{B}_{\rho\gamma}^{J}},$$ are identically zero, the rank of the Hessian is zero. Thus, we expect $10(N^{2}-1)$ primary constraints. From the definition of the momenta (\[momenta bf\]), we identify the next $10(N^{2}-1)$ primary constraints $$\begin{split} \phi^{0I}&:\Pi^{0I}\approx 0,\\ \phi^{iI}&:\Pi^{iI}+B^{0iI}\approx 0,\\ \phi^{0iI}&:\Pi^{0iI}\approx 0,\\ \phi^{ijI}&:\Pi^{ijI}\approx 0. \end{split}$$ The canonical Hamiltonian density for the system has the next form $$\begin{split} \mathcal{H}_{c}&=\dot{A}_{\mu}^{I}\Pi^{\mu I}+\dot{B}_{\mu\nu}^{I}\Pi^{\mu\nu I}-\mathcal{L}\\ &=\frac{1}{2}\Pi^{iI}\Pi_{i}^{I}-\frac{1}{4}B_{ij}^{I}B^{ijI}-A_{0}^{I}D_{i}\Pi^{iI}+\frac{1}{2}B^{ijI}F_{ij}^{I}. \end{split}$$ Thus the primary Hamiltonian is given by $$H_{P}=H_{c}+\int d^{3}x\left[\lambda_{0}^{I}\phi^{0I}+\lambda_{i}^{I}\phi^{iI}+\lambda_{0i}^{I}\phi^{0iI}+\lambda_{ij}^{I}\phi^{ijI}\right],$$ where $\lambda_{0}^{I},\lambda_{i}^{I},\lambda_{0i}^{I},\lambda_{ij}^{I}$ are Lagrange multipliers enforcing the constraints. The fundamental Poisson brackets for our theory are given by $$\begin{split} \{A_{\alpha}^{I}(x),\Pi^{\mu J}(y)\}&=\delta^{\mu}_{\alpha}\delta^{IJ}\delta^{3}(x-y),\\ \{B_{\alpha\beta}^{I}(x),\Pi^{\mu\nu J}(y)\}&=\frac{1}{2}\left(\delta_{\alpha}^{\mu}\delta_{\beta}^{\nu}-\delta_{\beta}^{\mu}\delta_{\alpha}^{\nu}\right)\delta^{IJ}\delta^{3}(x-y). \end{split}$$ The $10(N^{2}-1)\times 10(N^{2}-1)$ matrix whose entries are the Poisson brackets among the primary constraints are given by $$\begin{aligned} \{ \phi^{0P}(x),\phi^{0I}(y) \}&=&0, \qquad \{ \phi^{0P}(x),\phi^{iI}(y) \} = 0, \nonumber \\ \{ \phi^{0P}(x),\phi^{0iI}(y) \}&=&0, \qquad \{ \phi^{0P}(x),\phi^{ijI}(y) \} = 0, \nonumber \\ \{ \phi^{lP}(x),\phi^{iI}(y) \}&=&0, \qquad \{ \phi^{lP}(x),\phi^{0iI}(y) \} = -\frac{1}{2}\delta^{l}_{i}\delta^{PI}\delta^{3}(x-y), \nonumber \\ \{ \phi^{lP}(x),\phi^{ijI}(y) \}&=&0, \qquad \{ \phi^{0lP}(x),\phi^{0iI}(y) \} = 0, \nonumber \\ \{ \phi^{0lP}(x),\phi^{ijI}(y)\}&=&0, \qquad \{\phi^{lmP}(x),\phi^{ijI}(y) \}= 0, \nonumber \end{aligned}$$ has rank $6(N^{2}-1)$ and $4(N^{2}-1)$ null vectors. Thus by using the null vectors, consistency conditions yield the following $4(N^{2}-1)$ secondary constraints $$\begin{split} \dot{\phi}^{0I}&=\{\phi^{0I},H_{P}\}\approx 0 \;\; \Rightarrow \;\; \psi^{I}:=D_{i}\Pi^{iI}\approx 0,\\ \dot{\phi}^{ijI}&=\{\phi^{ijI},H_{P}\}\approx 0 \;\; \Rightarrow \;\; \psi^{ijI}:=B^{ijI}-F^{ijI}\approx 0, \end{split}$$ and the rank yields fix the following values for the Lagrange multipliers $$\begin{split} \dot{\phi}^{0iI}&=\{\phi^{0iI},H_{P}\}\approx 0 \;\; \Rightarrow\;\; \lambda_{i}^{I}=0,\\ \dot{\phi}^{ijI}&=\{\phi^{ijI},H_{P}\}\approx 0 \;\; \Rightarrow\;\; \lambda_{0i}^{I}=2D_{j}B^{jiI}-2f^{IJK}A_{0}^{J}\Pi^{iK}. \end{split}$$ For this theory there are not third constraints, instead we obtain the following Lagrange multipliers. $$\dot{\psi}^{lmP}=\{\psi^{lmP},H_{P}\}\approx 0 \;\; \Rightarrow\;\; \alpha_{lm}^{P}=0,\;\;\; \lambda_{lm}^{P}=D_{l}\Pi^{mP}-D_{m}\Pi^{lP}-f^{PKI}F_{lm}^{K}A_{0}^{I}$$ In this manner, with all the constraints at hand , we need identify those that are first and second class kind. For this purpose, we can observe that the $14(N^{2}-1)\times 14(N^{2}-1)$ matrix whose entries are the Poisson�s brackets among the primary and secondary constraints are given by $$\begin{aligned} \{ \phi^{0P}(x),\phi^{0I}(y) \}&=&0, \qquad \{ \phi^{0P}(x),\phi^{iI}(y) \} = 0, \nonumber \\ \{ \phi^{0P}(x),\phi^{0iI}(y) \}&=&0, \qquad \{ \phi^{0P}(x),\phi^{ijI}(y) \} = 0, \nonumber \\ \{ \phi^{lP}(x),\phi^{iI}(y) \}&=&0, \qquad \{ \phi^{lP}(x),\phi^{0iI}(y) \} = -\frac{1}{2}\delta^{i}_{l}\delta^{PI}\delta^{3}(x-y), \nonumber \\ \{ \phi^{lP}(x),\phi^{ijI}(y) \}&=&0, \qquad \{ \phi^{lmP}(x),\phi^{ijI}(y) \} = 0, \nonumber \\ \{ \phi^{0P}(x),\psi^{I}(y) \}&=&0, \qquad \{ \phi^{lP}(x),\psi^{I}(y) \} = f^{PIK}\Pi^{lK}\delta^{3}(x-y), \nonumber \\ \{ \phi^{0lP}(x),\psi^{I}(y) \}&=&0, \qquad \{ \phi^{lmP}(x),\psi^{I}(y) \} = 0, \nonumber \\ \{ \psi^{P}(x),\psi^{ijI}(y) \} &=& -f^{PIM}F_{ij}^{M}, \qquad \{ \psi^{P}(x),\psi^{I}(y) \}=f^{PIK}\psi^{K}=0 \nonumber \\ \{ \phi^{0P}(x),\psi^{ijI}(y) \}&=&0, \qquad \{ \phi^{lP}(x),\psi^{ijI}(y) \} =\left(\delta^{l}_{j}\delta^{PI}\partial_{i}-\delta^{l}_{i}\delta^{PI}\partial_{j}+f^{IPK}(\delta^{l}_{i}A_{j}^{K}+\delta^{l}_{j}A_{i}^{K})\right)\delta^{3}(x-y), \nonumber \\ \{ \psi^{lmP}(x),\phi^{ijI}(y) \}&=&(\delta^{i}_{l}\delta^{j}_{m}-\delta^{i}_{m}\delta^{j}_{l})\delta^{PI}\delta^{3}(x-y), \qquad \{ \phi^{0lP}(x),\psi^{ijI}(y) \} = 0, \nonumber \\ \{ \psi^{lmP}(x),\phi^{0I}(y) \}&=&0, \nonumber \end{aligned}$$ has rank=$12(N^{2}-1)$ and $2(N^{2}-1)$ null-vectors. From the null-vectors we identify the following $2(N^{2}-1)$ first class constraints $$\label{fc bf}\begin{split} \gamma^{0I}&=\Pi^{0I}\approx 0,\\ \gamma^{I}&=D_{i}\Pi^{iI}+2f^{IJK}B_{0i}^{J}\Pi^{0iK}+f^{IJK}B_{ij}^{J}\Pi^{ijK}\approx 0.\\ \end{split}$$ In particular, we would like to stress that the full structure of the Gauss constraint in (\[fc bf\]) has not been reported in the literature. On the other hand, that Gauss constraint is the full generator of $SU(N)$ transformations of the theory under study. Furthermore, the rank allow us to identify the following $12(N^{2}-1)$ second class constraints $$\label{sc bf}\begin{split} \chi^{iI}&=\Pi^{iI}+B^{0iI}\approx 0,\\ \chi^{0iI}&=\Pi^{0iI}\approx 0,\\ \chi^{ijI}&=\Pi^{ijI}\approx 0,\\ \phi^{ijI}&=(B^{ijI}-F^{ijI}) \approx 0. \end{split}$$ Therefore, the counting of degrees of freedom is performed as follows. There are $20(N^{2}-1)$ phase space variables, $2(N^{2}-1)$ independent first class constraints and $12(N^{2}-1)$ second class constraints, thus the theory given in (\[bf-ym\]) has $2(N^{2}-1)$ degrees of freedom just like \[YM\] theory.\ The algebra among the constraints (\[fc bf\]) and (\[sc bf\]) is given by $$\begin{aligned} \{ \gamma^{0P}(x),\gamma^{0I}(y) \}&=&0, \qquad \{ \chi^{lP}(x),\gamma^{I}(y) \} = f^{PIK}\chi^{lK}=0, \nonumber \\ \{ \gamma^{0P}(x),\chi^{iI}(y) \}&=&0, \qquad \{ \chi^{0lP}(x),\gamma^{I}(y) \} = f^{PIK}\chi^{0lP}=0, \nonumber \\ \{ \gamma^{0P}(x),\chi^{0iI}(y) \}&=&0, \qquad \{ \chi^{lmP}(x),\gamma^{I}(y) \} = f^{PIK}\chi^{lmK}=0, \nonumber \\ \{ \gamma^{0P}(x),\chi^{ijI}(y) \}&=&0, \qquad \{ \phi^{lmP}(x),\gamma^{I}(y) \} = f^{PIK}\phi^{lmK}=0, \nonumber \\ \{ \gamma^{0P}(x),\phi^{ijI}(y) \}&=&0, \qquad \{ \gamma^{P}(x),\gamma^{I}(y) \} = f^{PIK}\gamma^{K}=0, \nonumber \\ \{ \gamma^{0P}(x),\gamma^{I}(y) \}&=&0, \qquad \{ \chi^{lP}(x),\chi^{iI}(y) \} = 0, \nonumber \\ \{ \chi^{lP}(x),\chi^{0iI}(y) \} &=& -\frac{1}{2}\delta^{i}_{l}\delta^{PI}\delta^{3}(x-y), \qquad \{ \chi^{lmP}(x),\chi^{ijI}(y) \}=0 \nonumber \\ \{ \chi^{lP}(x),\chi^{ijI}(y) \}&=&0, \qquad \{ \chi^{lP}(x),\phi^{ijI}(y) \} =\left(\delta^{l}_{j}\delta^{PI}\partial_{i}-\delta^{l}_{i}\delta^{PI}\partial_{j}+f^{PIK}(\delta^{l}_{j}A_{i}^{K}-\delta^{l}_{i}A_{j}^{K}\right)\delta^{3}(x-y), \nonumber \\ \{ \chi^{lP}(x),\phi^{ijI}(y) \}&=&0, \qquad \{ \chi^{0lP}(x),\phi^{ijI}(y) \} = 0, \nonumber \\ \{ \chi^{0lP}(x),\chi^{0iI}(y) \}&=&0, \qquad \{ \chi^{lmP}(x),\phi^{ijI}(y) \} = -\frac{1}{2}\left(\delta^{l}_{i}\delta^{m}_{j}-\delta^{l}_{j}\delta^{m}_{i}\right)\delta^{PI}\delta^{3}(x-y) , \nonumber \\ \{ \phi^{lmP}(x),\phi^{ijI}(y) \}&=&0. \nonumber \end{aligned}$$ where we can appreciate that the algebra is closed.\ The identification of the constraints will allow us to identify the extended action. By using the first class constraints (\[fc bf\]), the second class constraints (\[sc bf\]), and the Lagrange multipliers we find that the extended action takes the form $$\begin{aligned} S_{E}& [A_{\mu}^{I},\Pi^{\mu I},B_{\mu\nu}^{I},\Pi^{\mu\nu I},\lambda_{0}^{I},\lambda^{I},u_{i}^{I},u_{0i}^{I},u_{ij}^{I},v_{ij}^{I}]=\int d^{4}x(\dot{A}_{\mu}^{I}\Pi^{\mu I}+\dot{B}_{\mu\nu}^{I}\Pi^{\mu\nu I}-\frac{1}{2}\Pi^{iI}\Pi_{i}^{I} +\frac{1}{4}B_{ij}^{I}B^{ijI} \nonumber\\ +A_{0}^{I}& D_{i}\Pi^{iI}-\frac{1}{2}B_{ij}^{I}F_{ij}^{I}-2D_{i}B^{ijI}\Pi^{0jI}+2f^{PKI}\Pi^{lK}A_{0}^{I}\Pi^{0lK} -2D_{l}\Pi^{mP}\Pi^{lmP}+f^{PKI}F_{lm}^{K}A_{0}^{I}\Pi^{lmP}\nonumber\\ -\lambda_{0}^{I}&\gamma^{0I}-\lambda^{I}\gamma^{I} -u_{i}^{I}\chi^{iI}-u_{0i}^{I}\chi^{0iI}-u_{ij}^{I}\chi^{ijI}-v_{ij}^{I}\phi^{ijI}).\end{aligned}$$ From the extended action we can identify the extended Hamiltonian given by $$H_{E}=H+\lambda_{0}^{I}\gamma^{0I}+\lambda^{I}\gamma^{I},$$ where $H$ is given by $$\begin{aligned} H& =\frac{1}{2}\Pi^{iI}\Pi_{i}^{I} +\frac{1}{4}B_{ij}^{I}B^{ijI} +A_{0}^{I} D_{i}\Pi^{iI}-\frac{1}{2}B_{ij}^{I}F_{ij}^{I}-2D_{i}B^{ijI}\Pi^{0jI}+2f^{PKI}\Pi^{lK}A_{0}^{I}\Pi^{0lK} \nonumber\\ -& 2D_{l}\Pi^{mP}\Pi^{lmP}+f^{PKI}F_{lm}^{K}A_{0}^{I}\Pi^{lmP}.\end{aligned}$$ We will continue this section by computing the equations of motion obtained from the extended action, which are expressed by $$\begin{aligned} \delta A_{0}^{P}& :\dot{\Pi}^{0P}=D_{l}\Pi^{lP}+2f^{JKP}\left(\Pi^{lK}\Pi^{0J}+F_{lm}^{K}\Pi^{lmJ}\right),\nonumber\\ \delta\Pi^{0P}& :\dot{A}_{0}^{P}=\lambda_{0}^{P},\\ \delta A_{l}^{P}& :\dot{\Pi}^{lP}=f^{IPK}A_{0}^{I}\Pi^{lK}+D_{i}B_{il}^{P}-2f^{KPJ}B^{ljJ}\Pi^{0jK}-2f^{IPK}\Pi^{jK}\Pi^{ljI}-2D_{i}\left(f^{KPI}A_{0}^{I}\Pi^{ilK}\right) \nonumber\\ \quad & +f^{IPK}\Pi^{lK}\lambda^{I}+2D_{i}v^{ilP},\nonumber\\ \delta\Pi^{lP}& :\dot{A}_{l}^{P}=\Pi^{lP}-D_{l}A_{0}^{P}-2f^{KPI}A_{0}^{I}\Pi^{0lK}+u^{lP}\nonumber\\ \delta B_{0l}^{P}& :\dot{\Pi}^{0lP}=f^{PIK}\Pi^{0lK}\lambda^{I}-\frac{1}{2}u^{lP} \nonumber \\ \delta B_{lm}^{P}& :\dot{\Pi}^{lmP}=2D_{l}\Pi^{0mP}+f^{PIK}\Pi^{lmK}\lambda^{I}-\frac{1}{2}u_{lm}^{P}\nonumber\\ \delta \Pi^{0lP}& :\dot{B}_{0l}^{P}=D_{i}B^{ilP}-f^{PKI}\Pi^{lK}A_{0}^{I}+\frac{1}{2}u_{0l}^{P}\nonumber\\ \delta\Pi^{lmP}& :\dot{B}_{lm}^{P}=2D_{l}\Pi^{mP}-f^{PKI}F_{lm}^{K}A_{0}^{I}+f^{IJP}\lambda^{I}B_{lm}^{J}+u_{lm}^{P}\nonumber\\ \delta\lambda_{0}^{I}& :\gamma^{0I}=0\nonumber\\ \delta\lambda^{I}& :\gamma^{I}=0\nonumber\\ \delta u_{i}^{I}& :\chi_{i}^{I}=0\nonumber\\ \delta u_{0i}^{I}& :\chi_{0i}^{I}=0\nonumber\\ \delta u_{ij}^{I}& :\chi_{ij}^{I}=0\nonumber\\ \delta v_{i}^{I}& :\phi_{i}^{I}=0\nonumber\end{aligned}$$ By following with our analysis, we need to know the gauge transformations on the phase space of the theory under study. For this step, we shall use Castellani’s formalism which allow us to define the following gauge generator in terms of the first class constraints (\[fc bf\]) $$G=\int_{\Sigma}\left[D_{0}\epsilon_{0}^{I}\gamma^{0I}+\epsilon^{I}\gamma^{I}\right]d^{3}x,$$ thus, we find that the gauge transformations on the phase space are given by $$\begin{aligned} \delta_{0} A_{0}^{P}& =D_{0}\epsilon_{0}^{P},\nonumber\\ \delta_{0} A_{i}^{P}& =-D_{i}\epsilon^{P},\nonumber\\ \delta_{0} \Pi^{0P}& = -f^{PKI}\epsilon_{0}^{K}\Pi^{0I},\nonumber\\ \delta_{0} \Pi^{iP}& = f^{PIK}\Pi^{iK}\epsilon^{I},\nonumber\\ \delta_{0} B_{0i}^{P}& = f^{PIJ}\epsilon^{I}B_{0i}^{J},\nonumber\\ \delta_{0} \Pi^{0i}& = f^{PIK}\epsilon^{I}\Pi^{0iK},\nonumber\\ \delta_{0} B_{ij}^{P}& = f^{PIJ}\epsilon^{I}B_{ij}^{J},\nonumber\\ \delta_{0} \Pi^{ijP}& = f^{PIK}\epsilon^{I}\Pi^{ijK}.\end{aligned}$$ We can observe that by redefining the gauge parameters $\epsilon_{0}^{I}=\epsilon^{I}$, the gauge transformations take the form $$\begin{aligned} A_{\mu}^{'I}& \rightarrow A_{\mu}^{I}-D_{\mu}\epsilon^{I},\nonumber\\ B_{\mu\nu}^{'I}& \rightarrow B_{\mu\nu}^{I}-f^{I}{_{JK}}\epsilon^{J}B_{\mu\nu}^{K},\end{aligned}$$ where the first one transformation corresponds to the usual gauge transformations for \[YM\] theory, and the later one by using the equations of motion gives us the transformation of a valued compact Lie algebra curvature tensor field. In order to obtain the path integral quantization of the theory and its *uv* behaviour, it is straightforward to perform Senjanovic’s method [@senjanovic] by taking into account the full constraint algebra obtained above to define the corresponding non-abelian measure. After some integration over the second class constraints, and by using the first class constraints to identify an appropriate gauge fixing [@Weinberg], one finally gets the usual quantum effective action of the \[YM\] theory. Martellini’s model ==================== An interesting alternative model to express \[YM\] theory as a constrained $BF$-like theory has been reported by M. Martellini and M. Zeni [@Martellini; @1]. Martellini’s model is a deformation of a topological field theory, namely the pure $BF$ theory resulting in the first order formulation of \[YM\] theory. In this formulation, new non local observables can be introduced following the topological theory and giving an explicit realization of t’Hooft algebra, recovering at the end the standard *u-v* behaviour of the theory [@Martellini; @2]. So, the aim of this section is to perform the canonical analysis for Martellini’s model on the full phase space context, which is absent in the literature, then we compare the results obtained with those found in former sections.\ Let us start with the action proposed by Martellini et al [@Martellini; @1] $$S[A,B]= \int \frac{i}{2} \varepsilon^{\mu \nu \alpha \beta} B_{I \mu \nu}F{^{I}}_{\alpha \beta}+ g^2\int B{^{I}}_{\mu \nu}B{_{I}}^{\mu \nu}. \label{eq52}$$ The firs term in the r.h.s. of (\[eq52\]) is the usual $BF$ theory laking of local degrees of freedom, and has been analyzed within a smaller phase space context in [@20], and by using a pure Dirac’s analysis in [@4a]. As we shall see below, local degrees of freedom are restored by the $g^2 B{^{I}}_{\mu \nu}B{_{I}}^{\mu \nu}$ term of (\[eq52\]), allowing an explicit breaking of the topological sector as long as $\mathit{g}\neq 0$. Therefore, in Martellini’s formulation \[YM\] theory is expressed as a deformation of the topological $BF$ field theory.\ We are able to observe that the actions (\[bf-ym\]) and (\[eq52\]) differ in the first term. In (\[bf-ym\]) neither is present the imaginary number that provides the euclidean feature nor the space time indices are contracted with the epsilon tensor. However, because the physical relation among \[YM\] and the action (\[eq52\]), in this section we are interested in develop a complete Hamiltonian framework of the action (\[eq52\]) because is absent in the literature.\ By performing the $3+1$ decomposition of the action (\[eq52\]) we obtain $$S[A,B]=\int\int_{\Sigma}dtd^{3}xg^{2}\left(2B_{0i}^{I}B^{0iI}+B_{ij}^{I}B^{ijI}\right)+i\eta^{ijk}\left(B_{0i}^{I}F_{jk}^{I}+B_{ij}^{I}F_{0k}^{I}\right),$$ hence, by following the procedure developed in above section, we find the following results; there are the following $2(N^{2}-1)$ first class constraints $$\label{Mfcc}\begin{split} \gamma^{0I}&=\Pi^{0I}\approx 0,\\ \gamma^{I}&=D_{i}\Pi^{iI}+2f^{IJK}B_{0i}^{J}\Pi^{0iK}+f^{IJK}B_{ij}^{J}\Pi^{ijK}\approx 0, \end{split}$$ and $12(N^{2}-1)$ second class constraints $$\label{Mscc}\begin{split} \phi^{iI}&=\Pi^{iI}-i\eta^{ijk}B_{jk}^{I}\approx 0,\\ \phi^{0iI}&=\Pi^{0iI}\approx 0,\\ \phi^{ijI}&=\Pi^{ijI}\approx 0,\\ \psi^{0iI}&=2g^{2}B^{0iI}+\frac{i}{2}\eta^{ijk}F_{jk}^{I}\approx 0. \end{split}$$ Therefore, the counting of degrees of freedom is performed as follows. There are $20(N^{2}-1)$ phase space variables, $2(N^{2}-1)$ independent first class constraints and $12(N^{2}-1)$ second class constraints, thus the theory given in (\[eq52\]) has $2(N^{2}-1)$ degrees of freedom.\ Now, we observe that the algebra of the constraints is given by $$\begin{aligned} \{ \gamma^{0P}(x),\gamma^{0I}(y) \}&=&0, \qquad \{ \phi^{lP}(x),\gamma^{I}(y) \} = f^{PIK}\phi^{lK}=0, \nonumber \\ \{ \gamma^{0P}(x),\phi^{iI}(y) \}&=&0, \qquad \{ \phi^{0lP}(x),\gamma^{I}(y) \} = f^{PIK}\phi^{0lP}=0, \nonumber \\ \{ \gamma^{0P}(x),\phi^{0iI}(y) \}&=&0, \qquad \{ \phi^{lmP}(x),\gamma^{I}(y) \} = f^{PIK}\phi^{lmK}=0, \nonumber \\ \{ \gamma^{0P}(x),\phi^{ijI}(y) \}&=&0, \qquad \{ \psi^{0lP}(x),\gamma^{I}(y) \} = f^{PIK}\psi^{0lK}=0, \nonumber \\ \{ \gamma^{0P}(x),\psi^{0iI}(y) \}&=&0, \qquad \{ \gamma^{P}(x),\gamma^{I}(y) \} = f^{PIK}\gamma^{K}=0, \nonumber \\ \{ \gamma^{0P}(x),\gamma^{I}(y) \}&=&0, \qquad \{ \phi^{lP}(x),\phi^{iI}(y) \} = 0, \nonumber \\ \{ \phi^{lP}(x),\phi^{0iI}(y) \} &=& 0, \qquad \{ \phi^{lmP}(x),\phi^{ijI}(y) \}=0 \nonumber \\ \{ \phi^{lP}(x),\phi^{ijI}(y) \}&=&-i\eta^{lij}\delta^{PI}\delta^{3}(x-y), \qquad \{ \phi^{0lP}(x),\psi^{0iI}(y) \} =-g^{2}\delta^{l}_{i}{\delta}^{PI}\delta^{3}(x-y), \nonumber \\ \{ \phi^{lP}(x),\psi^{0iI}(y) \}&=& i\eta^{ijl}(\delta^{PI}\partial_{j}+f^{PIK}A_{j}^{K})\delta^{3}(x-y), \qquad \{ \phi^{0lP}(x),\phi^{ijI}(y) \} = 0, \nonumber \\ \{ \phi^{0lP}(x),\phi^{0iI}(y) \}&=&0, \qquad \{ \psi^{lmP}(x),\psi^{ijI}(y) \} = 0 , \nonumber \end{aligned}$$ where we can appreciate that the constraints form a set of first and second class constraints as is expected.\ On the other hand, the identification of the constraints will allow us to identify the extended action. By using those results, we find the extended action given by $$\begin{aligned} S_{E}& [A_{mu}^{I},\Pi^{\mu I},B_{\mu\nu}^{I},\Pi^{\mu\nu I},\lambda_{0}^{I},\lambda^{I},u_{i}^{I},u_{0i}^{I},u_{ij}^{I},v_{0i}^{I}]=\int d^{4}x(\dot{A}_{\mu}^{I}\Pi^{\mu I}+\dot{B}_{\mu\nu}^{I}\Pi^{\mu\nu I}-\frac{1}{2}\Pi^{iI}\Pi_{i}^{I} + g^{2}2B_{0i}^{I}B^{0iI} \nonumber\\ +A_{0}^{I}& D_{i}\Pi^{iI}+i\eta^{ijk}B_{0i}^{I}F_{jk}^{I} -\frac{1}{2g^{2}}\eta^{ijk}f^{PIJ}A_{0}^{I}F_{jk}^{J}\Pi^{0iP}+\frac{i}{2}\eta^{ijk}D_{j}\Pi^{kP}\Pi^{0iP}-2D_{i}B_{0j}^{I}\Pi^{ijI} \nonumber\\ -\frac{i}{2}&\eta_{ijk}f^{PIK}A_{0}^{I}\Pi^{kK}\Pi^{ijP}-\lambda_{0}^{I}\gamma^{0I}-\lambda^{I}\gamma^{I} -u_{i}^{I} \phi^{iI}-u_{0i}^{I}\phi^{0iI}-u_{ij}^{I}\phi^{ijI}-v_{0i}^{I}\psi^{0iI}).\end{aligned}$$ From the extended action we can identify the extended Hamiltonian given by $$H_{E}=H+\lambda_{0}^{I}\gamma^{0I}+\lambda^{I}\gamma^{I},$$ where $H$ has the following form $$\begin{aligned} H& =\frac{1}{2}\Pi^{iI}\Pi_{i}^{I} -2g^{2}B_{0i}^{I}B^{0iI}-A_{0}^{I} D_{i}\Pi^{iI}-i\eta^{ijk}B_{0i}^{I}F_{jk}^{I} +\frac{1}{2g^{2}}\eta^{ijk}f^{PIJ}A_{0}^{I}F_{jk}^{J}\Pi^{0iP}-\frac{i}{2}\eta^{ijk}D_{j}\Pi^{kP}\Pi^{0iP}\nonumber\\ &+2D_{i}B_{0j}^{I}\Pi^{ijI} +\frac{i}{2}\eta_{ijk}f^{PIK}A_{0}^{I}\Pi^{kK}\Pi^{ijP}.\end{aligned}$$ Hence, the following question rise; Are there differences among the action (\[bf-ym1\]) and Martellini’s propose?. The difference lies in the constraint algebra, in fact, we observe the algebra among the second class constraints for action (\[bf-ym1\]) and Martellini’s is different. Furthermore, in Martellini’s theory the algebra among the constraints is defined over the complex numbers, consequence of a Wick rotation, and the definition of the momenta gives dual expressions of the constraints defined for the action (\[bf-ym1\]). In particular note that in Martellini’s model, $B$ is proportional to the field strength and satisfies the Bianchi identities on-shell. This is no longer true off-shell and this fact has been related to the presence of monopole charges in the vacuum [@Martellini; @2] which should enter in the non perturbative sector of the theory. Moreover the action (\[eq52\]) has been used to define new non local observables related to the phase space structure of the theory [@Cattaneo; @2]. On the other side, it is mandatory to investigate the quantum behavior of the action (\[bf-ym1\]) at perturbative level for finding new local observables, and thus, compare with Martellini’s model possibles advantages of the action (\[bf-ym1\]); we remark that the action (\[bf-ym1\]) and Martellini’s model have different algebra among the second class constraints, and this fact will be important in the quantum treatment for instance, in the construction of Dirac’s brakets. In this respect, the present letter has the necessary tools for studying these subjects in forthcoming works. Conclusions and prospects ========================== In this paper, we have developed a consistent application of a pure Dirac’s method for constrained systems. By working with the original phase space we performed a complete Hamiltonian dynamics for two $BF$-like theories. The first one, was related with \[YM\] theory, and the second action was associated with Martellini’s model, which has been used in recently works for studying the non perturbative character of the QCD confinement. From the present analysis, we calculated for the theories under study, the extended action, the extended Hamiltonian and the full constraints program, which is considerably enlarged in comparison with the analysis performed on the reduce phase space. The correct identification of the constraints as first and second class, enabled us to carry out the counting of degrees of freedom, concluding that classically, the theories under study have the same number of degrees of freedom of \[YM\] theory. The full phase space framework, allowed us observe that the physical degrees of freedom emerge from the coupling of topological theories. The topological invariance is broken because there are not in the full action reducibility conditions among the constraints, which endow the theory with local dynamics. The nature of such conditions are closely related to the full phase space, and cannot be obtained from the reduce one. With regard to the quantum aspect, the application of the pure Dirac’s procedure provides the full structure of the constraints, and this fact give us a complete gauge information of the theory. It is worth mentioning that once the full set of constraints is calculated, our procedure could shed light on the search of observables in the context of covariant field theories specifically in the case of strong-Dirac observables, which must be defined in the complete phase space. Finally, we observed that the action (\[bf-ym1\]) and Martellini’s model yield \[YM\] equations of motion, however, the algebra of their constraints is different, thus, we expect different quantum scenarios for these theories, all those ideas are in progress and will be reported in forthcoming works.\ **Acknowledgements**\ This work was supported by CONACyT México under grant 157641. [100]{} C. Rovelli, Quantum Gravity (Cambridge University Press, Cambridge, England, 2004) T.Thiemann, Modern Canonical Quantum General Relativity (Cambridge, UK: Cambridge Univ. Pr. 2007 ) M. Montesinos and A. Perez, Phys.Rev.D77:104020,2008. G.T Horowitz, Commun. Math. Phys. 125, 417, (1989). G.T Horowitz, M. Srednicki, Commun. Math. Phys. 130, 83, (1990). J. Baez, Lect.Notes Phys. 543 (2000) 25-94. J. F. Plebanski, J. Math. Phys. 18, 2511 (1977). Derek K. Wise, [*Macdowell Mansouri Gravity and Cartan Geometry, Available from: hep-th/0501191*]{} A. Accardi, A. Belli, M. Martellini and M. Zeni, hepth/9703152. F. Fucito, M. Martellini and M. Zeni, hep-th/9605018. A.S. Cattaneo, P. Cotta Ramusino, A. Gamba and M. Martellini. Phys. Lett. B355 (1995) 245. F. Fucito, M. Martellini and M. Zeni, hep-th/9607044. B. Bertand, J. Govaerts, hep-th/0704.1512v1. B. Bertrand, J. Govaerts, hep-th/0705.3452v1. L. Freidel, K. Krasnov, arXiv:0708.1595 \[gr-qc\]. G. Amelino-Camelia, L. Freidel, J. Kowalski-Glikman, L. Smolin, hep-th:1101.093. C. Di Bartolo, R. Gambini, J. Pullin, Class. Quan. Grav. 19, 5475 (2002). A. G. Lisi, [*An Exceptionally Simple Theory of Everything*]{}, arXiv:0711.0770. L. Smolin, [*The Plebanski action extended to a unification of gravity and \[YM\] theory*]{}, arXiv:0712.0977v2. A. Escalante and Leopoldo Carbajal, Annals of Physics 326, 323-339, (2011). A. Escalante, Int. J. Theo. Phys, Vol 48, No. 9, 2473-2729. (2009). A. Escalante and Ira[í]{}s Rubalcava, [*A Pure Dirac’s method for 4-dimensional BF theories*]{}, to be published in Int. J. Geom. Methods Mod. Phys (2012). P. Peldan, [*Actions for Gravity, with Generalizations: A review* ]{}, arXiv:gr-qc/930511v1. E.Buffenoir, M.Henneaux, K.Noui, Ph.Roche, Class.Quant.Grav. 21 (2004) 5203-5220. A. M. Frolov, N. Kiriushcheva and S. V. Kuzmin, Gravitation and Cosmology, 16: 181-194, (2010); E. Witten, Nucl. Phys. B 311, 46-78, (1988) . S. Carlip, Phys. Rev. D 42 (1990) 2647-2654. A. Escalante, Phys.Lett.B, 676:105-11, (2009). M. Montesinos, Class.Quant.Grav.23:2267-2278, (2006). Senjanovic P. Ann. Phys. (N.Y.) 100, 227 (1976) S. Weinberg, The Quantum Theory of Fields Vol. II, Cambridge University Press, 1996. M. Mondragon and M. Montesinos, J. Math. Phys. 47, 022301 (2006). A. S. Cattaneo, P. Cotta-Ramusino, J. Frohlich and M. Martellini, J. Math. Phys. 36 (1995) 6137. M. Martellini and M. Zeni, hep-th/9610090. [^1]: In the paper we refer as a topological theory, a classical theory lacking of local degrees of freedom, even though it does possess global physical degrees of freedom, which are characteristic by means of the topological properties either of the internal field space or of the base spacetime manifold.
{ "pile_set_name": "ArXiv" }
--- address: | $^1$Theory Group, Department of Physics and Astronomy,\ Schuster Laboratory, University of Manchester, Manchester, M13 9PL, England.\ $^2$ LPTHE, Université Paris VI and VII and CNRS UMR 7589, Paris 75005, France author: - 'R. B. Appleby$^1$ and G. P. Salam$^2$' title: 'Theory and phenomenology of non-global logarithms[^1]' --- MC–TH–2003–5\ LPTHE–P03–09\ hep-ph/0305232 Introduction and Theory {#sec:intro} ======================= Recently a distinction has been introduced between so-called global and non-global QCD observables. [@Dasgupta:2001sh] The former are sensitive to emissions in all directions, while the latter are sensitive only to emissions in some restricted angular region, for an example a jet or a hemisphere. Obvious examples of non-global (NG) observables are properties of individual jets (invariant mass, numbers of subjets). Many other common QCD observables are also non-global, including definitions of diffraction based on rapidity gaps (whether in terms of particles or energy flow); isolation criteria for photons (or other particles); distributions of interjet energy flow; or even the original Sterman-Weinberg jet definition.[ $\!$[@Sterman:1977wj]]{} The question of globalness becomes of particular relevance whenever one places a severe restriction on the energy $E$ (or in some cases, transverse energy) flowing into the observed region. In such a situation the perturbative series develops logarithmically enhanced terms at all orders, at the very least single logs ${\alpha_s}^n \ln^n E/Q$, where $Q$ is the hard scale. For $E\ll Q$ such a series needs to be resummed. For global observables it has been shown[ $\!$[@Catani:1992ua]]{} that the resummation can be carried out to single logarithmic (SL) accuracy, essentially by using the approximation of independent emission off the underlying Born event. [0.4]{} Until recently it had universally been assumed[ $\!$[@Assume]]{} that this approximation could be used more generally. However it turns out that for non-global observables, SL resummation is more subtle.[ $\!$[@Dasgupta:2001sh]]{} This is illustrated in fig. \[fig:2gloop\], which shows left and right hemispheres (${\mathcal{H_L}}$, ${\mathcal{H_R}}$) of a 2-jet $e^+e^-$ event, and two emissions going into opposite hemispheres such that $Q \gg E_1 \gg E_2$. A global observable would for example measure the sum of the two gluon energies. Since $E_1 \gg E_2$, placing a restriction $E_{\max}$ on the sum is equivalent to placing it directly on $E_1$ and there is cancellation between the real production and virtual loop contribution for gluon 2. Because of this cancellation, one is free to ‘mistreat’ the way gluon $2$ is emitted and pretend it is emitted from the simpler $q{\bar q}$ system, ignoring the presence of gluon $1$ — in other words one can make an independent emission approximation. Now suppose we have an observable that measures the energy only in ${\mathcal{H_R}}$. The limit is placed just on gluon $2$, $E_2 < E_{\max}$. On the other hand the loop contribution has an effective limit $E_{2,\mathrm{virtual}} \lesssim E_1$ and the mismatch between these two limits leads to a logarithmic enhancement $\ln E_1/E_2$. After integrating over $E_1$ one finds an overall contribution ${\alpha_s}^2 \ln^2 Q/E_{\max}$. Making an independent emission approximation, one would obtain the wrong coefficient for this term. The difference between the true answer (based on the coherent emission of gluon $2$ from the $q{{\bar q}}g_1$ system) and the independent emission result is termed a ‘non-global logarithm’ (NGL). [0.4]{} At this two-gluon level, non-global logarithms are essentially an edge effect: it is only close to the boundary between measurement and non-measurement that one is sensitive to the difference between independent emission and the true two-gluon emission pattern. This is illustrated in figure \[fig:2gcol\] which shows (through the colour shading) the contribution to the NGL as a function of the two gluons’ rapidities $\eta$ and azimuths $\phi$ for the case in which the ‘measurement’ is carried out in a square patch $|\eta| < 1$, $|\phi| < 1$.[^2] One sees clearly that the largest contribution to the NG term is concentrated around the borders of the patch. In the region where ${\alpha_s}\ln Q/E_{\max}$ is of order $1$, one needs to understand non-global effects at all orders ${\alpha_s}^n \ln^n Q/E_{\max}$. This amounts to accounting for the coherent radiation into the observed region of soft gluons from arbitrarily complex ensembles of harder (but still energy-ordered) gluons in the non-observed region. It turns out that given a na[ï]{}ve resummed calculation based on an independent emission picture, non-global effects can be accounted for by a multiplicative correction factor ${\mathcal{S}}(t)$, where $t$ is the running-coupling generalisation of $\frac{{\alpha_s}}{2\pi} \ln Q/E_{\max}$: $$\label{eq:t} t = \int_{E_{\max}}^Q \frac{dE_t}{E_t} \frac{{\alpha_s}(E_t)}{2\pi}\,.$$ It is sometimes useful also to write the expansion of ${\mathcal{S}}$, ${\mathcal{S}}(t) = \sum_{i=2}^\infty \,{\mathcal{S}}_i \,t^i$, and for example the two-gluon contribution discussed above gives us ${\mathcal{S}}_2$. The higher order terms are more difficult to calculate because of the complicated colour structures that appear for emission from multi-gluon configurations, and also because of the geometry. The colour problem has so far only been partially solved: using the large-$N_c$ approximation one can restrict one’s attention to planar graphs, equivalent to considering emission from sets of independent colour dipoles.[ $\!$[@BCM]]{} \ \ \ Two equivalent approaches have been proposed to deal with the geometry dependence. In practice, the simplest way of solving the problem of the geometry dependence is through a Monte Carlo branching algorithm,[ $\!$[@Dasgupta:2001sh]]{} in which the original $q {{\bar q}}$ dipole branches into two dipoles $qg$ and $g {{\bar q}}$. As one increases the logarithm $t$ each new dipole can itself branch and one iteratively builds up an ensemble of energy-ordered gluons with the correct (large-$N_c$) angular distribution. This algorithm is similar to that of the Ariadne event generator.[ $\!$[@Ariadne]]{} One then determines ${\mathcal{S}}(t)$ by taking the number of events at scale $t$ that are free of emissions in the observed region and dividing it by the number of events that would have been expected on the basis of an independent emission picture. Results for ${\mathcal{S}}(t)$ are shown in fig. \[fig:St\] for various geometries of observed regions. From a theoretical point of view the most remarkable feature of these curves is that modulo normalisation, they all have the same $t$-dependence (in contrast, the fixed order ${\mathcal{S}}_2$ terms differ by more than a factor of two between the different geometries). The explanation proposed for this observation[ $\!$[@Dasgupta:2002dc]]{} was so the so-called ‘buffer mechanism’, illustrated in fig. \[fig:buffer\]: the hypothesis is that the easiest way of forbidding ‘secondary’ emissions into the observed region is actually to forbid primary emissions close to the observed region. Accordingly for intermediate scales $t' < t$ one finds a buffer region, free of emissions, surrounding the observed region. Larger differences $t-t'$ imply a larger buffer region. As a result the dynamics governing the large-$t$ behaviour of ${\mathcal{S}}$ mostly occurs far from the observed the region and is not affected by the geometry of the observed region. Assuming that the buffer region is of size $\delta \eta_\mathrm{buffer} \simeq c C_A (t - t')$ one comes to the conclusion that $$\label{eq:asympS} {\mathcal{S}}(t) \sim e^{-2 c C_A^2 t^2} .$$ These arguments were placed on a mathematically sound footing by Banfi, Marchesini and Smye[ $\!$[@Banfi:2002hw]]{}. Firstly they introduced an alternative but equivalent approach to the problem of the geometry, in terms of a non-linear integral equation: $$\label{eq:BMS} \partial_t \,{\mathcal{S}}_{ab}(t) = \int_{\mathrm{unmeasured}} \frac{d^2\Omega_k}{4\pi}\, w_{ab}(k) \left(U_{abk}^{(0)}(t)\, {\mathcal{S}}_{ak}(t)\, {\mathcal{S}}_{kb}(t) - {\mathcal{S}}_{ab}(t)\right),$$ where ${\mathcal{S}}_{ab}(t)$ is the non-global correction factor for a dipole $ab$, $w_{ab}(k)$ is the weight for emission of a soft gluon $k$ from $ab$ and $U^{(0)}_{abk}(t)$ accounts for the different ‘primary’ emission contributions for dipoles $ak,kb$ compared to dipole $ab$. They were then able to demonstrate the existence of scaling solutions to , proving the buffer mechanism, with $c$ evaluated numerically to be $2.5\pm0.25\%$. They also evaluated a number of subleading corrections. They pointed out however that in practice applies only for $t \gtrsim 0.5$, i.e. very asymptotic values of $t$. Phenomenologically, $t$ is limited to be $\lesssim 0.2$ and one should use the full solutions to ${\mathcal{S}}$. Warnings for practitioners, a.k.a. Zoology ------------------------------------------ The definition of a non-global observable, given above, is one that is sensitive only to emissions (from the Born event) in a restricted angular region. However the situations in which NGLs can appear are rather subtle. Firstly there exist observables that measure only a subset of the particles, but which nevertheless are global. An example is the heavy-hemisphere invariant squared mass in $e^+e^-$. Though only one hemisphere is measured, the observable is global because it is always the heavier hemisphere that is measured — in a situation such as fig. \[fig:2gloop\] the heavier hemisphere is always the one with the harder gluon (1) and one therefore never has the situation of a harder unobserved particle radiating into the observed region. Other examples include certain DIS event shapes (e.g. $B_{zE}$, $\tau_{zQ}$) where the measurement is carried out in one current hemisphere, but for which conservation of momentum causes an indirect sensitivity to emissions in the unobserved hemisphere. Such observables are known as indirectly global. Another subtle case is that of observables referred to as *discontinuously* global. Such observables have different *parametric* sensitivities to emissions in different directions, e.g. $v \sim E_t^2/Q^2$ for emissions in one hemisphere, $v \sim E_t/Q$ for the other. Placing a limit on $v$ corresponds to different limits on $E_t$ in the two hemisphere (e.g. $\sqrt{v}Q$ and $vQ$ respectively) and one finds NGLs as for simple non-global observables, but with the appropriate replacement of the integration limits for $t$ in eq. .[ $\!$[@Dasgupta:2002dc; @Dokshitzer:2003uw]]{} The most subtle situation is perhaps that of *dynamically* discontinuously global observables. These are observables which for a single emission appear global (typically indirectly global). However they involve non-linearities such that for the configurations of emissions that are most common (e.g. given a certain value of the observable) they develop different effective parametric dependence on emissions in different regions.[ $\!$[@Dasgupta:2002dc]]{} One example of such an observable is the broadening $B_{zE}$ in DIS. Phenomenological implications ============================= In this section we will examine the phenomenological impact of NGLs in a variety of non-global observables. We will look at the invariant-squared jet mass in DIS[ $\!$[@Dasgupta:2002dc]]{} (although the conclusions we draw will apply to $e^+e^-$ non-global event shapes as well) and at 2-jet energy flows.[ $\!$[@Dasgupta:2002bw]]{} In both kinds of observable we will see that by neglecting the non-global logarithm suppression factor, one overestimates the distributions and cross sections by a considerable, and certainly phenomenologically relevant, amount. NGLs in DIS and $e^+e^-$ event shapes ------------------------------------- Event shapes (ES), and their distributions,[ $\!$[@Dasgupta:2002dc]]{} are widely used to compare the predictions of perturbative QCD with experimental observation. As an example of a non-global event shape, we will use the DIS invariant squared jet mass $\rho$, defined as $$\rho=\frac{\left(\Sigma_{\mathcal{H}_C} P_i\right)^2}{4\left(\Sigma_{\mathcal{H}_C} |\vec{P}_i|\right)^2},$$ where the summation is over all particles in the current-hemisphere ($\mathcal{H}_C$). When calculating the cross section for $\rho$ to be smaller than some given value, one finds that in the exclusive limit, small $\rho$, each power of the coupling is accompanied by up to two powers of $\ln \rho$, associated with with soft and collinear divergences. These terms need to be resummed. The standard accuracy is next-to-leading logarithmic, which in this case implies the inclusion of all single-logarithmic terms, i.e. the same accuracy as NG contributions. In figure \[fig:rho\] we show the effect of the non-global logarithms on the resummed distribution for $\rho$. The broken line is the resummation without NGLs, and the solid line shows the result once they are included through the function $\mathcal{S}$ discussed above. Neglecting the non-global logarithms leads to an overestimation of the peak-height by around 40%, while the effect is smaller in other parts of the distribution. In practice, the resummed calculation is usually matched to a fixed-order calculation, which includes the NGLs to order ${\alpha_s}^2$ and so reduces the impact of neglecting the NGLs in the final result. NGLs in 2 jet energy flow observables ------------------------------------- Let us now examine the impact of non-global logarithms on energy flow observables.[ $\!$[@Dasgupta:2002bw]]{} Such observables have attracted considerable interest in recent time as an infrared-safe way of studying gaps-between-jets processes[ $\!$[@OdeSter]]{} and the underlying event in hadron-hadron collisions.[ $\!$[@MW]]{} We shall take as our observable the total amount of transverse energy flowing into a restricted region of phase space $\Omega$. The probability for the amount of transverse energy to be smaller than some value $Q_\Omega$ is $$\Sigma_{\Omega}(Q_\Omega) =\frac{1}{\sigma}\int_0^{Q_{\Omega}} d\,E_t \frac{d\sigma}{dE_t}\,.$$ As discussed in section \[sec:intro\] this can be written as the product of two contributions, $$\Sigma_{\Omega}(Q_\Omega) = \mathcal{S}(t)\times \Sigma_P(t)\,,$$ \ \ where the function $\Sigma_P$ is based on an independent gluon emission approximation (‘primary’ radiation into $\Omega$), while $\mathcal{S}(t)$ accounts for secondary coherently radiated emission into $\Omega$ and $t$ is defined as the integral of ${\alpha_s}/2\pi$ between $Q_\Omega$ and $Q$. We can understand the phenomenological implications of the non-global logarithms on this observable by comparing the result for $\Sigma_{\Omega}$, calculated with only primary emissions, with the full result for $\Sigma_{\Omega}$, which accounts also for non-global logarithms. We do this in figure \[fig:slice\], where $\Omega$ is a slice in rapidity of width $\Delta\eta=1.0$. The plot clearly show the phenomenological impact of the NGLs on this observable; the suppression is significant, particularly at larger values of the effective logarithm $t$. The typical energies of current colliders corresponds to about $t=0.15$ and if we take this as our reference value, the inclusion of non-global logarithms increases the suppression, relative to the primary-only case, by a factor of about 1.65. Similar results are found for other definitions of $\Omega$, for example a patch in phase space bounded in rapidity and azimuthal angle. Controlling non-global logarithms ================================= In this section we will describe ways of controlling, or taming, non-global logarithms. Such a study is important given that our methods for resumming NG logarithms are both approximate (large-$N_C$ limit) and cumbersome (numerical or asymptotic). We will look at two different approaches: minimising the numerical effect of the NGLs by clustering the final state, and also by examining associated distributions of pairs of observables. Clustering algorithms --------------------- The application of clustering algorithms[ $\!$[@Catani:1993hr]]{} to the final state is motivated by recent H1 and ZEUS analyses[ $\!$[@Adloff:2002em]]{} of gaps-between-jets processes at HERA. In these analyses, the inclusive $k_t$ algorithm is used to define the hadronic final state, and hence the rapidity gap $\Omega$, and a gap event is defined by a restriction of the total transverse energy into $\Omega$. This observable, sensitive to soft radiation into $\Omega$ only, is clearly non-global, and so sensitive to non-global logarithmic effects. However, as we will show, the clustering procedure reduces the numerical importance of the NGLs for this observable,[ $\!$[@Appleby:2002ke]]{} and we can see why by looking at how the $k_t$ clustering algorithm works. The essential feature is that in an iterative algorithm over all the final state particles, it merges particles of lower transverse momentum into particles with higher transverse momentum, to produce pseudo-particle or mini-jets. Consider applying this algorithm to two gluons with strongly ordered transverse momentum, $$E_{T,1}\gg E_{T,2},$$ where gluon one, directed outside of the region $\Omega$, then in turn radiates gluon two into the region $\Omega$. The strong ordering ensures this configuration produces a non-global logarithm. When we cluster this system, gluon two will be clustered into, or merged with, gluon one (and hence pulled out of the gap) if the two gluons are sufficiently close in the $(\eta,\phi)$ plane, $$(\eta_1-\eta_2)^2+(\phi_1-\phi_2)^2<R^2,$$ where $R$ is a parameter playing the role of a radius in the algorithm. Therefore to get a kinematical configuration which produces a NGL, the two gluons need to be sufficiently separated in the $(\eta,\phi)$ plane to avoid being merged. Hence the $k_t$ algorithm ‘cleans up’ the restricted region of phase space, $\Omega$, and pulls soft gluons out of the gap. Figure \[fig:s2\] shows the leading, ${\mathcal{S}}_2$ contribution (order $\alpha_s^2$) to the NGL function $\mathcal{S}(t)$, without clustering (solid line) and with clustering (broken line.) In both cases ${\mathcal{S}}_2$ rapidly saturates at high $\Delta\eta$, which results from the fact that the NGLs are an edge effect, and the saturation value for the clustered case is smaller than that of the non-clustered case. This follows from the fact that, although the clustering algorithm pulls gluons out of the gap, gluons can still be sufficiently separated in the $(\eta,\phi)$ plane to survive clustering and give a significant NGL contribution. Figure \[fig:sigmaslice\] shows the all-orders calculation of the full function $\Sigma$, where $\Omega$ is taken to be a slice in rapidity of $\Delta\eta=1.0$. These figures allow us to see the phenomenological impact of the non-global logarithms on these observables, as they show the function $\Sigma_P$ (only primary logarithms) and the full function $\Sigma$ (primary and non-global logarithms) with and without clustering. We recall that $t$ is around $0.15$ to $0.2$ at the energies of current colliders, and as before we will take $t=0.15$ as our reference value. The plot shows that the non-global logarithms cause a considerable suppression of $\Sigma$, relative to the primary-only result, and that by clustering the final state this suppression is reduced. At $t=0.15$ the full result without clustering is suppressed relative to the primary only result by $1.65$, and this suppression is reduced with clustering to around $1.2$. Event shape/Energy flow Correlations ------------------------------------ Another way of reducing the phenomenological impact of NGLs that has been proposed,[ $\!$[@Berger:2003iw]]{} is the study of associated distributions in two variables. In this work, one combines measurement of a jet shape $V$ in the whole of phase space (for example thrust, $V=1-T$) and that of the transverse away-from-jets energy flow $E_{\mathrm{out}}$. The former is a global measurement and the latter is a non-global measurement. If the observable $V$ selects 2-jet-like configurations, one measures the associated distribution, $$\Sigma_{\mathrm{2ng}}(Q,V,E_{out}),$$ where $Q$ is the hard scale. It has been shown that this distribution factorizes,[ $\!$[@Dokshitzer:2003uw]]{} $$\Sigma_{\mathrm{2ng}}(Q,V,E_{out})=\Sigma(Q,V) \cdot\Sigma_{\mathrm{out}}(VQ,E_{\mathrm{out}}),$$ where $\Sigma(Q,V)$ is the standard global distribution of $V$ and $\Sigma_{\mathrm{out}}(VQ,E_{\mathrm{out}})$ contains the logarithmic distribution in $E_{\mathrm{out}}$. This latter distribution, containing non-global logarithms is evaluated at the reduced scale $VQ$, and hence the logarithmic terms will be $(\alpha_s \log(VQ/E_{\mathrm{out}}))^n$. The work of Berger, Kúcs and Sterman[ $\!$[@Berger:2003iw]]{} considered the region in which $VQ$ and $E_{\mathrm{out}}$ were comparable, so that the NGLs give a negligible contribution. Thus, for a restricted subset of appropriately selected events, it is possible, to ‘tune out’ the non-global logarithmically enhanced terms in associated distributions. Conclusion ========== To summarise, non-global logarithms are recently discovered contributions that arise in the distributions of any QCD observable sensitive only to emissions in a restricted part of phase space. They are phenomenologically important and significant progress has been made in resumming them to all orders in the large-$N_c$ limit. One of the main directions of current work focuses on understanding ways of designing observables so as to reduce the impact of non-global contributions. Acknowledgements {#acknowledgements .unnumbered} ================ We are grateful to the organisers and secretarial staff of the QCD session of Moriond 2003 for a stimulating and enjoyable conference. RBA would like to acknowledge the University of Manchester for financial support. References {#references .unnumbered} ========== [99]{} M. Dasgupta and G. P. Salam, Phys. Lett. B [**512**]{} (2001) 323. G. Sterman and S. Weinberg, Phys. Rev. Lett.  [**39**]{} (1977) 1436. S. Catani, L. Trentadue, G. Turnock and B. R. Webber, Nucl. Phys. B [**407**]{} (1993) 3. For example in references 6, 10, 11 and 12 of M. Dasgupta and G. P. Salam, Acta Phys. Polon. B [**33**]{} (2002) 3311. A. Bassetto, M. Ciafaloni and G. Marchesini, Phys. Rept. [**100**]{} (1983) 201. L. Lönnblad, Comp. Phys. Comm. [**71**]{} (1992) 15. M. Dasgupta and G. P. Salam, JHEP [**0208**]{} (2002) 032. A. Banfi, G. Marchesini and G. Smye, JHEP [**0208**]{} (2002) 006. Yu. L. Dokshitzer and G. Marchesini, JHEP [**0303**]{} (2003) 040. M. Dasgupta and G. P. Salam, JHEP [**0203**]{} (2002) 017. G. Oderda and G. Sterman, Phys. Rev. Lett. [**81**]{} (1998) 3591. G. Marchesini and B.R. Webber, Phys. Rev. D [**38**]{} (1988) 3419. S. Catani, Yu. L. Dokshitzer, M. H. Seymour and B. R. Webber, Nucl. Phys. B [**406**]{} (1993) 187. C. Adloff [*et al.*]{} \[H1 Collaboration\], Eur. Phys. J. C [**24**]{} (2002) 517. R. B. Appleby and M. H. Seymour, JHEP [**0212**]{} (2002) 063. C. F. Berger, T. Kúcs and G. Sterman, hep-ph/0303051. [^1]: Based on talks presented at the XXXVIIIth Rencontres de Moriond ‘QCD and high-energy hadronic interactions’. [^2]: When viewed in colour, the blue and red shadings are therefore respectively for the unmeasured and measured gluons.
{ "pile_set_name": "ArXiv" }
--- abstract: | The present paper is the first one in the sequence of papers about a simple class of [*framed $4$-graphs*]{}; the goal of the present paper is to collect some well-known results on planarity and to reformulate them in the language of [*minors*]{}. The goal of the whole sequence is to prove analogues of the Robertson-Seymour-Thomas theorems for framed $4$-graphs: namely, we shall prove that many minor-closed properties are classified by finitely many excluded graphs. From many points of view, framed $4$-graphs are easier to consider than general graphs; on the other hand, framed $4$-graphs are closely related to many problems in graph theory. [**Keywords:**]{} graph, $4$-valent, minor, planarity, embedding, immersion, Wagner conjecture. author: - Vassily Olegovich Manturov title: 'Framed $4$-valent Graph Minor Theory I: Intoduction. A Planarity Criterion and Linkless Embeddability' --- [**AMS MSC**]{} 05C83, 57M25, 57M27 Some years ago, a milestone in graph theory was established: as a result of series of papers by Robertson, Seymour (and later joined by Thomas) [@RS20] proved the celebrated Wagner conjecture [@Wagner] which stated that if a class of graphs (considered up to homeomorphism) is minor-closed (i.e., it is closed under edge deletion, edge contraction and isolated node deletion), then it can be characterized by a finite number of excluded minors. For a beautiful review of the subject we refer the reader to L.Lovász [@Lovasz]. This conjecture was motivated by various evidences for concrete natural minor-closed properties of graphs, such as knotless or linkless embeddability in $\R^{3}$, planarity or embeddability in a standardly embedded $S_{g}\subset \R^{3}$. Here we say that a property $P$ is [*minor-closed*]{} if for every graph $X$ possessing this property every minor $Y$ of $G$ possesses $P$ as well. Later, we shall define the notion of [*minor*]{} in a way suitable for framed $4$-graphs. The most famous evidence of this conjecture is the [*Pontrjagin-Kuratowski*]{} planarity criterion which states (in a slightly different formulation) that a graph is not planar if and only if it contains one of the two graphs shown in Fig. \[PK\] as a minor. ![The two Kuratowski graphs, $K_{5}$ and $K_{3,3}$[]{data-label="PK"}](pk.eps){width="200pt"} Throughout the paper (and all subsequent papers in the series), all graphs are assumed to be finite; loops and multiple edges are allowed. Among all graphs, there is an important class of four-valent [*framed*]{} graphs (or [*framed*]{} regular $4$-graphs). Here by [*framing*]{} we mean a way of indicating which half-edges are opposite at every vertex. Whenever drawing a framed four-valent graph on the plain, we shall indicate its vertices by solid dots, (self)intersection points of edges will be encircled, and the framing is assumed to be induced from the plane: those half-edges which are drawn opposite in $\R^{2}$ are thought to be opposite. Half-edges of a framed four-valent graph incident to the same vertex are which are not [*opposite*]{}, are called [*adjacent*]{}. This class of graph is interesting because of its close connection to classical and virtual knot theory [@ManDkld; @ManHdlbg], homotopy classes of curves on surfaces, see also [@FM1; @FM2]; for more about virtual knot theory see [@MI]. From time to time we shall admit some broader class of objects than just framed four-valent graphs. By a [*$4$-graph*]{} we mean a finite $1$-complex with every component either being homeomorphic to a circle or being a graph with all vertices having valency $4$; components of a $4$-graph homeomorphic to circles will be called [*circular components*]{} or [*circular edges*]{}; by a [*vertex*]{} of a framed $4$-graph we mean a vertex of its non-circular component. By a ([*non-circular*]{}) edge of a $4$-graph we mean an edge of its non-circular component. A $4$-graph is [*framed*]{} if all non-circular components of it are framed and all circular components of it are oriented. There are some natural ways to extend the notion of [*minor-closed property*]{} to four-valent framed graphs. A framed $4$-valent graph $G'$ is a [*minor*]{} of a framed $4$-valent graph $G$ if $G'$ can be obtained from $G$ by a sequence of [*smoothing*]{} operations (${\raisebox{-0.25\height}{\includegraphics[width=0.5cm]{skcr.eps}}}\to {\raisebox{-0.25\height}{\includegraphics[width=0.5cm]{skcrv.eps}}}$ and ${\raisebox{-0.25\height}{\includegraphics[width=0.5cm]{skcr.eps}}}\to {\raisebox{-0.25\height}{\includegraphics[width=0.5cm]{skcrh.eps}}}$) and deletions of connected components. Whenever talking about embedding or immersion of a framed $4$-graph into any $2$-surface we always assume its framing to be preserved: opposite edges at every crossing should be locally opposite on the surface. We say that a framed $4$-graph $\Gamma$ admits a source-sink structure if there is an orientation of all edges $\Gamma$ such that at every vertex of $\Gamma$ some two opposite edges are incoming, and the other two are emanating. Certainly, for every connected framed four-valent graph, if a source-sink structure exists, then there are exactly two such structures. Moreover, it can be easily seen that if $\Gamma$ admits a source-sink structure then every minor $\Gamma'$ of $\Gamma$ admits a source-sink structure as well. Indeed, the smoothing operation can be arranged to preserve the source-sink structure. So, it is natural to ask many questions about graphs admitting a source-sink structure. In the present paper, we restrict ourselves to framed $4$-graphs admitting source-sink structures. Framed $4$-graphs not admitting source-sink structures will be considered in subsequent papers. Denote by $\Delta$ the following framed $4$-graph with $3$ vertices: it has $3$ vertices $P,Q,R$, and $6$ edges $a,a',b,b',c,c'$ such that at vertex $P$ the edges $a$ and $a'$ are opposite and both connect $P$ to $Q$ (in $Q$ they are opposite, as well); $b,b'$ constitute the other pair of opposite edges at $P$; they connect $P$ to $R$, and they are opposite at $R$ as well; finally, $c$ and $c'$ are edges connecting $Q$ and $R$; these edges are opposite both at $Q$ and at $R$. ![The Graph $\Delta$[]{data-label="Delta"}](delta.eps){width="200pt"} When drawn immersed in $\R^{2}$, the graph $\Delta$ contains three pairwise intersecting cycles $(a,a'),(b,b'),(c,c')$; each two of these three cycles intersect transverselly at one point; thus, an immersion requires at lease one intersection point for each pair of these two cycles. In Fig. \[Delta\] these three immersion points are encircled. For a framed $4$-graph $P$ by a [*loop*]{} we mean either a circular component (also treated as a map $S^{1}\to P$) or a map $f:S^{1}\to \Gamma$ which is a bijection everywhere except preimages of crossings of $\Gamma$. A [*loop*]{} is a circuit if its image is the whole graph $P$ (certainly, only connected framed $4$-graphs admit circuits). A [*loop*]{} (resp., circuit) is [*rotating*]{} if at every crossing $X$ which has two preimages $Y_{1}$ and $Y_{2}$, the neighbourhoods of $Y_{1}$ is mapped to two [*non-opposite*]{} edges. By abuse of notation, we shall say that a loop (a circuit) passes through edges if its image contains these edges. Let $L_{1}, L_{2}$ be two loops of a framed $4$-graph $P$; let $X$ be a crossing of $P$; we say that $L_{1}$ and $L_{2}$ intersect transversely at $X$ if $L_{1}$ passes through a pair of opposite edges at $X$ as well $L_{2}$. By a [*chord diagram*]{} we mean either an oriented circle ([*empty*]{}) chord diagram or a cubic graph $D$ consisting of an oriented cycle (the [*core*]{}) passing through all vertices of $D$ such that the complement to it is a disjoint union of edges ([*chords*]{}) of the diagram. An easy exercise (see, e.g. [@ManVasConj]) shows that [*every connected framed $4$-graph admits a rotating circuit*]{}. Having a circuit $C$ of a framed connected $4$-graph $G$, we define the chord diagram $D_{C}(G)$, as follows. If $G$ is a circle, then $D_{C}(G)$ is empty. Think of $C$ as a map $f:S^{1}\to D$; then we mark by points on $S^{1}$ preimages of vertices of $G$. Thinking of $S^{1}$ as a core circle and connecting the preimages by chords, we get the desired cubic graph. Chord diagrams are considered up to combinatorial equivalence. One can associate chord diagrams not only to [*rotating circuits*]{}, but for the present paper we restrict ourselves only with rotating circuits and framed $4$-graphs admitting a source-sink structure. The opposite operation (of restoring a framed $4$-graph with a source-sink structure from a chord diagram) is obtained by removing chords from the chord diagram and approaching two endpoints of each chord towards each other as shown in Fig. \[CDgr\]. ![Restoring a framed $4$-graph from a chord diagram[]{data-label="CDgr"}](cdgr.eps){width="200pt"} A chord diagram $D'$ is called a [*subdiagram*]{} of a chord diagram $D$ if $D$ can be obtained from $D$ by deleting some chords and their endpoints. It follows from the definition that the removal of a chord from a chord diagram results in a smoothing of a framed $4$-graph. Consequently, if $D'$ is a subdiagram of $D$, then the resulting framed $4$-graph $G(D')$ is a [*minor*]{} of $G(D)$. Every embedding $i:P\to \R^{3}$ gives rise to an embedding of every rotating circuit $C$ of $P$: at each vertex where $C$ touches itself we perform a smoothing. We say that two rotating circuits $C_{1},C_{2}$ sharing no edges are [*not transverse*]{} if at every vertex which belongs to both $C_{1}$ and to $C_{2}$ the edges incident to $C_{1}$ are not opposite at this vertex. Any embedding of a framed $4$-valent graph in $\R^{3}$ is assumed to be smooth in the following sense: in the neighbourhood of every vertex $X$ we require that tangent vectors of opposite half-edges are opposite. Thus, having a framed $4$-graph $P$ and an embedding $i:P\to \R^{3}$, we may assume without loss of generality that the small neighbourhood of every vertex $X$ of $P$ is mapped to a piece of a $2$-surface containing $X$. Thus, having two rotating loops $L_{1},L_{2}$ of $P$ with no transverse intersections we can define the associate the disjoint embedding of $L_{1}$ and $L_{2}$ in $\R^{3}$ obtained by local smoothing at some vertices. By abusing notation, we shall talk about [*images of loops or circuits*]{} in $\R^{3}$ meaning the cooresponding smoothings (which represent collection of disjoint curves in $\R^{3}$. An embedding $i$ of a framed $4$-graph $P$ in $\R^{3}$ with a source-sink structure is called [*linkless*]{} if for every two rotating loops $L_{1}, L_{2}$ without transverse intersection the linking number of their images is $0$. Analogously, an embedding $i$ of a framed $4$-graph $P$ in $\R^{3}$ with a source-sink structure is [*knotless*]{} if the image of the every rotating loop $L$ is unknotted. This means that in the neighbourhood of such a vertex we can perform a smoothing of $X$ and an embedding $i$ gives rise to embeddings of all minors of $P$ defined up to homotopy. Now we list some [*minor*]{} properties of framed $4$-valent graphs (the proof is left for the reader): 1. Planarity. 2. Existence of an immersion into a fixed surface $\Sigma$ with no more than $s$ transverse simple intersection points ($s$ fixed). 3. Linkless embeddability (in $\R^{3}$). 4. Knotless embeddability (in $\R^{3}$). The Main Theorem of the present paper sounds as follows Let $\Gamma$ be a framed $4$-graph admitting a source-sink structure. Then the following four conditions are equivalent: 1. Every generic immersion of $\Gamma$ in $\R^{2}$ requires at least $3$ additional crossings; 2. For every embedding of $\Gamma$, there exists a pair of rotating loops with odd linking number. 3. $\Gamma$ has no linkless embedding in $\R^{3}$; 4. $\Gamma$ is not planar; 5. $\Gamma$ contains $\Delta$ as a minor. Certainly, 1) yields 4) and 3) yields 4): a planar graph has a planar [*embedding*]{} which is an immersion with [*no additional points*]{}; moreover, a planar embedding is always linkless. Our goal is to prove that the non-planarity of a framed $4$-graph with a source-sink structure yields the existence of $\Delta$ as a minor. After that, we see that every immersion of $\Delta$ requires at least $3$ points, which is obvious, and prove that there for every embedding of $\Delta$ in $\R^{3}$, there exists a pair of rotating loops without crossing points having [*odd linking number*]{}. The latter automatically means that the embedding is not linkless. We follow the proof of Vassiliev’s conjectutre [@Vas] from [@ManVasConj]. Take a rotating circuit $C$ for $\Gamma$; by assumption, $\Gamma$ admits a source-sink structure, thus, the chord diagram $D_{C}(\Gamma)$ contains a $(2n+1)$-gon $\Delta_{2n+1}$ as a subdiagram, see Fig. \[2ngon\]. ![A $(2n+1)$-gon[]{data-label="2ngon"}](tngon.eps){width="200pt"} Consequently, the initial graph will have a minor which corresponds to the chord diagram $\Delta_{2n+1}$; we denote this framed $4$-graph by $Z_{2n+1}$. Now, we apply the following fact whose prove is left to the reader as an exercise: $\Delta$ is a minor of $Z_{2n+1}$ for every natural $n$. Thus, $\Delta$ is a minor of $\Gamma$, as required. Let us now prove that there is no linkless embedding of $\Delta$ in $\R^{3}$; consequently, none exits for $\Gamma$. Indeed, let us consider the immersion given in Fig. \[immer\]. ![An immersion of $\Delta$ in $\R^{3}$[]{data-label="immer"}](deltaimm.eps){width="200pt"} Let us consider the following four pairs of cycles $F_{1}=(a,b,c), F_{2}=({a',b',c'}), G_{1}=(a,b,c'),G_{2}=(a',b',c),H_{1}=(a,b',c),H_{2}=(a',b,c'), I_{1}=(a',b,c),I_{2}=(a,b',c')$. For the immersion given in Fig. \[immer\] we see that the linking numbers are $lk(F_{1},F_{2})=0$, whence all linking numbers $lk(G_{1},G_{2}), lk(H_{1},H_{2}),lk(I_{1},I_{2})$ are congruent to $1$ modulo $2$. Thus, the sum of these four linking numbers is odd. Now, linking numbers do not change under homotopy; thus, this sum remains odd when applying homotopy to the immersion given in Fig. \[immer\]. Besides homotopy, we can apply some crossing switches in $3$-space. The whole graph $\Delta$ consists of $6$ edges; if we apply a crossing switch to an edge with itself (say, $a$ with $a$), it will make no effect in any of the four summands. Now, if we apply a crossing switch for an edge with a dash and a corresponding edge without a prime (say, $a$ and $a'$), this will result in changes modulo $2$ for all four summands; thus, the total sum will remain odd. In the case when we have two letters either both without primes or both with primes (without loss of generality we may assume they are $a$ and $b$), two of four summands will remain the same and the other two will change. Consequently, the parity will remain the same. Finally, if we apply a crossing switch to some edges which are not opposite at some vertex (without loss of generality, we may assume we deal with $a$ and $b'$), this will change two of four summands: namely, $lk(F_{1},F_{2})$ and $lk(G_{1},G_{2})$ will change by one. Thus, the total parity of the sum of linking numbers will not change. Thus, we conclude that at least one of these four crossing numbers will be odd. I am grateful to Igor Mikhailovich Nikonov for valuable comments and to Denis Petrovich Ilyutko for useful discussions. [100]{} Friesen, T., Manturov, V.O., Embeddings of $*$-graphs into $2$-surfaces, [*Journal of Knot Theory and Its Ramifications*]{}, Vol. 22, No. 11 (2013), 1341005 (15 pages) Friesen, T.,Manturov, V.O., Checkerboard embeddings of \*-graphs into nonorientable surfaces, (2013) [*arxiv.Math: CO*]{}, 1312.6342 L.Lovász, Graph Minor Theory, [*Bull. AMS*]{} (New Series), Vol. 43, No. 1, pp. 75–86. Manturov, V. O. (2005), A proof of Vassiliev’s conjecture on the planarity of singular links *Izv. Ross. Akad. Nauk Ser. Mat.* **69** (5) 169-178 Manturov, V.O. (2008), Embeddings of $4$-valent framed graphs into $2$-surfaces, [*Doklady Mathematics*]{}, 2009, Vol. 79, No. 1, pp. 56–58. (Original Russian Text © V.O. Manturov, 2009, published in Doklady Akademii Nauk, 2009, Vol. 424, No. 3, pp. 308–310). Manturov, V.O., Embeddings of Four-valent Framed Graphs into 2-surfaces, [*The Mathematics of Kntos. Theory and Applications*]{}, Contributions in Mathematical And Computational Sciences 1., M.Banagl, D.Vogel, Eds., Springer, pp. 169-198. . Virtual Knots: The State of the Art, World Scientic, Singapore, 2012, 547 pp. I. Nikonov, A New Proof of Vassiliev’s Conjecture, (2013) [*arxiv.Math: CO*]{}, 1306.5521 N. Robertson, P.D. Seymour: Graph minors XX. Wagner’s Conjecture J. Combin. Theory Ser. B, [**92**]{}, (2), November 2004, pp. 325–357. V. A. Vassiliev, (2005), First-order invariants and cohomology of spaces of embeddings of self-intersecting curves, [*Izv. Math.*]{}, 69:5 (2005), pp. 865–912 K. Wagner: Graphentheorie, B.J. Hochschultaschenbucher 248/248a, Mannheim (1970), 61. MR0282850 (44:84)
{ "pile_set_name": "ArXiv" }
--- abstract: 'We prove a family of results regarding connectivity in the theory of chiral Koszul duality. This provides new examples of Koszul duality being an equivalence, even when the base category is not pro-nilpotent in the sense of [@francis_chiral_2011]. Based on ideas sketched in [@gaitsgory_contractibility_2011], we show that these results also offer a simpler alternative to one of the two main steps in the proof of the Atiyah-Bott formula given in [@gaitsgory_weils_2014] and [@gaitsgory_atiyah-bott_2015].' address: 'Department of Mathematics, University of Chicago, Illinois, USA' author: - 'Q.P. Hô' bibliography: - 'Connectivity\_chiral\_koszul.bib' title: 'The Atiyah-Bott formula and connectivity in chiral Koszul duality' --- Introduction ============ History ------- Let $X$ be a smooth and complete curve, and $G$ a simply-connected semi-simple algebraic group over an algebraically closed field $k$.[^1] Then we know that $$C^*(BG, \Lambda) \simeq \operatorname{Sym}V$$ for some finite dimensional vector space $V$, where $\Lambda$ is ${\mathbb{Q}_\ell}$ when $k={{\overline{\mathbb{F}}}_p}$ ($\ell \neq p$), and $\Lambda$ is any field of characteristic 0 when $k$ has characteristic 0. Let $\operatorname{Bun}_G$ denote the moduli stack of principal $G$-bundles over $X$. In the differential geometric setting, i.e. when $k=\mathbb{C}$, the cohomology ring of $\operatorname{Bun}_G$ was computed by Atiyah and Bott in [@atiyah_yang-mills_1983]. \[thm:intro:Atiyah-Bott\] We have the following equivalence $$C^*(\operatorname{Bun}_G, \Lambda) = \operatorname{Sym}_\Lambda(C^*(X, V\otimes \omega_X)),$$ where $\omega_X$ is the dualizing sheaf of $X$. In the recent work [@gaitsgory_weils_2014], Gaitsgory and Lurie gave a purely algebro-geometric proof of the theorem above in the framework of [étale]{} cohomology (see also [@gaitsgory_atiyah-bott_2015] for an alternative perspective). In the case where $X$ and $G$ come from objects over $k={\mathbb{F}_q}$, the isomorphism in Theorem \[thm:intro:Atiyah-Bott\] was proved to be compatible with the Frobenius actions on both sides. The Grothendieck-Lefschetz trace formula for $\operatorname{Bun}_G$ then gives an expression for the number of $k$-points on $\operatorname{Bun}_G$ and hence, confirms the conjecture of Weil that the Tamagawa number of $G$ is 1. Following ideas suggested in [@gaitsgory_contractibility_2011], this paper aims to provide an alternative (and simpler) proof of one of the two main steps in the original proofs, as given in [@gaitsgory_weils_2014] and [@gaitsgory_atiyah-bott_2015]. This is possible due to a family of new results regarding connectivity in the theory of chiral Koszul duality proved in this paper. Prerequisites and guides to the literature ------------------------------------------ For the reader’s convenience, we include a quick review of the necessary background as well as pointers to the existing literature in §\[sec:prelims\]. The readers who are unfamiliar with the language used in the introduction are encouraged to take a quick look at §\[sec:prelims\] before returning to the current section. A sketch of Gaitsgory and Lurie’s method ---------------------------------------- We will now provide a sketch of the method employed by [@gaitsgory_weils_2014] and [@gaitsgory_atiyah-bott_2015]. In both cases, the proofs utilize the theory of factorization algebras. Broadly speaking, there are two main steps: non-abelian [Poincaré]{} duality and Verdier duality on the $\operatorname{Ran}$ space. ### Non-abelian [Poincaré]{} duality For the first step, one constructs a factorizable sheaf ${\eur{A}}$ on $\operatorname{Ran}X$ from $f_! \omega_{\operatorname{Gr}_{\operatorname{Ran}X}}$ where $f$ is the natural map $$f: \operatorname{Gr}_{\operatorname{Ran}X} \to \operatorname{Ran}X,$$ and $\operatorname{Gr}_{\operatorname{Ran}X}$ is the factorizable affine Grassmannian. The crucial observation is that the natural map $$\operatorname{Gr}_{\operatorname{Ran}X} \to \operatorname{Bun}_G$$ has homologically contractible fibers, and hence, we get an equivalence $$C^*_c(\operatorname{Bun}_G, \omega_{\operatorname{Bun}_G}) \simeq C^*_c(\operatorname{Ran}X, {\eur{A}}). {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:intro_nonab_poincare}$$ ### Verdier duality The right hand side of  is, however, not directly computable. If one thinks of factorizable sheaves on $\operatorname{Ran}X$ as $E_2$-algebras, then one reason that makes it hard to compute the factorization homology of ${\eur{A}}$ is the fact that it’s not necessarily commutative (i.e. not $E_\infty$). ${\eur{A}}$, however, also has a commutative co-algebra structure, via the diagonal map[^2] $$\operatorname{Gr}\to \operatorname{Gr}\times \operatorname{Gr}.$$ Thus, its Verdier dual $D_{\operatorname{Ran}X}{\eur{A}}$ naturally has the structure of a commutative algebra. In fact, it’s proved that $D_{\operatorname{Ran}X}{\eur{A}}$ is a commutative factorization algebra. ### Computing the Verdier dual One can prove something even better: $D_{\operatorname{Ran}X}{\eur{A}}$ is isomorphic to the commutative factorization algebra ${\eur{B}}$ coming from $C^*(BG)$. Indeed, a natural map from one to the other is given by a certain pairing between ${\eur{A}}$ and ${\eur{B}}$. Since these are factorizable, showing that this map is an equivalence amounts to showing that its restriction to $X$ is also an equivalence. This is now a purely local problem, and hence, for example, one can reduce it to the case of $\mathbb{P}^1$ to prove it. ### Conclusion Recall that $${\eur{B}} \simeq C^*(BG) \simeq \operatorname{Sym}V$$ is a free commutative algebra, where $V$ is some explicit chain complex that we can compute. But factorization homology with coefficients in a free commutative factorization algebra is easy to compute. Hence, we conclude $$\begin{aligned} C^*(\operatorname{Bun}_G, {\mathbb{Q}_\ell}) &\simeq C^*_c(\operatorname{Bun}_G, \omega_{\operatorname{Bun}_G})^\vee \\ &\simeq C^*_c(\operatorname{Ran}X, {\eur{A}})^\vee \\ &\simeq C^*_c(\operatorname{Ran}X, D_{\operatorname{Ran}X}{\eur{A}}) {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:intro_verdier_homology} \\ &\simeq C^*_c(\operatorname{Ran}X, {\eur{B}}) \\ &\simeq C^*_c(\operatorname{Ran}X, \operatorname{Sym}V) \\ &\simeq \operatorname{Sym}C^*_c(X, V).\end{aligned}$$ What does this paper do? ------------------------ The main difference between [@gaitsgory_weils_2014] and [@gaitsgory_atiyah-bott_2015] is in the use of Verdier duality on the $\operatorname{Ran}$ space.[^3] The latter greatly simplifies and clarifies the former by formally introducing the concept of Verdier duality on a general prestack and then applying it to the case of the $\operatorname{Ran}$ space. Since the $\operatorname{Ran}$ space is a big object,[^4] its technical properties in relation to factorization homology and factorizability are difficult to establish. More precisely, it takes a lot of work to prove the (innocent looking) equivalence  and to a somewhat lesser extent, the fact that $D_{\operatorname{Ran}X} {\eur{A}}$ is factorizable. This results in a rather complicated technical heart of [@gaitsgory_atiyah-bott_2015]. In this paper, we prove a series of new results regarding connectivity in the theory of chiral Koszul duality. These are interesting in their own rights, since they give new examples of Koszul duality being an equivalence, even when the base category is not pro-nilpotent in the sense of [@francis_chiral_2011]. Based on the ideas sketched in [@gaitsgory_contractibility_2011], the results proved in this paper also further simplify the second step of the proof. More precisely, these results could be used to replace all of §8, §9, and part of §12 and §15 of [@gaitsgory_atiyah-bott_2015]. An outline of our results ------------------------- We will now state the main results proved in this paper. Many results that we prove require connectivity assumptions that are somewhat cumbersome to state. Since these are merely technical conditions irrelevant to the discussion of the general method, we will gloss over them in this section. Many results in this paper could be proved in a more general setting. We avoid doing so to keep the presentation simple. We will, however, provide remarks about this throughout the text. ### Koszul duality for $\operatorname{Lie}$ and $\operatorname{ComCoAlg}$ Let ${\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)$ and $\operatorname{Lie}^\star(\operatorname{Ran}X)$ denote the categories of commutative co-algebra objects and Lie algebra objects in $\operatorname{Shv}(\operatorname{Ran}X)$ with respect to the ${\otimes^\star}$-monoidal structure. The theory of Koszul duality developed in [@francis_chiral_2011] gives a pair of adjoint functors[^5] $$\operatorname{Chev}: \operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X) \rightleftarrows {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X): {\mathrm{Prim}}[-1]. {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:intro_koszul_duality}$$ which restricts to a pair of adjoint functors $$\operatorname{Chev}: \operatorname{\operatorname{Lie}^\star}(X) \rightleftarrows \operatorname{\operatorname{coFact}^\star}(X): {\mathrm{Prim}}[-1],$$ where $\operatorname{\operatorname{coFact}^\star}(X)$ is the category of commutative factorization co-algebras on $X$. Even though the pair of adjoint functors above are not mutually inverses of each other in general, they are when we impose certain connectivity constraints on both sides. \[thm:intro:Koszul\_duality\_connectivity\_on\_Ran\] We have the following commutative diagram $$\xymatrix{ \operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X)^{\leq c_L} \ar@{=}[rr]^<<<<<<<<<<{\operatorname{Chev}}_<<<<<<<<<<{{\mathrm{Prim}}[-1]} && {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)^{\leq c_{cA}} \\ \operatorname{\operatorname{Lie}^\star}(X)^{\leq c_L} \ar@{^(->}[u] \ar@{=}[rr]^{\operatorname{Chev}}_{{\mathrm{Prim}}[-1]} && \operatorname{\operatorname{coFact}^\star}(X)^{\leq c_{cA}} \ar@{^(->}[u] }$$ where $\leq c_L$ and $\leq c_{cA}$ denote some connectivity constraints, and where $\operatorname{Chev}$ and ${\mathrm{Prim}}[-1]$ are the functors coming from Koszul duality. ### Koszul duality for $\operatorname{coLie}$ and $\operatorname{ComAlg}$ Let $\operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)$ and $\operatorname{\operatorname{coLie}^\star}(\operatorname{Ran}X)$ denote the categories of commutative algebra objects and co-Lie algebra objects in $\operatorname{Shv}(\operatorname{Ran}X)$ with respect to the ${\otimes^\star}$-monoidal structure. As above, we have the following pair of adjoint functors[^6] $$\operatorname{coPrim}[1]: \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X) \rightleftarrows \operatorname{\operatorname{coLie}^\star}(\operatorname{Ran}X): \operatorname{coChev}.$$ Unlike the case of $\operatorname{\operatorname{Lie}^\star}$ and ${\operatorname{ComCoAlg}^\star}$, for a co-Lie algebra $\mathfrak{g} \in \operatorname{\operatorname{coLie}^\star}(X)$, $$\operatorname{coChev}(\mathfrak{g}) \in \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)$$ doesn’t necessarily live inside $\operatorname{\operatorname{Fact}^\star}(X)$. However, we have the following \[thm:intro:factorizability\_coChev\] Restricted to the full subcategory $\operatorname{\operatorname{coLie}^\star}(X)^{\geq 1}$, where we are using the perverse $t$-structure on $X$, the functor $\operatorname{coChev}$ factors through $\operatorname{\operatorname{Fact}^\star}$, i.e. $$\xymatrix{ \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1} \ar[dr]_{\operatorname{coChev}} \ar[rr]^{\operatorname{coChev}} && \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X) \\ & \operatorname{\operatorname{Fact}^\star}(X) \ar@{^(->}[ur] }$$ ### Interaction between $\operatorname{coChev}$ and factorization homology In [@francis_chiral_2011], it’s proved that the functor of taking factorization homology: $$C^*_c: \operatorname{Shv}(\operatorname{Ran}X) \to {\mathrm{Vect}}$$ commutes with $\operatorname{Chev}$. This is because $\operatorname{Chev}$ is computed as a colimit, and moreover, $C^*_c$ has the following two useful properties: 1. $C^*_c$ is symmetric monoidal with respect to the ${\otimes^\star}$-monoidal structure on $\operatorname{Shv}(\operatorname{Ran}X)$ and the usual monoidal structure on ${\mathrm{Vect}}$. 2. $C^*_c$ is continuous. The functor $\operatorname{coChev}$, however, is constructed as a limit, so we need some extra conditions to make it behave nicely with $C^*_c$. \[thm:intro:coChev\_and\_C\^\*\_c(Ran)\] Let $\mathfrak{g} \in \operatorname{\operatorname{coLie}^\star}(X)^{\geq c_{cL}}$, where $\geq c_{cL}$ denotes some co-connectivity constraint. Then we have a natural equivalence $$C^*_c(\operatorname{Ran}X, \operatorname{coChev}\mathfrak{g}) \simeq \operatorname{coChev}(C^*_c(\operatorname{Ran}X, \mathfrak{g})).$$ ### $\operatorname{Chev}$, $\operatorname{coChev}$ and Verdier duality Unsurprisingly, the functors $\operatorname{Chev}$ and $\operatorname{coChev}$ mentioned above are linked via the Verdier duality functor on $\operatorname{Ran}X$. \[thm:intro:Chev\_coChev\_and\_D\_Ran\] Let $\mathfrak{g} \in \operatorname{\operatorname{Lie}^\star}(X)^{\leq -1}$, where we are using the perverse $t$-structure on $X$. Then we have the following natural equivalence $$D_{\operatorname{Ran}X} \operatorname{Chev}\mathfrak{g} \simeq \operatorname{coChev}(D_X \mathfrak{g}).$$ As we shall see, the connectivity constraint $\operatorname{\operatorname{Lie}^\star}(X)^{\leq -1}$ is less strict than the connectivity constraint $\operatorname{\operatorname{Lie}^\star}(X)^{\leq c_L}$ required by Theorem \[thm:intro:Koszul\_duality\_connectivity\_on\_Ran\]. As a corollary of Theorem \[thm:intro:factorizability\_coChev\], we know that when $\mathfrak{g} \in \operatorname{\operatorname{Lie}^\star}(X)^{\leq c_L}$, $$D_{\operatorname{Ran}X} \operatorname{Chev}\mathfrak{g} \simeq \operatorname{coChev}(D_X \mathfrak{g})$$ is factorizable. Relation to the Atiyah-Bott formula ----------------------------------- ### The initial observation is that the sheaf ${\eur{A}}$ mentioned above lies in the essential image of $\operatorname{Chev}$, i.e. $${\eur{A}} \simeq \operatorname{Chev}(\mathfrak{a}), \qquad\text{for some } \mathfrak{a} \in \operatorname{\operatorname{Lie}^\star}(X)^{\leq c_L}.$$ This is a direct result of Theorem \[thm:intro:Koszul\_duality\_connectivity\_on\_Ran\] and the fact that ${\eur{A}}$ satisfies this connectivity constraint on the ${\operatorname{ComCoAlg}^\star}$ side. ### As in [@gaitsgory_atiyah-bott_2015], we have a pairing $${\eur{A}} \boxtimes {\eur{B}} \to \delta_! \omega_{\operatorname{Ran}X},$$ which induces a map $${\eur{B}} \to D_{\operatorname{Ran}X} {\eur{A}},$$ compatible with the commutative algebra structures on both sides. Thus, we get a map $${\eur{B}} \to D_{\operatorname{Ran}X}\operatorname{Chev}(\mathfrak{a}) \simeq \operatorname{coChev}(D_X \mathfrak{a}),$$ which we want to be an equivalence. Since both sides are factorizable, it suffices to show that they are over $X$, which is now a local problem, and the same proof as in [@gaitsgory_atiyah-bott_2015] applies. ### Conclusion Let $V\in {\mathrm{Vect}}$ such that $\operatorname{Sym}(V\otimes\omega_X) \simeq {\eur{B}}$ where $\operatorname{Sym}$ is taken inside $\operatorname{Shv}(\operatorname{Ran}X)$ using the ${\otimes^\star}$-monoidal structure. Then, we have $$\begin{aligned} \operatorname{Sym}C^*_c(X, V \otimes \omega_X) &\simeq C^*_c(\operatorname{Ran}X, \operatorname{Sym}(V\otimes \omega_X)) \\ &\simeq C^*_c(\operatorname{Ran}X, {\eur{B}}) \\ &\simeq C^*_c(\operatorname{Ran}X, \operatorname{coChev}_{\operatorname{Ran}X} D_X \mathfrak{a}) \\ &\simeq \operatorname{coChev}(C^*_c(X, D_X \mathfrak{a})) \\ &\simeq \operatorname{coChev}(C^*_c(X, \mathfrak{a})^\vee) \\ &\simeq \operatorname{Chev}(C^*_c(X, \mathfrak{a}))^\vee \\ &\simeq C^*_c(\operatorname{Ran}X, \operatorname{Chev}\mathfrak{a})^\vee \\ &\simeq C^*_c(\operatorname{Ran}X, {\eur{A}})^\vee \\ &\simeq C^*_c(\operatorname{Bun}_G, \omega_{\operatorname{Bun}_G})^\vee \\ &\simeq C^*(\operatorname{Bun}_G, {\mathbb{Q}_\ell}).\end{aligned}$$ It is interesting to note that many technical results about Verdier duality are proved only for the case of curves in [@gaitsgory_atiyah-bott_2015], while results stated here about Koszul duality are for arbitrary dimension (even though in the end, they serve a similar purpose regarding the Atiyah-Bott formula). This is in part because [@gaitsgory_atiyah-bott_2015] works with more general sheaves on the $\operatorname{Ran}$ space, whereas we mostly concern ourselves with sheaves of special shapes, i.e. they are all of the form $\operatorname{Chev}\mathfrak{g}$ or $\operatorname{coChev}\mathfrak{g}$. Acknowledgments --------------- The author would like to express his gratitude to D. Gaitsgory, without whose tireless guidance and encouragement in pursuing this problem, this work would not have been possible. The author is grateful to his advisor B.C. Ngô for many years of patient guidance and support. Preliminaries {#sec:prelims} ============= In this section, we will set up the language and conventions used throughout the paper. Since the material covered here are used in various places, the readers should feel free to skip it and backtrack when necessary. The mathematical content in this section has already been treated elsewhere. Hence, results are stated without any proof, and we will do our best to provide the necessary references. It is important to note that it is not our aim to be exhaustive. Rather, we try to familiarize the readers with the various concepts and results used in the text, as well as to give pointers to the necessary references for the background materials. Notation and conventions ------------------------ ### Category theory We will use ${\mathrm{DGCat}}$ to denote the $(\infty, 1)$-category of stable infinity categories, ${{\mathrm{DGCat}}_{\mathrm{pres}}}$ to denote the full subcategory of ${\mathrm{DGCat}}$ consisting of presentable categories, and ${{\mathrm{DGCat}}_{{\mathrm{pres}}, {\mathrm{cont}}}}$ the (non-full) subcategory of ${{\mathrm{DGCat}}_{\mathrm{pres}}}$ where we restrict to continuous functors, i.e. those commuting with colimits. ${\mathrm{Spc}}$ will be used to denote the category of spaces, or more precisely, $\infty$-groupoids. The main references for this subject are [@lurie_higher_2015] and [@lurie_higher_2014]. For a slightly different point of view, see also [@gaitsgory_study_????]. ### Algebraic geometry Throughout this paper, $k$ will be an algebraically closed ground field. We will denote by ${\mathrm{Sch}}$ the $\infty$-category obtained from the ordinary category of separated schemes of finite type over $k$. All our schemes will be objects of ${\mathrm{Sch}}$. In most cases, we will use the calligraphic font to denote prestacks, for eg. ${\eur{X}}, {\eur{Y}}$ etc., and the usual font to denote schemes, for eg. $X, Y$ etc. ### $t$-structures {#subsubsec:t-structure_convention} Let ${\eur{C}}$ be a stable infinity category, equipped with a $t$-structure. Then we have the following diagram of adjoint functors $$\xymatrix{ {\eur{C}}^{\leq 0} {\ar@<{0.36ex}>}[r]^>>>>>{i_{\leq 0}} & {\ar@<{0.36ex}>}[l]^>>>>>{\operatorname{tr}_{\leq 0}} {\eur{C}} {\ar@<{0.36ex}>}[r]^<<<<<{\operatorname{tr}_{\geq 1}} & {\ar@<{0.36ex}>}[l]^>>>>>{i_{\geq 1}} {\eur{C}}^{\geq 1} }$$ We use $\tau_{\leq 0}$ and $\tau_{\geq 1}$ to denote $$\tau_{\leq 0} = i_{\leq 0} \circ \operatorname{tr}_{\leq 0}: {\eur{C}} \to {\eur{C}}$$ and $$\tau_{\geq 1} = i_{\geq 1} \circ \operatorname{tr}_{\geq 1}: {\eur{C}} \to {\eur{C}}$$ respectively. Shifts of these functors, for e.g. $\tau_{\geq n}$ and $\tau_{\leq n}$, are defined in the obvious ways. Prestacks --------- The theory of sheaves on prestacks has been developed in [@gaitsgory_weils_2014] and [@gaitsgory_atiyah-bott_2015]. In this subsection and the next, we will give a brief review of this theory, including the definition of the category of sheaves as well as various pull and push functors. We will state them as facts, without any proof, which (unless otherwise specified), could all be found in [@gaitsgory_atiyah-bott_2015]. ### A prestack is a contravariant functor from ${\mathrm{Sch}}$ to ${\mathrm{Spc}}$, i.e. a prestack ${\eur{Y}}$ is a functor $${\eur{Y}}: {\mathrm{Sch}}^{{\mathrm{op}}} \to {\mathrm{Spc}}.$$ Let ${\mathrm{PreStk}}$ be the $\infty$-category of prestacks. Then by Yoneda’s lemma, we have a fully-faithful embedding $${\mathrm{Sch}}\hookrightarrow {\mathrm{PreStk}}.$$ ### Properties of prestacks Due to categorical reasons, any prestack ${\eur{Y}}$ can be written as a colimit of schemes $${\eur{Y}} \simeq \operatorname*{colim}_{i\in I} Y_i.$$ ### A prestack is said to be is a pseudo-scheme if it could be written as a colimit of schemes, where all morphisms are proper. ### A prestack is pseudo-proper if it could be written as a colimit of proper schemes. It is straightforward to see that pseudo-proper prestacks are pseudo-schemes. ### A prestack is said to be finitary if it could be expressed as a finite colimit of schemes. ### We also have relative versions of the definitions above in an obvious manner. Namely, we can speak of a morphism $f: {\eur{Y}} \to S$, where ${\eur{Y}}$ is a prestack and $S$ is a scheme, being pseudo-schematic (resp. pseudo-proper, finitary). ### More generally, a morphism $$f: {\eur{Y}}_1 \to {\eur{Y}}_2$$ is said to be pseudo-schematic (resp. pseudo-proper, finitary) if for any scheme $S$, equipped with a morphism $S \to {\eur{Y}}_2$, the morphism $f_S$ in the following pull-back diagram $$\xymatrix{ S \times_{{\eur{Y}}_2} {\eur{Y}}_1 \ar[d]_{f_S} \ar[r] & {\eur{Y}}_1 \ar[d] \\ S \ar[r] & {\eur{Y}}_2 }$$ is pseudo-schematic (resp. pseudo-proper, finitary). Sheaves on prestacks -------------------- As we mentioned above, proofs of all the results in mentioned in this section, unless otherwise specified, could be found in [@gaitsgory_atiyah-bott_2015]. ### Sheaves on schemes We will adopt the same conventions as in [@gaitsgory_atiyah-bott_2015], except that for simplicity, we will restrict ourselves to the “constructible setting.” Namely, for a scheme $S$, (i) when the ground field is $\mathbb{C}$, and $\Lambda$ is an arbitrary field of characteristic 0, we take $\operatorname{Shv}(S)$ to be the ind-completion of the category of constructible sheaves on $S$ with $\Lambda$-coefficients. (ii) for any ground field $k$ in general, and $\Lambda = {\mathbb{Q}_\ell}, {{\overline{\mathbb{Q}}}_\ell}$ with $\ell \neq \operatorname{char}k$, we take $\operatorname{Shv}(S)$ to be the ind-completion of the category of constructible $\ell$-adic sheaves on $S$ with $\Lambda$-coefficients. See also [@gaitsgory_weils_2014 §4], [@liu_enhanced_2012], and [@liu_enhanced_2014]. The theory of sheaves on schemes is equipped with the various pairs of adjoint functors $$f_! \dashv f^! \qquad\text{and}\qquad f^* \dashv f_*$$ for any morphism $$f: S_1 \to S_2$$ between schemes. Moreover, we also have box-product $\boxtimes$ and hence, also $\otimes$ and ${\overset{!}{\otimes}}$. ### Throughout the text, we will use the perverse $t$-structure on $\operatorname{Shv}(S)$, when $S$ is a scheme. ### We will also use ${\mathrm{Vect}}$ to denote the category of sheaves on a point, i.e. ${\mathrm{Vect}}$ denotes the (infinity derived) category of chain complexes in vector spaces over $\Lambda$. ### Sheaves on prestacks For a prestack ${\eur{Y}}$, the category $\operatorname{Shv}({\eur{Y}})$ is defined by $$\operatorname{Shv}({\eur{Y}}) = \lim _{S\in ({\mathrm{Sch}}_{/{\eur{Y}}}^{{\mathrm{op}}})} \operatorname{Shv}(S),$$ where the transition functor we use is the shriek-pullback. Thus, an object ${\eur{F}} \in \operatorname{Shv}({\eur{Y}})$ is the same as the following data (i) A sheaf ${\eur{F}}_{S, y} \in \operatorname{Shv}(S)$ for each $S\in {\mathrm{Sch}}$ and $y: S \to {\eur{Y}}$ (i.e. $y\in {\eur{Y}}(S)$). (ii) An equivalence of sheaves ${\eur{F}}_{S', f(y)} \to f^! {\eur{F}}_{S, y}$ for each morphism of schemes $f: S' \to S$. Moreover, we require that this assignment satisfies a homotopy-coherent system of compatibilities. ### More formally, one can define $\operatorname{Shv}({\eur{Y}})$ as the right Kan extension of $$\operatorname{Shv}: {\mathrm{Sch}}^{{\mathrm{op}}} \to {{\mathrm{DGCat}}_{{\mathrm{pres}}, {\mathrm{cont}}}}$$ along the Yoneda embedding $${\mathrm{Sch}}^{{\mathrm{op}}} \hookrightarrow {\mathrm{PreStk}}^{{\mathrm{op}}}.$$ Thus, by formal reasons, the functor $$\operatorname{Shv}: {\mathrm{PreStk}}^{{\mathrm{op}}} \to {{\mathrm{DGCat}}_{{\mathrm{pres}}, {\mathrm{cont}}}}$$ preserves limits. In other words, we have $$\operatorname{Shv}(\operatorname*{colim}_i {\eur{Y}}_i) \simeq \lim_i \operatorname{Shv}({\eur{Y}}_i).$$ In particular, if a prestack $${\eur{Y}} \simeq \operatorname*{colim}_{i\in I} Y_i$$ is a colimit of schemes, then $$\operatorname{Shv}({\eur{Y}}) \simeq \lim_{i\in I} \operatorname{Shv}(Y_i).$$ ### Now, if we replace all the transition functors by their left adjoints, namely the $!$-pushforward, then we have a diagram $$I^{\mathrm{op}}\to {{\mathrm{DGCat}}_{{\mathrm{pres}}, {\mathrm{cont}}}},$$ and we have a natural equivalence $$\operatorname{Shv}({\eur{Y}}) \simeq \operatorname*{colim}_{i\in I^{\mathrm{op}}} \operatorname{Shv}(Y_i)$$ where the colimit is taken inside ${{\mathrm{DGCat}}_{{\mathrm{pres}}, {\mathrm{cont}}}}$. ### Let $${\eur{Y}} = \operatorname*{colim}_i Y_i$$ be a prestack, and denote $$\operatorname{ins}_i: Y_i \to {\eur{Y}}$$ the canonical map. Then, for any sheaf ${\eur{F}} \in \operatorname{Shv}({\eur{Y}})$, we have the following natural equivalence $${\eur{F}} \simeq \operatorname*{colim}_i \operatorname{ins}_{i!} \operatorname{ins}_i^! {\eur{F}} {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:sheaves_colimits_rep}$$ ### $f_! \dashv f^!$ Let $$f: {\eur{Y}}_1 \to {\eur{Y}}_2$$ be a morphism between prestacks. Then by restriction, we get a functor $$f^!: \operatorname{Shv}({\eur{Y}}_2) \to \operatorname{Shv}({\eur{Y}}_1),$$ which commutes with both limits and colimits. In particular, $f^!$ admits a left adjoint $f_!$.[^7] The functor $f_!$ is generally not computable. However, there are a couple of cases where it is. ### The first instance is when the target of $f$ is a scheme $$f: {\eur{Y}} \to S,$$ and suppose that $${\eur{Y}} \simeq \operatorname*{colim}_i Y_i.$$ Then, by , we have $$f_! {\eur{F}} \simeq \operatorname*{colim}f_! \operatorname{ins}_{i!}\operatorname{ins}_i^! {\eur{F}} \simeq \operatorname*{colim}f_{i!} \operatorname{ins}_i^! {\eur{F}}.$$ where $$f_i: Y_i \to {\eur{Y}} \to S$$ is just a morphism between schemes. ### The second case is where $f$ is pseudo-proper, then $f_!$ satisfies the base change theorem with respect to the $(-)^!$-pullback. Namely, for any pull-back diagram of prestacks $$\xymatrix{ {\eur{Y}}'_1 \ar[d]_{f} \ar[r]^g & {\eur{Y}}_1 \ar[d]_{f} \\ {\eur{Y}}'_2 \ar[r]^g & {\eur{Y}}_2 }$$ and any sheaf ${\eur{F}} \in \operatorname{Shv}({\eur{Y}})$, we have a natural equivalence $$g^! f_! {\eur{F}} \simeq f_! g^! {\eur{F}}.$$ Thus, in particular, if we have a pull-back diagram $$\xymatrix{ S \times_{{\eur{Y}}_2} {\eur{Y}}_1 \ar[d]_{f_S} \ar[r]^{i_S} & {\eur{Y}}_1 \ar[d]_f \\ S \ar[r]^{i_S} & {\eur{Y}}_2 }$$ where $S$ is a scheme, then $$i_S^! f_! {\eur{F}} \simeq f_{S!} i_S^!{\eur{F}}$$ and as discussed above, $f_{S!}$ could be computed as an explicit colimit. ### Let ${\eur{F}} \in \operatorname{Shv}({\eur{Y}})$. Then we denote by $$C^*_c({\eur{Y}}, {\eur{F}}) = s_! {\eur{F}},$$ where $$s: {\eur{Y}} \to \operatorname{Spec}k$$ is the structural map of ${\eur{Y}}$ to a point. ### In case where ${\eur{F}} \simeq \omega_{{\eur{Y}}}$ is the dualizing sheaf on ${\eur{Y}}$ (characterized by the property that its $(-)^!$-pullback to any scheme is the dualizing sheaf on that scheme), then we write $$C_*({\eur{Y}}) = C^*_c({\eur{Y}}, \omega_{{\eur{Y}}}),$$ and $$C_*^{{\mathrm{red}}}({\eur{Y}}) = \operatorname{Fib}(C_*({\eur{Y}}) \to \Lambda).$$ ### $f^* \dashv f_*$ When $$f: {\eur{Y}}_1 \to {\eur{Y}}_2$$ is a schematic morphism between prestacks, one can also define a pair of adjoint functors (see [@gaitsgory_atiyah-bott_2015] where the functor $f_*$ is defined, and [@ho_free_2015] where the adjunction is constructed) $$f^*: \operatorname{Shv}({\eur{Y}}_2) \rightleftarrows \operatorname{Shv}({\eur{Y}}_1): f_*.$$ ### The behavior of $f_*$ is easy to describe, due to the fact that $f_*$ satisfies the base change theorem with respect to the $(-)^!$-pullback functor. Namely, suppose ${\eur{F}} \in \operatorname{Shv}({\eur{Y}}_1)$ and we have a pullback square where $S_2$ (and hence, $S_1$) is a scheme $$\xymatrix{ S_1 \ar[r]^g \ar[d]_{f_S} & {\eur{Y}}_1 \ar[d]_{f} \\ S_2 \ar[r]^g & {\eur{Y}}_2 }$$ Then, the pullback could be described in classical terms, since $$g^! f_* {\eur{F}} \simeq f_{S*} g^! {\eur{F}},$$ where $f_S$ is just a morphism between schemes. ### The functor $f^*$ is slightly more complicated to describe. However, when $$f: {\eur{Y}}_1 \to {\eur{Y}}_2$$ is [étale]{}, which is the case where we need, we have a natural equivalence (see [@ho_free_2015 Prop. 2.7.3]) $$f^! \simeq f^*. {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:etale_!_*_pullback_equiv}$$ ### We will also need the following fact in the definition of commutative factorizable co-algebras: let $$\xymatrix{ {\eur{U}} \ar[r]^f & {\eur{Z}} \ar[r]^g & {\eur{X}} }$$ be morphisms between prestacks, where $g$ is finitary pseudo-proper, $f$ and $h = g\circ f$ are schematic. Then we have a natural equivalence (see [@ho_free_2015 Prop. 2.10.4]) $$g_! \circ f_* \simeq (g\circ f)_* \simeq h_*. {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:pushforward_!circ*_equiv_*}$$ ### Monoidal structure The theory of sheaves on prestacks discussed so far naturally inherits the box-tensor structure from the theory of sheaves on schemes. Namely, let ${\eur{F}}_i\in \operatorname{Shv}({\eur{Y}}_i)$ where ${\eur{Y}}_i$’s are prestacks, for $i=1, 2$. Then, for any pair of schemes $S_1, S_2$ equipped with maps $$f_i: S_i\to {\eur{Y}}_i,$$ we have $$(f_1 \times f_2)^! ({\eur{F}}_1 \boxtimes {\eur{F}}_2) \simeq f_1^! {\eur{F}}_1 \boxtimes f_2^! {\eur{F}}_2.$$ Pulling back along the diagonal $$\delta: {\eur{Y}} \to {\eur{Y}} \times {\eur{Y}}$$ for any prestack ${\eur{Y}}$, we get the ${\overset{!}{\otimes}}$-symmetric monoidal structure on ${\eur{Y}}$ in the usual way. More explicitly, for ${\eur{F}}_1, {\eur{F}}_2 \in \operatorname{Shv}({\eur{Y}})$, we define $${\eur{F}}_1 {\overset{!}{\otimes}}{\eur{F}}_2 = \delta^! ({\eur{F}}_1 \boxtimes {\eur{F}}_2).$$ The $\operatorname{Ran}$ space/prestack --------------------------------------- The $\operatorname{Ran}$ space (or more precisely, prestack) of a scheme plays a central role in this paper. The $\operatorname{Ran}$ space, along with various objects on it, was first studied in the seminal book [@beilinson_chiral_2004] in the case of curves, and was generalized to higher dimensions in [@francis_chiral_2011]. In what follows, we will quickly review the main definitions and results. For proofs, unless otherwise specified, we refer the reader to [@gaitsgory_atiyah-bott_2015] and [@francis_chiral_2011]. The topologically inclined reader could also find an intuitive introduction in [@ho_free_2015 §1]. ### For a scheme $X \in {\mathrm{Sch}}$, we will use $\operatorname{Ran}X$ to denote the following prestack: for each scheme $S \in {\mathrm{Sch}}$, $$(\operatorname{Ran}X)(S) = \{\text{non-empty finite subsets of } X(S)\}$$ Alternatively, one has $$\operatorname{Ran}X \simeq \operatorname*{colim}_{I\in {\mathrm{fSet}}^{{\mathrm{surj}}, {\mathrm{op}}}} X^I$$ where ${\mathrm{fSet}}^{{\mathrm{surj}}}$ denotes the category of non-empty finite sets, where morphisms are surjections. Using the fact that $X$ is separated, one sees easily that $\operatorname{Ran}X$ is a pseudo-scheme. Moreover, when $X$ is proper, $\operatorname{Ran}X$ is pseudo-proper. ### The ${\otimes^\star}$ monoidal structure There is a special monoidal structure on $\operatorname{Ran}X$ which we will use throughout the text: the ${\otimes^\star}$-monoidal structure. Consider the following map $$\xymatrix{ {\mathrm{union}}: \operatorname{Ran}X \times \operatorname{Ran}X \to \operatorname{Ran}X }$$ given by the union of non-empty finite subsets of $X$. One can check that ${\mathrm{union}}$ is finitary pseudo-proper. Given two sheaves ${\eur{F}}, {\eur{G}} \in \operatorname{Shv}(\operatorname{Ran}X)$, we define $${\eur{F}} {\otimes^\star}{\eur{G}} = {\mathrm{union}}_!({\eur{F}} \boxtimes {\eur{G}}).$$ This defines the ${\otimes^\star}$-monoidal structure on $\operatorname{Shv}(\operatorname{Ran}X)$. ### Since ${\mathrm{union}}$ is pseudo-proper, ${\eur{F}}{\otimes^\star}{\eur{G}}$ has an easy presentation. Namely, for $${\eur{F}}_1, {\eur{F}}_2, \dots, {\eur{F}}_k \in \operatorname{Shv}(\operatorname{Ran}X),$$ and any non-empty finite set $I$, we have the following $$({\eur{F}}_1 {\otimes^\star}{\eur{F}}_2 {\otimes^\star}\cdots {\otimes^\star}{\eur{F}}_k)|_{{\overset{\circ}{X}{}^{I}}} \simeq \bigoplus_{I = \bigcup_{i=1}^k I_i} \Delta^!_{\sqcup_{i=1}^k I_i {\twoheadrightarrow}\cup_{i=1}^k I_i}({\eur{F}}_1 \boxtimes \cdots \boxtimes{\eur{F}}_k)|_{(\prod {\overset{\circ}{X}{}^{I_i}})_{{\mathrm{disj}}}} {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:otimesstar_explicit_formula}$$ where $(\prod {\overset{\circ}{X}{}^{I_i}})_{{\mathrm{disj}}}$ denotes the open subscheme of $\prod_i X^{I_i}$ where no two “coordinates” are equal, and where $$\Delta_{\sqcup_{i=1}^k I_i {\twoheadrightarrow}\cup_{i=1}^k I_i}: X^I \hookrightarrow \prod_i X^{I_i}$$ is the map induced by the surjection $$\bigsqcup_{i=1}^k I_i {\twoheadrightarrow}\bigcup_{i=1}^k I_i \simeq I.$$ ### Factorizable sheaves Using the ${\otimes^\star}$-monoidal structure on $\operatorname{Shv}(\operatorname{Ran}X)$, one can talk about various types of algebras/coalgebras in $\operatorname{Shv}(\operatorname{Ran}X)$. The ones that are of importance to us in this papers are $$\operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X), \operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X), {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X), \text{ and } \operatorname{\operatorname{coLie}^\star}(\operatorname{Ran}X).$$ As the name suggests, these are used, respectively, to denote the categories of commutative algebras, Lie algebras, commutative co-algebras and co-Lie algebras in $\operatorname{Shv}(\operatorname{Ran}X)$ with respect to the ${\otimes^\star}$-monoidal structure defined above. ### We use $\operatorname{\operatorname{Lie}^\star}(X)$ and $\operatorname{\operatorname{coLie}^\star}(X)$ to denote the full subcategories of $\operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X)$ and $\operatorname{\operatorname{coLie}^\star}(\operatorname{Ran}X)$ respectively, consisting of objects whose supports are inside the diagonal $$\operatorname{ins}_X: X \hookrightarrow \operatorname{Ran}X$$ of $\operatorname{Ran}X$. ### Let $$j: (\operatorname{Ran}X)^n_{\mathrm{disj}}\to (\operatorname{Ran}X)^n$$ where $(\operatorname{Ran}X)^n_{{\mathrm{disj}}}$ is the open sub-prestack of $(\operatorname{Ran}X)^n$ defined by the following condition: for each scheme $S$, $(\operatorname{Ran}X)^n(S)$ consists of $n$ non-empty subsets of $X(S)$, whose graphs are pair-wise disjoint. ### Let $${\eur{A}} \in {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X).$$ Then, by definition, we have the following map (which is the co-multiplication of the commutative co-algebra structure) $${\eur{A}} \to {\eur{A}} {\otimes^\star}{\eur{A}} {\otimes^\star}\cdots {\otimes^\star}{\eur{A}} \simeq {\mathrm{union}}_!({\eur{A}} \boxtimes \cdots \boxtimes {\eur{A}}).$$ Using the the unit map of the adjunction $j^* \dashv j_*$, we get the following map $${\mathrm{union}}_!({\eur{A}}\boxtimes \cdots \boxtimes {\eur{A}}) \to {\mathrm{union}}_! j_* j^* ({\eur{A}} \boxtimes \cdots \boxtimes {\eur{A}}) \simeq ({\mathrm{union}}\circ j)_* j^!({\eur{A}} \boxtimes \cdots \boxtimes {\eur{A}}),$$ where for the equivalence, we made use of  and . Altogether, we get a map $${\eur{A}} \to ({\mathrm{union}}\circ j)_* j^! ({\eur{A}} \boxtimes \cdots \boxtimes {\eur{A}})$$ and hence, by adjunction and , we get a map $$j^! {\mathrm{union}}^! {\eur{A}} \to j^! ({\eur{A}}\boxtimes \cdots \boxtimes {\eur{A}}). {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:commutative_cofac_map}$$ ${\eur{A}}$ is a commutative factorization algebra if the map  is an equivalence for all $n$’s. We use $\operatorname{\operatorname{coFact}^\star}(X)$ to denote the full subcategory of $\operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)$ consisting of commutative factorization co-algebras. ### Let $${\eur{B}} \in \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X).$$ Then, by definition, we have the following map (which is the multiplication of the commutative algebra structure) $${\mathrm{union}}_!({\eur{B}} \boxtimes {\eur{B}} \boxtimes \cdots \boxtimes {\eur{B}}) \simeq {\eur{B}} {\otimes^\star}{\eur{B}} {\otimes^\star}\cdots {\otimes^\star}{\eur{B}} \to {\eur{B}}.$$ This induces the following map of sheaves $${\eur{B}} \boxtimes \cdots \boxtimes {\eur{B}} \to {\mathrm{union}}^! {\eur{B}}$$ on $(\operatorname{Ran}X)^n$, and hence, a map of sheaves $$j^! ({\eur{B}} \boxtimes \cdots \boxtimes {\eur{B}}) \to j^!{\mathrm{union}}^! {\eur{B}}. {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:commutative_fac_map}$$ on $(\operatorname{Ran}X)^n_{\mathrm{disj}}$. ${\eur{B}}$ is a commutative factorization algebra if the map  is an equivalence for all $n$’s. We use $\operatorname{\operatorname{Fact}^\star}(X)$ to denote the full subcategory of $\operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)$ consisting of commutative factorization algebras. Koszul duality -------------- In this subsection, we will quickly review various concepts and results in the theory of Koszul duality that are relevant to us. This theory, initially developed in [@quillen_rational_1969], illuminates the duality between commutative co-algebras and Lie algebras. It was further developed and generalized in the operadic setting in [@ginzburg_koszul_1994]. In the chiral/factorizable setting, the paper [@francis_chiral_2011] provides us with necessary technical tools and language to carry out many topological arguments in the context of algebraic geometry. The results and definitions we review below could be found in [@francis_chiral_2011] and [@gaitsgory_study_????]. ### Symmetric sequences Let ${\mathrm{Vect}}^{\Sigma}$ denote the category of symmetric sequences. Namely, its objects are collections $${\eur{O}} = \{{\eur{O}}(n), n\geq 1\},$$ where each ${\eur{O}}(n)$ is an object of ${\mathrm{Vect}}$, acted on by the symmetric group $\Sigma_n$. The infinity category ${\mathrm{Vect}}^{\Sigma}$ is equipped with a natural monoidal structure which makes the functor $${\mathrm{Vect}}^{\Sigma} \to \operatorname{Fun}({\mathrm{Vect}}, {\mathrm{Vect}})$$ given by the following formula $${\eur{O}} \star V = \bigoplus_n ({\eur{O}}(n)\otimes V^{\otimes n})_{\Sigma_n}$$ symmetric monoidal. ### Operads and co-operads By an operad (resp. co-operad), we will mean an augmented associative algebra (resp. co-algebra) object in ${\mathrm{Vect}}^{\Sigma}$, with respect to the monoidal structure described above. We use $\operatorname{Op}$ (resp. $\operatorname{coOp}$) to denote the categories of operads (resp. co-operads). In general, the Bar and coBar construction gives us the following pair of adjoint functors $$\operatorname{Bar}: \operatorname{Op}\rightleftarrows \operatorname{coOp}: \operatorname{coBar}.$$ For an operad ${\eur{O}}$ (resp. co-operad ${\eur{P}}$), we also use ${\eur{O}}^\vee$ (resp. ${\eur{P}}^\vee$) to denote $\operatorname{Bar}({\eur{O}})$ (resp. $\operatorname{coBar}({\eur{P}})$). In what follows, we will adopt the following convention: all our operads/co-operads will have the property that the augmentation map is an equivalence, when restricted to ${\eur{O}}(1)$ (resp. ${\eur{P}}(1)$). And under this restriction, one can show that the following unit map is an equivalence $${\eur{O}} \to \operatorname{coBar}\circ \operatorname{Bar}({\eur{O}})$$ or in a slightly different notation $${\eur{O}} \to ({\eur{O}}^\vee)^\vee$$ when ${\eur{O}} \in \operatorname{Op}$ satisfying the assumption above. ### Algebras and co-algebras Let ${\eur{C}}$ be a stable presentable symmetric monoidal $\infty$-category compatibly tensored over ${\mathrm{Vect}}$. Then, an operad ${\eur{O}}$ (resp. co-operad ${\eur{P}}$) naturally defines a monad (resp. comonad) on ${\eur{C}}$. Thus, for an operad ${\eur{O}}$ (resp. co-operad ${\eur{P}}$), one can talk about the category of algebras ${\eur{O}}\operatorname{-alg}({\eur{C}})$ (resp. co-algebras ${\eur{P}}\operatorname{-coalg}({\eur{C}})$) in ${\eur{C}}$ with respect to the operad ${\eur{O}}$ (resp. co-operad ${\eur{P}}$). As usual (as for any augmented monad), one has the following pairs of adjoint functors $$\operatorname{Free}_{{\eur{O}}}: {\eur{C}} \rightleftarrows {\eur{O}}\operatorname{-alg}({\eur{C}}):{\mathrm{oblv}}_{{\eur{O}}} \qquad \text{and}\qquad \operatorname{Bar}_{{\eur{O}}}: {\eur{O}}\operatorname{-coalg}({\eur{C}}) \rightleftarrows {\eur{C}} :\operatorname{triv}_{{\eur{O}}}$$ for an operad ${\eur{O}}$, and similarly, the following pairs of adjoint functors $${\mathrm{oblv}}_{{\eur{P}}}: {\eur{P}}\operatorname{-coalg}({\eur{C}}) \rightleftarrows {\eur{C}} :\operatorname{coFree}_{{\eur{P}}} \qquad\text{and}\qquad \operatorname{cotriv}_{{\eur{P}}}: {\eur{C}} \rightleftarrows {\eur{P}}\operatorname{-coalg}({\eur{C}}) :\operatorname{coBar}_{{\eur{P}}}$$ for a co-operad ${\eur{P}}$. ### Koszul duality The functors mentioned above could be lifted to get a pair of adjoint functors $$\operatorname{Bar}^{\mathrm{enh}}: {\eur{O}}\operatorname{-alg}({\eur{C}}) \rightleftarrows {\eur{P}}\operatorname{-coalg}({\eur{C}}): \operatorname{coBar}^{\mathrm{enh}}{\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:adjunction_Koszul_dual}$$ where $${\mathrm{oblv}}_{{\eur{P}}}\circ \operatorname{Bar}_{\eur{O}}^{\mathrm{enh}}\simeq \operatorname{Bar}_{\eur{O}} \qquad\text{and}\qquad {\mathrm{oblv}}_{{\eur{O}}}\circ \operatorname{coBar}_{\eur{P}}^{\mathrm{enh}}\simeq \operatorname{coBar}_{\eur{P}}.$$ ### Turning Koszul duality into an equivalence In general, the pair of adjoint functors at  is not an equivalence. One of the main achievements of [@francis_chiral_2011] is to formulate a precise condition on the base category ${\eur{C}}$, namely the pro-nilpotent condition,[^8] which turns  into an equivalence. One of the main technical points of our paper is to prove another case where Koszul duality is still an equivalence, even when the categories involved are not pro-nilpotent. The two main instances of Koszul duality that are important in this paper are the duality between $\operatorname{Lie}$-algebras and $\operatorname{ComCoAlg}$-algebras, and $\operatorname{coLie}$-algebras and $\operatorname{ComAlg}$-algebras. ### The case of $\operatorname{Lie}$ and $\operatorname{ComCoAlg}$ We have the following equivalence of co-operads (see [@francis_chiral_2011]): $$\operatorname{Lie}^\vee \simeq \operatorname{ComCoAlg}[1],$$ where $$\operatorname{ComCoAlg}[1](n) \simeq k[n-1]$$ is equipped with the sign action of the symmetric group $\Sigma_n$. ### Equivalently, the functor $$[1]: {\eur{C}} \to {\eur{C}}$$ gives rise to an equivalence of categories $$[1]: \operatorname{ComCoAlg}[1]({\eur{C}}) \simeq \operatorname{ComCoAlg}({\eur{C}}).$$ ### This gives us the following diagram $$\xymatrix{ \operatorname{Lie}({\eur{C}}) {\ar@<{0.72ex}>}[ddrr]^{\operatorname{Chev}} {\ar@<{0.36ex}>}[dd]^{[1]} {\ar@<{0.36ex}>}[rr]^<<<<<<<<<<<<{\operatorname{Bar}_{\operatorname{Lie}}} && {\ar@<{0.36ex}>}[ll]^<<<<<<<<<<<{\operatorname{coBar}_{\operatorname{ComCoAlg}[1]}} \operatorname{ComCoAlg}[1]({\eur{C}}) {\ar@<{0.36ex}>}[dd]^{[1]} \\ \\ \operatorname{Lie}[-1]({\eur{C}}) {\ar@<{0.36ex}>}[uu]^{[-1]} {\ar@<{0.36ex}>}[rr]^{\operatorname{Bar}_{\operatorname{Lie}[-1]}} && {\ar@<{0.36ex}>}[ll]^{\operatorname{coBar}_{\operatorname{ComCoAlg}}} \operatorname{ComCoAlg}({\eur{C}}) {\ar@<{0.36ex}>}[uu]^{[-1]} \ar[uull]^{{\mathrm{Prim}}[-1]} }$$ We usually use $\operatorname{Chev}$ to denote $$[1]\circ \operatorname{Bar}_{\operatorname{Lie}} \simeq \circ \operatorname{Bar}_{\operatorname{Lie}[-1]} \circ [1]$$ and ${\mathrm{Prim}}[-1]$ to denote $$\operatorname{coBar}_{\operatorname{ComCoAlg}[1]}\circ [-1] \simeq [-1]\circ \operatorname{coBar}_{\operatorname{ComCoAlg}}. {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:Prim[-1]_and_coBar}$$ ### The case of $\operatorname{coLie}$ and $\operatorname{ComAlg}$ {#subsubsec:prelim_Koszul_coLie_ComAlg} Dually, we have the following equivalence of co-operads $$\operatorname{ComAlg}^\vee \simeq \operatorname{coLie}[1],$$ and similar to the above, the functor $$[1]: {\eur{C}} \to {\eur{C}}$$ gives rise to an equivalence of categories $$[1]: \operatorname{coLie}[1]({\eur{C}}) \simeq \operatorname{coLie}({\eur{C}}).$$ ### This gives us the following diagram $$\xymatrix{ \operatorname{ComAlg}({\eur{C}}) {\ar@<{0.72ex}>}[ddrr]^{\operatorname{coPrim}[1]} {\ar@<{0.36ex}>}[dd]^{[1]} {\ar@<{0.36ex}>}[rr]^<<<<<<<<<<<<<{\operatorname{Bar}_{\operatorname{ComAlg}}} && {\ar@<{0.36ex}>}[ll]^>>>>>>>>>>>>>{\operatorname{coBar}_{\operatorname{coLie}[1]}} \operatorname{coLie}[1]({\eur{C}}) {\ar@<{0.36ex}>}[dd]^{[1]} \\ \\ \operatorname{ComAlg}[-1]({\eur{C}}) {\ar@<{0.36ex}>}[uu]^{[-1]} {\ar@<{0.36ex}>}[rr]^>>>>>>>>>>>{\operatorname{Bar}_{\operatorname{ComAlg}[-1]}} && {\ar@<{0.36ex}>}[ll]^<<<<<<<<<<<{\operatorname{coBar}_{\operatorname{coLie}}} \operatorname{coLie}({\eur{C}}) {\ar@<{0.36ex}>}[uu]^{[-1]} \ar[uull]^{\operatorname{coChev}} }$$ As above, we usually use $\operatorname{coChev}$ to denote $$[-1] \circ \operatorname{coBar}_{\operatorname{coLie}} \simeq \operatorname{coBar}_{\operatorname{coLie}[1]} \circ [-1]$$ and $\operatorname{coPrim}[1]$ to denote $$[1] \circ \operatorname{Bar}_{\operatorname{ComAlg}} \simeq \operatorname{Bar}_{\operatorname{ComAlg}[-1]} \circ [1].$$ Turning Koszul duality into an equivalence ========================================== The goal of this section is to prove Theorem \[thm:intro:Koszul\_duality\_connectivity\_on\_Ran\]. We will start by examining the special case where $X$ is just a point, i.e. $\operatorname{Shv}(\operatorname{Ran}X) \simeq \operatorname{Shv}(X) \simeq {\mathrm{Vect}}$, and prove that Koszul duality provides a natural equivalence of categories $$\operatorname{Chev}:\operatorname{Lie}({\mathrm{Vect}}^{\leq -1}) \simeq \operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2}): {\mathrm{Prim}}[-1].$$ Even though this case is not strictly needed in the proof of the general case, it is interesting in its own right, as it allows us to predict the correct connectivity condition needed in the general case, whose precise statement and proof are presented in the final subsection. We recommend the reader to first read the case of ${\mathrm{Vect}}$, since it shares the same strategy as the main proof without the additional numerical complexity. The case of $\operatorname{Lie}$- and $\operatorname{ComCoAlg}$-algebras inside ${\mathrm{Vect}}$ {#subsec:Koszul_equivalence_Lie_ComCoAlg_in_Vect} ------------------------------------------------------------------------------------------------- We will now prove the following \[thm:equivalence\_koszul\_vec\_case\] $\operatorname{Chev}$ and ${\mathrm{Prim}}[-1]$ give rise to a pair of mutually inverse functors $$\operatorname{Chev}: \operatorname{Lie}({\mathrm{Vect}}^{\leq -1}) \rightleftarrows \operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2}): {\mathrm{Prim}}[-1]$$ Since $\operatorname{Chev}$ is defined as a colimit, it is easy to see that $\operatorname{Chev}|_{\operatorname{Lie}({\mathrm{Vect}}^{\leq -1})}$ lands in the correct subcategory cut out by the connectivity assumption ${\mathrm{Vect}}^{\leq -2}$. However, a priori, the same is not obvious for ${\mathrm{Prim}}[-1]$, being defined as a limit. It is, however, clear from the proof below that this in fact holds. Unless otherwise specified, our functors will be automatically restricted to the subcategories with the appropriate connectivity conditions. For example, we will write $\operatorname{Chev}$ instead of $\operatorname{Chev}|_{\operatorname{Lie}({\mathrm{Vect}}^{\leq -1})}$ in most cases. Note that Theorem \[thm:equivalence\_koszul\_vec\_case\] can be proved more generally for a presentable symmetric monoidal stable infinity category with a $t$-structure satisfying some mild properties. The pair of operad and co-operad $\operatorname{Lie}$ and $\operatorname{ComCoAlg}$ could also be made more general. The curious readers could take a look at the remarks at the end of this subsection. ### We follow a similar strategy as in [@francis_chiral_2011]. Namely, to prove that $\operatorname{Chev}$ and ${\mathrm{Prim}}[-1]$ are mutually inverse functors, it suffices to show that the left adjoint functor, $\operatorname{Chev}$, is fully-faithful, and the right adjoint functor, ${\mathrm{Prim}}[-1]$ is conservative. We start with the following \[lem:Prim\[-1\]\_is\_good\_Vec\_case\] The functor ${\mathrm{Prim}}[-1]|_{\operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2})}$ satisfies the following conditions (i) \[lem:Prim\[-1\]\_is\_good\_Vec\_case:sifted\_colims\] ${\mathrm{Prim}}[-1]$ commutes with sifted colimits. (ii) \[lem:Prim\[-1\]\_is\_good\_Vec\_case:monad\_isom\] The natural map $$\operatorname{Free}_{\operatorname{Lie}} \to {\mathrm{Prim}}[-1]\circ \operatorname{triv}_{\operatorname{ComCoAlg}}$$ is an equivalence. As in [@francis_chiral_2011 §4.1.8], this immediately implies the following corollary. For the sake of completeness, we include the proof here. \[cor:vect\_case\_Chev\_fully\_faithful\] $\operatorname{Chev}|_{\operatorname{Lie}({\mathrm{Vect}}^{\leq -1})}$ is fully faithful. It suffices to show that the unit map $${\mathrm{id}}\to {\mathrm{Prim}}[-1]\circ \operatorname{Chev}$$ is an equivalence. Since ${\mathrm{Prim}}[-1]$ commutes with sifted colimits by part  of Lemma \[lem:Prim\[-1\]\_is\_good\_Vec\_case\], it suffices to show that the following is an equivalence $$\operatorname{Free}_{\operatorname{Lie}} \to {\mathrm{Prim}}[-1]\circ \operatorname{Chev}\circ \operatorname{Free}_{\operatorname{Lie}},$$ since any $\operatorname{Lie}$-algebra could be written as a sifted colimit of the free ones.[^9] However, we know that (even without the connectivity condition) $$\operatorname{Chev}\circ \operatorname{Free}_{\operatorname{Lie}} \simeq \operatorname{triv}_{\operatorname{Lie}}$$ and hence, it suffices to show that $$\operatorname{Free}_{\operatorname{Lie}} \to {\mathrm{Prim}}[-1]\circ \operatorname{Chev}.$$ But now, we are done due to part  of Lemma \[lem:Prim\[-1\]\_is\_good\_Vec\_case\]. ### Before proving Lemma \[lem:Prim\[-1\]\_is\_good\_Vec\_case\], we start with a couple of preliminary observations. In essence, the lemma is a statement about commuting limits and colimits. In a stable infinity category, if, for instance, the limit is a finite one, then one can always do that. In our situation, $\operatorname{coBar}$ is causing troubles because it is defined as an infinite limit. The main idea of the proof is that when $$c\in \operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2}),$$ then even though $$\operatorname{coBar}_{\eur{\operatorname{ComCoAlg}}}(c)$$ is computed as an infinite limit, each of its cohomological degree will be controlled by finitely many of terms in the limit. ### For brevity’s sake, we will use ${\eur{P}}$ to denote the co-operad $\operatorname{ComCoAlg}$. Recall that in general, for any $$c\in \operatorname{ComCoAlg}({\mathrm{Vect}}),$$ we have $$\operatorname{coBar}_{{\eur{P}}}(c) = \operatorname{Tot}(\operatorname{coBar}^\bullet_{{\eur{P}}}(c))$$ where $\operatorname{coBar}^\bullet_{\eur{P}}(c)$ is a co-simplicial object. Let $$\operatorname{coBar}_{{\eur{P}}}^n(c) = \operatorname{Tot}(\operatorname{coBar}_{\eur{P}}^\bullet(c)|_{\Delta^{\leq n}})$$ be the limit over the restriction of the co-simplicial object to $\Delta^{\leq n}$. Then we have the following tower $$c\simeq \operatorname{coBar}_{\eur{P}}^0(c) \leftarrow \operatorname{coBar}_{\eur{P}}^1(c) \leftarrow \cdots \leftarrow \operatorname{coBar}_{{\eur{P}}}^n(c) \leftarrow \cdots$$ and $$\operatorname{coBar}_{\eur{P}}(c) \simeq \lim_n \operatorname{coBar}_{\eur{P}}^n(c).$$ \[lem:vec\_case\_stability\] Let $$c\in \operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2}).$$ Then, for all $n \geq 0$, the following natural map $$\operatorname{tr}_{\geq -2^{n+1}+n+1} \operatorname{coBar}_{\eur{P}}^n(c) \to \operatorname{tr}_{\geq -2^{n+1}+n+1} \operatorname{coBar}_{\eur{P}}^{n-1}(c).$$ is an equivalence. Let $F^n(c)$ denote the difference between $\operatorname{coBar}^n_{\eur{P}}(c)$ and $\operatorname{coBar}^{n-1}_{\eur{P}}(c)$, $$F^n(c) = \operatorname{Fib}(\operatorname{coBar}^n_{\eur{P}}(c) \to \operatorname{coBar}^{n-1}_{\eur{P}}(c)).$$ Then for $$c\in \operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2}),$$ we see that $$F^n(c) \in {\mathrm{Vect}}^{\leq -2\cdot 2^n + n} \simeq {\mathrm{Vect}}^{\leq -2^{n+1} +n}.$$ Indeed, this is because of the fact that $c\in {\mathrm{Vect}}^{\leq -2}$ and in the direct sum $$\operatorname{coBar}^\bullet_{\eur{P}}(c)([n]) = \bigoplus_{m\geq 1} {\eur{P}}^{\star n}(m) \otimes_{S_m} c^{\otimes m},$$ $m=2^n$ is the first summand where we have non-degenerate “(co-)cells.” As a consequence, $$\operatorname{tr}_{\geq -2^{n+1}+n+1} \operatorname{coBar}_{\eur{P}}^n(c) \to \operatorname{tr}_{\geq -2^{n+1}+n+1} \operatorname{coBar}_{\eur{P}}^{n-1}(c)$$ is an equivalence and we are done. Using the fact that infinite products preserve ${\mathrm{Vect}}^{\leq 0}$, the lemma above directly implies the following \[cor:vec\_case\_from\_coBar\_to\_coBarm\] Let $$c\in \operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2}).$$ Then, for any $n$, the following natural map $$\operatorname{tr}_{\geq -n} \operatorname{coBar}_{\eur{P}}(c) \to \operatorname{tr}_{\geq -n} \operatorname{coBar}_{\eur{P}}^m(c)$$ is an equivalence for all $m \gg 0$, where the bound depends only on $n$. The proof is now simple. In fact, we will only prove part , as the other one is almost identical. Note that due to , what we prove about $\operatorname{coBar}_{\eur{P}}$ implies the corresponding statement of ${\mathrm{Prim}}[-1]$, up to a shift. It suffices to show that for all $n$, we have $$\operatorname{tr}_{\geq -n} \operatorname{coBar}_{\eur{P}} (\operatorname*{colim}_\alpha c_\alpha) \simeq \operatorname{tr}_{\geq -n} \operatorname*{colim}_{\alpha} \operatorname{coBar}_{\eur{P}} (c_\alpha)$$ where $\alpha$ runs over some sifted diagram. But now, from Corollary \[cor:vec\_case\_from\_coBar\_to\_coBarm\], for all $m\gg 0$, we have $$\begin{aligned} \operatorname{tr}_{\geq -n} \operatorname{coBar}_{\eur{P}}(\operatorname*{colim}_\alpha c_\alpha) &\simeq \operatorname{tr}_{\geq -n} \operatorname{coBar}_{\eur{P}}^m (\operatorname*{colim}_\alpha c_\alpha) \simeq \operatorname{tr}_{\geq -n} \operatorname*{colim}_\alpha \operatorname{coBar}_{\eur{P}}^m(c_\alpha) \simeq \operatorname*{colim}_\alpha \operatorname{tr}_{\geq -n} \operatorname{coBar}_{\eur{P}}^m(c_\alpha) \\ &\simeq \operatorname*{colim}_\alpha \operatorname{tr}_{\geq -n} \operatorname{coBar}_{\eur{P}}(c_\alpha) \simeq \operatorname{tr}_{\geq -n} \operatorname*{colim}_\alpha \operatorname{coBar}_{\eur{P}}(c_\alpha).\end{aligned}$$ The cohomological estimate done above implies that $$\operatorname{coBar}_{\operatorname{ComCoAlg}}(c) \in \operatorname{Lie}[-1]({\mathrm{Vect}}^{\leq -2}),$$ or equivalently $${\mathrm{Prim}}[-1](c) \in \operatorname{Lie}({\mathrm{Vect}}^{\leq -1}),$$ when $$c\in \operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2}).$$ Indeed, from Corollary \[cor:vec\_case\_from\_coBar\_to\_coBarm\], we know that for some $m\gg 0$, $$\operatorname{tr}_{\geq -1} \operatorname{coBar}_{{\eur{P}}}(c) \simeq \operatorname{tr}_{\geq -1} \operatorname{coBar}_{{\eur{P}}}^m(c),$$ and moreover, a downward induction using Lemma \[lem:vec\_case\_stability\] shows that $$\operatorname{tr}_{\geq -1} \operatorname{coBar}_{{\eur{P}}}^m(c) \simeq \operatorname{tr}_{\geq -1} \operatorname{coBar}^0_{\eur{P}}(c) \simeq \operatorname{tr}_{\geq -1} c \simeq 0.$$ ### The following result will conclude the proof of Theorem \[thm:equivalence\_koszul\_vec\_case\]. \[lem:vec\_case\_Prim\[-1\]\_conservative\] The functor $${\mathrm{Prim}}[-1]: \operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2}) \to \operatorname{Lie}({\mathrm{Vect}}^{\leq -1})$$ is conservative. It suffices to show that $$\operatorname{coBar}_{\eur{P}}: \operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2}) \to \operatorname{Lie}[-1]({\mathrm{Vect}}^{\leq -2})$$ is conservative, and we will prove that by contradiction. Namely, let $$f: c_1\to c_2$$ be a morphism in $\operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2})$ such that $f$ is not an equivalence. Suppose that $$\operatorname{coBar}_{\eur{P}}(f): \operatorname{coBar}_{\eur{P}}(c_1) \to \operatorname{coBar}_{\eur{P}}(c_2)$$ is an equivalence, we will derive a contradiction. Let $k$ be the smallest number such that $$\operatorname{tr}_{\geq -k}(f): \operatorname{tr}_{\geq -k} c_1 \to \operatorname{tr}_{\geq -k} c_2$$ is not an equivalence. Now, by Corollary \[cor:vec\_case\_from\_coBar\_to\_coBarm\], we know that there is some $m\gg 0$ such that $$\operatorname{tr}_{\geq -k} \operatorname{coBar}_{\eur{P}}(c_i) \simeq \operatorname{tr}_{\geq -k} \operatorname{coBar}_{\eur{P}}^m(c_i)$$ for $i\in \{1, 2\}$. Thus, we know that $$\operatorname{tr}_{\geq -k} \operatorname{coBar}_{\eur{P}}^m(c_1) \to \operatorname{tr}_{\geq -k} \operatorname{coBar}_{\eur{P}}^m(c_2)$$ is an equivalence. By an estimate similar to that of Lemma \[lem:vec\_case\_stability\], we see that $$\operatorname{tr}_{\geq -k} F^n(c_1) \simeq \operatorname{tr}_{\geq -k} F^n(c_2)$$ for all $n \geq 1$. Indeed, the difference between $F^n(c_1)$ and $F^n(c_2)$ lies in cohomological degrees $$\leq -2(2^n-1) - k + n = -2^{n+1} - k + n + 2 < -k, \quad \forall n \geq 1.$$ And hence, a downward induction, starting from $n=m$, using the diagram $$\xymatrix{ F^n(c_1) \ar[d] \ar[r] & \operatorname{coBar}_{\eur{P}}^n(c_1) \ar[d] \ar[r] & \operatorname{coBar}_{\eur{P}}^{n-1}(c_1) \ar[d] \\ F^n(c_2) \ar[r] & \operatorname{coBar}_{\eur{P}}^n(c_2) \ar[r] & \operatorname{coBar}_{\eur{P}}^{n-1}(c_2) }$$ implies that $$\tau_{\geq -k} c_1 \simeq \tau_{\geq -k} c_2,$$ which contradicts our original assumption. Hence, we are done. Note that the proof we gave above could be carried out in a more general setting. Namely, the only properties of ${\mathrm{Vect}}$ that we used are (i) The symmetric monoidal structure is right exact (namely, it preserved ${\mathrm{Vect}}^{\leq 0}$). (ii) The $t$-structure on ${\mathrm{Vect}}$ is left separated. (iii) Infinite products preserve ${\mathrm{Vect}}^{\leq 0}$. We can also replace the operad $\operatorname{Lie}$ by any operad ${\eur{O}}$ such that (i) ${\eur{O}}$ is classical, i.e. it lies in the heart of the $t$-structure of ${\mathrm{Vect}}$. (ii) ${\eur{O}}^\vee[-1]$ is also classical. (iii) ${\eur{O}}(1) \simeq \Lambda$ (as we already assume throughout this paper). Higher enveloping algebras -------------------------- This subsection serves as the topological analogue of the results proved in the next one. The main reference of this part is [@gaitsgory_study_????]. ### Let $$\mathfrak{g} \in \operatorname{Lie}({\mathrm{Vect}}).$$ Then one can form its $E_n$-universal enveloping algebra $$U_{E_n}(\mathfrak{g}) \in E_n({\mathrm{Vect}})$$ by applying the following sequence of functors $$\xymatrix{ \operatorname{Lie}({\mathrm{Vect}}) \ar[rr]^<<<<<<<<<<{\Omega^{\times n} \simeq [-n]} && E_n(\operatorname{Lie}({\mathrm{Vect}})) \ar[rr]^<<<<<<<<<<{E_n(\operatorname{Chev})} && E_n(\operatorname{ComCoAlg}({\mathrm{Vect}})) \ar[rr]^<<<<<<<<<<{{\mathrm{oblv}}_{\operatorname{ComCoAlg}}} && E_n({\mathrm{Vect}}) }$$ where $E_n(\operatorname{Lie}({\mathrm{Vect}}))$ and $E_n(\operatorname{ComCoAlg}({\mathrm{Vect}}))$ are categories of $E_n$-algebras with respect to the Cartesian monoidal structure on $\operatorname{Lie}({\mathrm{Vect}})$ and $\operatorname{ComCoAlg}({\mathrm{Vect}})$ respectively (note that the latter on is just the given by $\otimes$ in ${\mathrm{Vect}}$). ### It is proved in [@gaitsgory_study_????] that $[-n]$ induces an equivalence $$[-n]: \operatorname{Lie}({\mathrm{Vect}}) \simeq E_n(\operatorname{Lie}({\mathrm{Vect}})) : [n].$$ Moreover, we know from Theorem \[thm:equivalence\_koszul\_vec\_case\] that $$E_n(\operatorname{Chev}): E_n(\operatorname{Lie}({\mathrm{Vect}}^{\leq -1})) \to E_n(\operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2})).$$ Thus, we get the following equivalence of categories $$\operatorname{Lie}({\mathrm{Vect}}^{\leq -n-1}) \simeq E_n(\operatorname{ComCoAlg}({\mathrm{Vect}}^{\leq -2})). {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:higher_enveloping_alg_equivalence}$$ ### The equivalence  is precisely what we are looking for in the context of factorization algebras on the Ran space in the following subsection. One part of the work is to find connectivity assumptions on $\operatorname{Shv}(\operatorname{Ran}X)$ which mirror those in ${\mathrm{Vect}}^{\leq -n-1}$ and ${\mathrm{Vect}}^{\leq -2}$ respectively. The case of $\operatorname{\operatorname{Lie}^\star}$- and ${\operatorname{ComCoAlg}^\star}$-algebras on $\operatorname{Ran}X$ {#subsec:koszul_equivalence_Lie_ComCoAlg_Ran_case} ------------------------------------------------------------------------------------------------------------------------------ We come to the precise formulation and the proof of Theorem \[thm:intro:Koszul\_duality\_connectivity\_on\_Ran\]. \[defn:connectivity\_constraints\_Lie\_ComCoAlg\_Ran\] Let $\operatorname{Shv}(\operatorname{Ran}X)^{\leq c_{cA}}$ and $\operatorname{Shv}(\operatorname{Ran}X)^{\leq c_L}$ denote the full subcategory of $\operatorname{Shv}(\operatorname{Ran}X)$ consisting of sheaves ${\eur{F}}$ such that for all non-empty finite sets $I$, $${\eur{F}}|_{{\overset{\circ}{X}{}^{I}}} \in \operatorname{Shv}({\overset{\circ}{X}{}^{I}})^{\leq (-1-d)|I| - 1},$$ and respectively, $${\eur{F}}|_{{\overset{\circ}{X}{}^{I}}} \in \operatorname{Shv}({\overset{\circ}{X}{}^{I}})^{\leq (-1-d)|I|}.$$ Here, we use the perverse $t$-structure, and $X$ is a scheme of pure dimension $d$. We will use $$\operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X)^{\leq c_L} \qquad\text{and}\qquad {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)^{\leq c_{cA}}$$ to denote $$\operatorname{\operatorname{Lie}^\star}(\operatorname{Shv}(\operatorname{Ran}X)^{\leq c_L}) \qquad\text{and}\qquad {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)^{\leq c_{cA}}$$ respectively. With these connectivity assumptions in mind, we will prove the following \[thm:Koszul\_duality\_connectivity\_on\_Ran\] We have the following commutative diagram $$\xymatrix{ \operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X)^{\leq c_L} \ar@{=}[rr]^<<<<<<<<<<{\operatorname{Chev}}_<<<<<<<<<<{{\mathrm{Prim}}[-1]} && {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)^{\leq c_{cA}} \\ \operatorname{\operatorname{Lie}^\star}(X)^{\leq c_L} \ar@{^(->}[u] \ar@{=}[rr]^{\operatorname{Chev}}_{{\mathrm{Prim}}[-1]} && \operatorname{\operatorname{coFact}^\star}(X)^{\leq c_{cA}} \ar@{^(->}[u] } {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:diagram_in_thm:Koszul_duality_connectivity_on_Ran}$$ where $\leq c_L$ and $\leq c_{cA}$ denote the connectivity constraints given in Definition \[defn:connectivity\_constraints\_Lie\_ComCoAlg\_Ran\], and where $\operatorname{Chev}$ and ${\mathrm{Prim}}[-1]$ are the functors coming from Koszul duality. \[rmk:Ran\_case\_supress\_conn\_cond\] As in the case of ${\mathrm{Vect}}$, our functors will be automatically restricted to the subcategories with the appropriate connectivity conditions, unless otherwise specified. For example, we will write $\operatorname{Chev}$ instead of $\operatorname{Chev}|_{\operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X)^{\leq c_L}}$ in most cases. As in the case of ${\mathrm{Vect}}$, the pair of operad and co-operad $\operatorname{Lie}$ and $\operatorname{ComCoAlg}$ could be replaced by an operad ${\eur{O}}$ and its Koszul dual ${\eur{O}}^\vee$ such that[^10] (i) ${\eur{O}}$ is classical, i.e. it lies in the heart of the $t$-structure of ${\mathrm{Vect}}$. (ii) ${\eur{O}}^\vee[-1]$ is also classical. (iii) ${\eur{O}}(1) \simeq \Lambda$ (as we already assume throughout this paper). We start with a preliminary lemma, which ensures that the categories $$\operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X)^{\leq c_L} \qquad\text{and}\qquad {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)^{\leq c_{cA}}$$ are actually well-defined. \[lem:otimesstar\_preserves\_cL\_ccA\] The subcategories $\operatorname{Shv}(\operatorname{Ran}X)^{\leq c_L}$ and $\operatorname{Shv}(\operatorname{Ran}X)^{\leq c_{cA}}$ are preserved under the ${\otimes^\star}$-monoidal structure on $\operatorname{Shv}(\operatorname{Ran}X)$. Recall from  that if $${\eur{F}}_1, \dots, {\eur{F}}_k \in \operatorname{Shv}(\operatorname{Ran}X),$$ then from the definition of ${\otimes^\star}$, we have $$({\eur{F}}_1 {\otimes^\star}\cdots {\otimes^\star}{\eur{F}}_k)|_{{\overset{\circ}{X}{}^{I}}} \simeq \bigoplus_{I = \cup_{i=1}^k I_i} \Delta^!_{\sqcup_{i=1}^k I_i {\twoheadrightarrow}\cup_{i=1}^k I_i} ({\eur{F}}_1 \boxtimes\cdots\boxtimes {\eur{F}}_k)|_{(\prod_{i=1}^k {\overset{\circ}{X}{}^{I_i}})_{\mathrm{disj}}}. {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:otimesstar_as_a_sum}$$ Now, suppose that $${\eur{F}}_1, \dots, {\eur{F}}_k \in \operatorname{Shv}(\operatorname{Ran}X)^{\leq c_L},$$ then we see that each summand in  lies in perverse cohomological degrees $$\begin{aligned} &\leq (-1-d)\sum_{i=1}^k |I_i| + d\left(\sum_{i=1}^k |I_i| - |I|\right) \\ &\leq -\sum_{i=1}^k |I_i| - d|I| \\ &\leq (-1-d)|I|.\end{aligned}$$ Here, the first inequality is due to the fact that the map $${\overset{\circ}{X}{}^{I}} \to \prod_{i=1}^k ({\overset{\circ}{X}{}^{I_i}})_{{\mathrm{disj}}}$$ is a regular embedding, and that the (perverse) cohomological amplitude of the $!$-pullback along a regular embedding is equal to the codimension. Thus, this implies that $${\eur{F}}_1{\otimes^\star}\cdots{\otimes^\star}{\eur{F}}_k \in \operatorname{Shv}(\operatorname{Ran}X)^{\leq c_L}.$$ Similarly, suppose that $${\eur{F}}_1, \dots, {\eur{F}}_k \in \operatorname{Shv}(\operatorname{Ran}X)^{\leq c_{cA}},$$ then each summand in  lies in perverse cohomological degrees $$\begin{aligned} &\leq (-1-d)\sum_{i=1}^k |I_i| - k + d\left(\sum_{i=1}^k |I_i| - |I|\right) {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:cohomological_estimates_cA_otimesstar} \\ &\leq -\sum_{i=1}^k |I_i| - k - d|I| \\ &\leq (-1-d)|I| - 1.\end{aligned}$$ Thus, $${\eur{F}}_1 {\otimes^\star}\cdots {\otimes^\star}{\eur{F}}_k \in \operatorname{Shv}(\operatorname{Ran}X)^{\leq c_{cA}},$$ which concludes the proof. ### Back to Theorem \[thm:Koszul\_duality\_connectivity\_on\_Ran\]. First, we will prove the equivalence on the top row of . And then, we will show that it induces an equivalence between the corresponding sub-categories on the bottom row. As in the case of ${\mathrm{Vect}}$, to prove that $\operatorname{Chev}$ and ${\mathrm{Prim}}[-1]$ are mutually inverse functors, it suffices to show that $\operatorname{Chev}$ is fully-faithful, and ${\mathrm{Prim}}[-1]$ is conservative. The following lemma will help us achieve this goal. \[lem:Prim\[-1\]\_is\_good\_Ran\_case\] The functor ${\mathrm{Prim}}[-1]|_{{\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)^{\leq c_{cA}}}$ satisfies the following conditions (see Remark \[rmk:Ran\_case\_supress\_conn\_cond\]) (i) ${\mathrm{Prim}}[-1]$ commutes with sifted colimits. (ii) The natural map $$\operatorname{Free}_{\operatorname{Lie}} \to {\mathrm{Prim}}[-1]\circ \operatorname{triv}_{\operatorname{ComCoAlg}}$$ is an equivalence. As in Corollary \[cor:vect\_case\_Chev\_fully\_faithful\], this immediately implies the following \[cor:Ran\_case\_Chev\_fully\_faithful\] $\operatorname{Chev}|_{\operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X)^{\leq c_L}}$ is fully faithful. ### In essence, the strategy we follow here is identical to that of the ${\mathrm{Vect}}$ case even though the actual execution might seem somewhat more involved. The main observation (which is new compared to the case of ${\mathrm{Vect}}$) is that to prove the equivalences involved in Lemma \[lem:Prim\[-1\]\_is\_good\_Ran\_case\], it suffices to prove them after after pulling back to ${\overset{\circ}{X}{}^{I}}$ for each non-empty finite set $I$. ### In general, for any $${\eur{A}} \in {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)^{\leq c_{cA}},$$ we have $$\operatorname{coBar}_{\operatorname{ComCoAlg}}({\eur{A}}) = \operatorname{Tot}(\operatorname{coBar}^\bullet_{\operatorname{ComCoAlg}}({\eur{A}})),$$ where $\operatorname{coBar}^\bullet_{\operatorname{ComCoAlg}}({\eur{A}})$ is a co-simplicial object. Let $$\operatorname{coBar}^n_{\operatorname{ComCoAlg}}({\eur{A}}) = \operatorname{Tot}(\operatorname{coBar}^\bullet_{\operatorname{ComCoAlg}}({\eur{A}})|_{\Delta^{\leq n}}).$$ Then, we have the following tower $${\eur{A}} \simeq \operatorname{coBar}^0_{\operatorname{ComCoAlg}}({\eur{A}}) \leftarrow \operatorname{coBar}^1_{\operatorname{ComCoAlg}}({\eur{A}}) \leftarrow \cdots$$ and $$\operatorname{coBar}_{\operatorname{ComCoAlg}}({\eur{A}}) \simeq \lim_n \operatorname{coBar}^n_{\operatorname{ComCoAlg}}({\eur{A}}).$$ ### Let $$F^n({\eur{A}}) = \operatorname{Fib}(\operatorname{coBar}^n_{\operatorname{ComCoAlg}}({\eur{A}}) \to \operatorname{coBar}^{n-1}_{\operatorname{ComCoAlg}}({\eur{A}})),$$ and let $I$ be a non-empty finite set. Using the same argument as in the case of ${\mathrm{Vect}}$ in combination with the cohomological estimate , we see that $F^n({\eur{A}})|_{{\overset{\circ}{X}{}^{I}}}$ lives in cohomological degrees $$\begin{aligned} &\leq (-1-d)\sum_{i=1}^{2^n} |I_i| - 2^n + d\left(\sum_{i=1}^{2^n} |I_i| -|I|\right) + n \\ &= -\sum_{i=1}^{2^n} |I_i| - 2^n - d|I| + n \\ &\leq -2^{n+1} - d|I| + n\end{aligned}$$ which goes to $-\infty$ when $n \to \infty$. This gives us the following analog of Lemma \[lem:vec\_case\_stability\]. \[lem:ran\_case\_stability\] Let $${\eur{A}} \in {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)^{\leq c_{cA}}.$$ Then, for any $n$ and $I$, the following natural map $$\operatorname{tr}_{\geq -2^{n+1} - d|I| + n + 1} (\operatorname{coBar}^n_{\operatorname{ComCoAlg}}({\eur{A}})|_{{\overset{\circ}{X}{}^{I}}}) \to \operatorname{tr}_{\geq -2^{n+1} - d|I| + n + 1} (\operatorname{coBar}^{n-1}_{\operatorname{ComCoAlg}}({\eur{A}})|_{{\overset{\circ}{X}{}^{I}}})$$ is an equivalence. This implies the following result, which is parallel to Corollary \[cor:vec\_case\_from\_coBar\_to\_coBarm\]. \[cor:Ran\_case\_from\_coBar\_to\_coBarm\] Let $${\eur{A}} \in {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)^{\leq c_{cA}}.$$ Then, for any $n$ and $I$, the following natural map $$\operatorname{tr}_{\geq -n} (\operatorname{coBar}_{\operatorname{ComCoAlg}}({\eur{A}})|_{{\overset{\circ}{X}{}^{I}}}) \to \operatorname{tr}_{\geq -n} (\operatorname{coBar}_{\operatorname{ComCoAlg}}^m({\eur{A}})|_{{\overset{\circ}{X}{}^{I}}})$$ is an equivalence, when $m \gg 0$ depending only on $n$ and $I$. ### Now, Lemma \[lem:ran\_case\_stability\] and Corollary \[cor:Ran\_case\_from\_coBar\_to\_coBarm\] allow us to conclude the proof of Lemma \[lem:Prim\[-1\]\_is\_good\_Ran\_case\], and hence, Corollary \[cor:Ran\_case\_Chev\_fully\_faithful\], as in the ${\mathrm{Vect}}$ case. Note that when $X$ is a point, namely when $d = \dim X = 0$, the cohomological estimates in Lemma \[lem:ran\_case\_stability\] recover those of Lemma \[lem:vec\_case\_stability\]. To finish with the top equivalence in , we need the following \[lem:Ran\_case\_Prim\[-1\]\_conservative\] The functor $${\mathrm{Prim}}[-1]: {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)^{\leq c_{cA}} \to \operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X)^{\leq c_L}$$ is conservative. It suffices to show that $$\operatorname{coBar}_{\operatorname{ComCoAlg}}: {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)^{\leq c_{cA}} \to \operatorname{\operatorname{Lie}^\star}[-1](\operatorname{Ran}X)^{\leq c_{cA}}$$ is conservative, and we will do so by contradiction. Namely, let $$f: {\eur{A}}_1 \to {\eur{A}}_2$$ be a morphism in ${\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)^{\leq c_{cA}}$ that is not an equivalence. Suppose that $$\operatorname{coBar}_{\operatorname{ComCoAlg}}(f): \operatorname{coBar}_{\operatorname{ComCoAlg}}({\eur{A}}_1) \to \operatorname{coBar}_{\operatorname{ComCoAlg}}({\eur{A}}_2)$$ is an equivalence, we will derive a contradiction. Let $I$ be the smallest set such that the map $$f|_{{\overset{\circ}{X}{}^{I}}}: {\eur{A}}_1|_{{\overset{\circ}{X}{}^{I}}} \to {\eur{A}}_2|_{{\overset{\circ}{X}{}^{I}}}$$ is not an equivalence. Let $k \geq 0$ be the smallest number such that $$\operatorname{tr}_{\geq (-1-d)|I|-1-k}({\eur{A}}_1|_{{\overset{\circ}{X}{}^{I}}}) \to \operatorname{tr}_{\geq (-1-d)|I|-1-k} ({\eur{A}}_2|_{{\overset{\circ}{X}{}^{I}}})$$ is not an equivalence. By Corollary \[cor:Ran\_case\_from\_coBar\_to\_coBarm\], we know that there exists some $m\gg 0$ such that $$\operatorname{tr}_{\geq (-1-d)|I|-1-k} (\operatorname{coBar}_{\operatorname{ComCoAlg}}({\eur{A}}_i)|_{{\overset{\circ}{X}{}^{I}}}) \simeq \operatorname{tr}_{\geq (-1-d)|I|-1-k} (\operatorname{coBar}^m_{\operatorname{ComCoAlg}}({\eur{A}}_i)|_{{\overset{\circ}{X}{}^{I}}})$$ for $i \in \{1, 2\}$. Thus, we get the following equivalence $$\operatorname{tr}_{\geq (-1-d)|I|-1-k}(\operatorname{coBar}^m_{\operatorname{ComCoAlg}}({\eur{A}}_1)|_{{\overset{\circ}{X}{}^{I}}}) \simeq \operatorname{tr}_{\geq (-1-d)|I|-1-k} (\operatorname{coBar}^m_{\operatorname{ComCoAlg}}({\eur{A}}_2)|_{{\overset{\circ}{X}{}^{I}}}).$$ But observe that if we let $$F^n({\eur{A}}_i) = \operatorname{Fib}(\operatorname{coBar}_{\operatorname{ComCoAlg}}^n({\eur{A}}_i) \to \operatorname{coBar}_{\operatorname{ComCoAlg}}^{n-1}({\eur{A}}_i))$$ then the difference between $F^n({\eur{A}}_1)|_{{\overset{\circ}{X}{}^{I}}}$ and $F^n({\eur{A}}_2)|_{{\overset{\circ}{X}{}^{I}}}$ lies in cohomological degrees $$\begin{aligned} &\leq (-1-d)|I| - 1 - k + (-1-d)\sum_{i=1}^{2^n-1} |I_i| - (2^n - 1) + n + d\left(|I| + \sum_{i=1}^{2^n - 1}|I_i| - |I|\right) \\ &\leq (-1-d)|I| - 1 -k - \sum_{i=1}^{2^n-1} |I_i| - 2^n +1 + n \\ &< (-1-d)|I| - 1 - k.\end{aligned}$$ This implies that for $n\geq 1$, $$\operatorname{tr}_{\geq (-1-d)|I|-1-k} (F^n({\eur{A}}_1)|_{{\overset{\circ}{X}{}^{I}}}) \simeq \operatorname{tr}_{\geq (-1-d)|I|-1-k} (F^n({\eur{A}}_2)|_{{\overset{\circ}{X}{}^{I}}}).$$ Thus, as in the case of ${\mathrm{Vect}}$, a downward induction implies that $$\operatorname{tr}_{\geq (-1-d)|I|-1-k}({\eur{A}}_1|_{{\overset{\circ}{X}{}^{I}}}) \simeq \operatorname{tr}_{\geq (-1-d)|I|-1-k}({\eur{A}}_2|_{{\overset{\circ}{X}{}^{I}}}),$$ which contradicts our original assumption, and we are done. ### Corollary \[cor:Ran\_case\_Chev\_fully\_faithful\] and Lemma \[lem:Ran\_case\_Prim\[-1\]\_conservative\] together prove the equivalence on the top row of diagram . For the equivalence in the bottom row, it suffices to show that for $$\mathfrak{g} \in \operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X)^{\leq c_L},$$ $\operatorname{Chev}(\mathfrak{g})$ is factorizable if and only if $\mathfrak{g} \in \operatorname{\operatorname{Lie}^\star}(X)^{\leq c_L}$. ### For the “if” direction, recall that as a consequence of [@francis_chiral_2011 Thm. 6.4.2 and 5.2.1], we know that the functor $$\operatorname{Chev}: \operatorname{\operatorname{Lie}^\star}(X) \to {\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X)$$ lands inside the full-subcategory $\operatorname{\operatorname{coFact}^\star}(X)$ of factorizable co-algebras. We thus get a functor $$\operatorname{Chev}: \operatorname{\operatorname{Lie}^\star}(X)^{\leq c_L} \to \operatorname{\operatorname{coFact}^\star}(X)^{\leq c_{cA}},$$ which settles the “if” direction. ### For the “only if” direction, let $$\mathfrak{g} \in \operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X)^{\leq c_L}$$ whose support does not lie in $X$. We will show that $\operatorname{Chev}\mathfrak{g}$ is not factorizable. Using the $\operatorname{ass-gr}\circ \operatorname{addFil}$ trick (see §\[sec:appendix:addFil\_trick\]), it suffices to prove for the case where $\mathfrak{g}$ is a trivial (i.e. abelian) Lie algebra. In that case, we know that $$\operatorname{Chev}\mathfrak{g} = \operatorname{Sym}^{>0}(\mathfrak{g}[1]),$$ where $\operatorname{Sym}$ is taken using the ${\otimes^\star}$-monoidal structure. Let $I$ be the smallest set, with $|I| > 1$, such that $\mathfrak{g}|_{{\overset{\circ}{X}{}^{I}}} \not\simeq 0$. Now, it’s easy to see that $\operatorname{Sym}^{>0}(\mathfrak{g}[1])$ fails the factorizability condition at ${\overset{\circ}{X}{}^{I}}$, which concludes the “only if” direction. Factorizability of $\operatorname{coChev}$ ========================================== In this section, we will prove Theorem \[thm:intro:factorizability\_coChev\], which asserts that when $$\mathfrak{g} \in \operatorname{\operatorname{coLie}^\star}(X)$$ satisfies a certain co-connectivity constraint, the commutative algebra $$\operatorname{coChev}(\mathfrak{g}) \in \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)$$ is factorizable. Note that an analog of this result, where $\operatorname{coChev}$ is replaced by $\operatorname{Chev}$, has been proved in [@francis_chiral_2011] (and in fact, we used this result in the previous section). The main difficulties of the $\operatorname{coChev}$ case stem from the fact that, unlike $\operatorname{Chev}$, $\operatorname{coChev}$ is defined as a limit, and most of the functors that we want it to interact with don’t generally commute with limits. As above, our main strategy is to introduce a certain co-connectivity condition to ensure that when one takes the limit of a diagram involving objects satisfying it, the answer, in some sense, converges instead of running off to infinity, so we still have a good control over it. We start with the precise statement of the theorem. Then, after a quick digression on the various notions related to the convergence of a limit, we will present the main strategy. Finally, the proof itself will be given. The statement ------------- We start with the co-connectivity conditions. \[defn:connectivity\_constraints\_coLie\_ComAlg\_Ran\] Let $\operatorname{Shv}(\operatorname{Ran}X)^{\geq n}$ denote the full subcategory of $\operatorname{Shv}(\operatorname{Ran}X)$ consisting of sheaves ${\eur{F}}$ such that for all non-empty finite sets $I$, $${\eur{F}}|_{{\overset{\circ}{X}{}^{I}}} \in \operatorname{Shv}({\overset{\circ}{X}{}^{I}})^{\geq n},$$ As before, we use the perverse $t$-structure. We will use $$\operatorname{\operatorname{coLie}^\star}(\operatorname{Ran}X)^{\geq n} \qquad\text{and}\qquad \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq n}$$ to denote $$\operatorname{\operatorname{coLie}^\star}(\operatorname{Shv}(\operatorname{Ran}X)^{\geq n}) \qquad\text{and}\qquad \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Shv}(\operatorname{Ran}X)^{\geq n})$$ respectively. We will prove the following \[thm:factorizability\_coChev\] Restricted to the full subcategory $\operatorname{\operatorname{coLie}^\star}(X)^{\geq 1}$, the functor $\operatorname{coChev}$ factors through $\operatorname{\operatorname{Fact}^\star}$, i.e. $$\xymatrix{ \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1} \ar[dr]_{\operatorname{coChev}} \ar[rr]^{\operatorname{coChev}} && \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X) \\ & \operatorname{\operatorname{Fact}^\star}(X) \ar@{^(->}[ur] }$$ In other words, $\operatorname{coChev}\mathfrak{g}$ is factorizable when $\mathfrak{g} \in \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1}$. Stabilizing co-filtrations and decaying sequences (a digression) ---------------------------------------------------------------- In this subsection, we describe a condition on co-filtered and graded objects which make them behave nicely with respect to taking limits. Let ${\eur{C}}$ be a stable infinity category equipped with a $t$-structure. Then, a co-filtered object $c\in {\eur{C}}^{\operatorname{coFil}^{>0}}$ (see §\[sec:appendix:cofiltration\_addCoFil\]) is said to stabilize if for all $n$, $$\operatorname{tr}_{\leq n} c_m \rightarrow \operatorname{tr}_{\leq n} c_{m+1}$$ is an equivalence for all $m\gg 0$. A graded object in $c\in {\eur{C}}^{{\mathrm{gr}}^{>0}}$ is said to be decaying if for all $n$ if $$\operatorname{tr}_{\leq n} c_m \simeq 0$$ for all $m\gg 0$. We use ${\eur{C}}^{\operatorname{coFil}^{>0}, {\mathrm{stab}}}$ and ${\eur{C}}^{{\mathrm{gr}}^{>0}, {\mathrm{decay}}}$ to denote the subcategories of ${\eur{C}}^{\operatorname{coFil}^{>0}}$ and ${\eur{C}}^{{\mathrm{gr}}^{>0}}$ consisting of stabilizing and decaying objects respectively. We have the following lemmas, whose proofs are straightforward. Let $c\in {\eur{C}}^{\operatorname{coFil}^{>0}}$. Then $$c\in {\eur{C}}^{\operatorname{coFil}^{>0} , {\mathrm{stab}}}$$ if and only if $$\operatorname{ass-gr}c \in {\eur{C}}^{{\mathrm{gr}}^{>0}, {\mathrm{decay}}}.$$ \[lem:limit\_of\_stabilizing\_cofiltration\] If $c \in {\eur{C}}^{\operatorname{coFil}^{>0}, {\mathrm{stab}}}$, then for each $n$, the natural map $$\tau_{\leq n} {\mathrm{oblv}}_{\operatorname{coFil}} c \to \tau_{\leq n} c_m$$ is an equivalence when $m\gg 0$. By throwing away finitely many terms at the beginning, without loss of generality, we can assume that the natural maps $$\tau_{\leq n+1} c_i \to \tau_{\leq n+1} c_j, \qquad \forall i\geq j>0$$ are all equivalences. Now, it suffices to show that the following map is an equivalence $$\tau_{\leq n} \lim_i c_i \to \tau_{\leq n} c_1.$$ Equivalently, it suffices to show that $$\operatorname{Fib}(\lim_i c_i \to c_1) \in {\eur{C}}^{\geq n+1}.$$ However, $$\operatorname{Fib}(\lim_i c_i \to c_1) \simeq \lim_i(\operatorname{Fib}(c_i \to c_1)) \in {\eur{C}}^{\geq n+1}$$ because $$\operatorname{Fib}(c_i \to c_1) \in {\eur{C}}^{\geq n+1}, \qquad\forall i.$$ Hence, we are done, since $$i_{\geq n+1}: {\eur{C}}^{\geq n+1} \to {\eur{C}}$$ commutes with limits (see §\[subsubsec:t-structure\_convention\]). The natural transformation $$\bigoplus \to \prod$$ between functors $${\eur{C}}^{{\mathrm{gr}}^{>0}, {\mathrm{decay}}} \to {\eur{C}}$$ is an equivalence. Note that $$\prod_i c_i \simeq \lim_k \bigoplus_{i\leq k} c_i.$$ Moreover, since the sequence we are taking the limit over stabilizes, the result follows as a direct corollary of Lemma \[lem:limit\_of\_stabilizing\_cofiltration\]. ### The various definitions and observations above have straightforward analogues in the case of sheaves on the $\operatorname{Ran}$ space. A co-filtered sheaf ${\eur{F}} \in \operatorname{Shv}(\operatorname{Ran}X)^{\operatorname{coFil}^{>0}}$ is said to stabilize if for any non-empty finite set $I$, $${\eur{F}}|_{{\overset{\circ}{X}{}^{I}}} \in \operatorname{Shv}({\overset{\circ}{X}{}^{I}})^{\operatorname{coFil}^{>0}, {\mathrm{stab}}}.$$ Similarly, a graded sheaf ${\eur{F}} \in \operatorname{Shv}(\operatorname{Ran}X)^{{\mathrm{gr}}^{>0}}$ is said to be decaying if for any non-empty finite set $I$, $${\eur{F}}|_{{\overset{\circ}{X}{}^{I}}} \in \operatorname{Shv}({\overset{\circ}{X}{}^{I}})^{{\mathrm{gr}}^{>0}, {\mathrm{decay}}}.$$ We use $\operatorname{Shv}(\operatorname{Ran}X)^{\operatorname{coFil}^{>0}, {\mathrm{stab}}}$ and $\operatorname{Shv}(\operatorname{Ran}X)^{{\mathrm{gr}}^{>0}, {\mathrm{decay}}}$ to denote the sub-categories of $\operatorname{Shv}(\operatorname{Ran}X)^{\operatorname{coFil}^{>0}}$ and $\operatorname{Shv}(\operatorname{Ran}X)^{{\mathrm{gr}}^{>0}}$ consisting of stabilizing and decaying objects, respectively. It’s straightforward to see that the following analogs of the lemmas above still hold in this setting. \[lem:stab\_decay\_equivalence\_Ran\_case\] Let ${\eur{F}} \in \operatorname{Shv}(\operatorname{Ran}X)^{\operatorname{coFil}^{>0}}$. Then $${\eur{F}} \in \operatorname{Shv}(\operatorname{Ran}X)^{\operatorname{coFil}^{>0}, {\mathrm{stab}}}$$ if and only if $$\operatorname{ass-gr}{\eur{F}} \in \operatorname{Shv}(\operatorname{Ran}X)^{{\mathrm{gr}}^{>0}, {\mathrm{decay}}}.$$ If ${\eur{F}} \in \operatorname{Shv}(\operatorname{Ran}X)^{\operatorname{coFil}^{>0}, {\mathrm{stab}}}$, then for each $I$ and $n$, the natural map[^11] $$\tau_{\leq n} {\mathrm{oblv}}_{\operatorname{coFil}} {\eur{F}}|_{{\overset{\circ}{X}{}^{I}}} \to \tau_{\leq n} {\eur{F}}_m|_{{\overset{\circ}{X}{}^{I}}}$$ is an equivalence when $m\gg 0$. \[lem:direct\_sum\_and\_products\_are\_the\_same\_Ran\] The natural transformation $$\bigoplus \to \prod$$ between functors $$\operatorname{Shv}(\operatorname{Ran}X)^{{\mathrm{gr}}^{>0}, {\mathrm{decay}}} \to \operatorname{Shv}(\operatorname{Ran}X)$$ is an equivalence. Strategy -------- To prove that $\operatorname{Chev}\mathfrak{g}$ is factorizable when $\mathfrak{g}\in \operatorname{\operatorname{Lie}^\star}(X)$, [@francis_chiral_2011] uses the $\operatorname{addFil}$ trick (see §\[sec:appendix:addFil\_trick\]) to reduce to the case where $\mathfrak{g}$ is a trivial. In that case, $$\operatorname{Chev}\mathfrak{g} \simeq \operatorname{Sym}^{>0}\mathfrak{g},$$ and the result can be seen directly. In the case of $\operatorname{coChev}$, while the core strategy remains the same, it is more complicated to carry out since many commutative diagrams needed for the $\operatorname{addFil}$ trick to work (see ) don’t commute in general in this new setting. The co-connectivity constraints are what needed to make these diagrams commute and hence, to allow us to reduce to the trivial case. ### Now, suppose for the moment that we have the following commutative diagram, which is analogous to , except for the extra conditions $$\xymatrix{ \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1} \ar[d]_{\operatorname{addCoFil}} \ar[rr]^{\operatorname{coChev}} && \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2} \\ \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1, \operatorname{coFil}^{>0}, {\mathrm{stab}}} \ar[rr]^{\operatorname{coChev}_{\operatorname{coFil}}} \ar[d]_{\operatorname{ass-gr}} && \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2, \operatorname{coFil}^{>0}, {\mathrm{stab}}} \ar[d]_{\operatorname{ass-gr}} \ar[u]^{{\mathrm{oblv}}_{\operatorname{coFil}}} \\ \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1, {\mathrm{gr}}^{>0}, {\mathrm{decay}}} \ar[rr]^{\operatorname{coChev}_{{\mathrm{gr}}}} \ar[d]_{\prod} && \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2, {\mathrm{gr}}^{>0}, {\mathrm{decay}}} \ar[d]_{\prod} \\ \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1} \ar[rr]^{\operatorname{coChev}} && \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2} } {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:addCoFil_trick_main_diagram}$$ Suppose also that ${\mathrm{oblv}}_{\operatorname{coFil}}$ preserves factorizability, and that $\operatorname{ass-gr}$ and $\prod$ are conservative with respect to factorizability. Then by the same reasoning as in the $\operatorname{addFil}$ trick, to prove that $\operatorname{coChev}\mathfrak{g}$ is factorizable, it suffices to assume that $\mathfrak{g}$ has a trivial $\operatorname{coLie}$-structure. In that case, $$\operatorname{coChev}\mathfrak{g} \simeq \operatorname{Sym}^{>0}(\mathfrak{g}[-1]),$$ and as in the $\operatorname{Chev}$ case, we are done. ### The rest of this section will be devoted to the execution of the strategy outlined above. Well-definedness of functors ---------------------------- Before proving that the diagram commutes, we need to first make sense of it. A priori, the functors written in the diagram are not necessarily well-defined. For instance, we haven’t shown that all the four instances of $\operatorname{coChev}$ land in the correct target categories. Moreover, we also don’t know that ${\mathrm{oblv}}_{\operatorname{coFil}}, \operatorname{ass-gr}$, and $\prod$ preserve the algebra/co-algebra structures. We start with the following straight-forward observation which settles the latter question. \[lem:oblvCoFil\_ass-gr\_prod\_are\_monoidal\] For any $n$, the functors $$\begin{aligned} {\mathrm{oblv}}_{\operatorname{coFil}}&: \operatorname{Shv}(\operatorname{Ran}X)^{\geq n, \operatorname{coFil}^{>0}, {\mathrm{stab}}} \to \operatorname{Shv}(\operatorname{Ran}X)^{\geq n} \\ \operatorname{ass-gr}&: \operatorname{Shv}(\operatorname{Ran}X)^{\geq n, \operatorname{coFil}^{>0}} \to \operatorname{Shv}(\operatorname{Ran}X)^{\geq n, {\mathrm{gr}}^{>0}} \\ \prod \simeq \bigoplus&: \operatorname{Shv}(\operatorname{Ran}X)^{\geq n, {\mathrm{gr}}^{>0}, {\mathrm{decay}}} \to \operatorname{Shv}(\operatorname{Ran}X)^{\geq n}\end{aligned}$$ are symmetric monoidal with respect to the ${\otimes^\star}$-monoidal structure on $\operatorname{Ran}X$. ### We will now tackle the former question: namely, the various instances of the functor $\operatorname{coChev}$ appeared in  land in the correct target categories. ### The top and bottom $\operatorname{coChev}$ are the same, and it’s easy to see that they land in the correct category using the fact that the shriek-pullback functor is left exact and ${\eur{C}}^{\geq n}$ is preserved under limits for any stable infinity category ${\eur{C}}$ with a $t$-structure (since $i_{\geq n}$ commutes with limits, see §\[subsubsec:t-structure\_convention\]). ### By the same token, we know that the essential images of $\operatorname{coChev}_{\operatorname{coFil}}$ and $\operatorname{coChev}_{{\mathrm{gr}}}$ satisfy the co-connectivity assumption (i.e. live in (perverse) cohomological degree $\geq 1$). Thus, it remains to show that they also satisfy the ${\mathrm{stab}}$ and ${\mathrm{decay}}$ conditions respectively. First, observe that the assertion about $\operatorname{ass-gr}$ in Lemma \[lem:oblvCoFil\_ass-gr\_prod\_are\_monoidal\], combined with the fact that $\operatorname{ass-gr}$ commutes with limits, gives us a weakened version of the middle square of . \[cor:middle\_diagram\_cofiltrick\_commutes\_weakened\] We have the following commutative diagram $$\xymatrix{ \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1, \operatorname{coFil}^{>0}, {\mathrm{stab}}} \ar[rr]^<<<<<<<<<<{\operatorname{coChev}_{\operatorname{coFil}}} \ar[d]_{\operatorname{ass-gr}} && \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2, \operatorname{coFil}^{>0}} \ar[d]_{\operatorname{ass-gr}} \\ \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1, {\mathrm{gr}}^{>0}, {\mathrm{decay}}} \ar[rr]^<<<<<<<<<<{\operatorname{coChev}_{{\mathrm{gr}}}} && \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2, {\mathrm{gr}}^{>0}} }$$ Now, by Lemma \[lem:stab\_decay\_equivalence\_Ran\_case\], to show that $\operatorname{coChev}_{\operatorname{coFil}}$ and $\operatorname{coChev}_{{\mathrm{gr}}}$ satisfy the ${\mathrm{stab}}$ and ${\mathrm{decay}}$ conditions respectively, it suffices to show that $\operatorname{coChev}_{{\mathrm{gr}}}$ satisfies the ${\mathrm{decay}}$ condition. However, this is also a direct consequence of the fact that the shriek-pullback functor is left exact and ${\eur{C}}^{>n}$ is preserved under limits (for any stable infinity category ${\eur{C}}$ with a $t$-structure), and we are done. Commutative diagrams -------------------- We will now proceed to prove that the diagram  commutes. First note that we have just settled the commutativity of the middle diagram of  at the end of the previous subsection. ### The commutativity of the bottom diagram of  is clear if we know that $\prod$ is symmetric monoidal. However, by Lemma \[lem:direct\_sum\_and\_products\_are\_the\_same\_Ran\], we have $$\prod \simeq \bigoplus$$ and we know that $\bigoplus$ is symmetric monoidal. ### Finally, to show that the top diagram of  commutes, it suffices to show that the following diagram commutes $$\xymatrix{ \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1} \ar[rr]^{\operatorname{coChev}} && \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2} \\ \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1, \operatorname{coFil}^{>0}, {\mathrm{stab}}} \ar[u]^{{\mathrm{oblv}}_{\operatorname{coFil}}} \ar[rr]^{\operatorname{coChev}_{\operatorname{coFil}}} && \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2, \operatorname{coFil}^{>0}, {\mathrm{stab}}} \ar[u]^{{\mathrm{oblv}}_{\operatorname{coFil}}} } {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:coChev_and_oblv_coFil}$$ since the composition $$\operatorname{\operatorname{coLie}^\star}(X)^{\geq 1} \overset{\operatorname{addCoFil}}{\longrightarrow} \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1, \operatorname{coFil}^{>0}, {\mathrm{stab}}} \overset{{\mathrm{oblv}}_{\operatorname{coFil}}}{\longrightarrow} \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1}$$ is the identity functor (see also §\[sec:appendix:addFil\_oblvFil\_commutative\]). However, this is clear since the functor $${\mathrm{oblv}}_{\operatorname{coFil}}: \operatorname{Shv}(\operatorname{Ran}X)^{\geq n, \operatorname{coFil}^{>0}, {\mathrm{stab}}} \to \operatorname{Shv}(\operatorname{Ran}X)^{\geq n}$$ commutes with limit for any $n$, and moreover it is symmetric monoidal with respect to the ${\otimes^\star}$-monoidal structure on $\operatorname{Shv}(\operatorname{Ran}X)$ by Lemma \[lem:oblvCoFil\_ass-gr\_prod\_are\_monoidal\]. Relation to factorizability --------------------------- It is easy to see that $$\operatorname{ass-gr}: \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2, \operatorname{coFil}^{>0}, {\mathrm{stab}}} \to \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2, {\mathrm{gr}}^{>0}, {\mathrm{decay}}}$$ reflects factorizability. Moreover, as we’ve discussed above, we have the equivalence $$\prod \simeq \bigoplus$$ as functors $$\operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2, {\mathrm{gr}}^{>0}, {\mathrm{decay}}} \to \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2}.$$ But now it’s clear that $\prod$ reflects factorizability, since $\bigoplus$ does. Finally, since $${\mathrm{oblv}}_{\operatorname{coFil}}: \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2, \operatorname{coFil}^{>0}, {\mathrm{stab}}} \to \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)^{\geq 2}$$ is compatible with $\boxtimes$ (for the same reason that it is compatible with ${\otimes^\star})$, and moreover $(-)^!$ commutes with limits (being a right adjoint), we see easily that ${\mathrm{oblv}}_{\operatorname{coFil}}$ preserves factorizability. Thus, we conclude the proof of Theorem \[thm:factorizability\_coChev\]. Relation to $\operatorname{\operatorname{coLie}^!}(X)$ and $\operatorname{\operatorname{ComAlg}^!}(X)$ ------------------------------------------------------------------------------------------------------ In this subsection, we will discuss the various links between objects defined on $X$ such as $\operatorname{\operatorname{coLie}^!}(X)$ and $\operatorname{\operatorname{ComAlg}^!}(X)$ and objects defined on $\operatorname{Ran}X$ such as $\operatorname{\operatorname{coLie}^\star}(\operatorname{Ran}X)$, $\operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)$ and $\operatorname{\operatorname{Fact}^\star}(X)$. This subsection is not used anywhere in the paper. We include it here for the sake of completeness. ### Recall that on a scheme $X$, there are two symmetric monoidal structures, $\otimes$ and ${\overset{!}{\otimes}}$. Thus, we could talk about various algebra/co-algebra objects defined on it $$\operatorname{\operatorname{Lie}^\ast}(X), \operatorname{\operatorname{coLie}^!}(X), \operatorname{\operatorname{ComAlg}^!}(X),$$ where $\operatorname{\operatorname{Lie}^\ast}(X)$ (not to be confused with $\operatorname{\operatorname{Lie}^\star}(X)$) is the category of Lie-algebra objects in $\operatorname{Shv}(X)$ with respect to the $\otimes$-monoidal structure, and $\operatorname{\operatorname{coLie}^!}(X)$ (resp. $\operatorname{\operatorname{ComAlg}^!}(X)$) is the category of coLie-algebra (resp. commutative algebra) objects in $\operatorname{Shv}(X)$ with respect to the ${\overset{!}{\otimes}}$-monoidal structure. ### The following observations are straightforward, and are both based on the fact that the functors $$\operatorname{ins}_X^*: \operatorname{Shv}(\operatorname{Ran}X)^{{\otimes^\star}} \to \operatorname{Shv}(X)^{\otimes}$$ and $$\operatorname{ins}_X^!: \operatorname{Shv}(\operatorname{Ran}X)^{{\otimes^\star}} \to \operatorname{Shv}(X)^{{\overset{!}{\otimes}}}$$ are symmetric monoidal, where $$\operatorname{ins}_X: X \to \operatorname{Ran}X$$ is the diagonal embedding. We have a pair of adjoint functors $$\operatorname{ins}_{X}^* :\operatorname{\operatorname{Lie}^\star}(\operatorname{Ran}X) \rightleftarrows \operatorname{\operatorname{Lie}^\ast}(X): \operatorname{ins}_{X*}$$ which induces an equivalence of categories $$\operatorname{\operatorname{Lie}^\star}(X) \simeq \operatorname{\operatorname{Lie}^\ast}(X).$$ We have a pair of adjoint functors $$\operatorname{ins}_{X!}: \operatorname{\operatorname{coLie}^!}(X) \rightleftarrows \operatorname{\operatorname{coLie}^\star}(\operatorname{Ran}X): \operatorname{ins}_X^!$$ which induces an equivalence of categories $$\operatorname{\operatorname{coLie}^!}(X) \simeq \operatorname{\operatorname{coLie}^\star}(X).$$ ### We also have the following functor $$\operatorname{ins}_X^!: \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X) \to \operatorname{\operatorname{ComAlg}^!}(X)$$ which commutes with limits. Thus, we get a pair of adjoint functors $$\operatorname{ins}_{X?}: \operatorname{\operatorname{ComAlg}^!}(X) \rightleftarrows \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X):\operatorname{ins}_X^!. {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:adjunction_ComAlg_X_RanX}$$ We have the following result from [@gaitsgory_weils_2014 Thm. 5.6.4]. \[thm:GL\_equiv\_cats\_comalg\] The pair of adjoint functors in  induces an equivalence of categories $$\operatorname{\operatorname{ComAlg}^!}(X) \simeq \operatorname{\operatorname{Fact}^\star}(X).$$ ### The first link between $\operatorname{\operatorname{coLie}^!}(X), \operatorname{\operatorname{coLie}^\star}(X), \operatorname{\operatorname{ComAlg}^!}(X), \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)$ and $\operatorname{\operatorname{Fact}^\star}(X)$ is given by the following diagram $$\xymatrix{ \operatorname{\operatorname{coLie}^!}(X) \ar[d]_{\operatorname{coChev}} & \operatorname{\operatorname{coLie}^\star}(X) \ar[l]^\simeq_{\operatorname{ins}_X^!} \ar[d]_{\operatorname{coChev}} \\ \operatorname{\operatorname{ComAlg}^!}(X) & \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X) \ar[l]_{\operatorname{ins}_X^!} } {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:coChev_commutes_uppershriek}$$ whose commutativity is straightforward due to the fact that $\operatorname{ins}_X^!$ commutes with limits and that it’s monoidal. ### The second link, and also the more interesting one, is given by the following We have the following commutative diagram $$\xymatrix{ \operatorname{\operatorname{coLie}^!}(X)^{\geq 1} \ar[d]_{\operatorname{coChev}} \ar[r]_\simeq^{\operatorname{ins}_{X!}} & \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1} \ar[d]_{\operatorname{coChev}} \\ \operatorname{\operatorname{ComAlg}^!}(X) \ar[r]^{\operatorname{ins}_{X?}} & \operatorname{\operatorname{Fact}^\star}(X) }$$ For any $\mathfrak{g} \in \operatorname{\operatorname{coLie}^!}(X)$, we have a natural map $$\operatorname{ins}_{X?} \circ \operatorname{coChev}\to \operatorname{coChev}\circ \operatorname{ins}_{X!}.$$ of objects in $\operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)$. Now, we know from Theorem \[thm:GL\_equiv\_cats\_comalg\] that the LHS is factorizable. Moreover, when $\mathfrak{g}\in \operatorname{\operatorname{coLie}^!}(X)^{\geq 1}$, we know from Theorem \[thm:factorizability\_coChev\] that the RHS is also factorizable. Thus, to show that the map above is an equivalence when $\mathfrak{g} \in \operatorname{\operatorname{coLie}^!}(X)^{\geq 1}$, it suffices to show that they are equivalence on the diagonal. However, that is clear from  and we are done. Interactions between various functors on the Ran space ====================================================== In this section, we tie together the links between the various functors on the $\operatorname{Ran}$ spaces: $\operatorname{Chev}, \operatorname{coChev}, C^*_c(\operatorname{Ran}X, -)$, and $D_{\operatorname{Ran}}$, the functor of Verdier duality on the Ran space. $C^*_c(\operatorname{Ran}X, -)$ and $\operatorname{coChev}$ ----------------------------------------------------------- In this subsection, we will prove Theorem \[thm:intro:coChev\_and\_C\^\*\_c(Ran)\], which gives us a criterion for the commutativity of the functor $\operatorname{coChev}$ and the functor $C^*_c(\operatorname{Ran}X, -)$. Note that it has been proved in [@francis_chiral_2011] that $\operatorname{Chev}$ always commutes with $C^*_c(\operatorname{Ran}X, -)$. The main reason is that $C^*_c(\operatorname{Ran}X, -)$ is continuous and monoidal with respect to the ${\otimes^\star}$-monoidal structure on $\operatorname{Shv}(\operatorname{Ran}X)$ and the usual monoidal structure on ${\mathrm{Vect}}$. As before, our main difficulty comes from the fact that $\operatorname{coChev}$ is defined as a limit, and for that to behave well with respect to $C^*_c(\operatorname{Ran}X, -)$, we need to impose a certain co-connectivity assumption. ### Throughout this subsection, $X$ will be assumed to be a proper scheme of pure dimension $d$. \[thm:coChev\_and\_C\^\*\_c(Ran)\] For any $\mathfrak{g} \in \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1+d}$, the natural map $$C^*_c(\operatorname{Ran}X, \operatorname{coChev}\mathfrak{g}) \to \operatorname{coChev}(C^*_c(X, \mathfrak{g}))$$ is an equivalence.[^12] ### The proof of Theorem \[thm:coChev\_and\_C\^\*\_c(Ran)\] is essentially the dual of the proofs of Lemma \[lem:Prim\[-1\]\_is\_good\_Vec\_case\] and Lemma \[lem:Prim\[-1\]\_is\_good\_Ran\_case\]. There, we express the limit (i.e. $\operatorname{coBar}_{{\eur{P}}}$) as a sequential limit, and then establish a certain stability condition on the sequence we take the limit over. The main point is to show that for any $n$, $\operatorname{tr}_{\geq -n}$ of our limit is just $\operatorname{tr}_{\geq -n}$ of the terms when we go sufficiently far in the sequence. And at a finite step, commuting with a colimit is automatic. ### Our current situation is the dual of that. Namely, we will express $$C^*_c(\operatorname{Ran}X, -)$$ as the colimit of a sequence satisfying a certain stability condition, which allows us, after truncating on the right via $\operatorname{tr}_{\leq n}$ for each $n$, to commute it with the limit defining $\operatorname{coChev}$. ### We start with a general remark: in general, the limit (resp. colimit) of a diagram $${\eur{K}} \to {\eur{C}}$$ could be written as a sequential limit (resp. colimit) if we have a functor ${\eur{K}} \to \mathbb{Z}$. We can then use left (resp. right) Kan extension to produce a new diagram $$\mathbb{Z} \to {\eur{C}},$$ and the original limit (resp. colimit) could be written as a sequential limit (resp. colimit) of this new diagram. For example, in the case of limit over a co-simplicial object, the functor to $\mathbb{Z}$ is simply $$\begin{aligned} \Delta &\to \mathbb{Z}^{\geq 0} \\ [n] &\mapsto n.\end{aligned}$$ And in the case of the Ran space, the functor is $$\begin{aligned} {\mathrm{fSet}}_{{\mathrm{surj}}} &\to \mathbb{Z}^{>0} \\ I &\mapsto |I|.\end{aligned}$$ ### Truncated Ran space Now we can apply the remark above to the case of the $\operatorname{Ran}$ space. For any scheme $X$ and any positive integer $n$, we define $$\operatorname{Ran}^{\leq n} X \simeq \operatorname*{colim}_{\substack{I \in {\mathrm{fSet}}^{{\mathrm{surj}}} \\ |I| \leq n}} X^I.$$ Then $$\operatorname{Ran}X \simeq \operatorname*{colim}\operatorname{Ran}^{\leq n} X \simeq \operatorname*{colim}(X \to \operatorname{Ran}^{\leq 2} X \to \operatorname{Ran}^{\leq 3} X \to \cdots),$$ and hence, for any ${\eur{F}} \in \operatorname{Shv}(\operatorname{Ran}X)$, $$C^*_c(\operatorname{Ran}X, {\eur{F}}) \simeq \operatorname*{colim}_n C^*_c(\operatorname{Ran}^{\leq n}, {\eur{F}}|_{\operatorname{Ran}^{\leq n} X}).$$ The following observation, which gives the link among the cohomology groups $$C^*_c(\operatorname{Ran}^{\leq n}X, {\eur{F}}|_{\operatorname{Ran}^{\leq n}X})$$ for various $n$’s, comes from [@gaitsgory_atiyah-bott_2015 Cor. 9.1.4]. \[lem:difference\_in\_XcircI\] We have the following natural equivalence $$C^*({\overset{\circ}{X}{}^{I}}, {\eur{F}}|_{{\overset{\circ}{X}{}^{I}}})_{\Sigma_I} \simeq \operatorname{coFib}(C^*_c(\operatorname{Ran}^{\leq |I|-1} X, {\eur{F}}|_{\operatorname{Ran}^{\leq |I|-1} X}) \to C^*_c(\operatorname{Ran}^{\leq |I|} X, {\eur{F}}|_{\operatorname{Ran}^{\leq |I|} X})).$$ ### When $$\mathfrak{g} \in \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1+d},$$ using the $\operatorname{addCoFil}$ trick , we can also express $\operatorname{coChev}\mathfrak{g}$ as a sequential limit $$\operatorname{coChev}\mathfrak{g} \simeq {\mathrm{oblv}}_{\operatorname{coFil}}\operatorname{coChev}_{\operatorname{coFil}} \operatorname{addCoFil}\mathfrak{g} \simeq \lim_i (\operatorname{coChev}_{\operatorname{coFil}} \operatorname{addCoFil}\mathfrak{g})_i.$$ Where $$(\operatorname{coChev}_{\operatorname{coFil}} \operatorname{addCoFil}\mathfrak{g})_i$$ is the $i$-th step in the co-filtration. ### For brevity’s sake, we will denote $$\operatorname{coChev}^i \mathfrak{g} = (\operatorname{coChev}_{\operatorname{coFil}} \operatorname{addCoFil}\mathfrak{g})_i$$ and so we have $$\operatorname{coChev}\mathfrak{g} \simeq \lim_i \operatorname{coChev}^i \mathfrak{g}.$$ ### The advantage of using this co-filtration (instead of the usual one coming from the co-simplicial object defining $\operatorname{coChev}$) lies in the fact that both the supports and cohomological estimates of $\operatorname{coChev}^i$ vary nicely with respect to $i$. Namely, for any non-negative integer $i$, $$\operatorname{Supp}\operatorname{coChev}^i \mathfrak{g} \subset \operatorname{Ran}^{\leq i} X$$ and for all non-empty finite set $I$ such that $|I| \leq i$, $$(\operatorname{coChev}^i \mathfrak{g})|_{{\overset{\circ}{X}{}^{I}}}$$ lives in perverse cohomological degrees $\geq i(d+1)+1$. This gives us the following observations. For any $\mathfrak{g} \in \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1+d}$ and any non-empty finite set $I$, $$(\operatorname{coChev}\mathfrak{g})|_{{\overset{\circ}{X}{}^{I}}}$$ lives in cohomological degrees $\geq (1+d)|I|+1$. \[cor:stabilization\_coChev\_C\^\*\] For any $\mathfrak{g} \in \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1+d}$ and any non-empty finite set $I$, $$C^*({\overset{\circ}{X}{}^{I}}, (\operatorname{coChev}\mathfrak{g})|_{{\overset{\circ}{X}{}^{I}}})_{\Sigma_I}$$ lives in cohomological degrees $\geq |I| + 1$. \[lem:stabilization\_coChev\^i\_C\^\*\] For any $\mathfrak{g} \in \operatorname{\operatorname{coLie}^\star}(X)^{\geq 1+d}$, any positive integer $i$, and any non-empty finite set $I$, $$C^*({\overset{\circ}{X}{}^{I}}, (\operatorname{coChev}^i\mathfrak{g})|_{{\overset{\circ}{X}{}^{I}}})_{\Sigma_I}$$ lives in cohomological degrees $\geq \max(|I| + 1, i+1)$. With these observations, we are ready for the proof of Theorem \[thm:coChev\_and\_C\^\*\_c(Ran)\]. For each $i$, we know that $\operatorname{coChev}^i \mathfrak{g}$ is computed as a finite limit. Thus, we have the following natural equivalence $$C^*_c(\operatorname{Ran}X, \operatorname{coChev}^i \mathfrak{g}) \simeq \operatorname{coChev}^i(C^*_c(\operatorname{Ran}X, \mathfrak{g})).$$ Taking the limit over $i$ on both sides, we observe that it suffices to prove that $$\lim_i C^*_c(\operatorname{Ran}X, \operatorname{coChev}^i\mathfrak{g}) \simeq C^*_c(\operatorname{Ran}X, \lim_i \operatorname{coChev}^i \mathfrak{g}).$$ For that, it suffices to show that for each $m$, we have an equivalence $$\operatorname{tr}_{\leq m} \lim_i C^*_c(\operatorname{Ran}X, \operatorname{coChev}^i \mathfrak{g}) \simeq \operatorname{tr}_{\leq m} C^*_c(\operatorname{Ran}X, \lim_i \operatorname{coChev}^i \mathfrak{g}).$$ But now, for some $M \gg 0$, depending only on $m$, we have $$\begin{aligned} \operatorname{tr}_{\leq m} C^*_c(\operatorname{Ran}X, \lim_i \operatorname{coChev}^i \mathfrak{g}) &\simeq \operatorname{tr}_{\leq m} \operatorname*{colim}_n C^*_c(\operatorname{Ran}^{\leq n} X, \lim_i \operatorname{coChev}^i \mathfrak{g}|_{\operatorname{Ran}^{\leq n} X}) \\ &\simeq \operatorname{tr}_{\leq m} C^*_c(\operatorname{Ran}^{\leq M} X, \lim_i \operatorname{coChev}^i \mathfrak{g}|_{\operatorname{Ran}^{\leq M} X}) {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:used_stabilization_coChev_C^*} \\ &\simeq \operatorname{tr}_{\leq m} \lim_i C^*_c(\operatorname{Ran}^{\leq M} X, \operatorname{coChev}^i \mathfrak{g}|_{\operatorname{Ran}^{\leq M} X}) {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:commutes_C^*_c_truncated_Ran_lim_i_coChev_i} \\ &\simeq \lim_i \operatorname{tr}_{\leq m} C^*_c(\operatorname{Ran}^{\leq M} X, \operatorname{coChev}^i \mathfrak{g}|_{\operatorname{Ran}^{\leq M} X})\\ &\simeq \lim_i \operatorname{tr}_{\leq m} C^*_c(\operatorname{Ran}X, \operatorname{coChev}^i \mathfrak{g}) {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:used_stabilization_coChev^i_C^*} \\ &\simeq \operatorname{tr}_{\leq m} \lim_i C^*_c(\operatorname{Ran}X, \operatorname{coChev}^i \mathfrak{g}).\end{aligned}$$ Here, we used Lemma \[lem:difference\_in\_XcircI\] in both and . Moreover, and  use Corollary \[cor:stabilization\_coChev\_C\^\*\] and Lemma \[lem:stabilization\_coChev\^i\_C\^\*\] respectively. Finally, is due to the fact that $$C^*_c(\operatorname{Ran}^{\leq M} X, -)$$ commutes with limits. Indeed, this functor is computed as a finite colimit of functors of the form $$C^*_c(X^I, -).$$ But these functors commute with limits since $X^I$ are all complete due to our assumption on $X$, i.e. $$C^*_c(X^I, -) \simeq C^*(X^I, -).$$ Verdier duality --------------- Before studying the link between $\operatorname{Chev}$ and $\operatorname{coChev}$, we start with a quick recollection of Verdier duality on prestacks along with various useful properties. The main reference is [@gaitsgory_atiyah-bott_2015]. However, since we only use basic properties of $D_{\operatorname{Ran}}$, we’ll provide the complete proof in most cases. ### Let ${\eur{Y}}$ be a prestack such that the diagonal map $$\operatorname{diag}_{{\eur{Y}}}: {\eur{Y}} \to {\eur{Y}} \times {\eur{Y}}$$ is pseudo-proper. Given ${\eur{F}}, {\eur{G}} \in \operatorname{Shv}({\eur{Y}})$, by a pairing between them, we shall mean a map $${\eur{F}}\boxtimes {\eur{G}} \to \operatorname{diag}_{{\eur{Y}}!} \omega_{{\eur{Y}}}.$$ We define the Verdier dual, $D_{\eur{Y}} {\eur{G}}$, of ${\eur{G}}$ to represent the functor $${\eur{F}} \mapsto \operatorname{Hom}({\eur{F}} \boxtimes {\eur{G}}, \operatorname{diag}_{{\eur{Y}}!} \omega_{{\eur{Y}}}).$$ Namely, we have the following natural equivalence $$\operatorname{Hom}({\eur{F}}, D_{{\eur{Y}}} {\eur{G}}) \simeq \operatorname{Hom}({\eur{F}} \boxtimes {\eur{G}}, \operatorname{diag}_{{\eur{Y}}!} \omega_{{\eur{Y}}}).$$ The following lemma is immediate from the definition. \[lem:D\_turns\_colimits\_to\_limits\] Let ${\eur{F}} \in \operatorname{Shv}({\eur{Y}})$, such that $${\eur{F}} \simeq \operatorname*{colim}_{i \in {\eur{I}}} {\eur{F}}_i.$$ Then $$D_{{\eur{Y}}} {\eur{F}} \simeq \lim_{i \in {\eur{I}}^{{\mathrm{op}}}} D_{{\eur{Y}}} {\eur{F}}_i.$$ ### We will now study the link between Verdier duality and $\boxtimes$. Let ${\eur{Y}}_1$ and ${\eur{Y}}_2$ be finitary pseudo-schemes, and $${\eur{F}}_i \in \operatorname{Shv}({\eur{Y}}_i).$$ Then, we have a natural equivalence $$D_{{\eur{Y}}_1} {\eur{F}}_1 \boxtimes D_{{\eur{Y}}_2} {\eur{F}}_2 \simeq D_{{\eur{Y}}_1 \times {\eur{Y}}_2} ({\eur{F}}_1 \boxtimes {\eur{F}}_2).$$ First, note that the result holds when both ${\eur{Y}}_1$ and ${\eur{Y}}_2$ are schemes. For the general case of finitary pseudo-schemes, we write $${\eur{Y}}_1 \simeq \operatorname*{colim}_i Y_{1i}\qquad\text{and}\qquad {\eur{Y}}_2 \simeq \operatorname*{colim}_j Y_{2j}.$$ Then, $${\eur{F}}_1 \simeq \operatorname*{colim}_{i} \operatorname{ins}_{1i!} \operatorname{ins}_{1i}^! {\eur{F}}_1 \qquad\text{and}\qquad {\eur{F}}_2 \simeq \operatorname*{colim}_j \operatorname{ins}_{2j!} \operatorname{ins}_{2j}^! {\eur{F}}_2.$$ Thus, $$\begin{aligned} D_{{\eur{Y}}_1 \times {\eur{Y}}_2}({\eur{F}}_1 \boxtimes {\eur{F}}_2) &\simeq D_{{\eur{Y}}_1 \times {\eur{Y}}_2} \operatorname*{colim}_{i, j} (\operatorname{ins}_{1i} \times \operatorname{ins}_{2j})_! (\operatorname{ins}_{1i}\times \operatorname{ins}_{2j})^! ({\eur{F}}_1 \boxtimes {\eur{F}}_2) \\ &\simeq \lim_{i, j} (\operatorname{ins}_{1i} \times \operatorname{ins}_{2j})_! D_{Y_{1i} \times Y_{2j}} (\operatorname{ins}_{1i}^! {\eur{F}}_1 \boxtimes \operatorname{ins}_{2j}^! {\eur{F}}_2) {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:D_and_tensor_finitary_!pushforward_1} \\ &\simeq \lim_{i, j} (\operatorname{ins}_{1i} \times \operatorname{ins}_{2j})_! (D_{Y_{1i}} \operatorname{ins}_{1i}^! {\eur{F}}_1 \boxtimes D_{Y_{2j}} \operatorname{ins}_{2j}^! {\eur{F}}_2) \label{eq:D_and_tensor_scheme_case} {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\\ &\simeq (\lim_i \operatorname{ins}_{1i!} D_{Y_{1i}} \operatorname{ins}_{1i}^! {\eur{F}}_1) \boxtimes (\lim_j \operatorname{ins}_{2j!} D_{Y_{2j}} \operatorname{ins}_{1j}^! {\eur{F}}_2) \label{eq:D_and_tensor_finite_limit} {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\\ &\simeq (D_{{\eur{Y}}_1} \operatorname*{colim}_i \operatorname{ins}_{1i!} \operatorname{ins}_{1i}^! {\eur{F}}_1)\boxtimes (D_{{\eur{Y}}_2} \operatorname{ins}_{2j!} \operatorname{ins}_{2j}^! {\eur{F}}_2) {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:D_and_tensor_finitary_!pushforward_2} \\ &\simeq D_{{\eur{Y}}_1} {\eur{F}}_1 \boxtimes D_{{\eur{Y}}_2}{\eur{F}}_2.\end{aligned}$$ Here, 1. is due to the fact that the statement we are trying to prove holds for the case of schemes. 2. is due to the fact that the limits we are taking are all finite (due to the finitary assumption). 3. and are both due to Lemma \[lem:D\_turns\_colimits\_to\_limits\] and Proposition \[prop:D\_finitary\_!pushforward\] below. \[prop:D\_finitary\_!pushforward\] Let $f: {\eur{Y}}_1 \to {\eur{Y}}_2$ be a finitary pseudo-proper map between pseudo-schemes, each having a finitary diagonal. Then, the natural transformation $$f_!\circ D_{{\eur{Y}}_1} \to D_{{\eur{Y}}_2}\circ f_!$$ is an equivalence. See [@gaitsgory_atiyah-bott_2015 Cor. 7.5.6]. One direct corollary of this proposition is the fact that for any sheaf ${\eur{F}} \in \operatorname{Shv}(X)$, we have the following natural equivalence $$\delta_{X!} D_X {\eur{F}} \simeq D_{\operatorname{Ran}X} \delta_{X!} {\eur{F}}.$$ \[cor:D\_Ran\_commutes\_otimesstar\_finite\_support\] Let ${\eur{F}}_1, {\eur{F}}_2, \cdots, {\eur{F}}_k \in \operatorname{Shv}(\operatorname{Ran}X)$ with finite supports, i.e. there exists an $n$ such that all the ${\eur{F}}_i$’s come from $\operatorname{Shv}(\operatorname{Ran}^{\leq n} X)$. Then, we have the following natural equivalence $$D_{\operatorname{Ran}X} ({\eur{F}}_1 {\otimes^\star}{\eur{F}}_2 \otimes \cdots {\otimes^\star}{\eur{F}}_k) \simeq (D_{\operatorname{Ran}X} {\eur{F}}_1) {\otimes^\star}(D_{\operatorname{Ran}X} {\eur{F}}_2) {\otimes^\star}\cdots {\otimes^\star}(D_{\operatorname{Ran}X} {\eur{F}}_k).$$ $\operatorname{Chev}$, $\operatorname{coChev}$, and $D_{\operatorname{Ran}X}$ ----------------------------------------------------------------------------- We will now turn to Theorem \[thm:intro:Chev\_coChev\_and\_D\_Ran\], which provides the link between the two functors $\operatorname{Chev}$ and $\operatorname{coChev}$ via the functor of taking Verdier duality on the $\operatorname{Ran}$ space. \[thm:Chev\_coChev\_and\_D\_Ran\] Let $\mathfrak{g}\in \operatorname{\operatorname{Lie}^\star}(X)^{\leq -1}$. Then we have a natural equivalence $$\operatorname{coChev}(D_{X} \mathfrak{g}) \simeq D_{\operatorname{Ran}X} \operatorname{Chev}(\mathfrak{g}),$$ of objects in $\operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X)$, where $D_{\operatorname{Ran}X}$ is the functor of taking Verdier duality on $\operatorname{Ran}X$. Note that this is the only place we use Verdier duality on the Ran space. However, we essentially use it in a rather minimal way: not much besides the definition itself. We will employ ideas originated from the $\operatorname{addFil}$ and $\operatorname{addCoFil}$ tricks (see also §\[sec:appendix:addFil\_trick\]). First, observe that for any $\mathfrak{g} \in \operatorname{\operatorname{Lie}^\star}(X)$, we have a canonical equivalence $$\operatorname{addCoFil}D_{\operatorname{Ran}X} \mathfrak{g} \simeq D_{\operatorname{Ran}X} \operatorname{addFil}\mathfrak{g}.$$ We denote $$\operatorname{Chev}^i \mathfrak{g} \qquad\text{and}\qquad \operatorname{coChev}^i D_{\operatorname{Ran}X}\mathfrak{g}$$ to be the $i$-th piece in the filtration/co-filtration of $$\operatorname{Chev}(\operatorname{addFil}\mathfrak{g}) \qquad\text{and}\qquad \operatorname{coChev}(\operatorname{addCoFil}D_{\operatorname{Ran}X} \mathfrak{g})$$ respectively. From §\[sec:appendix:addFil\_trick\] and the top part of the commutative diagram , we have the following natural equivalences $$\begin{aligned} \operatorname{Chev}\mathfrak{g} &\simeq \operatorname*{colim}_i \operatorname{Chev}^i \mathfrak{g}, \\ \operatorname{coChev}(D_{\operatorname{Ran}X}\mathfrak{g}) &\simeq \lim_i \operatorname{coChev}^i (D_{\operatorname{Ran}X} \mathfrak{g}).\end{aligned}$$ At the same time, by Lemma \[lem:D\_turns\_colimits\_to\_limits\], we know that $$D_{\operatorname{Ran}X} \operatorname*{colim}_i \operatorname{Chev}^i \mathfrak{g} \simeq \lim_i D_{\operatorname{Ran}X} \operatorname{Chev}^i \mathfrak{g}.$$ Thus, it suffices to show that $$D_{\operatorname{Ran}X} \operatorname{Chev}^i \mathfrak{g} \simeq \operatorname{coChev}^i D_{\operatorname{Ran}X}\mathfrak{g}.$$ Now, it’s an immediate consequence of Corollary \[cor:D\_Ran\_commutes\_otimesstar\_finite\_support\]. Let $\mathfrak{g}\in \operatorname{\operatorname{Lie}^\star}(X)^{\leq -1}$. Then $$D_{\operatorname{Ran}X} \operatorname{Chev}(\mathfrak{g})$$ is a factorizable commutative algebra on $\operatorname{Ran}X$. This is a direct consequence of Theorem \[thm:Chev\_coChev\_and\_D\_Ran\] and Theorem \[thm:intro:factorizability\_coChev\]. $\operatorname{coChev}$ and open embeddings ------------------------------------------- We end the section with the following easy observation. \[prop:coChev\_open\_embedding\] Let $$j: X' \to X$$ be an open embedding of schemes, which induces an open embedding of prestacks $$j_{\operatorname{Ran}}: \operatorname{Ran}X' \to \operatorname{Ran}X.$$ Then for any $\mathfrak{g}' \in \operatorname{\operatorname{coLie}^\star}(X')$, we have the following natural equivalence $$(j_{\operatorname{Ran}})_* \operatorname{coChev}(\mathfrak{g}') \simeq \operatorname{coChev}(j_* \mathfrak{g}').$$ The result is a direct consequence of the fact that $f_*$, being a right adjoint, commutes with limits for any schematic morphism $f$ between prestacks. Moreover, if $f_i: X'_i \to X_i$ are open embeddings of schemes, and ${\eur{F}}_i \in \operatorname{Shv}(X'_i)$ for $i=1, 2$, then we have a natural equivalence $$(f_1 \times f_2)_*({\eur{F}}_1 \boxtimes {\eur{F}}_2) \simeq f_{1*}{\eur{F}}_1 \boxtimes f_{2*} {\eur{F}}_2.$$ An application to the Atiyah-Bott formula ========================================= We will now give an application of the results proved so far to the Atiyah-Bott formula. As mentioned in the introduction, these results allow us to simplify the second of the two main steps in the original proofs given in [@gaitsgory_weils_2014] and [@gaitsgory_atiyah-bott_2015]. In what follows, §\[subsec:statement\]–§\[subsec:pairing\] are intended to orient the readers with the existing results proved in [@gaitsgory_weils_2014] and [@gaitsgory_atiyah-bott_2015],[^13] whereas the purpose of the last part, §\[subsec:conclusion\], is to explain how the results we’ve proved so far fit in with the rest. The statement {#subsec:statement} ------------- From now on, $X$ is a smooth and complete curve over an algebraically closed field $k$, and $G$ a smooth, fiber-wise connected group-scheme over $X$, whose generic fiber is semi-simple simply connected. Due to [@gaitsgory_weils_2014 Lem. 7.1.1 and Prop A.3.11], we can (and from now on we will) assume that $G$ is semi-simple simply connected over an open dense subset $$j: X' \hookrightarrow X,$$ and moreover, the fibers of $G$ over any point in $X-X'$ are homologically trivial. We will also use $$j_{\operatorname{Ran}}: \operatorname{Ran}X' \to \operatorname{Ran}X$$ to denote the corresponding open embedding on the $\operatorname{Ran}$ space and $$\Gamma_{j_{\operatorname{Ran}}}: \operatorname{Ran}X' \to \operatorname{Ran}X'\times \operatorname{Ran}X$$ to denote its graph. ### Let $G_0$ be the split form of $G$. Then it is well-known that $$C^*(BG_0) \simeq \operatorname{Sym}M_0 {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:C^*(BG_0_is_free)}$$ is a free commutative algebra, for some $M_0\in {\mathrm{Vect}}$. In the case of $\ell$-adic sheaves in positive characteristic setting, this equivalence is compatible with the geometric Frobenius action, where $$M_0 \simeq \bigoplus_e \Lambda[-2e](-e),$$ and $e$’s are the exponents of $G_0$. The assignment $G_0 \mapsto M_0$ is functorial with respect to automorphisms of $G_0$, and hence, for a general $G$ (subject to the assumptions mentioned above), we get a local system $$M \in \operatorname{Shv}(X'),$$ whose $!$-fiber at each geometric point $x \in X$ is equivalent to $M_0$. Below is the statement of the Atiyah-Bott formula. \[thm:Atiyah-Bott\] Let $G, X$ as above. Then (a) We have an equivalence $$C^*(\operatorname{Bun}_G) \simeq \operatorname{Sym}(C^*(X', M)).$$ (b) When $k={{\overline{\mathbb{F}}}_q}$, and $X$ and $G$ are defined over ${\mathbb{F}_q}$, the above equivalence can be chosen to be compatible with the Frobenius actions. $BG$ and the sheaf ${\eur{B}}$ {#subsec:BG_and_sheaf_B} ------------------------------ ### The sheaf ${\eur{B}}$ that we will describe now encodes the reduced cohomology $BG$, the classifying stack of $G$. For each $I \in \operatorname{Ran}_X(S)$, let $D_I \subset S\times X$ be the corresponding Cartier divisor. Let $BG_I$ denote the Artin stack classifying $G$-bundles over $D_I$ and $f_I: BG_I \to S$ the forgetful map. Then, we define $${\widetilde{{\eur{B}}}}_{S, I} = D_S(\operatorname{Fib}(f_{I!}f_I^! \Lambda_S \to \Lambda_S)),$$ where $D_S$ is the functor of taking Verdier duality on $S$. These sheaves, assembled together, give rise to a sheaf (see also [@gaitsgory_weils_2014 Prop. 5.4.3]) $${\widetilde{{\eur{B}}}} \in \operatorname{Shv}(\operatorname{Ran}X).$$ ### Note that for any finite set of points $\{x_1, \dots, x_n\} \in (\operatorname{Ran}X)(k)$, the $!$-fiber of ${\widetilde{{\eur{B}}}}$ at this point is $$\operatorname{coFib}\left(\Lambda \to \bigotimes_{i=1}^n C^*(BG_{x_i})\right). {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:B'_!_fiber}$$ ### Using a variant of the diagonal map $$BG \to BG \times BG,$$ we can equip ${\widetilde{{\eur{B}}}}$ with the structure of an object in $$\operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X).$$ However, we see easily from  that ${\widetilde{{\eur{B}}}}$ is not factorizable. The functor $\operatorname{TakeOut}$ developed in [@gaitsgory_atiyah-bott_2015] allows us to remove all the extra components in it and construct out of it a new object ${\eur{B}} \in \operatorname{\operatorname{Fact}^\star}(X)$ with the correct $!$-fibers at a point $\{x_1, \dots, x_n\} \in (\operatorname{Ran}X)(k)$ $$\bigotimes_{i=1}^n C^*_{{\mathrm{red}}}(BG_{x_i}).$$ Moreover, ${\eur{B}}$ has the same cohomology along $\operatorname{Ran}X$ as the original sheaf ${\widetilde{{\eur{B}}}}$ (see also [@gaitsgory_atiyah-bott_2015 Cor. 5.3.5]) $$C^*_c(\operatorname{Ran}X, {\eur{B}}) \simeq C^*_c(\operatorname{Ran}X, {\widetilde{{\eur{B}}}}).$$ ### ${\eur{B}}$ and $\operatorname{Bun}_G$ For every $S \in {\mathrm{Sch}}$ and $I\in (\operatorname{Ran}X)(S)$, we have a map of prestacks over $S$ by restricting the bundle to the divisor $D_I$ $$S \times \operatorname{Bun}_G \to BG_I. {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:ev_Bun_G_BGI}$$ This induces a map $${\widetilde{{\eur{B}}}}_{S, I} \to \omega_S \otimes C^*_{\mathrm{red}}(\operatorname{Bun}_G)$$ and hence, also a map $${\widetilde{{\eur{B}}}} \to \omega_{\operatorname{Ran}X} \otimes C^*_{{\mathrm{red}}}(\operatorname{Bun}_G).$$ Applying the functor $C^*_c(\operatorname{Ran}X, -)$ and using the fact that $\operatorname{Ran}X$ is homologically contractible, we get a map $$C^*_c(\operatorname{Ran}X, {\eur{B}}) \simeq C^*_c(\operatorname{Ran}X, {\widetilde{{\eur{B}}}}) \to C^*_{{\mathrm{red}}}(\operatorname{Bun}_G). {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:atiyah-bott-canonical-map}$$ ### Using  and the assumption we have on $G$, i.e. homologically contractible fibers outside of $X$’, one gets an equivalence $${\eur{B}} \simeq (j_{\operatorname{Ran}X})_* {\eur{B}}' \simeq \operatorname{Sym}^{>0}(j_* M) {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:B_is_pushforward_of_B'}$$ where ${\eur{B}}'$ is the restriction of ${\eur{B}}$ to $\operatorname{Ran}X'$ and, the symmetric algebra is taken inside $\operatorname{Shv}(\operatorname{Ran}X)$ using the ${\otimes^\star}$-monoidal structure. ### Using the equivalence  and the fact that $C^*_c(\operatorname{Ran}X, -)$ commutes with $\operatorname{Sym}$,[^14] we get an explicit presentation of the LHS of  $$C^*_c(\operatorname{Ran}X, {\eur{B}}) \simeq \operatorname{Sym}^{>0} C^*_c(X, j_* M) \simeq \operatorname{Sym}^{>0} C^*(X', M), {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:explicit_C^*_c(Ran X, B)}$$ which appears in the statement of the Atiyah-Bott formula as stated in Theorem \[thm:Atiyah-Bott\]. ### Now, we are done if we could show that the map in  is an equivalence. Affine Grassmannian and the sheaf ${\eur{A}}$ {#subsec:affine_grassmannian_and_sheaf_A} --------------------------------------------- Unfortunately, one does not know how to directly prove that  is an equivalence. Instead, [@gaitsgory_weils_2014] proceeds with an equivalence of a dual nature, which we will now briefly recall. ### The main player in this step is the affine Grassmannian, or more precisely, a factorizable version of the affine Grassmannian. Let $G, X$ as above. The factorizable affine Grassmannian of $G$, denoted by $\operatorname{Gr}_{\operatorname{Ran}X'}$, is the prestack such that for each scheme $S$, $$\operatorname{Gr}_{\operatorname{Ran}X'}(S) = \{({\eur{P}}, I, \alpha)\},$$ where (i) ${\eur{P}}$ is a $G$-bundle over $S\times X$ (ii) $I$ is a non-empty finite subset of $X'(S)$ (iii) $\alpha$ is a trivialization of ${\eur{P}}$ on the complement of the graph of $I$. ### From the definition, we have the following natural morphism $$g: \operatorname{Gr}_{\operatorname{Ran}X'} \to \operatorname{Ran}X',$$ where we forget everything, except for the set $I$, and similarly another natural morphism $$u: \operatorname{Gr}_{\operatorname{Ran}X'} \to \operatorname{Bun}_G,$$ we we only remember the bundle ${\eur{P}}$. ### The map $g$ allows us to define $${\widetilde{{\eur{A}}}}' \simeq \operatorname{Fib}(g_!(\omega_{\operatorname{Gr}_{\operatorname{Ran}X'}})\to \omega_{\operatorname{Ran}X'}) \in \operatorname{Shv}(\operatorname{Ran}X'),$$ and the map $u$ induces a map at the cohomology level, namely $$C_*^{\mathrm{red}}(\operatorname{Gr}_{\operatorname{Ran}X'}) \to C_*^{\mathrm{red}}(\operatorname{Bun}_G). {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:non-abelian-poincare-uniformizing}$$ Together, we get the following map $$C^*_c(\operatorname{Ran}X', {\widetilde{{\eur{A}}}}') \to C_*^{\mathrm{red}}(\operatorname{Bun}_G). {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:non-abelian-poincare-sheaf-A'}$$ ### Note that since $$\operatorname{Gr}_{\operatorname{Ran}X'} \to \operatorname{Ran}X'$$ is pseudo-proper, ${\widetilde{{\eur{A}}}}'$ is easy to describe. Namely for any finite set of points $\{x_1, x_2, \dots, x_n\} \subset X(k)$, the $!$-fiber of ${\widetilde{{\eur{A}}}}'$ at this point is $$\operatorname{Fib}\left(\bigotimes_{i=1}^n C_*(\operatorname{Gr}_{G_{x_i}}) \to \Lambda\right). {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:!-fiber_A'}$$ ### ${\eur{A}}$ and $\operatorname{Bun}_G$ The equivalence of a dual nature that we alluded to earlier is given by the following important result (see [@gaitsgory_weils_2014 Thm. 3.2.13]). The map , and hence , is an equivalence. ### Using a variant of the diagonal map $$\operatorname{Gr}\to \operatorname{Gr}\times \operatorname{Gr},$$ one can equip ${\widetilde{{\eur{A}}}}'$ with the structure of an object in $${\operatorname{ComCoAlg}^\star}(\operatorname{Ran}X').$$ However, note that the sheaf ${\widetilde{{\eur{A}}}}'$ is not factorizable, since its $!$-fiber, as described in , is too big, i.e. it’s not equivalent to $$\bigotimes_{i=1}^n C_*^{\mathrm{red}}(\operatorname{Gr}_{G_{x_i}}). {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:correct-!-fiber-A}$$ Using a similar reasoning as in the case of ${\widetilde{{\eur{B}}}}$ and ${\eur{B}}$, we can construct an object ${\eur{A}}' \in \operatorname{\operatorname{coFact}^\star}(X')$ with the correct $!$-fiber as given in , and moreover, ${\eur{A}}'$ has the property that $$C^*_c(\operatorname{Ran}X', {\widetilde{{\eur{A}}}}') \simeq C^*_c(\operatorname{Ran}X', {\eur{A}}').$$ ### Altogether, we have the following \[prop:non-abelian-poincare\_sheaf\_A\] We have a natural equivalence $$C^*_c(\operatorname{Ran}X', {\eur{A}}') \simeq C^{\mathrm{red}}_*(\operatorname{Bun}_G).$$ Pairing {#subsec:pairing} ------- We will now describe how the equivalence given by Proposition \[prop:non-abelian-poincare\_sheaf\_A\] helps us show that is an equivalence. ### For any schemes $S, S'\in {\mathrm{Sch}}$ and any non-empty finite subsets $I \subset X(S)$ and $I'\subset X'(S')$, we have a natural map (which is just a more elaborate variant of ) $$\operatorname{Gr}_{I'} \times S \to \operatorname{Bun}_G \times S' \times S \to S'\times BG_{I},$$ which induces a map $${\eur{A}}' \boxtimes {\eur{B}} \to \omega_{\operatorname{Ran}X' \times \operatorname{Ran}X},$$ and hence, a pairing (using $\operatorname{TakeOut}$) $${\eur{A}}' \boxtimes {\eur{B}} \to \Gamma_{j_{\operatorname{Ran}}!}\omega_{\operatorname{Ran}X'}.$$ ### Restricting this map to $\operatorname{Ran}X' \times \operatorname{Ran}X'$ gives us the following map $${\eur{A}}' \boxtimes {\eur{B}}' \to (\delta_{\operatorname{Ran}X'})_! \omega_{\operatorname{Ran}X'},$$ and hence, using the definition of Verdier duality, a map $${\eur{B}}' \to D_{\operatorname{Ran}X'}{\eur{A}}' {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:B'_to_DA'}$$ between objects in $\operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X')$. ### It is proved, in fact twice (using different methods), in §17 and §18 of [@gaitsgory_atiyah-bott_2015], that the restriction of  to the diagonal $X'$ of $\operatorname{Ran}X'$ is an equivalence. Namely, we have $${\eur{B}}'|_{X'} \simeq (D_{\operatorname{Ran}X'} {\eur{A}}')|_{X'} {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:B'_eq_DA'_on_X'}.$$ The last steps {#subsec:conclusion} -------------- The results that we have just proved in this paper appear in two places in the concluding steps, which are given by Proposition \[prop:DA’\_factorizable\] and \[prop:local\_dual\_implies\_global\_dual\]. Together, they imply the Atiyah-Bott formula. \[prop:DA’\_factorizable\] $D_{\operatorname{Ran}X'} {\eur{A}}'$ is factorizable, i.e. $$D_{\operatorname{Ran}X'} {\eur{A}}' \in \operatorname{\operatorname{Fact}^\star}(X') \subset \operatorname{\operatorname{ComAlg}^\star}(\operatorname{Ran}X').$$ It is well-known that for a split semi-simple simply connected group $G_0$, $C_*^{\mathrm{red}}(\operatorname{Gr}_{G_0}, \Lambda)$ lives in cohomological degrees $\leq -2$. Using the fact that $$\operatorname{Gr}_{\operatorname{Ran}X'} \to \operatorname{Ran}X'$$ is pseudo-proper and that ${\eur{A}}'$ is factorizable, we see that for each non-empty finite set $I$, ${\eur{A}}'|_{{\overset{\circ}{X'}{}^{I}}}$ lives in (perverse) cohomological degrees $\leq -3|I|$. Now, by Theorem \[thm:Koszul\_duality\_connectivity\_on\_Ran\], we know that there exists an object $$\mathfrak{a}' \in \operatorname{\operatorname{Lie}^\star}(X')^{\leq c_{L}}$$ such that $${\eur{A}}' \simeq \operatorname{Chev}(\mathfrak{a'}).$$ Theorem \[thm:Chev\_coChev\_and\_D\_Ran\] then implies that $$D_{\operatorname{Ran}X'} \operatorname{Chev}(\mathfrak{a}') \simeq \operatorname{coChev}(D_{\operatorname{Ran}X'} \mathfrak{a}'),$$ which is known to be factorizable by Theorem \[thm:factorizability\_coChev\] \[cor:B’\_and\_DRan\_A’\] The map given in  is an equivalence, i.e. $${\eur{B}}' \simeq D_{\operatorname{Ran}X'} {\eur{A}}' {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:B'_eq_DA'},$$ and hence $${\eur{B}} \simeq (j_{\operatorname{Ran}})_*\operatorname{coChev}D_{X'}\mathfrak{a}' \simeq \operatorname{coChev}j_* D_{X'} \mathfrak{a}'.$$ The first statement is a direct consequence of the proposition above and the equivalence , where as the second statement is the result of Proposition \[prop:coChev\_open\_embedding\]. \[prop:local\_dual\_implies\_global\_dual\] We have the following equivalence induced by Proposition \[prop:DA’\_factorizable\] $$C^*_c(\operatorname{Ran}X, {\eur{B}}) \simeq C^*_c(\operatorname{Ran}X', {\eur{A}}')^\vee.$$ We have the following equivalences $$\begin{aligned} C^*_c(\operatorname{Ran}X, {\eur{B}}) &\simeq C^*_c(\operatorname{Ran}X, \operatorname{coChev}j_* D_{X'} \mathfrak{a}') {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:use_cor:B'_and_DRan_A'} \\ &\simeq \operatorname{coChev}C^*_c(X, j_*D_{X'} \mathfrak{a}') {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:use_thm:coChev_and_C^*_c(Ran)} \\ &\simeq \operatorname{coChev}C^*(X, D_{X'} \mathfrak{a}')\\ &\simeq \operatorname{coChev}(C^*_c(X, \mathfrak{a}')^{\vee}) \\ &\simeq (\operatorname{Chev}(C^*_c(X', \mathfrak{a}')))^\vee {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:use_thm:Chev_coChev_and_D_Ran} \\ &\simeq C^*_c(\operatorname{Ran}X', \operatorname{Chev}\mathfrak{a}')^\vee \\ &\simeq C^*_c(\operatorname{Ran}X', {\eur{A}}')^\vee.\end{aligned}$$ Here, , and  are due to Corollary \[cor:B’\_and\_DRan\_A’\], Theorem \[thm:coChev\_and\_C\^\*\_c(Ran)\] and Theorem \[thm:Chev\_coChev\_and\_D\_Ran\] (applied to a point) respectively. ### Finally, as a corollary, we have the Atiyah-Bott formula. Indeed, we have $$C_*^{\mathrm{red}}(\operatorname{Bun}_G)^\vee \simeq C^*_c(\operatorname{Ran}X', {\eur{A}}')^\vee \simeq C^*_c(\operatorname{Ran}X, {\eur{B}}) \simeq \operatorname{Sym}^{>0} C^*(X', M)$$ where the first, second and third equivalences are due to Proposition \[prop:non-abelian-poincare\_sheaf\_A\], Proposition \[prop:local\_dual\_implies\_global\_dual\], and  respectively. The $\operatorname{addFil}$ trick {#sec:appendix:addFil_trick} ================================= In this appendix, we will quickly recall, without proof, a useful construction taken from [@gaitsgory_study_???? §IV.2], which allows us to reduce many statements about ${\eur{P}}$-algebras to trivial ${\eur{P}}$-algebras, where ${\eur{P}}$ is an operad in ${\mathrm{Vect}}$. Throughout this subsection, all categories without any further description will be assumed to be presentable, symmetric monoidal stable infinity over a field $k$ of characteristic 0. Moreover, functors between these categories are assumed to be continuous. All such categories, along with continuous functors between them, form a category, which we will use $${{\mathrm{DGCat}}_{{\mathrm{pres}}, {\mathrm{cont}}}}^{\operatorname{SymMon}},$$ to denote, or for simplicity $${\mathrm{DGCat}}^{\operatorname{SymMon}}.$$ Notations --------- For a symmetric monoidal category ${\eur{C}}$, we denote the category of filtered objects in ${\eur{C}}$ $${\eur{C}}^{\mathrm{Fil}}= \operatorname{Fun}(\mathbb{Z}, {\eur{C}}),$$ the category of functors from $\mathbb{Z}$ to ${\eur{C}}$. Here, $\mathbb{Z}$ is a ordered set, viewed as a category. Similarly, we denote the category of graded objects $${\eur{C}}^{\mathrm{gr}}= \operatorname{Fun}(\mathbb{Z}^{{\mathrm{set}}}, {\eur{C}}),$$ where $\mathbb{Z}^{\mathrm{set}}$ is a the discrete category, whose underlying underlying objects are the integers.[^15] Functors -------- Now, we will recall several familiar functors between ${\eur{C}}$, ${\eur{C}}^{\mathrm{Fil}}$, and ${\eur{C}}^{\mathrm{gr}}$. ### Let $$V = \cdots \to V_{n-1} \to V_n \to V_{n+1} \to \cdots,$$ be an object in ${\eur{C}}^{\mathrm{Fil}}$. Then, we define $$\operatorname{ass-gr}: {\eur{C}}^{\mathrm{Fil}}\to {\eur{C}}^{\mathrm{gr}}$$ to be the functor of taking the associated graded object $$\operatorname{ass-gr}(V)_n = \operatorname{coFib}(V_{n-1} \to V_n),$$ and $${\mathrm{oblv}}_{{\mathrm{Fil}}}: {\eur{C}}^{\mathrm{Fil}}\to {\eur{C}}$$ to be the left Kan extension along $$\mathbb{Z} \to {\mathrm{pt}}.$$ Namely $${\mathrm{oblv}}_{\mathrm{Fil}}(V) = \operatorname*{colim}_{n\in \mathbb{Z}} V_n.$$ ### We also use $$({\mathrm{gr}}\to{\mathrm{Fil}}): {\eur{C}}^{\mathrm{gr}}\to {\eur{C}}^{\mathrm{Fil}}$$ and $$\bigoplus: {\eur{C}}^{\mathrm{gr}}\to {\eur{C}}$$ to denote the functor obtained by taking the left Kan extension along $$\mathbb{Z}^{\mathrm{set}}\to \mathbb{Z},$$ and $$\mathbb{Z}^{\mathrm{set}}\to {\mathrm{pt}}$$ respectively. ### Note that the categories ${\eur{C}}^{\mathrm{Fil}}$ and ${\eur{C}}^{\mathrm{gr}}$ are equipped with a natural symmetric monoidal structure coming from ${\eur{C}}$, and moreover, the functors $\operatorname{ass-gr}$, ${\mathrm{oblv}}_{\mathrm{Fil}}$, ${\mathrm{gr}}\to {\mathrm{Fil}}$, and $\bigoplus$ are naturally symmetric monoidal. ### Adding a filtration Let $$\operatorname{addFil}: {\eur{C}} \to {\eur{C}}^{\mathrm{Fil}}$$ be the functor defined as follows: for an object $V$ in ${\eur{C}}$, $$\operatorname{addFil}(V)_n = \begin{cases} V, &\text{when } n \geq 1, \\ 0, &\text{otherwise.} \end{cases}$$ It’s easy to see that $$\bigoplus \circ \operatorname{ass-gr}\circ \operatorname{addFil}\simeq {\mathrm{oblv}}_{\mathrm{Fil}}\circ \operatorname{addFil}\simeq {\mathrm{id}}_{\eur{C}}.$$ Interactions with algebras over an operad ----------------------------------------- Let ${\eur{P}}$ be an operad in ${\mathrm{Vect}}$. Then we have the following pair of functors $$\operatorname{addFil}: {\eur{P}}\operatorname{-alg}({\eur{C}}) \to {\eur{P}}\operatorname{-alg}({\eur{C}}^{{\mathrm{Fil}}^{>0}}) \qquad \text{and} \qquad {\mathrm{oblv}}_{\mathrm{Fil}}: {\eur{P}}\operatorname{-alg}({\eur{C}}^{{\mathrm{Fil}}^{>0}}) \to {\eur{P}}\operatorname{-alg}({\eur{C}}).$$ ### {#sec:appendix:addFil_oblvFil_commutative} Let $$F: {\mathrm{DGCat}}^{\operatorname{SymMon}} \to {{\eur{C}}\mathrm{at}}_\infty$$ be a functor, where ${{\eur{C}}\mathrm{at}}_\infty$ is the $\infty$-category of all $\infty$-categories. Suppose we have a continuous natural transformation $$\Phi: {\eur{P}}\operatorname{-alg}(-) \to F(-),$$ i.e. morphisms between two objects in $$\operatorname{Fun}({\mathrm{DGCat}}^{\operatorname{SymMon}}, {{\eur{C}}\mathrm{at}}_\infty).$$ Then from what we’ve discussed above, we have the following commutative diagram $$\xymatrix{ {\eur{P}}\operatorname{-alg}({\eur{C}}) \ar[r]^>>>>>>>\Phi & F({\eur{C}}) \\ {\eur{P}}\operatorname{-alg}({\eur{C}}^{\mathrm{Fil}}) \ar[u]^{{\mathrm{oblv}}_{\mathrm{Fil}}} \ar[r]^>>>>>\Phi & F({\eur{C}}^{\mathrm{Fil}}) \ar[u]^{{\mathrm{oblv}}_{\mathrm{Fil}}} }$$ which, combined with the fact that $${\mathrm{oblv}}_{\mathrm{Fil}}\circ \operatorname{addFil}\simeq {\mathrm{id}}_{\eur{C}},$$ implies that the following diagram also commutes $$\xymatrix{ {\eur{P}}\operatorname{-alg}({\eur{C}}) \ar[d]_{\operatorname{addFil}} \ar[r]^>>>>>>>\Phi & F({\eur{C}}) \\ {\eur{P}}\operatorname{-alg}({\eur{C}}^{\mathrm{Fil}}) \ar[r]^>>>>>\Phi & F({\eur{C}}^{\mathrm{Fil}}) \ar[u]^{{\mathrm{oblv}}_{\mathrm{Fil}}} }$$ ### {#section-86} Further composing the diagram above with $\operatorname{ass-gr}$ and $\bigoplus$ gives us the following commutative diagram $$\xymatrix{ {\eur{P}}\operatorname{-alg}({\eur{C}}) \ar[r]^>>>>>>>\Phi \ar[d]_{\operatorname{addFil}} & F({\eur{C}}) \\ {\eur{P}}\operatorname{-alg}({\eur{C}}^{{\mathrm{Fil}}^{>0}}) \ar[d]_{\operatorname{ass-gr}} \ar[r]^>>>>>{\Phi^{{\mathrm{Fil}}}} & F({\eur{C}}^{{\mathrm{Fil}}^{>0}}) \ar[u] ^{{\mathrm{oblv}}_{\mathrm{Fil}}}\ar[d]_{\operatorname{ass-gr}} \\ {\eur{P}}\operatorname{-alg}({\eur{C}}^{{\mathrm{gr}}^{>0}}) \ar[d]_{\bigoplus} \ar[r]^>>>>>{\Phi^{{\mathrm{gr}}}} & F({\eur{C}}^{{\mathrm{gr}}^{>0}}) \ar[d]_{\bigoplus} \\ {\eur{P}}\operatorname{-alg}({\eur{C}}) \ar[r]^>>>>>>>\Phi & F({\eur{C}}) } {\addtocounter{subsubsection}{1}\tag{\thesubsubsection}}\label{eq:addFil_trick_main_diagram}$$ We will refer to this as the *fundamental commutative diagram of the $\operatorname{addFil}$ trick*. ### {#section-87} Now, suppose there are two natural transformations $$\Phi_1, \Phi_2: {\eur{P}}\operatorname{-alg}(-) \to F(-)$$ equipped with a morphism between them $$\alpha: \Phi_1 \to \Phi_2.$$ Or more concretely, we have a compatible family of morphisms in $F({\eur{C}})$ $$\Phi_1(c) \to \Phi_2(c)$$ parametrized by pairs $({\eur{C}}, c)$ where $c\in {\eur{C}}$ and ${\eur{C}} \in {\mathrm{DGCat}}^{\operatorname{SymMon}}$, and we want to prove that $\alpha$ is an equivalence. ### {#section-88} The top square of the commutative diagram above implies that it suffices to show that $$\Phi_1 ^{\mathrm{Fil}}\circ \operatorname{addFil}\to \Phi_2^{\mathrm{Fil}}\circ \operatorname{addFil}$$ is an equivalence. But since $\operatorname{ass-gr}$ and $\bigoplus$ are conservative, it suffices to show that $$\bigoplus \circ \operatorname{ass-gr}\circ \Phi_1^{{\mathrm{Fil}}} \circ \operatorname{addFil}\to \bigoplus \circ \operatorname{ass-gr}\circ \Phi_2^{{\mathrm{Fil}}}\circ \operatorname{addFil}$$ is an equivalence, which, due to the commutativity of the diagrams, is equivalent to $$\Phi_1 \circ \bigoplus \circ \operatorname{ass-gr}\circ \operatorname{addFil}\to \Phi_2 \circ \bigoplus \circ \operatorname{ass-gr}\circ \operatorname{addFil}$$ being an equivalence. ### {#section-89} The crucial observation of [@gaitsgory_study_???? Prop. IV.2.1.4.6] is the following The functor $$\bigoplus \circ \operatorname{ass-gr}\circ\operatorname{addFil}: {\eur{P}}\operatorname{-alg}({\eur{C}}) \to F({\eur{C}})$$ is canonically equivalent to $\operatorname{triv}_{{\eur{P}}} \circ {\mathrm{oblv}}_{{\eur{P}}}$, i.e. $${\eur{P}}\operatorname{-alg}({\eur{C}}) \overset{{\mathrm{oblv}}_{{\eur{P}}}}{\longrightarrow} {\eur{C}} \overset{\operatorname{triv}_{{\eur{P}}}}{\longrightarrow} {\eur{P}}\operatorname{-alg}({\eur{C}}).$$ ### {#section-90} This implies that it suffices to prove that $$\Phi_1(c) \to \Phi_2(c)$$ is an equivalence only for the case where $c$ is a trivial algebra. A general principle ------------------- More generally, suppose we want to prove a property of $\Phi(c)$ for some $c\in {\eur{P}}\operatorname{-alg}({\eur{C}})$. Moreover, suppose this property is preserved under under ${\mathrm{oblv}}_{{\mathrm{Fil}}}$, and is conservative under $\bigoplus$ and $\operatorname{ass-gr}$. Then, it suffices to prove the case where $c$ has a trivial algebra structure. Co-filtration and $\operatorname{addCoFil}$ {#sec:appendix:cofiltration_addCoFil} =========================================== In this appendix, we will collect various notions that are dual to the one in §\[sec:appendix:addFil\_trick\]. These are used in the body of the paper to give a proof of the $\operatorname{addCoFil}$ trick in a special case. Notations --------- For a symmetric monoidal category ${\eur{C}}$, we denote the category of co-filtered objects in ${\eur{C}}$ $${\eur{C}}^{\operatorname{coFil}} = \operatorname{Fun}(\mathbb{Z}^{\mathrm{op}}, {\eur{C}}).$$ Functors -------- As in the case of filtration, there are several familiar functors between ${\eur{C}}, {\eur{C}}^{\operatorname{coFil}}$, and ${\eur{C}}^{{\mathrm{gr}}}$. ### {#section-91} Let $$V = \cdots \to V_{n+1} \to V_n \to V_{n-1} \to \cdots,$$ be an object in ${\eur{C}}^{\operatorname{coFil}}$. Then we define $$\operatorname{ass-gr}: {\eur{C}}^{\operatorname{coFil}} \to {\eur{C}}^{{\mathrm{gr}}}$$ to be the functor of taking the associated graded object $$\operatorname{ass-gr}(V)_n = \operatorname{Fib}(V_n \to V_{n-1}),$$ and $${\mathrm{oblv}}_{\operatorname{coFil}}: {\eur{C}}^{\operatorname{coFil}} \to {\eur{C}}$$ to be the right Kan extension along $$\mathbb{Z}^{\mathrm{op}}\to {\mathrm{pt}}.$$ Namely $${\mathrm{oblv}}_{\operatorname{coFil}}(V) = \lim_{n\in \mathbb{Z}^{{\mathrm{op}}}} V_n.$$ ### {#section-92} Note that the category ${\eur{C}}^{\operatorname{coFil}}$ naturally inherits the monoidal structure coming from ${\eur{C}}$. Moreover, the functor $\operatorname{ass-gr}$ is monoidal. ### {#section-93} We also use $$\prod: {\eur{C}}^{{\mathrm{gr}}} \to {\eur{C}}$$ to denote the right Kan extension along $$\mathbb{Z}^{{\mathrm{set}}} \to {\mathrm{pt}}.$$ Namely $$\prod((V_n)_{n\in \mathbb{Z}}) = \prod_{n\in \mathbb{Z}} V_n.$$ ### Adding a co-filtration We will use $$\operatorname{addCoFil}: {\eur{C}} \to {\eur{C}}^{\operatorname{coFil}}$$ to denote a functor defined as follows: for an object $V$ in ${\eur{C}}$, $$\operatorname{addCoFil}(V)_n = \begin{cases} V, & \text{when } n\geq 1, \\ 0, & \text{otherwise}. \end{cases}$$ [^1]: This corresponds to the case of constant group $G\times X$ over $X$. For simplicity’s sake, we will restrict ourselves to this case in the introduction. [^2]: We are eliding a minor, but technical, point about unital vs. non-unital here. [^3]: [@gaitsgory_atiyah-bott_2015] doesn’t reprove non-abelian [Poincaré]{} duality. [^4]: In the terminology of [@gaitsgory_atiyah-bott_2015], it’s not finitary. [^5]: \[fn:indnilp\_vs\_ord\] Strictly speaking, we are using the category $\operatorname{ComCoAlg}^{\operatorname{ind-nilp}}$ of ind-nilpotent commutative co-algebras. However, we will see easily that, subject to an appropriate connectivity assumption of sheaves on $\operatorname{Ran}X$, this category coincides with the category $\operatorname{ComCoAlg}$. [^6]: See also footnote \[fn:indnilp\_vs\_ord\]. [^7]: It also admits a right adjoint. However, we do not make use of it in this paper. [^8]: The interested reader could read more about this in [@francis_chiral_2011], since we do not need this fact in the current work. [^9]: This fact applies to the category of algebras over any operad in general. [^10]: Note that for a general operad ${\eur{O}}$, only the first row of  makes sense. [^11]: Note that ${\mathrm{oblv}}_{\operatorname{coFil}}$ commutes with restricting to ${\overset{\circ}{X}{}^{I}}$ for any non-empty, finite set $I$. Thus, the LHS is free of ambiguity. [^12]: Since $\operatorname{Supp}\mathfrak{g} \subset X \subset \operatorname{Ran}X$, $$C^*_c(\operatorname{Ran}X, \mathfrak{g}) \simeq C^*_c(X, \mathfrak{g}).$$ [^13]: Namely, all the results stated in these subsections could be found in [@gaitsgory_weils_2014] or [@gaitsgory_atiyah-bott_2015]. The readers should be warned that we provide a mere overview of the development given in these two papers, with many technical points elided. [^14]: Note that this is a special case of the fact that $C^*_c(\operatorname{Ran}X, -)$ commutes with $\operatorname{Chev}$. And in fact, both are due to the same reasons: that $C^*_c(\operatorname{Ran}X, -)$ is continuous and that it’s symmetric monoidal. [^15]: In [@gaitsgory_study_????], it’s called $\mathbb{Z}^{\mathrm{Spc}}$.
{ "pile_set_name": "ArXiv" }
--- abstract: 'For a Dynkin quiver $Q$ of type $\mathsf{ADE}$ and a sum $\beta$ of simple roots, we construct a bimodule over the quantum loop algebra and the quiver Hecke algebra of the corresponding type via equivariant $K$-theory, imitating Ginzburg-Reshetikhin-Vasserots geometric realization of the quantum affine Schur-Weyl duality. Our construction is based on Hernandez-Leclercs isomorphism between a certain graded quiver variety and the space of representations of the quiver $Q$ of dimension vector $\beta$. We identify the functor induced from our bimodule with Kang-Kashiwara-Kim’s generalized quantum affine Schur-Weyl duality functor. As a by-product, we verify a conjecture by Kang-Kashiwara-Kim on the simpleness of some poles of normalized $R$-matrices for any quiver $Q$ of type $\mathsf{ADE}$.' author: - 'Ryo Fujita[^1]' title: 'Geometric realization of Dynkin quiver type quantum affine Schur-Weyl duality' --- Introduction ============ For a fixed pair $(n, d)$ of positive integers, we have the following two fundamental objects: the complex simple Lie algebra $\mathfrak{sl}_{n+1}$ of type $\mathsf{A}_{n}$ and the symmetric group ${\mathfrak{S}}_{d}$ of degree $d$. The natural $(\mathfrak{sl}_{n+1}, {\mathfrak{S}}_{d})$-bimodule structure on the tensor power $({\mathbb{C}}^{n+1})^{\otimes d}$ produces a close relationship between their representation theories. This phenomenon is known as the classical Schur-Weyl duality and has many interesting variants. The quantum affine Schur-Weyl duality is a variant involving their quantum affinizations: the quantum loop algebra $U_{q}(L\mathfrak{sl}_{n+1})$ of $\mathfrak{sl}_{n+1}$ and the affine Hecke algebra ${H^{\mathrm{af}}}_d(q)$ of $GL_d$. Both algebras are defined over ${\Bbbk}:= {\mathbb{Q}}(q)$. Here we equip the tensor power ${\mathbb{V}}^{\otimes d}$ of the natural representation ${\mathbb{V}}:= {\Bbbk}^{n+1}[z^{\pm 1}]$ of $U_{q}(L \mathfrak{sl}_{n+1})$ with a commuting right action of ${H^{\mathrm{af}}}_{d}(q^{2})$ using the $R$-matrices. Chari-Pressley [@CP96] proved that the induced functor $${H^{\mathrm{af}}}_{d}(q^{2}) {\text{-}\mathrm{mod}}\to U_{q}(L \mathfrak{sl}_{n+1}) {\text{-}\mathrm{mod}}; \quad M \mapsto {\mathbb{V}}^{\otimes d} \otimes_{{H^{\mathrm{af}}}_{d}(q^{2})} M$$ gives an equivalence between suitable subcategories of finite-dimensional modules. The quantum affine Schur-Weyl duality has a beautiful geometric realization due to Ginzburg-Reshetikhin-Vasserot [@GRV94]. Here we recall their construction briefly. Let $\mu_{d} : {\mathcal{F}}_{d} \to \mathcal{N}_{d}$ be the Springer resolution of the nilpotent cone $\mathcal{N}_{d}$ of $\mathfrak{gl}_{d}({\mathbb{C}})$, where ${\mathcal{F}}_{d}$ is the cotangent bundle of the full flag variety of $GL_{d}({\mathbb{C}})$. The morphism $\mu_{d}$ is equivariant with respect to a natural action of the group ${\mathbb{G}}_{d} := GL_{d}({\mathbb{C}}) \times {\mathbb{C}}^{\times}$, where ${\mathbb{C}}^{\times}$ acts as the scalar multiplication on the cone $\mathcal{N}_{d}$. Due to Ginzburg and Kazhdan-Lusztig, the affine Hecke algebra ${H^{\mathrm{af}}}_{d}(q^{2})$ is isomorphic to the convolution algebra $K^{{\mathbb{G}}_{d}}({\mathcal{Z}}_{d}) \otimes_{A} {\Bbbk}$ of the equivariant $K$-group of the Steinberg variety $ {\mathcal{Z}}_{d} := {\mathcal{F}}_{d} \times_{\mathcal{N}_{d}} {\mathcal{F}}_{d}$, where $A = R({\mathbb{C}}^{\times}) = {\mathbb{Z}}[v^{\pm 1}]$ is the representation ring of ${\mathbb{C}}^{\times}$ and $- \otimes_{A} {\Bbbk}$ means the specialization $v \mapsto q$. On the other hand, we consider another Steinberg type variety $Z_{d} := {\mathfrak{M}}_{d} \times_{\mathcal{N}_{d}} {\mathfrak{M}}_{d}$. Here ${\mathfrak{M}}_d$ is the cotangent bundle of the variety of partial flags in ${\mathbb{C}}^{d}$ of length $\le n+1$. Due to Ginzburg-Vasserot, there is an algebra homomorphism $\Phi : U_{q}(L \mathfrak{sl}_{n+1}) \to K^{{\mathbb{G}}_{d}}(Z_{d}) \otimes_{A}{\Bbbk}$ with some good properties. Based on these facts, Ginzburg-Reshetikhin-Vasserot considered the intermediary fiber product ${\mathfrak{M}}_{d} \times_{\mathcal{N}_{d}} {\mathcal{F}}_{d}$ and identified its equivariant $K$-group with the bimodule ${\mathbb{V}}^{\otimes d}$. More precisely, they established an isomorphism ${\mathbb{V}}^{\otimes d} \cong K^{{\mathbb{G}}_{d}}({\mathfrak{M}}_{d} \times_{\mathcal{N}_{d}} {\mathcal{F}}_{d}) \otimes_{A} {\Bbbk}$ making the following diagram commute: $$\xy \xymatrix{ U_{q}(L\mathfrak{sl}_{n+1}) \ar[r] \ar[d]^{\Phi} & {\mathop{\mathrm{End}}\nolimits}\left( {\mathbb{V}}^{\otimes d} \right) \ar[d]^-{\cong} & {H^{\mathrm{af}}}_{d}(q^{2}) \ar[l] \ar[d]^-{\cong} \\ K^{{\mathbb{G}}_{d}}(Z_{d})\otimes_{A} {\Bbbk}\ar[r] & {\mathop{\mathrm{End}}\nolimits}\left( K^{{\mathbb{G}}_{d}}({\mathfrak{M}}_{d} \times_{\mathcal{N}_{d}} {\mathcal{F}}_{d}) \otimes_{A} {\Bbbk}\right) & K^{{\mathbb{G}}_{d}}({\mathcal{Z}}_{d})\otimes_{A}{\Bbbk}, \ar[l] } \endxy$$ where horizontal arrows denote the bimodule structures. Recently, in a series of papers [@KKK18; @KKK15; @KKKO15; @KKKO16], Kang, Kashiwara, Kim and Oh established some interesting generalized versions of the quantum affine Schur-Weyl duality. One of them (treated in [@KKK15] by Kang-Kashiwara-Kim) is associated with a pair $(Q, \beta)$ of a Dynkin quiver $Q$ of type $\mathsf{ADE}$ and a sum $\beta=\sum_{i} d_{i} \alpha_{i}$ of simple roots, which plays a similar role as the pair $(n, d)$ in the previous paragraphs. One player is the quantum loop algebra $U_{q}(L{\mathfrak{g}})$ of the complex simple Lie algebra ${\mathfrak{g}}$ whose Dynkin diagram is the underlying graph of $Q$. The other is the quiver Hecke (KLR) algebra $H_{Q}(\beta)$ associated with $(Q, \beta)$, or actually its completion ${\widehat{H}}_{Q}(\beta)$ along the grading. The quiver Hecke algebra $H_{Q}(\beta)$ is regarded as a generalization of the affine Hecke algebra ${H^{\mathrm{af}}}_{d}(q)$ from the viewpoint of the categorification of the quantum group. Inspired by the work of Hernandez-Leclerc [@HL15], Kang-Kashiwara-Kim [@KKK15] constructed on a left $U_{q}(L{\mathfrak{g}})$-module ${\widehat{V}}^{\otimes \beta}$ which is a direct sum of some tensor products of affinized fundamental modules a commuting right action of the algebra ${\widehat{H}}_{Q}(\beta)$ by using the normalized $R$-matrices. However, to make the ${\widehat{H}}_{Q}(\beta)$-action well-defined, we need to assume the simpleness of some poles of the normalized $R$-matrices. This assumption was verified for type $\mathsf{AD}$ in [@KKK15] by an explicit computation of the denominators of the normalized $R$-matrices. On the other hand, for type $\mathsf{E}$, this remains a conjecture. Under this well-definedness assumption, Kang-Kashiwara-Kim [@KKK15] also proved that the induced functor $${\widehat{H}}_{Q}(\beta) {\text{-}\mathrm{mod}_{\mathrm{fd}}}\to U_{q}(L {\mathfrak{g}}){\text{-}\mathrm{mod}_{\mathrm{fd}}}; \quad M \mapsto {\widehat{V}}^{\otimes \beta} \otimes_{{\widehat{H}}_{Q}(\beta)} M$$ is exact, factors through the $\beta$-block ${\mathcal{C}}_{Q, \beta}$ of a monoidal full subcategory ${\mathcal{C}}_{Q}$ of $U_{q}(L{\mathfrak{g}}){\text{-}\mathrm{mod}_{\mathrm{fd}}}$ introduced by Hernandez-Leclerc [@HL15] and gives a bijection between the simple isomorphism classes. More recently, the author [@Fujita17] proved that it actually gives an equivalence ${\widehat{H}}_{Q}(\beta) {\text{-}\mathrm{mod}_{\mathrm{fd}}}\simeq {\mathcal{C}}_{Q, \beta}$ by using the notion of affine highest weight category. Note that here we forget the gradings by working with the completion ${\widehat{H}}_{Q}(\beta)$. In the present paper, we give a geometric realization of the bimodule ${\widehat{V}}^{\otimes \beta}$ imitating Ginzburg-Reshetikhin-Vasserots realization. In our case, the nilpotent cone $\mathcal{N}_{d}$ is replaced by the space $E_{\beta}$ of representations of the quiver $Q$ over ${\mathbb{C}}$ of dimension vector $\beta$. The group $G_{\beta}:= \prod_{i} GL_{d_{i}}({\mathbb{C}})$ naturally acts on $E_{\beta}$. Instead of the Springer resolution ${\mathcal{F}}_{d} \to \mathcal{N}_d$, we consider a proper morphism ${\mathcal{F}}_{\beta} \to E_{\beta}$ from a “quiver flag variety” ${\mathcal{F}}_{\beta}$ introduced by Lusztig in order to construct the canonical basis of the quantum group. Varagnolo-Vasserot [@VV11] proved that the quiver Hecke algebra $H_{Q}(\beta)$ is isomorphic to the convolution algebra of the equivariant Borel-Moore homology $H^{G_{\beta}}_{*}({\mathcal{Z}}_{\beta}, {\Bbbk})$, where ${\mathcal{Z}}_{\beta} := {\mathcal{F}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta}$. After completion, it is isomorphic to the completed equivariant $K$-group ${\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}}$. On the $U_{q}(L{\mathfrak{g}})$-side, we consider a canonical $G_{\beta}$-equivariant proper morphism ${\mathfrak{M}^{\bullet}}_{\beta} \to {\mathfrak{M}^{\bullet}}_{0, \beta}$ between certain graded quiver varieties. By Nakajima [@Nakajima01], we have an algebra homomorphism $ \widehat{\Phi}_{\beta} : U_{q}(L {\mathfrak{g}}) \to {\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}}, $ where ${Z^{\bullet}}_{\beta} := {\mathfrak{M}^{\bullet}}_{\beta} \times_{{\mathfrak{M}^{\bullet}}_{0, \beta}} {\mathfrak{M}^{\bullet}}_{\beta}$. The key of our construction is a $G_{\beta}$-equivariant isomorphism ${\mathfrak{M}^{\bullet}}_{0, \beta} \cong E_{\beta}$ due to Hernandez-Leclerc [@HL15], which was originally established in order to give a geometric interpretation to their isomorphism between the Grothendieck ring $K({\mathcal{C}}_{Q})$ and the coordinate ring of the maximal unipotent subgroup (see Remark \[Rem:Grotisom\]). This allows us to form the intermediary fiber product ${\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta}$. There is an isomorphism $${\widehat{V}}^{\otimes \beta} \cong {\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta})_{{\Bbbk}}$$ such that the following diagram commutes (up to a twist): $$\xy \xymatrix{ U_{q}(L{\mathfrak{g}}) \ar[r] \ar[d]^-{\widehat{\Phi}_{\beta}} & {\mathop{\mathrm{End}}\nolimits}\left({\widehat{V}}^{\otimes \beta}\right) \ar[d]^-{\cong} & {\widehat{H}}_{Q}(\beta) \ar[l] \ar[d]^-{\cong} \\ {\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}} \ar[r] & {\mathop{\mathrm{End}}\nolimits}\left( {\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta})_{{\Bbbk}} \right) & {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}}, \ar[l] } \endxy$$ where the horizontal arrows denote the bimodule structures. Actually, our geometric construction of the ${\widehat{H}}_{Q}( \beta )$-action is independent of that of [@KKK15], which shares the same characterization of the actions. Therefore, their comparison yields a uniform proof of: \[Thm:KKKconj\] Kang-Kashiwara-Kim’s conjecture [@KKK15 Conjecture 4.3.2] on the simpleness of some specific poles of normalized $R$-matrices for tensor products of fundamental modules is true for any quiver $Q$ of type $\mathsf{ADE}$. Besides, a discussion involving geometric extension algebras yields another proof of the equivalence ${\widehat{H}}_{Q}(\beta) {\text{-}\mathrm{mod}_{\mathrm{fd}}}\simeq {\mathcal{C}}_{Q, \beta}$ given by the bimodule ${\widehat{V}}^{\otimes \beta}$ without using affine highest weight categories (Theorem \[Thm:equiv\]). We would also remark that we do not use any results of [@KKK18], [@KKK15] for our proofs. The present paper is organized as follows. In Section \[Sec:HL\], we recall the definition of graded quiver varieties ${\mathfrak{M}^{\bullet}}_{\beta}$ and ${\mathfrak{M}^{\bullet}}_{0, \beta}$, and Hernandez-Leclercs isomorphism ${\mathfrak{M}^{\bullet}}_{0, \beta} \cong E_{\beta}$. In Section \[Sec:Conv\], we study the convolution algebra ${\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}}$ (resp. ${\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}}$) and recall its relation to the quiver Hecke algebra $H_{Q}(\beta)$ (resp. the quantum loop algebra $U_{q}(L{\mathfrak{g}})$). In the final section \[Sec:SW\], we study the structure of the bimodule ${\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta})_{{\Bbbk}}$. While the author was writing this paper, there appeared a preprint by Oh-Scrimshaw [@OS18] in arXiv which also proves Theorem \[Thm:KKKconj\] by a different approach. They compute the denominators of the normalized $R$-matrices for type $\mathsf{E}$ explicitly with a computer. #### Acknowledgment. The author thanks Syu Kato for helpful discussions and comments. He also thanks Ryosuke Kodera for suggesting him to study the geometric realization of the generalized quantum affine Schur-Weyl duality. #### Convention. An algebra $A$ is associative and unital. We denote by $A^{\mathrm{op}}$ (resp. $A^{\times}$) the opposite algebra (resp. the set of invertible elements) of $A$ and by $A \text{-}\mathrm{mod}$ the category of left $A$-modules. Working over a base field $\mathbb{F}$, the symbol $\otimes$ (resp. ${\mathop{\mathrm{Hom}}\nolimits}$) stands for $\otimes_{\mathbb{F}}$ (resp. ${\mathop{\mathrm{Hom}}\nolimits}_{\mathbb{F}}$) if there is no other clarification. If $A$ is an $\mathbb{F}$-algebra, we denote by $A {\text{-}\mathrm{mod}_{\mathrm{fd}}}$ the category of finite-dimensional left $A$-modules. Hernandez-Leclercs isomorphism {#Sec:HL} ============================== Notation {#Ssec:Notation} -------- Throughout this paper, we fix a finite-dimensional complex simple Lie algebra ${\mathfrak{g}}$ of type $\mathsf{ADE}$ and a quiver $Q=(I, \Omega)$ whose underlying graph is the Dynkin diagram of ${\mathfrak{g}}$, where $I=\{ 1,2, \ldots,n \}$ (resp. $\Omega$) is the set of vertices (resp. arrows). For an arrow $h \in \Omega$, let $h^{\prime}, h^{{\prime \prime}} \in I$ denote its origin and goal respectively. We write $i \sim j$ (resp. $i \to j$) if there is an arrow $h \in \Omega$ such that $\{i,j\}=\{h^{\prime}, h^{{\prime \prime}}\}$ (resp. $(i, j) = (h^{\prime}, h^{{\prime \prime}})$). Then the Cartan matrix $(a_{ij})_{i,j \in I}$ of ${\mathfrak{g}}$ is given by $$a_{ij} = \begin{cases} 2 & \text{if $i=j$}; \\ -1 & \text{if $i \sim j$}; \\ 0 & \text{otherwise}. \end{cases}$$ Let ${\mathsf{P}}^{\vee} = \bigoplus_{i \in I} {\mathbb{Z}}h_{i}$ be the coroot lattice of ${\mathfrak{g}}$. The fundamental weights $\{ \varpi_{i}\}_{i \in I}$ form a basis of the weight lattice ${\mathsf{P}}= {\mathop{\mathrm{Hom}}\nolimits}_{{\mathbb{Z}}}({\mathsf{P}}^{\vee}, {\mathbb{Z}})$ which is dual to $\{h_{i} \}_{i \in I}$. Let $\alpha_{i} = \sum_{j \in I} a_{ij} \varpi_{j}$ be the $i$-th simple root and ${\mathsf{Q}}= \bigoplus_{i \in I} {\mathbb{Z}}\alpha_{i} \subset {\mathsf{P}}$ be the root lattice. We put ${\mathsf{P}}^{+} = \sum_{i \in I}{\mathbb{Z}}_{\ge 0} \varpi_{i}$ and ${\mathsf{Q}}^{+} = \sum_{i \in I} {\mathbb{Z}}_{\ge 0} \alpha_{i}$. The Weyl group is the finite group $W$ of linear transformations on ${\mathsf{P}}$ generated by the set $\{r_{i}\}_{ i \in I}$ of simple reflections, which are given by $r_{i}(\lambda) := \lambda - \lambda(h_{i}) \alpha_{i}$ for $\lambda \in {\mathsf{P}}$. The set ${\mathsf{R}}^{+}$ of positive roots is defined by ${\mathsf{R}}^{+} = (W\{ \alpha_{i}\}_{ i \in I}) \cap {\mathsf{Q}}^{+}$. Representations of Dynkin quiver {#Ssec:Rep_quiver} -------------------------------- For an element $\beta \in {\mathsf{Q}}^{+}$, we fix an $I$-graded ${\mathbb{C}}$-vector space $D=\bigoplus_{i \in I} D_{i}$ such that ${\underline{\dim}\,}D := \sum_{i \in I} (\dim D_{i}) \alpha_{i} = \beta$. Let us consider the space $$E_{\beta} := \bigoplus_{h \in \Omega} {\mathop{\mathrm{Hom}}\nolimits}(D_{h^{\prime}}, D_{h^{{\prime \prime}}})$$ of representations of the quiver $Q$ of dimension vector $\beta$. On the space $E_{\beta}$, the group $G_{\beta} := \prod_{i \in I}\mathop{GL}(D_{i})$ acts by conjugation. The set $G_{\beta} \backslash E_{\beta}$ of $G_{\beta}$-orbits is naturally in bijection with the set of isomorphism classes of representations of the quiver $Q$ of dimension vector $\beta$. By Gabriels theorem, for each $\alpha \in {\mathsf{R}}^{+}$ there exists an indecomposable representation $M_\alpha$ such that ${\underline{\dim}\,}M_\alpha = \alpha$ uniquely up to isomorphism. The correspondence $\alpha \mapsto M_{\alpha}$ gives a bijection between the set ${\mathsf{R}}^{+}$ of positive roots and the set of isomorphism classes of indecomposable objects of the category ${\mathop{\mathrm{Rep}}\nolimits}Q$ of finite-dimensional representations of $Q$. Hence, the set $${\mathrm{KP}}(\beta) := \left\{ (m_{\alpha}) \in ({\mathbb{Z}}_{\ge 0})^{\mathsf{R}^{+}} \; \middle| \; \textstyle \sum_{\alpha \in {\mathsf{R}}^{+}} m_{\alpha} \alpha = \beta \right\}$$ of Kostant partitions of $\beta$ labels the set of $G_{\beta}$-orbits: $G_{\beta} \backslash E_{\beta} = \{ \mathbb{O}_{{\mathbf{m}}} \}_{{\mathbf{m}}\in {\mathrm{KP}}(\beta)}$, where for each ${\mathbf{m}}= (m_\alpha) \in {\mathrm{KP}}(\beta)$, the $G_{\beta}$-orbit $\mathbb{O}_{{\mathbf{m}}}$ corresponds to the isomorphism class of the representation $\bigoplus_{\alpha \in {\mathsf{R}}^{+}} M_{\alpha}^{\oplus m_{\alpha}}$. We have the natural $G_{\beta}$-orbit stratification $$\label{Eq:decE} E_{\beta} = \bigsqcup_{{\mathbf{m}}\in {\mathrm{KP}}(\beta)} \mathbb{O}_{{\mathbf{m}}}.$$ Repetition quiver {#Ssec:Repet_quiver} ----------------- We fix a height function $\xi : I \to {\mathbb{Z}}; i \mapsto \xi_{i}$ of the quiver $Q$ i.e. it satisfies $\xi_{i} = \xi_{j} +1$ if $i \to j$. Such a function $\xi$ is determined up to adding a constant. Choose a total ordering $I=\{i_{1}, i_{2}, \ldots , i_{n} \}$ such that $\xi_{i_1} \ge \xi_{i_2} \ge \cdots \ge \xi_{i_n}$ and define the corresponding Coxeter element $c := r_{i_1}r_{i_2} \cdots r_{i_n} \in W$. The repetition quiver $\widehat{Q} = (\widehat{I}, \widehat{\Omega})$ is an infinite quiver defined by $$\begin{aligned} \widehat{I} &:= \{(i,p) \in I \times {\mathbb{Z}}\mid p-\xi_{i} \in 2{\mathbb{Z}}\}, \\ \widehat{\Omega} &:= \{(i,p) \to (j, p+1) \mid (i, p), (j, p+1) \in \widehat{I}, \; i \sim j \}.\end{aligned}$$ It is well-known (cf. [@Happel88]) that there exists an isomorphism $\phi$ from the Auslander-Reiten quiver of the derived category $D^{b}({\mathop{\mathrm{Rep}}\nolimits}Q)$ to the repetition quiver $\widehat{Q}$, which depends on the choice of $\xi$ and is described as follows. Since each indecomposable object of $D^{b}({\mathop{\mathrm{Rep}}\nolimits}Q)$ is isomorphic to a unique stalk complex $M_{\alpha}[k]$ for some $(\alpha, k) \in {\mathsf{R}}^{+} \times {\mathbb{Z}}$, we have a bijection between the sets of vertices $${\mathsf{R}}^{+} \times {\mathbb{Z}}\ni (\alpha, k) \mapsto \phi(M_\alpha [k]) \in \widehat{I},$$ which we denote by the same symbol $\phi$. This bijection $\phi : {\mathsf{R}}^{+} \times {\mathbb{Z}}\to \widehat{I}$ is determined inductively as follows: - For each $i \in I$, we put $\gamma_{i} := \sum_{j} \alpha_{j}$ where $j$ runs all the vertices $j \in I$ such that there is a path in $Q$ from $j$ to $i$. Then $M_{\gamma_{i}}$ is an injective hull of the $1$-dimensional representation $M_{\alpha_{i}}$. We define $\phi (\gamma_{i}, 0) := (i, \xi_{i})$; - Inductively, if $\phi(\alpha, k) = (i, p)$ for $(\alpha, k) \in \mathsf{R}^{+} \times {\mathbb{Z}}$, then we define as: $$\begin{aligned} \phi(c^{\pm1}(\alpha), k) &:= (i, p\mp 2) & \text{if $c^{\pm 1}(\alpha) \in \mathsf{R}^{+}$}, \\ \phi(-c^{\pm1}(\alpha), k \mp 1) &:= (i, p\mp 2) & \text{if $c^{\pm 1}(\alpha) \in - \mathsf{R}^{+}$}. \end{aligned}$$ In the followings, we only consider the restriction of the bijection $\phi$ on ${\mathsf{R}}^{+} = {\mathsf{R}}^{+} \times \{0\}$, which we denote by the same symbol, i.e. we define $\phi(\alpha) := \phi(\alpha, 0)$ for $\alpha \in {\mathsf{R}}^{+}$. Graded quiver varieties {#Ssec:Quiver_var} ----------------------- In this subsection, we recall the definition of the graded quiver varieties. A basic reference is [@Nakajima01]. For elements $\nu = \sum_{i \in I} n_{i} \alpha_{i} \in {\mathsf{Q}}^{+}$ and $\lambda = \sum_{i \in I} l_{i} \varpi_{i} \in {\mathsf{P}}^{+}$, we fix $I$-graded ${\mathbb{C}}$-vector spaces $ V = \bigoplus_{i \in I} V_{i}, W = \bigoplus_{i \in I} W_{i} $ such that $\dim V_{i} = n_{i}, \dim W_{i} = l_{i}$ for each $i \in I$. We form the following space of linear maps: $$\mathbf{M}(\nu, \lambda) := \left( \bigoplus_{i \sim j} {\mathop{\mathrm{Hom}}\nolimits}(V_{j}, V_{i}) \right) \oplus \left( \bigoplus_{i \in I} {\mathop{\mathrm{Hom}}\nolimits}(W_{i}, V_{i}) \right) \oplus \left( \bigoplus_{i \in I} {\mathop{\mathrm{Hom}}\nolimits}(V_{i}, W_{i}) \right)$$ On the ${\mathbb{C}}$-vector space $\mathbf{M}(\nu, \lambda)$, the groups $G(\nu) := \prod_{i \in I} \mathop{GL}(V_{i})$, $G(\lambda) := \prod_{i \in I} \mathop{GL}(W_{i})$ act by conjugation and the $1$-dimensional torus ${\mathbb{C}}^{\times}$ acts by the scalar multiplication. We write an element of $\mathbf{M}(\nu, \lambda)$ as a triple $(B,a,b)$ of linear maps $B = \bigoplus B_{ij}$, $a = \bigoplus a_{i}$ and $b = \bigoplus b_{i}$. Let $\mu = \bigoplus_{i \in I} \mu_{i} : \mathbf{M}(\nu, \lambda) \to \bigoplus_{i \in I} \mathfrak{gl}(V_{i})$ be the map given by $$\mu_{i}(B, a, b) = a_{i} b_{i} +\sum_{j \sim i} \varepsilon(i,j) B_{ij}B_{ji},$$ where $\varepsilon(i,j) := 1$ (resp. $-1$) if $j \to i$ (resp. $i \to j$). A point $(B, a, b) \in \mu^{-1}(0)$ is said to be stable if there exists no non-zero $I$-graded subspace $V^{\prime} \subset V$ such that $B(V^{\prime}) \subset V^{\prime}$ and $V^{\prime} \subset {\mathop{\text{Ker}}}b$. Let $\mu^{-1}(0)^{\mathrm{st}}$ be the set of stable points, on which $G(\nu)$ acts freely. Then we consider a set-theoretic quotient $${\mathfrak{M}}(\nu, \lambda) := \mu^{-1}(0)^{\mathrm{st}} / G(\nu).$$ It is known that this quotient has a structure of a non-singular quasi-projective variety which is isomorphic to a quotient in the geometric invariant theory. We also consider the affine algebro-geometric quotient $${\mathfrak{M}}_{0}(\nu, \lambda) := \mu^{-1}(0)/\!/G(\nu) = \mathrm{Spec}\, {\mathbb{C}}[\mu^{-1}(0)]^{G(\nu)},$$ together with a canonical projective morphism ${\mathfrak{M}}(\nu, \lambda) \to {\mathfrak{M}}_{0}(\nu, \lambda)$. These quotients ${\mathfrak{M}}(\nu, \lambda)$, ${\mathfrak{M}}_{0}(\nu, \lambda)$ naturally inherit the actions of the group ${\mathbb{G}}(\lambda) := G(\lambda) \times {\mathbb{C}}^{\times}$, which makes the canonical projective morphism into a ${\mathbb{G}}(\lambda)$-equivariant morphism. For $\nu, \nu^{\prime} \in {\mathsf{Q}}^{+}$ such that $\nu^{\prime} - \nu \in {\mathsf{Q}}^{+}$, there is a natural closed embedding ${\mathfrak{M}}_{0} (\nu , \lambda) \hookrightarrow {\mathfrak{M}}_{0}(\nu^{\prime},\lambda). $ With respect to these embeddings, the family $\{ {\mathfrak{M}}_{0}(\nu, \lambda)\}_{\nu \in {\mathsf{Q}}^{+}}$ forms an inductive system, which stabilizes at some $\nu \in {\mathsf{Q}}^{+}$. We consider the union (inductive limit) and obtain the following combined ${\mathbb{G}}(\lambda)$-equivariant morphism: $$\pi : {\mathfrak{M}}(\lambda) := \bigsqcup_{\nu} {\mathfrak{M}}(\nu, \lambda) \to {\mathfrak{M}}_{0}(\lambda) := \bigcup_{\nu} {\mathfrak{M}}_{0}(\nu, \lambda).$$ We denote the fiber $\pi^{-1}(0)$ of the origin $0 \in {\mathfrak{M}}_{0}(\lambda)$ by ${\mathfrak{L}}(\lambda) = \bigsqcup_{\nu \in {\mathsf{Q}}^{+}} {\mathfrak{L}}(\nu, \lambda)$. Note that ${\mathfrak{M}}(0,\lambda) = {\mathfrak{L}}(0, \lambda)$ consists of a single point. Next we consider a free abelian monoid ${\mathscr{P}}^{+} = \mathbb{Z}_{\ge 0} \widehat{I}$ with the free generating set $\widehat{I}$. Define a homomorphism $\mathsf{cl} : {\mathscr{P}}^{+} \to {\mathsf{P}}^{+}$ by ${\mathsf{cl}}(i, p)=\varpi_{i}$. For an element ${{\boldsymbol \lambda}}= \sum l_{i, p} (i,p) \in {\mathscr{P}}^{+}$ with ${\mathsf{cl}}({{\boldsymbol \lambda}}) = \lambda$, we fix a decomposition $W_{i} = \bigoplus_{p} W_{i, p}$ such that $\dim W_{i, p} = l_{i,p}$ for each $(i,p) \in \widehat{I}$. Define a group homomorphism $f_{i} : {\mathbb{C}}^{\times} \to \prod_{p} GL(W_{i, p}) \subset GL(W_{i})$ by $f_{i} (t) |_{W_{i, p}} := t^{p} \cdot \mathrm{id}_{W_{i, p}}$ for $t \in {\mathbb{C}}^{\times}$. We put $T({{\boldsymbol \lambda}}) := (\prod_{i \in I}f_{i} \times \mathrm{id}) ({\mathbb{C}}^{\times}) \subset {\mathbb{G}}(\lambda)$ and consider the subvarieties of $T({{\boldsymbol \lambda}})$-fixed points: $$\pi^{\bullet} := \pi^{T({{\boldsymbol \lambda}})} : {\mathfrak{M}^{\bullet}}({{\boldsymbol \lambda}}) := {\mathfrak{M}}(\lambda)^{T({{\boldsymbol \lambda}})} \to {\mathfrak{M}^{\bullet}}_{0}({{\boldsymbol \lambda}}) := {\mathfrak{M}}_{0}(\lambda)^{T({{\boldsymbol \lambda}})}.$$ We refer these varieties as the graded quiver varieties. We put ${\mathfrak{L}^{\bullet}}({{\boldsymbol \lambda}}) := {\mathfrak{L}}(\lambda)^{T({{\boldsymbol \lambda}})} = (\pi^{\bullet})^{-1}(0)$. The centralizer of $T({{\boldsymbol \lambda}})$ inside ${\mathbb{G}}(\lambda)$ is $${\mathbb{G}}({{\boldsymbol \lambda}}) \equiv G({{\boldsymbol \lambda}}) \times {\mathbb{C}}^{\times} := \prod _{(i,p) \in \widehat{I}} \mathop{GL}(W_{i, p}) \times {\mathbb{C}}^{\times} \subset {\mathbb{G}}(\lambda),$$ which naturally acts on the varieties ${\mathfrak{M}^{\bullet}}({{\boldsymbol \lambda}})$, ${\mathfrak{M}^{\bullet}}_{0}({{\boldsymbol \lambda}})$, ${\mathfrak{L}^{\bullet}}({{\boldsymbol \lambda}})$. The morphism $\pi^{\bullet}$ is ${\mathbb{G}}({{\boldsymbol \lambda}})$-equivariant. Hernandez-Leclercs isomorphism {#Ssec:HL} ------------------------------ Let ${\mathscr{P}}_{0}^{+} \subset {\mathscr{P}}^{+}$ be the submonoid generated by the subset $\phi({\mathsf{R}}^{+}) \subset \widehat{I}$. For an element $\beta := \sum_{i \in I} d_{i} \alpha_{i} \in {\mathsf{Q}}^{+}$, we define ${{\boldsymbol \lambda}}_{\beta} := \sum_{i \in I} d_{i} \phi(\alpha_{i}) \in {\mathscr{P}}^{+}_{0}$. In this case, we write $\pi_{\beta} : {\mathfrak{M}^{\bullet}}_{\beta} \to {\mathfrak{M}^{\bullet}}_{0, \beta}$ instead of $\pi^{\bullet} : {\mathfrak{M}^{\bullet}}({{\boldsymbol \lambda}}_{\beta}) \to {\mathfrak{M}^{\bullet}}_{0}({{\boldsymbol \lambda}}_{\beta})$ for simplicity. For each $i \in I$, we identify the vector space $D_{i}$ in Subsection \[Ssec:Rep\_quiver\] with the vector space $W_{\phi(\alpha_{i})}$ in Subsection \[Ssec:Quiver\_var\]. This induces the identification $G_{\beta} = G({{\boldsymbol \lambda}}_{\beta})$. We write ${\mathbb{G}}_{\beta}$, $T_{\beta}$ instead of ${\mathbb{G}}({{\boldsymbol \lambda}}_\beta)$, $T({{\boldsymbol \lambda}}_{\beta})$ respectively. By the inclusion $G_{\beta} = G_{\beta} \times \{ 1\} \subset G_{\beta} \times {\mathbb{C}}^{\times} = {\mathbb{G}}_{\beta}$, the group $G_{\beta}$ is regarded as a subgroup of the group ${\mathbb{G}}_{\beta}$. Then the multiplication map $G_{\beta} \times T_{\beta} \to {\mathbb{G}}_{\beta}$ gives an isomorphism of algebraic groups $$\label{Eq:group} G_{\beta} \times T_{\beta} \cong {\mathbb{G}}_{\beta}.$$ We equip an action of the group ${\mathbb{G}}_{\beta}$ on the space $E_{\beta}$ via the projection ${\mathbb{G}}_{\beta} \cong G_{\beta} \times T_{\beta} \twoheadrightarrow G_{\beta}.$ \[Thm:HL\] There exists a ${\mathbb{G}}_{\beta}$-equivariant isomorphism of varieties $${\mathfrak{M}}_{0, \beta}^{\bullet} \xrightarrow{\cong} E_{\beta}.$$ Henceforth, we identify the graded quiver variety ${\mathfrak{M}^{\bullet}}_{0, \beta}$ with the space $E_{\beta}$ under the isomorphism in Theorem \[Thm:HL\]. We recall some properties of fibers of the ${\mathbb{G}}_{\beta}$-equivariant morphism $\pi_{\beta}:{\mathfrak{M}^{\bullet}}_{\beta} \to E_{\beta}$. By the injective map $${\mathrm{KP}}(\beta) \ni (m_{\alpha}) \mapsto \sum_{\alpha} m_{\alpha} \phi(\alpha) \in {\mathscr{P}}^{+}_{0},$$ we regard ${\mathrm{KP}}(\beta)$ as a subset of ${\mathscr{P}}^{+}_{0}$. Then we have a disjoint union decomposition $${\mathscr{P}}^{+}_{0} = \bigsqcup_{\beta \in {\mathsf{Q}}^{+}} {\mathrm{KP}}(\beta).$$ \[Prop:fiber\] Let ${\mathbf{m}}\in {\mathrm{KP}}(\beta)$ and pick a point $x_{{\mathbf{m}}} \in \mathbb{O}_{{\mathbf{m}}}$. 1. \[Prop:fiber:isom\] We have an isomorphism $\pi_{\beta}^{-1}(x_{{\mathbf{m}}}) \cong {\mathfrak{L}^{\bullet}}({\mathbf{m}})$. 2. \[Prop:fiber:stab\] The maximal reductive quotient of the stabilizer ${\mathop{\mathrm{Stab}}\nolimits}_{G_{\beta}}(x_{{\mathbf{m}}}) \subset G_{\beta}$ of the point $x_{{\mathbf{m}}}$ is isomorphic to $G({\mathbf{m}})$. 3. The isomorphism in (\[Prop:fiber:isom\]) induces the following commutative diagram: $$\xy \xymatrix{ {\mathop{\mathrm{Aut}}\nolimits}(\pi_{\beta}^{-1}(x_{{\mathbf{m}}})) \ar[r]^-{\cong} & {\mathop{\mathrm{Aut}}\nolimits}({\mathfrak{L}^{\bullet}}({\mathbf{m}})) \\ {\mathop{\mathrm{Stab}}\nolimits}_{G_\beta}(x_{{\mathbf{m}}}) \ar[u] \ar@{->>}[r] & G({\mathbf{m}}), \ar[u] } \endxy$$ where the vertical arrows are the action maps and the lower horizontal arrow is the canonical quotient map in (\[Prop:fiber:stab\]). Convolution and geometric extension algebras {#Sec:Conv} ============================================ Let ${\Bbbk}$ be a field of characteristic zero. Later in Subsection \[Ssec:Nakajima\], we specialize ${\Bbbk}= {\mathbb{Q}}(q)$. Preliminary on equivariant geometry {#Ssec:Pre} ----------------------------------- For the materials in this subsection, we refer [@CG97] and [@EG00]. Let $G$ be a complex linear algebraic group. A $G$-variety $X$ is a quasi-projective complex algebraic variety equipped with an algebraic action of the group $G$. We set $\mathrm{pt} := \mathop{\mathrm{Spec}} {\mathbb{C}}$ with the trivial $G$-action. The equivariant $K$-group $K^{G}(X)$ is defined to be the Grothendieck group of the abelian category of $G$-equivariant coherent sheaves on $X$ which is a module over the representation ring $R(G)=K^{G}(\mathrm{pt})$. We put $$K^{G}(X)_{{\Bbbk}} := K^{G}(X)\otimes_{{\mathbb{Z}}}{\Bbbk}, \quad R(G)_{{\Bbbk}} := R(G)\otimes_{{\mathbb{Z}}}{\Bbbk}.$$ Let $I \subset R(G)_{{\Bbbk}}$ be the augmentation ideal, i.e. the ideal generated by virtual representations of dimension $0$. We define the $I$-adic completions by $${\widehat{K}}^{G}(X)_{{\Bbbk}} := \varprojlim_{k} K^{G}(X)_{{\Bbbk}}/I^{k} K^{G}(X)_{{\Bbbk}}, \quad {\widehat{R}}(G)_{{\Bbbk}}:= \varprojlim_{k} R(G)_{{\Bbbk}} / I^{k}.$$ The completed $K$-group ${\widehat{K}}^{G}(X)_{{\Bbbk}}$ is a module over the algebra ${\widehat{R}}(G)_{{\Bbbk}}$. On the other hand, the $G$-equivariant Borel-Moore homology with ${\Bbbk}$-coefficients $$H^{G}_{*}(X, {\Bbbk}) = \bigoplus_{k \in {\mathbb{Z}}}H_{k}^{G}(X, {\Bbbk}),$$ is a module over the $G$-equivariant cohomology ring $H_{G}^{*}(\mathrm{pt}, {\Bbbk})$ of $\mathrm{pt}$ (with the cup product). Let us define the completion of a ${\mathbb{Z}}$-graded ${\Bbbk}$-vector space $V=\bigoplus_{k \in {\mathbb{Z}}}V_{k}$ by $ V^{\wedge} := \prod_{k \in {\mathbb{Z}}} V_{k}. $ The completion $H^{*}_{G}(\mathrm{pt}, {\Bbbk})^{\wedge}$ naturally becomes a ${\Bbbk}$-algebra and the completion $H^{G}_{*}(X, {\Bbbk})^{\wedge}$ becomes a module over $H^{*}_{G}(\mathrm{pt}, {\Bbbk})^{\wedge}$. Assume that our $G$-variety $X$ is a $G$-stable closed subvariety of a non-singular ambient $G$-variety $M$. Then we have the $G$-equivariant local Chern character map $$({\mathrm{ch}}^{G})^{M}_{X} : {\widehat{K}}^{G}(X)_{{\Bbbk}} \to H^{G}_{*}(X, {\Bbbk})^{\wedge}.$$ relative to $M$. We simply write ${\mathrm{ch}}^{G}$ instead of $({\mathrm{ch}}^{G})^{M}_{X}$ if the pair $(M, X)$ is obvious from the context. When $X=M=\mathrm{pt}$, the corresponding Chern character map induces an isomorphism of ${\Bbbk}$-algebras $${\widehat{R}}(G)_{{\Bbbk}} ={\widehat{K}}^{G}(\mathrm{pt})_{{\Bbbk}} \cong H^{G}_{*}(\mathrm{pt}, {\Bbbk})^{\wedge} = H_{G}^{*}(\mathrm{pt}, {\Bbbk})^{\wedge}.$$ We identify $ H^{G}_{*}(\mathrm{pt}, {\Bbbk})^{\wedge} $ with ${\widehat{R}}(G)_{{\Bbbk}}$ via this isomorphism. Then $({\mathrm{ch}}^{G})^{M}_{X}$ is regarded as an ${\widehat{R}}(G)_{{\Bbbk}}$-homomorphism. For a $G$-equivariant vector bundle $E$ on a non-singular $M$, let ${\mathrm{Td}}^{G}(E) \in H^{*}_{G}(M, {\Bbbk})^{\wedge}$ be the $G$-equivariant Todd class. This is an invertible element with respect to the cup product. For the tangent bundle $T_{M}$ of $M$, we put ${\mathrm{Td}}^{G}_{M} := {\mathrm{Td}}^{G}(T_{M})$. \[Thm:EG\] For $i=1,2$, let $X_{i}$ be a $G$-variety which is a $G$-stable closed subvariety of a non-singular ambient $G$-variety $M_{i}$. Assume that a $G$-equivariant morphism $\tilde{f}:M_{1}\to M_{2}$ restricts to a proper morphism $f : X_{1} \to X_{2}$. Then we have $$f_{*} \left({\mathrm{Td}}_{M_{1}}^{G} \cdot ({\mathrm{ch}}^{G})_{X_{1}}^{M_{1}}(\zeta) \right) = {\mathrm{Td}}^{G}_{M_{2}} \cdot ({\mathrm{ch}}^{G})_{X_{2}}^{M_{2}}(f_{*} \zeta), \quad \zeta \in {\widehat{K}}^{G}(X_{1})_{{\Bbbk}}.$$ The following proposition is standard. \[Prop:chex\] Let $M$ be a non-singular $G$-variety. Let $Y \subset X \subset M$ be $G$-stable closed subvarieties, and $i : Y \hookrightarrow X$, $j : X\setminus Y \hookrightarrow X$ be inclusions. Then we have the following commutative diagram: $$\xy \xymatrix{ {\widehat{K}}^{G}(Y)_{{\Bbbk}} \ar[r]^-{i_{*}} \ar[d]^-{({\mathrm{ch}}^{G})^{M}_{Y}} & {\widehat{K}}^{G}(X)_{{\Bbbk}} \ar[r]^-{j^{*}} \ar[d]^-{({\mathrm{ch}}^{G})^{M}_{X}} & {\widehat{K}}^{G}(X\setminus Y)_{{\Bbbk}} \ar[d]^-{({\mathrm{ch}}^{G})^{M\setminus Y}_{X\setminus Y}} \\ H_{*}^{G}(Y, {\Bbbk})^{\wedge} \ar[r]^-{i_{*}} & H_{*}^{G}(X, {\Bbbk})^{\wedge} \ar[r]^-{j^{*}} & H_{*}^{G}(X\setminus Y, {\Bbbk})^{\wedge}. } \endxy$$ Next we consider the convolution products. Let $M_i$ be non-singular $G$-varieties for $i=1,2,3$. We denote by $p_{ij} : M_1 \times M_2 \times M_3 \to M_i \times M_j$ the projection to the $(i,j)$-factors for $(i,j) = (1,2), (2,3), (1,3)$. Let $Z_{12} \subset M_{1} \times M_{2}$ and $Z_{23} \subset M_{2} \times M_{3}$ be $G$-stable closed subvarieties such that the morphism $$p_{13} : p_{12}^{-1}(Z_{12}) \cap p_{23}^{-1}(Z_{23}) \to Z_{13} := p_{13}(p_{12}^{-1}(Z_{12}) \cap p_{23}^{-1}(Z_{23}) )$$ is proper. Then we define the convolution product $ * : K^{G}(Z_{12}) \otimes_{R(G)} K^{G}(Z_{23}) \to K^{G}(Z_{13}) $ relative to $M_{1} \times M_{2} \times M_{3}$ by $$\zeta * \eta := p_{13*} (p_{12}^{*}\zeta \otimes^{\mathbb{L}}_{M_{1} \times M_{2} \times M_{3}} p_{23}^{*}\eta ), \quad \zeta \in K^{G}(Z_{12}), \eta \in K^{G}(Z_{23}).$$ This naturally induces the convolution product on the completed $G$-equivariant $K$-groups ${\widehat{K}}^{G}(Z_{12})_{{\Bbbk}} \otimes_{{\widehat{R}}(G)_{{\Bbbk}}} {\widehat{K}}^{G}(Z_{23})_{{\Bbbk}} \to {\widehat{K}}^{G}(Z_{13})_{{\Bbbk}} $. Similarly, we have the convolution product on the $G$-equivariant Borel-Moore homologies $H_{*}^{G}(Z_{12}, {\Bbbk}) \otimes_{H^{*}_{G}(\mathrm{pt}, {\Bbbk})} H_{*}^{G}(Z_{23}, {\Bbbk}) \to H^{G}_{*}(Z_{13}, {\Bbbk})$ relative to $M_{1} \times M_{2} \times M_{3}$ and its completed version $H_{*}^{G}(Z_{12}, {\Bbbk})^{\wedge} \otimes_{{\widehat{R}}(G)_{{\Bbbk}}} H_{*}^{G}(Z_{23}, {\Bbbk})^{\wedge} \to H_{*}^{G}(Z_{13}, {\Bbbk})^{\wedge}. $ Under the situation in the previous paragraph, for each $(i,j)= (1,2), (2,3), (1,3)$, we also define the $G$-equivariant Riemann-Roch homomorphism ${\mathrm{RR}}^{G} : {\widehat{K}}^{G}(Z_{ij})_{{\Bbbk}} \to H_{*}^{G}(Z_{ij}, {\Bbbk})^{\wedge}$ relative to $M_{i} \times M_{j}$ by $${\mathrm{RR}}^{G}(\zeta) := (p_{i}^{*} {\mathrm{Td}}^{G}_{M_{i}}) \cdot ({\mathrm{ch}}^{G})_{Z_{ij}}^{M_{i} \times M_{j}} (\zeta), \quad \zeta \in {\widehat{K}}^{G}(Z_{ij})_{{\Bbbk}},$$ where $p_{i} : M_{i} \times M_{j} \to M_{i}$ is the projection. By a completely similar discussion as in [@CG97 5.11.11], we can prove the following. \[Prop:RR\] The $G$-equivariant Riemann-Roch homomorphisms are compatible with the convolution product, i.e. we have $${\mathrm{RR}}^{G}(\zeta * \eta) = {\mathrm{RR}}^{G}(\zeta) * {\mathrm{RR}}^{G}(\eta), \quad \zeta \in {\widehat{K}}^{G}(Z_{12})_{{\Bbbk}}, \eta \in {\widehat{K}}^{G}(Z_{23})_{{\Bbbk}}.$$ Quiver Hecke algebra {#Ssec:KLR} -------------------- Fix an element $\beta = \sum_{i \in I} d_{i} \alpha_{i} \in {\mathsf{Q}}^{+}$ and put $d := \sum_{i \in I} d_{i}$. Let $$I^{\beta} := \{ {\mathbf{i}}= (i_{1}, \ldots, i_{d}) \in I^{d} \mid \alpha_{i_{1}} + \cdots + \alpha_{i_{d}} = \beta \}.$$ The symmetric group ${\mathfrak{S}}_{d}$ of degree $d$ acts on the set $I^{\beta}$ from the right by $(i_{1}, \ldots, i_{d}) \cdot w := (i_{w(1)}, \ldots , i_{w(d)}).$ Let $s_{k} \in {\mathfrak{S}}_{d}$ denote the transposition of $k$ and $k+1$ for $1 \le k < d$. \[Def:KLR\] The quiver Hecke algebra $H_{Q}(\beta)$ is defined to be a ${\Bbbk}$-algebra with the generating set $ \{ e( {\mathbf{i}}) \mid {\mathbf{i}}\in I^{\beta}\} \cup \{x_{1}, \ldots, x_{d} \} \cup \{\tau_{1}, \ldots, \tau_{d-1} \}, $ satisfying the following relations: $$e({\mathbf{i}}) e({\mathbf{i}}^{\prime}) = \delta_{{\mathbf{i}}, {\mathbf{i}}^{\prime}} e({\mathbf{i}}), \quad \sum_{{\mathbf{i}}\in I^{\beta}} e({\mathbf{i}}) = 1, \quad x_{k} x_{l} = x_{l} x_{k}, \quad x_k e({\mathbf{i}}) = e({\mathbf{i}}) x_{k},$$ $$\tau_{k} e({\mathbf{i}}) = e({\mathbf{i}}\cdot s_k) \tau_k, \quad \tau_k \tau_l = \tau_l \tau_k \quad \text{if $|k-l| > 1$},$$ $$\tau_{k}^{2} e({\mathbf{i}}) = \begin{cases} (x_{k} - x_{k+1})e({\mathbf{i}}), & \text{if $i_{k} \leftarrow i_{k+1}$},\\ (x_{k+1} - x_{k})e({\mathbf{i}}), & \text{if $i_{k} \to i_{k+1}$},\\ e({\mathbf{i}}) & \text{if $a_{i_{k}, i_{k+1}} = 0$},\\ 0 & \text{if $i_{k}=i_{k+1}$}, \end{cases}$$ $$(\tau_k x_l - x_{s_{k}(l)} \tau_k)e({\mathbf{i}}) = \begin{cases} - e({\mathbf{i}}) & \text{if $l=k, i_{k} = i_{k+1}$}, \\ e({\mathbf{i}}) & \text{if $l=k+1, i_{k} = i_{k+1}$}, \\ 0 & \text{otherwise}, \end{cases}$$ $$(\tau_{k+1} \tau_{k} \tau_{k+1} - \tau_{k} \tau_{k+1} \tau_{k}) e({\mathbf{i}}) = \begin{cases} e({\mathbf{i}}) & \text{if $i_{k} = i_{k+2}, i_{k} \leftarrow i_{k+1}$}, \\ -e({\mathbf{i}}) & \text{if $i_{k} = i_{k+2}, i_{k} \to i_{k+1}$}, \\ 0 & \text{otherwise}. \end{cases}$$ The quiver Hecke algebra $H_{Q}(\beta)$ is equipped with a ${\mathbb{Z}}$-grading given by $$\deg e({\mathbf{i}}) = 0, \quad \deg x_k = 2, \quad \deg \tau_{k} e({\mathbf{i}}) = -a_{i_k, i_{k+1}}.$$ Since the grading is bounded from below (see [@KL09 Theorem 2.5]), the completion $\widehat{H}_{Q}(\beta) := H_{Q}(\beta)^{\wedge}$ inherits a natural structure of ${\Bbbk}$-algebra. We recall the faithful polynomial right representation of $H_{Q}(\beta)$ from [@KL09 Section 2.3]. We set $$P_{\beta} := \bigoplus_{{\mathbf{i}}\in I^{\beta}} {\Bbbk}[x_{1}, \ldots, x_{d}] 1_{{\mathbf{i}}}$$ with a commutative ${\Bbbk}[x_{1}, \ldots, x_{d}]$-algebra structure $1_{{\mathbf{i}}} \cdot 1_{{\mathbf{i}}^{\prime}} = \delta_{{\mathbf{i}}{\mathbf{i}}^{\prime}} 1_{{\mathbf{i}}}$. We define $f^{w}(x_{1}, \ldots, x_{d}) := f(x_{w(1)}, \ldots, x_{w(d)})$ for $f \in {\Bbbk}[x_{1}, \ldots, x_{d}]$ and $w \in {\mathfrak{S}}_{d}$. \[Thm:KL\] The following formulas give a faithful right $H_{Q}(\beta)$-module structure on the ${\Bbbk}$-vector space $P_{\beta}$: $$\begin{aligned} a \cdot e({\mathbf{i}}) &= a 1_{{\mathbf{i}}}, \\ a \cdot x_{k} &= a x_{k}, \\ (f 1_{{\mathbf{i}}}) \cdot \tau_{k} &= \begin{cases} \displaystyle \frac{f^{s_{k}} - f}{x_{k} - x_{k+1}} 1_{{\mathbf{i}}} & \text{if $i_{k} = i_{k+1}$},\\ (x_{k+1} - x_{k})f^{s_{k}} 1_{{\mathbf{i}}\cdot s_{k}} & \text{if $i_{k} \leftarrow i_{k+1}$},\\ f^{s_{k}} 1_{{\mathbf{i}}\cdot s_{k}} & \text{otherwise}, \end{cases} \end{aligned}$$ where $a \in P_{\beta}$ and $f 1_{{\mathbf{i}}} \in {\Bbbk}[x_{1}, \ldots, x_{d}]1_{{\mathbf{i}}}$. Replacing the polynomial ring ${\Bbbk}[x_{1}, \ldots, x_{d}]$ with the ring ${\Bbbk}[\![ x_{1}, \ldots, x_{d} ]\!]$ of formal power series, we get the completion of the representation $P_{\beta}$: $$\label{Eq:Phat} \widehat{P}_{\beta} := \bigoplus_{{\mathbf{i}}\in I^{\beta}} {\Bbbk}[\![ x_{1}, \ldots, x_{d} ]\!] 1_{{\mathbf{i}}} = P_{\beta} \otimes_{H_{Q}(\beta)} \widehat{H}_{Q}(\beta).$$ Varagnolo-Vasserots realization {#Ssec:VV} ------------------------------- Fix an $I$-graded ${\mathbb{C}}$-vector space $D=\bigoplus_{i \in I} D_{i}$ with ${\underline{\dim}\,}D = \beta$, i.e. $\dim D_{i} = d_i$ as in Subsection \[Ssec:Rep\_quiver\]. We consider the following two non-singular $G_{\beta}$-varieties: $$\begin{aligned} {\mathcal{B}}_{\beta} &= \{F^{\bullet} = (D=F^{0} \supsetneq F^{1} \supsetneq \cdots \supsetneq F^{d}=0) \mid \text{$F^{k}$ is an $I$-graded subspace of $D$} \}, \\ {\mathcal{F}}_{\beta} &= \{(F^{\bullet}, x) \in {\mathcal{B}}_{\beta} \times E_{\beta} \mid x(F^{k}) \subset F^{k} \; \text{for any $1\le k \le d$} \}.\end{aligned}$$ The $G_\beta$-action on ${\mathcal{F}}_{\beta}$ is defined so that the projections $\mathrm{pr}_{1} : {\mathcal{F}}_{\beta} \to {\mathcal{B}}_{\beta}$ and $\mu_{\beta} := \mathrm{pr}_{2}: {\mathcal{F}}_{\beta} \to E_{\beta}$ are $G_{\beta}$-equivariant. They decompose into connected components as $${\mathcal{B}}_{\beta} = \bigsqcup_{{\mathbf{i}}\in I^{\beta}} {\mathcal{B}}_{{\mathbf{i}}}, \qquad {\mathcal{F}}_{\beta} = \bigsqcup_{{\mathbf{i}}\in I^{\beta}} {\mathcal{F}}_{{\mathbf{i}}},$$ where we put $${\mathcal{B}}_{{\mathbf{i}}} := \{ F^{\bullet} \in {\mathcal{B}}_{\beta} \mid {\underline{\dim}\,}F^{k-1} = {\underline{\dim}\,}F^{k} + \alpha_{i_k}, \; \forall k \}, \quad {\mathcal{F}}_{{\mathbf{i}}} := (\mathrm{pr}_{1})^{-1}({\mathcal{B}}_{{\mathbf{i}}})$$ for ${\mathbf{i}}=(i_{1}, \ldots, i_{d}) \in I^{\beta}$. We fix a basis $\{ v_{k} \}_{1 \le k\le d}$ of the vector space $D$ so that the set $\{ v_{i,j}\}_{1 \le j \le d_{i}}$ forms a basis of the vector space $D_{i}$ for each $i \in I$, where we put $ v_{i,j} := v_{d_{1} + \cdots + d_{i-1} + j}. $ Let $H_{i} \subset GL(D_{i})$ be the maximal torus fixing the lines $\{{\mathbb{C}}v_{i,j}\}_{1\le j \le d_{i}}$ for each $i \in I$. We set $H_{\beta} := \prod_{i \in I} H_{i} \subset G_{\beta}$. Let $ F_{0}^{\bullet} \in {\mathcal{B}}_{\beta} $ be the flag defined by $ F_{0}^{k} := \bigoplus_{l > k} {\mathbb{C}}v_{l}, $ which belongs to the component ${\mathcal{B}}_{{\mathbf{i}}_0}$ with ${\mathbf{i}}_0 := (1^{d_{1}}, 2^{d_{2}}, \ldots, n^{d_{n}}) \in I^{\beta}$. For each ${\mathbf{i}}\in I^{\beta}$, we fix an element $w_{{\mathbf{i}}} \in {\mathfrak{S}}_{d}$ such that ${\mathbf{i}}= {\mathbf{i}}_{0} \cdot w_{{\mathbf{i}}}$. The set $\{ w_{{\mathbf{i}}} \}_{{\mathbf{i}}\in I^{\beta}}$ forms a complete system of coset representatives for the quotient ${\mathfrak{S}}_{\beta} \backslash {\mathfrak{S}}_{d}$, where ${\mathfrak{S}}_{\beta} := {\mathop{\mathrm{Stab}}\nolimits}_{{\mathfrak{S}}_{d}}({\mathbf{i}}_0) = {\mathfrak{S}}_{d_1} \times \cdots \times {\mathfrak{S}}_{d_n}$. For each $w \in {\mathfrak{S}}_{d}$, we define the flag $F^{\bullet}_{w}$ by $ F_{w}^{k} := \bigoplus_{l > k} {\mathbb{C}}v_{w(l)} $ which belongs to the component ${\mathcal{B}}_{{\mathbf{i}}_{0} \cdot w}$. Let $F^{\bullet}_{{\mathbf{i}}} := F^{\bullet}_{w_{{\mathbf{i}}}} \in {\mathcal{B}}_{{\mathbf{i}}}$ for ${\mathbf{i}}\in I^{\beta}$. Then we have ${\mathcal{B}}_{{\mathbf{i}}} \cong G_{\beta}/B_{{\mathbf{i}}}$ with $B_{{\mathbf{i}}} := {\mathop{\mathrm{Stab}}\nolimits}_{G_{\beta}}(F_{{\mathbf{i}}}^{\bullet}) \subset G_{\beta}$ being the Borel subgroup fixing the flag $F^{\bullet}_{{\mathbf{i}}}$, which contains the maximal torus $H_{\beta}$. Then we have $$\label{Eq:isomRH} H^{G_{\beta}}_{*}({\mathcal{B}}_{{\mathbf{i}}}, {\Bbbk}) \cong H^{B_{{\mathbf{i}}}}_{*}(\mathrm{pt}, {\Bbbk}) \cong H_{H_{\beta}}^{*}(\mathrm{pt}, {\Bbbk}) \cong {\Bbbk}[x_{1}, \ldots, x_{d}] 1_{{\mathbf{i}}},$$ where the last isomorphism sends the $1$st $H_{\beta}$-equivariant Chern class of the line ${\mathbb{C}}v_{w_{{\mathbf{i}}}(k)}$ to the element $x_{k}1_{{\mathbf{i}}}$. Thus we get an isomorphism $$\label{Eq:isomP} H_{*}^{G_{\beta}}({\mathcal{B}}_{\beta}, {\Bbbk}) = \bigoplus_{{\mathbf{i}}\in I^{\beta}} H_{*}^{G_{\beta}}({\mathcal{B}}_{{\mathbf{i}}}, {\Bbbk}) \cong \bigoplus_{{\mathbf{i}}\in I^{\beta}} {\Bbbk}[x_{1}, \ldots, x_{d}] 1_{{\mathbf{i}}} = P_{\beta}.$$ We consider the Steinberg type variety $ {\mathcal{Z}}_\beta := {\mathcal{F}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta} $ associated with the morphism $\mu_{\beta}: {\mathcal{F}}_{\beta} \to E_{\beta}$. Its $G_{\beta}$-equivariant Borel-Moore homology group $ H^{G_{\beta}}_{*}({\mathcal{Z}}_{\beta}, {\Bbbk})$ becomes a ${\Bbbk}$-algebra with respect to the convolution product relative to ${\mathcal{F}}_{\beta} \times {\mathcal{F}}_{\beta} \times {\mathcal{F}}_{\beta}$. We identify the variety ${\mathcal{B}}_{\beta}$ with the fiber product $\{ 0 \} \times_{E_{\beta}} {\mathcal{F}}_{\beta}$. Then the convolution product relative to $\{ 0 \} \times {\mathcal{F}}_{\beta} \times {\mathcal{F}}_{\beta}$ makes the space $H_{*}^{G_{\beta}}({\mathcal{B}}_{\beta}, {\Bbbk})$ into a right $H_{*}^{G_{\beta}}({\mathcal{Z}}_{\beta}, {\Bbbk})$-module. Let $\mu_{{\mathbf{i}}}$ denote the restriction of the proper morphism $\mu_{\beta}: {\mathcal{F}}_{\beta} \to E_{\beta}$ to the component ${\mathcal{F}}_{{\mathbf{i}}}$ for ${\mathbf{i}}\in I^{\beta}$. We put $$\mathcal{L}_{\beta} := \bigoplus_{{\mathbf{i}}\in I^{\beta}} (\mu_{{\mathbf{i}}})_{*} \underline{{\Bbbk}} [\dim {\mathcal{F}}_{{\mathbf{i}}}],$$ where $\underline{{\Bbbk}}[\dim {\mathcal{F}}_{{\mathbf{i}}}]$ is the trivial local system (i.e. the constant ${\Bbbk}$-sheaf of rank $1$) on ${\mathcal{F}}_{{\mathbf{i}}}$ homologically shifted by $\dim {\mathcal{F}}_{{\mathbf{i}}}$. By the decomposition theorem, we have $$\mathcal{L}_{\beta} \cong \bigoplus_{{\mathbf{m}}\in {\mathrm{KP}}(\beta)} L_{{\mathbf{m}}} \otimes_{{\Bbbk}} \mathcal{IC}_{{\mathbf{m}}} = \bigoplus_{{\mathbf{m}}\in {\mathrm{KP}}(\beta)} \bigoplus_{k \in {\mathbb{Z}}} L_{{\mathbf{m}}, k} \otimes_{{\Bbbk}} \mathcal{IC}_{{\mathbf{m}}}[k],$$ where $\mathcal{IC}_{{\mathbf{m}}}$ denotes the intersection cohomology complex associated with the trivial local system on the orbit $\mathbb{O}_{{\mathbf{m}}}$ and $L_{{\mathbf{m}}} = \bigoplus_{k \in {\mathbb{Z}}} L_{{\mathbf{m}}, k}[k]$ is a self-dual finite-dimensional graded ${\Bbbk}$-vector space for each ${\mathbf{m}}\in {\mathrm{KP}}(\beta)$. The vector space $L_{{\mathbf{m}}}$ is known to be non-zero for all ${\mathbf{m}}\in {\mathrm{KP}}(\beta)$ (see [@Kato14 Corollary 2.8]). We consider the Yoneda algebra $${\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*} (\mathcal{L}_{\beta}, \mathcal{L}_{\beta}) = \bigoplus_{k \in {\mathbb{Z}}} {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{k}( \mathcal{L}_{\beta}, \mathcal{L}_{\beta})$$ in the derived category of $G_{\beta}$-equivariant constructible complexes on $E_{\beta}$. This is a ${\mathbb{Z}}$-graded ${\Bbbk}$-algebra whose grading is bounded from below. By a standard argument (see [@CG97 Section 8.6]), we have an isomorphism of ${\Bbbk}$-algebras $$\label{Eq:isomG1} {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}, \mathcal{L}_{\beta}) \cong H^{G_{\beta}}_{*}({\mathcal{Z}}_{\beta}, {\Bbbk}).$$ Note that this is not compatible with the ${\mathbb{Z}}$-grading. Let $\mathcal{L}_{{\mathbf{i}}}(k)$ be the $G_{\beta}$-equivariant line bundle on ${\mathcal{F}}_{{\mathbf{i}}}$ whose fiber at the point $(F^{\bullet}, x) \in {\mathcal{F}}_{{\mathbf{i}}}$ is $F^{k-1} / F^{k}$ for ${\mathbf{i}}\in I^{\beta}$ and $1 \le k \le d$. \[Thm:VV\] There is a unique isomorphism of ${\mathbb{Z}}$-graded ${\Bbbk}$-algebras $$\label{Eq:isomVV} H_{Q}(\beta) \xrightarrow{\cong} {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*} (\mathcal{L}_{\beta}, \mathcal{L}_{\beta})$$ which satisfies the following properties: 1. \[Thm:VV:elem\] The composition $H_{Q}(\beta) \xrightarrow{\cong} H^{G_{\beta}}({\mathcal{Z}}_{\beta}, {\Bbbk})$ of the isomorphisms (\[Eq:isomVV\]) and (\[Eq:isomG1\]) sends the element $e({\mathbf{i}})$ (resp. $x_{k} e({\mathbf{i}})$) to the push-forward of the fundamental class $[{\mathcal{F}}_{{\mathbf{i}}}]$ (resp. the $1$st $G_{\beta}$-equivariant Chern class of the line bundle $\mathcal{L}_{{\mathbf{i}}}(k)$) with respect to the diagonal embedding ${\mathcal{F}}_{{\mathbf{i}}} \to {\mathcal{F}}_{{\mathbf{i}}} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}}$; 2. We have the following commutative diagram: $$\xy \xymatrix{ H_{Q}(\beta) \ar[r]^-{\cong} \ar[d] & H_{*}^{G_{\beta}}({\mathcal{Z}}_{\beta}, {\Bbbk}) \ar[d] \\ {\mathop{\mathrm{End}}\nolimits}\left(P_{\beta} \right)^{\mathrm{op}} \ar[r]^-{\cong} & {\mathop{\mathrm{End}}\nolimits}\left( H_{*}^{G_{\beta}}({\mathcal{B}}_{\beta}, {\Bbbk}) \right)^{\mathrm{op}}, } \endxy$$ where the lower horizontal arrow denotes the isomorphism induced from (\[Eq:isomP\]) and the vertical arrows denote the right module structures. Because our convention of the flag variety ${\mathcal{B}}_{\beta}$ differs from Varagnolo-Vasserots [@VV11], we need a modification. Actually, our isomorphism (\[Eq:isomVV\]) is obtained by twisting the original isomorphism $H_{Q}(\beta) \cong {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}( \mathcal{L}_{\beta}, \mathcal{L}_{\beta})$ in [@VV11] by a ${\Bbbk}$-algebra involution on $H_{Q}(\beta)$ given by $$e({\mathbf{i}}) \mapsto e({\mathbf{i}}^{\mathrm{op}}), \quad x_{k} \mapsto x_{d-k+1}, \quad \tau_{k}e({\mathbf{i}}) \mapsto \begin{cases} -\tau_{d-k} e({\mathbf{i}}^{\mathrm{op}}) & \text{if $i_k = i_{k+1}$}; \\ \tau_{d-k} e({\mathbf{i}}^{\mathrm{op}}) & \text{if $i_k \neq i_{k+1}$}, \end{cases}$$ where ${\mathbf{i}}^{\mathrm{op}} := (i_{d}, \ldots, i_{2}, i_{1})$ for ${\mathbf{i}}= (i_{1}, i_{2}, \ldots, i_{d}) \in I^{\beta}$. Similarly to the case of the $G_{\beta}$-equivariant Borel-Moore homologies, the $K$-group $K^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}}$ becomes an $R(G_{\beta})_{{\Bbbk}}$-algebra and the $K$-group $K^{G_{\beta}}({\mathcal{B}}_{\beta})_{{\Bbbk}}$ becomes a right $K^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}}$-module with respect to the convolution products. For each ${\mathbf{i}}\in I^{\beta}$, we have $$K^{G_{\beta}}({\mathcal{B}}_{{\mathbf{i}}})_{{\Bbbk}} \cong K^{B_{{\mathbf{i}}}}(\mathrm{pt})_{{\Bbbk}} \cong K^{H_{\beta}}(\mathrm{pt})_{{\Bbbk}} = R(H_{\beta})_{{\Bbbk}} \cong {\Bbbk}[y^{\pm 1}_{1}, \ldots, y^{\pm 1}_{d}] 1_{{\mathbf{i}}}$$ where the last isomorphism sends the class $[{\mathbb{C}}v_{w_{{\mathbf{i}}}(k)}]$ of the $1$-dimensional $H_{\beta}$-module ${\mathbb{C}}v_{w_{{\mathbf{i}}}(k)}$ to the element $y_{k} 1_{{\mathbf{i}}}$. The $G_{\beta}$-equivariant Chern character map $ ({\mathrm{ch}}^{G_{\beta}})^{{\mathcal{B}}_{{\mathbf{i}}}}_{{\mathcal{B}}_{{\mathbf{i}}}} $ gives an isomorphism of ${\Bbbk}$-algebras $${\widehat{K}}^{G_{\beta}}({\mathcal{B}}_{{\mathbf{i}}})_{{\Bbbk}} \cong {\Bbbk}[\![ y_{1} -1, \ldots, y_{d}-1]\!] 1_{{\mathbf{i}}} \xrightarrow{\cong} {\Bbbk}[\![ x_{1}, \ldots, x_{d} ]\!] 1_{{\mathbf{i}}} \cong H_{*}^{G_{\beta}}({\mathcal{B}}_{{\mathbf{i}}}, {\Bbbk})^{\wedge},$$ where the middle arrow sends the element $y_{k} 1_{{\mathbf{i}}}$ to the exponential $e^{x_{k}} 1_{{\mathbf{i}}}$ for $1 \le k \le d$. Applying the equivariant Riemann-Roch theorem (=Theorem \[Thm:EG\]) to the inclusion ${\mathcal{B}}_{{\mathbf{i}}} \hookrightarrow {\mathcal{F}}_{{\mathbf{i}}}$, we have $$\label{Eq:const} ({\mathrm{ch}}^{G_{\beta}})^{{\mathcal{F}}_{{\mathbf{i}}}}_{{\mathcal{B}}_{{\mathbf{i}}}} = C_{{\mathbf{i}}} \cdot ({\mathrm{ch}}^{G_{\beta}})^{{\mathcal{B}}_{{\mathbf{i}}}}_{{\mathcal{B}}_{{\mathbf{i}}}}, \quad C_{{\mathbf{i}}} := ({\mathrm{Td}}^{G_{\beta}}_{{\mathcal{F}}_{{\mathbf{i}}}})^{-1} {\mathrm{Td}}^{G_{\beta}}_{{\mathcal{B}}_{{\mathbf{i}}}} \cdot 1_{{\mathbf{i}}} \in {\Bbbk}[\![ x_{1}, \ldots, x_{d} ]\!] 1_{{\mathbf{i}}}$$ and hence the map $({\mathrm{ch}}^{G_{\beta}})^{{\mathcal{F}}_{{\mathbf{i}}}}_{{\mathcal{B}}_{{\mathbf{i}}}}$ is an isomorphism of ${\widehat{R}}(G_{\beta})_{{\Bbbk}}$-modules. Summing up over ${\mathbf{i}}\in I^{\beta}$, we obtain an isomorphism of ${\widehat{R}}(G_{\beta})_{{\Bbbk}}$-modules $$\label{Eq:isomRRB} ({\mathrm{ch}}^{G_{\beta}})^{{\mathcal{F}}_{\beta}}_{{\mathcal{B}}_{\beta}} : {\widehat{K}}^{G_{\beta}}({\mathcal{B}}_{\beta})_{{\Bbbk}} \xrightarrow{\cong} H_{*}^{G_{\beta}}({\mathcal{B}}_{\beta}, {\Bbbk})^{\wedge}.$$ \[Prop:RRflag\] The Riemann-Roch homomorphism gives an isomorphism of $\widehat{R}(G_{\beta})_{{\Bbbk}}$-algebras: $${\mathrm{RR}}^{G_{\beta}} : {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}} \xrightarrow{\cong} H_{*}^{G_{\beta}}({\mathcal{Z}}_{\beta}, {\Bbbk})^{\wedge},$$ which makes the following diagram commute: $$\label{Diag:sqKH} \xy \xymatrix{ {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}} \ar[r]^-{\cong} \ar[d] & H_{*}^{G_{\beta}}({\mathcal{Z}}_{\beta}, {\Bbbk})^{\wedge} \ar[d] \\ {\mathop{\mathrm{End}}\nolimits}\left({\widehat{K}}^{G_{\beta}}({\mathcal{B}}_{\beta})_{{\Bbbk}} \right)^{\mathrm{op}} \ar[r]^-{\cong} & {\mathop{\mathrm{End}}\nolimits}\left( H_{*}^{G_{\beta}}({\mathcal{B}}_{\beta}, {\Bbbk})^{\wedge} \right)^{\mathrm{op}}, } \endxy$$ where the lower horizontal arrow denotes the isomorphism induced from (\[Eq:isomRRB\]) and the vertical arrows denote the right module structures. By Proposition \[Prop:RR\], the map ${\mathrm{RR}}^{G_{\beta}} : {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}} \to H_{*}^{G_{\beta}}({\mathcal{Z}}_{\beta}, {\Bbbk})^{\wedge}$ is an algebra homomorphism and the diagram (\[Diag:sqKH\]) commutes. To prove that the map ${\mathrm{RR}}^{G_{\beta}} : {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}} \to H_{*}^{G_{\beta}}({\mathcal{Z}}_{\beta}, {\Bbbk})^{\wedge}$ is an isomorphism, it suffices to check that the equivariant Chern character map $({\mathrm{ch}}^{G_{\beta}} )^{{\mathcal{F}}_{\beta} \times {\mathcal{F}}_{\beta}}_{{\mathcal{Z}}_{\beta}} : {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}} \to H_{*}^{G_{\beta}}({\mathcal{Z}}_{\beta}, {\Bbbk})^{\wedge}$ gives an isomorphism of $\widehat{R}(G_{\beta})_{{\Bbbk}}$-modules since ${\mathrm{RR}}^{G_{\beta}}$ is obtained from $({\mathrm{ch}}^{G_{\beta}} )^{{\mathcal{F}}_{\beta} \times {\mathcal{F}}_{\beta}}_{{\mathcal{Z}}_{\beta}}$ by multiplying the $G_{\beta}$-equivariant Todd class $p_{1}^{*}{\mathrm{Td}}^{G_{\beta}}_{{\mathcal{F}}_{\beta}}$, which is an invertible element. Because we have the connected component decomposition $${\mathcal{Z}}_{\beta} = \bigsqcup_{{\mathbf{i}}, {\mathbf{i}}^{\prime} \in I^{\beta}} {\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}, \quad {\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}} := {\mathcal{F}}_{{\mathbf{i}}} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}^{\prime}},$$ we focus on a connected component $${\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}} = \{ (F^{\bullet}, F^{\prime \bullet}, x) \in {\mathcal{B}}_{{\mathbf{i}}} \times {\mathcal{B}}_{{\mathbf{i}}^{\prime}} \times E_{\beta} \mid x(F^{k}) \subset F^{k}, x(F^{\prime k}) \subset F^{\prime k}, \; \forall k \}.$$ For each $w \in {\mathfrak{S}}_{\beta} w_{{\mathbf{i}}^{\prime}}$, we define a locally closed $G_{\beta}$-subvariety $${\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{w} = G_{\beta} \times^{B_{{\mathbf{i}}}} \{ (F_{{\mathbf{i}}}^{\bullet}, F^{\prime \bullet}, x ) \in {\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}} \mid F^{\prime \bullet} \in B_{{\mathbf{i}}} F_{w}^{\bullet} \}$$ which is a $G_{\beta}$-equivariant affine bundle over ${\mathcal{B}}_{{\mathbf{i}}}$. They give a $G_{\beta}$-stable stratification $ {\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}} := \bigsqcup_{w \in {\mathfrak{S}}_{\beta} w_{{\mathbf{i}}^{\prime}}} Z_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{w}. $ Fix a total ordering ${\mathfrak{S}}_{\beta}w_{{\mathbf{i}}^{\prime}} = \{w_{1}, w_{2}, \ldots, w_{m} \}$ such that we have $w_{k}w_{{\mathbf{i}}}^{-1} < w_{l} w_{{\mathbf{i}}}^{-1}$ in the Bruhat ordering only if $k<l$. We simply write ${\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{k} := {\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{w_k}$ and set ${\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{\le k} := \bigsqcup_{j \le k} {\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{j}$. Then for each $k$, the variety ${\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{\le k-1}$ is closed in ${\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{\le k}$ and its complement is ${\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{k}$. Since ${\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{k}$ is a $G_{\beta}$-equivariant affine bundle over ${\mathcal{B}}_{{\mathbf{i}}}$, its homology of odd degree vanishes: $H_{\mathrm{odd}}^{G_{\beta}}({\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{k}, {\Bbbk})=0$. Therefore an inductive argument with respect to $k$ yields $H_{\mathrm{odd}}^{G_{\beta}}({\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{\le k}, {\Bbbk}) =0$. Using the cellular fibration lemma [@CG97 5.5.1] for equivariant $K$-groups and Proposition \[Prop:chex\], we obtain the following commutative diagram with exact rows for each $k$: $$\xy \xymatrix{ 0 \ar[r] & {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{\le k-1})_{{\Bbbk}} \ar[r] \ar[d]^-{{\mathrm{ch}}^{G_{\beta}}} & {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{\le k})_{{\Bbbk}} \ar[r] \ar[d]^-{{\mathrm{ch}}^{G_{\beta}}} & {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{k})_{{\Bbbk}} \ar[r] \ar[d]^-{{\mathrm{ch}}^{G_{\beta}}} & 0 \\ 0 \ar[r] & H_{*}^{G_{\beta}}({\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{\le k-1}, {\Bbbk})^{\wedge} \ar[r] & H_{*}^{G_{\beta}}({\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{\le k}, {\Bbbk})^{\wedge} \ar[r] & H_{*}^{G_{\beta}}({\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{k}, {\Bbbk})^{\wedge} \ar[r] & 0. } \endxy$$ Note that the map ${\mathrm{ch}}^{G_{\beta}} : {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}^{k}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}})_{{\Bbbk}} \to H_{*}^{G_{\beta}}({\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{k}, {\Bbbk})^{\wedge}$ is an isomorphism for any $k$ since again the variety $Z^{k}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}$ is an affine bundle over ${\mathcal{B}}_{{\mathbf{i}}}$. Hence, by induction on $k$, we conclude that ${\mathrm{ch}}^{G_{\beta}} : {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}^{\le k}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}})_{{\Bbbk}} \to H_{*}^{G_{\beta}}({\mathcal{Z}}_{{\mathbf{i}}, {\mathbf{i}}^{\prime}}^{\le k}, {\Bbbk})^{\wedge}$ is an isomorphism for all $k$. Note that the isomorphism (\[Eq:isomG1\]) induces an isomorphism between the completions: $${\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}, \mathcal{L}_{\beta})^{\wedge} \cong H_{*}^{G_{\beta}}({\mathcal{Z}}_{\beta}, {\Bbbk})^{\wedge}.$$ As a summary of this subsection, we have the following. We have the following isomorphisms of ${\Bbbk}$-algebras: $$\widehat{H}_{Q}(\beta) \cong {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}, \mathcal{L}_{\beta})^{\wedge} \cong H_{*}^{G_{\beta}}({\mathcal{Z}}_{\beta}, {\Bbbk})^{\wedge} \cong {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}}.$$ Nakajimas homomorphism and the category ${\mathcal{C}}_{Q, \beta}$ {#Ssec:Nakajima} ------------------------------------------------------------------ Henceforth, we specialize ${\Bbbk}$ to be the field ${\mathbb{Q}}(q)$ of rational functions in an indeterminate $q$. In this subsection, we consider the quantum loop algebra $U_{q} \equiv U_{q}(L {\mathfrak{g}})$ defined over ${\Bbbk}$. The quantum loop algebra $U_{q}(L {\mathfrak{g}})$ is isomorphic to the level zero quotient of the quantum affine algebra $U_{q}^{\prime}(\widehat{{\mathfrak{g}}})$ without the degree operator. We do not recall the definitions here. See e.g. [@Fujita17], [@KKK15], [@Nakajima01] for the precise definitions of $U_{q}(L{\mathfrak{g}})$ or $U^{\prime}_{q}(\widehat{{\mathfrak{g}}})$. Recall the quiver varieties with proper ${\mathbb{G}}(\lambda)$-equivariant morphism $\pi : {\mathfrak{M}}(\lambda) \to {\mathfrak{M}}_{0}(\lambda)$ for each $\lambda \in {\mathsf{P}}^{+}$ (see Subsection \[Ssec:Quiver\_var\]). We consider the Steinberg type variety $Z(\lambda) := {\mathfrak{M}}(\lambda) \times_{{\mathfrak{M}}_{0}(\lambda)} {\mathfrak{M}}(\lambda)$. Then its ${\mathbb{G}}(\lambda)$-equivariant $K$-group $K^{{\mathbb{G}}(\lambda)}(Z(\lambda))$ becomes an $R({\mathbb{G}}(\lambda))$-algebra with respect to the convolution product relative to ${\mathfrak{M}}(\lambda) \times {\mathfrak{M}}(\lambda) \times {\mathfrak{M}}(\lambda)$. We identify the fiber ${\mathfrak{L}}(\lambda) = \pi^{-1}(0)$ with the fiber product ${\mathfrak{M}}(\lambda) \times_{{\mathfrak{M}}_{0}(\lambda)} \{0\}.$ Then the convolution product relative to ${\mathfrak{M}}(\lambda) \times {\mathfrak{M}}(\lambda) \times \{ 0 \}$ makes the $K$-group $K^{{\mathbb{G}}(\lambda)}({\mathfrak{L}}(\lambda))$ into a left $K^{{\mathbb{G}}(\lambda)}(Z(\lambda))$-module. Recall that ${\mathbb{G}}(\lambda) = G(\lambda) \times {\mathbb{C}}^{\times}$. We set $A := R({\mathbb{C}}^{\times})$ and identify $A = {\mathbb{Z}}[v^{\pm 1}]$ in the standard way. Specializing $v$ to $q$, we regard ${\Bbbk}$ as an $A$-algebra. \[Thm:Nakajima\] There exists a ${\Bbbk}$-algebra homomorphism $$\Phi_{\lambda} : U_{q}(L{\mathfrak{g}}) \to K^{{\mathbb{G}}(\lambda)}(Z(\lambda))\otimes_{A} {\Bbbk}$$ such that the pull-back $${\mathbb{W}}(\lambda) := \Phi_{\lambda}^{*} \left( K^{{\mathbb{G}}(\lambda)}({\mathfrak{L}}(\lambda))\otimes_{A}{\Bbbk}\right)$$ is a cyclic $U_{q}(L{\mathfrak{g}})$-module generated by an extremal weight vector $w_{\lambda} := [\mathcal{O}_{{\mathfrak{L}}(0, \lambda)}] \in K^{{\mathbb{G}}(\lambda)}({\mathfrak{L}}(0,\lambda))\otimes_{A} {\Bbbk}$ of weight $\lambda$. Moreover the module ${\mathbb{W}}(\lambda)$ is free of finite rank over ${\mathop{\mathrm{End}}\nolimits}_{U_{q}}({\mathbb{W}}(\lambda)) \cong R({\mathbb{G}}(\lambda)) \otimes_{A} {\Bbbk}$. The module ${\mathbb{W}}(\lambda)$ is known to be isomorphic to the global Weyl module defined by Chari-Pressley [@CP01] and also to the level $0$ extremal weight module defined by Kashiwara [@Kashiwara94]. In particular, if $\lambda = \varpi_{i}$ for some $i \in I$, the module ${\mathbb{W}}(\varpi_{i})$ is isomorphic to the affinization of the fundamental module $W(\varpi_{i})$ (see [@Kashiwara02]). Take an element ${{\boldsymbol \lambda}}\in {\mathscr{P}}^{+}$ with ${\mathsf{cl}}({{\boldsymbol \lambda}}) = \lambda$ and recall the $1$-dimensional subtorus $T({{\boldsymbol \lambda}}) \subset {\mathbb{G}}({{\boldsymbol \lambda}}) \subset {\mathbb{G}}(\lambda)$. We identify $R(T({{\boldsymbol \lambda}})) = A$ via the isomorphism $\prod_{i \in I}f_{i} \times \mathrm{id} : {\mathbb{C}}^{\times} \xrightarrow{\cong} T({{\boldsymbol \lambda}})$. Let $\mathfrak{m}_{{{\boldsymbol \lambda}}}$ be the kernel of the restriction $ R({\mathbb{G}}(\lambda))\otimes_{A} {\Bbbk}\to R(T({{\boldsymbol \lambda}}))\otimes_{A} {\Bbbk}= {\Bbbk}. $ The corresponding specialization $ {\mathbb{W}}(\lambda) / \mathfrak{m}_{{{\boldsymbol \lambda}}} {\mathbb{W}}(\lambda) $ (known as the local Weyl module defined in [@CP01]) has a unique simple quotient $L({{\boldsymbol \lambda}})$ in $U_{q} {\text{-}\mathrm{mod}_{\mathrm{fd}}}$. We define the category ${\mathcal{C}}_{Q}$ (resp. ${\mathcal{C}}_{Q, \beta}$ for each $\beta \in {\mathsf{Q}}^{+}$) to be the minimal Serre full subcategory of the category $U_{q} {\text{-}\mathrm{mod}_{\mathrm{fd}}}$ of finite-dimensional $U_{q}(L{\mathfrak{g}})$-modules containing the simple objects $\{ L({{\boldsymbol \lambda}}) \mid {{\boldsymbol \lambda}}\in {\mathscr{P}}^{+}_{0} \}$ (resp. $\{ L({\mathbf{m}}) \mid {\mathbf{m}}\in {\mathrm{KP}}(\beta) \}$), where ${\mathscr{P}}^{+}_{0} = \bigsqcup_{\beta \in {\mathsf{Q}}^{+}} {\mathrm{KP}}(\beta) \subset {\mathscr{P}}^{+}$ is as in Subsection \[Ssec:HL\]. \[Rem:Grotisom\] Let $G$ be a linear algebraic group whose Lie algebra is ${\mathfrak{g}}$ and $N$ be the maximal unipotent subgroup of $G$ corresponding to the positive roots. Hernandez-Leclerc [@HL15] proved that the category ${\mathcal{C}}_{Q}$ is a monoidal subcategory and there is an isomorphism from the complexified Grothendieck ring $K({\mathcal{C}}_{Q})_{{\mathbb{C}}}$ to the coordinate ring ${\mathbb{C}}[N]$, which sends the classes of simple objects to the elements of the dual canonical basis bijectively. Actually, Hernandez-Leclerc established an isomorphism between their quantizations. We have a block decomposition ${\mathcal{C}}_{Q} = \bigoplus_{\beta \in {\mathsf{Q}}^{+}} {\mathcal{C}}_{Q, \beta}$ satisfying ${\mathcal{C}}_{Q, \beta} \otimes {\mathcal{C}}_{Q, \beta^{\prime}} \subset {\mathcal{C}}_{Q, \beta + \beta^{\prime}}$ (see [@Fujita17 Section 2.6]). This decomposition corresponds to the weight decomposition ${\mathbb{C}}[N] = \bigoplus_{\beta \in {\mathsf{Q}}^{+}}{\mathbb{C}}[N]_{\beta}$. The isomorphism $ {\mathfrak{M}}_{0, \beta}^{\bullet} \cong E_{\beta}$ in Theorem \[Thm:HL\] was originally established in order to give a geometric interpretation to the isomorphism $K({\mathcal{C}}_{Q, \beta})_{{\mathbb{C}}} \cong {\mathbb{C}}[N]_{\beta}$. Now we fix an element $\beta \in {\mathsf{Q}}^{+}$. In Subsection \[Ssec:HL\], we defined the graded quiver variety ${\mathfrak{M}^{\bullet}}_{\beta}$ with a canonical ${\mathbb{G}}_{\beta}$-equivariant proper morphism $\pi_{\beta}: {\mathfrak{M}^{\bullet}}_{\beta} \to E_{\beta}$, which is obtained from $\pi : {\mathfrak{M}}(\lambda) \to {\mathfrak{M}}_{0}(\lambda)$ with $\lambda = {\mathsf{cl}}({{\boldsymbol \lambda}}_{\beta})$ by taking the fixed locus with respect to the action of the $1$-dimensional torus $T_{\beta} \subset {\mathbb{G}}_{\beta} \subset {\mathbb{G}}(\lambda)$. We form the Steinberg type variety ${Z^{\bullet}}_{\beta} := {\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathfrak{M}^{\bullet}}_{\beta} = Z(\lambda)^{T_{\beta}}$. Let ${\mathfrak{r}}_{\beta}$ be the kernel of the restriction $R({\mathbb{G}}_{\beta})\otimes_{A} {\Bbbk}\to R(T_{\beta}) \otimes_{A}{\Bbbk}= {\Bbbk}$. Note that the decomposition (\[Eq:group\]) ${\mathbb{G}}_{\beta} \cong G_{\beta} \times T_{\beta}$ yields an isomorphism $$K^{{\mathbb{G}}_{\beta}}(X)\otimes_{A} {\Bbbk}\cong K^{G_{\beta}}(X)_{{\Bbbk}}$$ for any ${\mathbb{G}}_{\beta}$-variety $X$ with a trivial $T_\beta$-action. In particular, we have an isomorphism $R({\mathbb{G}}_{\beta})\otimes_{A} {\Bbbk}\cong R(G_{\beta})_{{\Bbbk}}$ of ${\Bbbk}$-algebras, via which the maximal ideal ${\mathfrak{r}}_{\beta} \subset R({\mathbb{G}}_{\beta})\otimes_{A} {\Bbbk}$ corresponds to the augmentation ideal $I \subset R(G_{\beta})_{{\Bbbk}}$. Therefore we have an isomorphism $$\label{Eq:complG} \left[ K^{{\mathbb{G}}_{\beta}}(X)\otimes_{A} {\Bbbk}\right]_{{\mathfrak{r}}_{\beta}}^{\wedge} \cong {\widehat{K}}^{G_{\beta}}(X)_{{\Bbbk}},$$ where $[ - ]_{{\mathfrak{r}}_{\beta}}^{\wedge}$ denotes the ${\mathfrak{r}}_{\beta}$-adic completion. We define the ${\Bbbk}$-algebra homomorphism $\widehat{\Phi}_{\beta} : U_{q}(L{\mathfrak{g}}) \to {\widehat{K}}^{G_{\beta}} (Z^{\bullet}_{\beta})_{{\Bbbk}}$ as the following composition: $$\begin{aligned} U_{q}(L{\mathfrak{g}}) & \xrightarrow{\Phi_{\lambda}} K^{{\mathbb{G}}(\lambda)}(Z(\lambda)) \otimes_{A} {\Bbbk}\\ &\to K^{{\mathbb{G}}_{\beta}}(Z(\lambda)) \otimes_{A} {\Bbbk}&& \text{(restriction to ${\mathbb{G}}_{\beta} \subset {\mathbb{G}}(\lambda)$)} \\ & \to \left[ K^{{\mathbb{G}}_{\beta}}(Z(\lambda))\otimes_{A} {\Bbbk}\right]_{{\mathfrak{r}}_{\beta}}^{\wedge} && \text{(${\mathfrak{r}}_{\beta}$-adic completion)} \\ &\cong \left[ K^{{\mathbb{G}}_{\beta}}({Z^{\bullet}}_{\beta})\otimes_{A} {\Bbbk}\right]_{{\mathfrak{r}}_{\beta}}^{\wedge} && \text{(localization theorem)} \\ &\cong {\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}}. && \text{(isomorphism (\ref{Eq:complG}))}\end{aligned}$$ \[Thm:mine\] The pull-back along the homomorphism $\widehat{\Phi}_{\beta} : U_{q}(L {\mathfrak{g}}) \to {\widehat{K}}^{G_{\beta}} ({Z^{\bullet}}_{\beta})_{{\Bbbk}}$ induces an equivalence $$\widehat{\Phi}_{\beta}^{*}: {\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}} {\text{-}\mathrm{mod}_{\mathrm{fd}}}\xrightarrow{\simeq} {\mathcal{C}}_{Q, \beta}$$ between the category ${\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}} {\text{-}\mathrm{mod}_{\mathrm{fd}}}$ of finite-dimensional ${\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}}$-modules and the category ${\mathcal{C}}_{Q, \beta} \subset U_{q}{\text{-}\mathrm{mod}_{\mathrm{fd}}}$. The next proposition is a counterpart of Proposition \[Prop:RRflag\]. \[Prop:RRquiver\] The Riemann-Roch homomorphism gives an isomorphism of $\widehat{R}(G_{\beta})_{{\Bbbk}}$-algebras: $${\mathrm{RR}}^{G_{\beta}} : {\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}} \xrightarrow{\cong} H_{*}^{G_{\beta}}({Z^{\bullet}}_{\beta}, {\Bbbk})^{\wedge}.$$ As in the proof of Proposition \[Prop:RRflag\], it suffices to prove that the equivariant Chern character map $({\mathrm{ch}}^{G_{\beta}})^{{\mathfrak{M}^{\bullet}}_{\beta} \times {\mathfrak{M}^{\bullet}}_{\beta}}_{Z^{\bullet}_{\beta}}: {\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}} \to H_{*}^{G_{\beta}}({Z^{\bullet}}_{\beta}, {\Bbbk})^{\wedge}$ is an isomorphism. Note that the $G_{\beta}$-orbit stratification (\[Eq:decE\]) yields a stratification of ${Z^{\bullet}}_{\beta}$: $$Z^{\bullet}_{\beta} = \bigsqcup_{{\mathbf{m}}\in {\mathrm{KP}}(\beta)} Z^{\bullet}_{\beta} |_{\mathbb{O}_{{\mathbf{m}}}}, \quad Z^{\bullet}_{\beta} |_{\mathbb{O}_{{\mathbf{m}}}} \cong G_{\beta} \times^{{\mathop{\mathrm{Stab}}\nolimits}_{G_{\beta}}(x_{{\mathbf{m}}})} \left( \pi_{\beta}^{-1}(x_{{\mathbf{m}}}) \times \pi_{\beta}^{-1}(x_{{\mathbf{m}}}) \right).$$ Fix a total ordering ${\mathrm{KP}}(\beta) = \{{\mathbf{m}}_{1}, {\mathbf{m}}_{2}, \ldots, {\mathbf{m}}_{s}\}$ such that we have $\mathbb{O}_{k} \subset \overline{\mathbb{O}}_{l}$ only if $k < l$. Set $Z_{\beta}^{k} := {Z^{\bullet}}_{\beta}|_{\mathbb{O}_{{\mathbf{m}}_k}}$ and $Z_{\beta}^{\le k} := \bigsqcup_{j \le k} Z_{\beta}^{j}$. Then the variety $Z_{\beta}^{\le k-1}$ is a closed subvariety of $Z_{\beta}^{\le k}$ whose complement is $Z_{\beta}^{k}$. By Proposition \[Prop:fiber\] and the reduction, we have $$\begin{aligned} K^{G_{\beta}}(Z^{k}_{\beta}) &\cong K^{G({\mathbf{m}}_k)}({\mathfrak{L}^{\bullet}}({\mathbf{m}}_k) \times {\mathfrak{L}^{\bullet}}({\mathbf{m}}_k)), \\ H_{*}^{G_{\beta}}(Z^{k}_{\beta}, {\Bbbk}) & \cong H_{*}^{G({\mathbf{m}}_k)}({\mathfrak{L}^{\bullet}}({\mathbf{m}}_k) \times {\mathfrak{L}^{\bullet}}({\mathbf{m}}_k), {\Bbbk})\end{aligned}$$ for each $k$. Then, using [@Nakajima01 Theorem 7.4.1], we can prove that the equivariant Chern character map gives an isomorphism ${\mathrm{ch}}^{G_{\beta}} : {\widehat{K}}^{G_{\beta}} (Z_{\beta}^{k})_{{\Bbbk}} \xrightarrow{\cong} H_{*}^{G_{\beta}}(Z_{\beta}^{k}, {\Bbbk})^{\wedge} $ for each $k$. Moreover, we obtain the following commutative diagram with exact rows for each $k$: $$\xy \xymatrix{ 0 \ar[r] & {\widehat{K}}^{G_{\beta}}(Z_{\beta}^{\le k-1})_{{\Bbbk}} \ar[r] \ar[d]^-{{\mathrm{ch}}^{G_{\beta}}} & {\widehat{K}}^{G_{\beta}}(Z_{\beta}^{\le k})_{{\Bbbk}} \ar[r] \ar[d]^-{{\mathrm{ch}}^{G_{\beta}}} & {\widehat{K}}^{G_{\beta}}(Z_{\beta}^{k})_{{\Bbbk}} \ar[r] \ar[d]^-{{\mathrm{ch}}^{G_{\beta}}} & 0 \\ 0 \ar[r] & H_{*}^{G_{\beta}}(Z_{\beta}^{\le k-1}, {\Bbbk})^{\wedge} \ar[r] & H_{*}^{G_{\beta}}(Z_{\beta}^{\le k}, {\Bbbk})^{\wedge} \ar[r] & H_{*}^{G_{\beta}}(Z_{\beta}^{k}, {\Bbbk})^{\wedge} \ar[r] & 0. } \endxy$$ By induction on $k$, the equivariant Chern character map gives an isomorphism ${\mathrm{ch}}^{G_{\beta}} : {\widehat{K}}^{G_{\beta}}(Z_{\beta}^{\le k})_{{\Bbbk}} \xrightarrow{\cong} H_{*}^{G_{\beta}}(Z_{\beta}^{\le k}, {\Bbbk})^{\wedge}$ for all $k$. We consider the proper push-forward $$\mathcal{L}_{\beta}^{\bullet} := (\pi_{\beta})_{*} \underline{{\Bbbk}}$$ of the trivial local system $ \underline{{\Bbbk}}$ on ${\mathfrak{M}^{\bullet}}_\beta$. By the decomposition theorem, we have $$\mathcal{L}_{\beta}^{\bullet} \cong \bigoplus_{{\mathbf{m}}\in {\mathrm{KP}}(\beta)} L_{{\mathbf{m}}}^{\bullet} \otimes_{{\Bbbk}} \mathcal{IC}_{{\mathbf{m}}} = \bigoplus_{{\mathbf{m}}\in {\mathrm{KP}}(\beta)} \bigoplus_{k \in {\mathbb{Z}}} L^{\bullet}_{{\mathbf{m}}, k} \otimes_{{\Bbbk}} \mathcal{IC}_{{\mathbf{m}}}[k],$$ where $L_{{\mathbf{m}}}^{\bullet} = \bigoplus_{k} L^{\bullet}_{{\mathbf{m}}, k}$ is a finite-dimensional graded ${\Bbbk}$-vector space, which is known to be non-zero for each ${\mathbf{m}}$ (see [@Nakajima01 Theorem 14.3.2]). Similarly to the previous subsection, we have a standard isomorphism of ${\Bbbk}$-algebras $$\label{Eq:isomG2} {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}^{\bullet}, \mathcal{L}_{\beta}^{\bullet}) \cong H_{*}^{G_{\beta}}({Z^{\bullet}}_{\beta}, {\Bbbk}),$$ which also induces an isomorphism between completions. We have the following isomorphisms of ${\Bbbk}$-algebras: $${\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}^{\bullet}, \mathcal{L}_{\beta}^{\bullet})^{\wedge} \cong H_{*}^{G_{\beta}}({Z^{\bullet}}_{\beta}, {\Bbbk})^{\wedge} \cong {\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}}.$$ Dynkin quiver type quantum affine Schur-Weyl duality {#Sec:SW} ==================================================== Geometric construction of a bimodule and a Morita equivalence ------------------------------------------------------------- We keep the notation in the previous sections. In particular, ${\Bbbk}= {\mathbb{Q}}(q).$ We fix an element $\beta = \sum_{i \in I} d_{i} \alpha_{i} \in {\mathsf{Q}}^{+}$ and put $\lambda := {\mathsf{cl}}({{\boldsymbol \lambda}}_{\beta}) \in {\mathsf{P}}^{+}$. From the two $G_{\beta}$-equivariant proper morphisms $ \pi_{\beta} : {\mathfrak{M}^{\bullet}}_{\beta} \to E_{\beta} $ and $ \mu_{\beta} : {\mathcal{F}}_{\beta} \to E_{\beta}, $ we form the fiber product $ {\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta} $. The convolution products make its completed $G_{\beta}$-equivariant $K$-group $ {\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta})_{{\Bbbk}}$ into a $({\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}}, {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}})$-bimodule. More precisely, the convolution products give ${\Bbbk}$-algebra homomorphisms $${\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}} \to {\mathop{\mathrm{End}}\nolimits}\left({\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta})_{{\Bbbk}}\right) \leftarrow {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}}^{\mathrm{op}},$$ whose images commute with each other. In the rest of this subsection, we prove that this bimodule induces a Morita equivalence. For a moment, we focus on a component ${\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}}$ for a fixed ${\mathbf{i}}\in I^{\beta}$. Using the isomorphism ${\mathcal{B}}_{{\mathbf{i}}} \cong G_{\beta}/B_{{\mathbf{i}}}$ with $B_{{\mathbf{i}}} = {\mathop{\mathrm{Stab}}\nolimits}_{G_{\beta}}(F^{\bullet}_{{\mathbf{i}}})$, we have $$\begin{aligned} {\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}} &\cong {\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} \left( G_{\beta} \times^{B_{{\mathbf{i}}}} \mathrm{pr}_{1}^{-1}(F_{{\mathbf{i}}}^{\bullet}) \right) \nonumber \\ &\cong G_{\beta} \times^{B_{{\mathbf{i}}}} \left( {\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} \mathrm{pr}_{1}^{-1}(F_{{\mathbf{i}}}^{\bullet}) \right), \label{Eq:isomatr}\end{aligned}$$ where $\mathrm{pr}_{1}$ is the projection ${\mathcal{F}}_{{\mathbf{i}}} \ni (F^{\bullet}, x) \mapsto F^{\bullet} \in {\mathcal{B}}_{{\mathbf{i}}}$. We define a $1$-parameter subgroup $ \rho_{{\mathbf{i}}} : {\mathbb{C}}^{\times} \to H_{\beta} $ by $\rho_{{\mathbf{i}}}(t) v_{w_{{\mathbf{i}}}(k)} := t^{k} v_{w_{{\mathbf{i}}}(k)}$ for $t \in {\mathbb{C}}^{\times}$. Note that this depends on the choice of $w_{{\mathbf{i}}} \in {\mathfrak{S}}_{d}$ fixed in Subsection \[Ssec:VV\]. We observe that $$\mathrm{pr}_{1}^{-1}(F_{{\mathbf{i}}}^{\bullet}) \cong \{ x \in E_{\beta} \mid x(F_{{\mathbf{i}}}^{k}) \subset F_{{\mathbf{i}}}^{k}, \, \forall k \} = \left\{ x \in E_{\beta} \; \middle| \; \lim_{t \to 0} \rho_{{\mathbf{i}}}(t) x = 0 \right\}.$$ Therefore we get $${\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} \mathrm{pr}_{1}^{-1}(F_{{\mathbf{i}}}^{\bullet}) \cong \left\{ x \in {\mathfrak{M}^{\bullet}}_{\beta}\; \middle| \; \lim_{t \to 0} \rho_{{\mathbf{i}}}(t) \pi_{\beta}(x) = 0 \right\}.$$ Since the morphism $\pi_{\beta} : {\mathfrak{M}^{\bullet}}_{\beta} \to E_{\beta}$ is the $T_{\beta}$-fixed part of $\pi : {\mathfrak{M}}(\lambda) \to {\mathfrak{M}}_{0}(\lambda)$, it is natural to consider the following subvariety of ${\mathfrak{M}}(\lambda)$: $$\widetilde{\mathfrak{Z}}(\lambda ; w_{{\mathbf{i}}}) := \left\{ x \in {\mathfrak{M}}(\lambda) \; \middle| \; \lim_{t \to 0} \rho_{{\mathbf{i}}}(t) \pi(x) = 0 \in {\mathfrak{M}}_{0}(\lambda) \right\},$$ which turns out to be the tensor product variety introduced by Nakajima [@Nakajima01t]. Since the subgroups $T_{\beta}$ and $\rho_{{\mathbf{i}}}({\mathbb{C}}^{\times})$ commute with each other, we have $$\label{Eq:obs2} {\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} \mathrm{pr}_{1}^{-1}(F_{{\mathbf{i}}}^{\bullet}) \cong \widetilde{\mathfrak{Z}}(\lambda ; w_{{\mathbf{i}}})^{T_{\beta}}.$$ Using (\[Eq:isomatr\]), (\[Eq:obs2\]) and the reduction, we obtain $$\begin{aligned} K^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}}) &\cong K^{H_{\beta}}({\widetilde{\mathfrak{Z}}}(\lambda ; w_{{\mathbf{i}}})^{T_{\beta}}), \label{Eq:redK} \\ H_{*}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}}, {\Bbbk}) &\cong H_{*}^{H_{\beta}}({\widetilde{\mathfrak{Z}}}(\lambda; w_{{\mathbf{i}}})^{T_{\beta}}, {\Bbbk}). \label{Eq:redH}\end{aligned}$$ \[Prop:chbimodule\] The $G_{\beta}$-equivariant Chern character map gives an isomorphism: $${\mathrm{ch}}^{G_{\beta}} : {\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}})_{{\Bbbk}} \xrightarrow{\cong} H_{*}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}}, {\Bbbk})^{\wedge}.$$ Thanks to (\[Eq:redK\]) and (\[Eq:redH\]), it is enough to show that the $H_{\beta}$-equivariant Chern character map $${\mathrm{ch}}^{H_{\beta}}: {\widehat{K}}^{H_{\beta}}({\widetilde{\mathfrak{Z}}}(\lambda ; w_{{\mathbf{i}}})^{T_{\beta}})_{{\Bbbk}} \to H_{*}^{H_{\beta}}({\widetilde{\mathfrak{Z}}}(\lambda, w_{{\mathbf{i}}})^{T_{\beta}}, {\Bbbk})^{\wedge}$$ is an isomorphism. This latter assertion follows from a $T_{\beta}$-fixed part analogue of [@Nakajima01t Theorem 3.10. (1)]. The $G_{\beta}$-equivariant Borel-Moore homology $H_{*}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta}, {\Bbbk})$ becomes a $(H_{*}^{G_{\beta}}({Z^{\bullet}}_{\beta}, {\Bbbk}), H_{*}^{G_{\beta}}({\mathcal{Z}}_{\beta}, {\Bbbk}))$-bimodule by the convolution products, similarly to the case of $K$-groups. On the other hand, the ${\mathop{\mathrm{Ext}}\nolimits}$-group ${\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}^{\bullet}, \mathcal{L}_{\beta})$ becomes a $( {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}^{\bullet}, \mathcal{L}_{\beta}^{\bullet}), {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}, \mathcal{L}_{\beta} ))$-bimodule by the Yoneda products. This bimodule ${\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}^{\bullet}, \mathcal{L}_{\beta})$ gives a Morita equivalence between ${\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}^{\bullet}, \mathcal{L}_{\beta}^{\bullet})$ and $ {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}, \mathcal{L}_{\beta})$ because $\mathcal{IC}_{{\mathbf{m}}}$ appears as a non-zero direct summand of both $\mathcal{L}_{\beta}$ and $\mathcal{L}^{\bullet}_{\beta}$ for each ${\mathbf{m}}\in {\mathrm{KP}}(\beta)$. Moreover, we have a standard isomorphism $$\label{Eq:isomG3} H_{*}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta}, {\Bbbk}) \cong {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}^{\bullet}, \mathcal{L}_{\beta})$$ \[Thm:Morita\] We have the following commutative diagram: $$\xy \xymatrix{ {\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}} \ar[r] \ar[d]^-{\cong}_{{\mathrm{RR}}^{G_{\beta}}} & {\mathop{\mathrm{End}}\nolimits}\left( {\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta})_{{\Bbbk}} \right) \ar[d]^-{\cong}_{{\mathrm{RR}}^{G_{\beta}}} & {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}}^{\mathrm{op}} \ar[l] \ar[d]^-{\cong}_{{\mathrm{RR}}^{G_{\beta}}} \\ H_{*}^{G_{\beta}}({Z^{\bullet}}_{\beta}, {\Bbbk})^{\wedge} \ar[r] \ar[d]^-{\cong}_-{(\ref{Eq:isomG2})} & {\mathop{\mathrm{End}}\nolimits}\left( H_{*}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta},{\Bbbk})^{\wedge} \right) \ar[d]^-{\cong}_-{(\ref{Eq:isomG3})} & H_{*}^{G_{\beta}}({\mathcal{Z}}_{\beta}, {\Bbbk})^{\wedge \mathrm{op}} \ar[d]^-{\cong}_-{(\ref{Eq:isomG1})} \ar[l] \\ {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}^{\bullet}, \mathcal{L}_{\beta}^{\bullet})^{\wedge} \ar[r] & {\mathop{\mathrm{End}}\nolimits}\left( {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}^{\bullet}, \mathcal{L}_{\beta})^{\wedge} \right) & {\mathop{\mathrm{Ext}}\nolimits}_{G_{\beta}}^{*}(\mathcal{L}_{\beta}, \mathcal{L}_{\beta})^{\wedge \mathrm{op}}, \ar[l] } \endxy$$ where each row denotes the bimodule structure defined above. In particular, the bimodule ${\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta} )_{{\Bbbk}}$ gives a Morita equivalence between two convolution algebras ${\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}}$ and ${\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}}$. The commutativity of the upper half (resp. lower half) of the diagram follows from Proposition \[Prop:RR\] (resp. an equivariant version of [@CG97 Theorem 8.6.7]). The left action of $U_{q}(L{\mathfrak{g}})$ ------------------------------------------- In this subsection, we fix ${\mathbf{i}}= (i_1, \ldots, i_d) \in I^{\beta}$ and investigate the $U_{q}(L{\mathfrak{g}})$-module structure of the pull-back $\widehat{\Phi}_{\beta}^{*} ( {\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}})_{{\Bbbk}})$. We use the following notation. For each $i \in I$, we define $\lambda_{i} := {\mathsf{cl}}(\phi(\alpha_i)) = \varpi_{j}$ and $a_{i} := q^{p}$ if $\phi(\alpha_{i}) = (j, p) \in \widehat{I}$. Recall from Theorem \[Thm:Nakajima\] that we have $$\label{Eq:isomRfund} {\mathop{\mathrm{End}}\nolimits}_{U_{q}} ( {\mathbb{W}}(\lambda_{i})) \cong R({\mathbb{G}}(\lambda_{i})) \otimes_{A}{\Bbbk}= R(G(\lambda_{i}))_{{\Bbbk}} \cong {\Bbbk}[z^{\pm 1}_{\lambda_{i}}],$$ where $z_{\lambda_i}$ denotes the class of the $1$-dimensional representation of $G(\lambda_{i}) = {\mathbb{C}}^{\times}$ of weight $1$. We recall some properties of the tensor product variety ${\widetilde{\mathfrak{Z}}}(\lambda ; w_{{\mathbf{i}}})$. Let $${\mathbb{H}}_{\beta} := H_{\beta} \times {\mathbb{C}}^{\times} \subset G_{\beta} \times {\mathbb{C}}^{\times} = {\mathbb{G}}_{\beta} \subset {\mathbb{G}}(\lambda)$$ be a maximal torus. By construction, the subvariety ${\widetilde{\mathfrak{Z}}}(\lambda ; w_{{\mathbf{i}}}) \subset {\mathfrak{M}}(\lambda)$ is stable under the action of ${\mathbb{H}}_{\beta}$. The convolution product makes the ${\mathbb{H}}_{\beta}$-equivariant $K$-group $K^{{\mathbb{H}}_{\beta}}(\widetilde{\mathfrak{Z}}(\lambda; w_{{\mathbf{i}}}))$ into a left $K^{{\mathbb{H}}_{\beta}}(Z(\lambda))$-module. Via the composition of the homomorphisms $$U_{q}(L{\mathfrak{g}}) \xrightarrow{\Phi_{\lambda}} K^{{\mathbb{G}}(\lambda)}(Z(\lambda))\otimes_{A} {\Bbbk}\to K^{{\mathbb{H}}_{\beta}}(Z(\lambda)) \otimes_{A} {\Bbbk},$$ where the latter one is the restriction to ${\mathbb{H}}_{\beta} \subset {\mathbb{G}}(\lambda)$, we regard the ${\mathbb{H}}_{\beta}$-equivariant $K$-group $K^{{\mathbb{H}}_{\beta}}(\widetilde{\mathfrak{Z}}(\lambda; w_{{\mathbf{i}}})) \otimes_{A} {\Bbbk}$ as a $U_{q}(L{\mathfrak{g}})$-module. \[Thm:tensor\] There is a $U_{q}(L{\mathfrak{g}})$-module isomorphism $$K^{{\mathbb{H}}_{\beta}}(\widetilde{\mathfrak{Z}}(\lambda; w_{{\mathbf{i}}})) \otimes_{A} {\Bbbk}\cong {\mathbb{V}}^{\otimes {\mathbf{i}}} := {\mathbb{W}}(\lambda_{i_1}) \otimes \cdots \otimes {\mathbb{W}}(\lambda_{i_d}),$$ where the action of $R({\mathbb{H}}_{\beta}) \otimes_{A} {\Bbbk}$ on the LHS is translated into the action on the RHS via the isomorphism $$\begin{aligned} \label{Eq:isomO} R({\mathbb{H}}_{\beta}) \otimes_{A} {\Bbbk}& \xrightarrow{\cong} \mathcal{O}_{{\mathbf{i}}} := {\Bbbk}[X_{1}^{\pm 1}, \ldots, X_{d}^{\pm 1}] \subset {\mathop{\mathrm{End}}\nolimits}_{U_{q}}({\mathbb{V}}^{\otimes {\mathbf{i}}}); \\ [{\mathbb{C}}v_{w_{{\mathbf{i}}}(k)}] & \mapsto X_{k}, \nonumber\end{aligned}$$ where we set $X_{k}:= z_{\lambda_{i_k}}$ using the notation in (\[Eq:isomRfund\]). The decomposition (\[Eq:group\]) ${\mathbb{G}}_{\beta} \cong G_{\beta} \times T_{\beta}$ induces the decomposition ${\mathbb{H}}_{\beta} \cong H_{\beta} \times T_{\beta}$ of the maximal torus ${\mathbb{H}}_{\beta}$. Similarly to the case of ${\mathbb{G}}_{\beta}$-equivariant $K$-groups in Subsection \[Ssec:Nakajima\], this decomposition yields a natural isomorphism $$K^{{\mathbb{H}}_{\beta}}(X) \otimes_{A} {\Bbbk}\cong K^{H_{\beta}}(X)_{{\Bbbk}}$$ for any ${\mathbb{H}}_{\beta}$-variety $X$ with a trivial $T_{\beta}$-action. When $X = \mathrm{pt}$, we have the following commutative diagram: $$\label{Diag:R} \xy \xymatrix{ R({\mathbb{H}}_{\beta})\otimes_{A} {\Bbbk}\ar[r]^-{\cong} \ar[d]^-{\cong}_{(\ref{Eq:isomO})} & R(H_{\beta})_{{\Bbbk}} \ar[d]^-{\cong} \\ \mathcal{O}_{{\mathbf{i}}} = {\Bbbk}[X_{1}^{\pm 1}, \ldots, X_{d}^{\pm1}] \ar[r]^-{\cong} & {\Bbbk}[y_{1}^{\pm 1}, \ldots, y_{d}^{\pm 1} ] 1_{{\mathbf{i}}}, } \endxy$$ where the bottom horizontal arrow sends the element $a_{i_k}^{-1}X_{k}$ to $y_{k} 1_{{\mathbf{i}}}$ for $1 \le k \le d$. Under this isomorphism, the maximal ideal ${\mathfrak{r}}^{\prime}_{\beta} \subset R({\mathbb{H}}_{\beta}) \otimes_{A} {\Bbbk}$ defined as the kernel of the restriction $R({\mathbb{H}}_{\beta})\otimes_{A}{\Bbbk}\to R(T_{\beta})\otimes_{A} {\Bbbk}= {\Bbbk}$ corresponds to the augmentation ideal of $R(H_{\beta})_{{\Bbbk}}$. Therefore we have a natural isomorphism $$\label{Eq:complH} \left[ K^{{\mathbb{H}}_{\beta}}(X) \otimes_{A} {\Bbbk}\right]^{\wedge}_{{\mathfrak{r}}^{\prime}_{\beta}} \cong {\widehat{K}}^{H_{\beta}}(X)_{{\Bbbk}},$$ where $[-]_{{\mathfrak{r}}^{\prime}_{\beta}}^{\wedge}$ denotes the ${\mathfrak{r}}^{\prime}_{\beta}$-adic completion. In particular, completing the diagram (\[Diag:R\]), we get $$\xy \xymatrix{ \left[ R({\mathbb{H}}_{\beta})\otimes_{A} {\Bbbk}\right]^{\wedge}_{{\mathfrak{r}}^{\prime}_{\beta}} \ar[r]^-{\cong} \ar[d]^-{\cong} & \widehat{R}(H_{\beta})_{{\Bbbk}} \ar[d]^-{\cong} \\ {\widehat{\mathcal{O}}}_{{\mathbf{i}}} := {\Bbbk}[\![X_{1} -a_{i_1}, \ldots, X_{d}-a_{i_d}]\!] \ar[r]^-{\cong} & {\Bbbk}[\![y_{1} -1 , \ldots, y_{d}-1]\!] 1_{{\mathbf{i}}}. } \endxy$$ \[Thm:left\] We have the following isomorphism of $U_{q}(L{\mathfrak{g}})$-modules: $$\widehat{\Phi}_{\beta}^{*} \left( {\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}})_{{\Bbbk}} \right) \cong {\widehat{V}}^{\otimes {\mathbf{i}}} := {\mathbb{V}}^{\otimes {\mathbf{i}}} \otimes_{\mathcal{O}_{{\mathbf{i}}}} {\widehat{\mathcal{O}}}_{{\mathbf{i}}}.$$ Actually, there is the following isomorphism: $$\begin{aligned} {\widehat{V}}^{\otimes {\mathbf{i}}} & \cong \left[ K^{{\mathbb{H}}_{\beta}}({\widetilde{\mathfrak{Z}}}(\lambda; w_{{\mathbf{i}}}))\otimes_{A} {\Bbbk}\right]_{{\mathfrak{r}}^{\prime}_{\beta}}^{\wedge} && (\text{Theorem~\ref{Thm:tensor}}) \\ & \cong \left[ K^{{\mathbb{H}}_{\beta}}({\widetilde{\mathfrak{Z}}}(\lambda ; w_{{\mathbf{i}}})^{T_{\beta}})\otimes_{A} {\Bbbk}\right]_{{\mathfrak{r}}^{\prime}_{\beta}}^{\wedge} && (\text{localization theorem})\\ & \cong {\widehat{K}}^{H_{\beta}}({\widetilde{\mathfrak{Z}}}(\lambda; w_{{\mathbf{i}}})^{T_{\beta}})_{{\Bbbk}} && (\text{isomorphism (\ref{Eq:complH})})\\ &\cong {\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}})_{{\Bbbk}}. && (\text{isomorphism (\ref{Eq:redK})})\end{aligned}$$ We need to show that this is a $U_{q}(L{\mathfrak{g}})$-homomorphism. By construction, the following diagram of ${\Bbbk}$-algebras commutes: $$\xy \xymatrix{ K^{{\mathbb{G}}(\lambda)}(Z(\lambda)) \otimes_{A} {\Bbbk}\ar[r] \ar[d] & \left[ K^{{\mathbb{G}}_{\beta}}({Z^{\bullet}}_{\beta}) \otimes_{A} {\Bbbk}\right]^{\wedge}_{{\mathfrak{r}}_{\beta}} \ar[r]^-{\cong}_-{(\ref{Eq:complG})} \ar[d] & {\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}} \ar[d] \\ K^{{\mathbb{H}}_{\beta}}(Z(\lambda)) \otimes_{A}{\Bbbk}\ar[r] & \left[ K^{{\mathbb{H}}_{\beta}}({Z^{\bullet}}_{\beta}) \otimes_{A} {\Bbbk}\right]^{\wedge}_{{\mathfrak{r}}_{\beta}^{\prime}} \ar[r]^-{\cong}_-{(\ref{Eq:complH})} & {\widehat{K}}^{H_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}}, } \endxy$$ where the vertical arrows denote the restrictions to the maximal tori. Moreover, by using an $H_{\beta}$-equivariant version of [@Nakajima01 Proposition 8.2.3], we can see that the following diagram also commutes: $$\xy \xymatrix{ K^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}} \otimes K^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}})_{{\Bbbk}} \ar[r]^-{*} \ar[d]_{(\text{restriction to $H_{\beta}$}) \otimes (\ref{Eq:redK}) } & K^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}})_{{\Bbbk}} \ar[d]^-{\cong}_-{(\ref{Eq:redK})} \\ K^{H_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}} \otimes K^{H_{\beta}}({\widetilde{\mathfrak{Z}}}(\lambda; w_{{\mathbf{i}}})^{T_{\beta}})_{{\Bbbk}} \ar[r]^-{*} & K^{H_{\beta}}({\widetilde{\mathfrak{Z}}}(\lambda; w_{{\mathbf{i}}})^{T_{\beta}})_{{\Bbbk}}, } \endxy$$ where the horizontal arrows denote the convolution products. From these commutative diagrams, combined with the definition of $\widehat{\Phi}_{\beta}$ and Theorem \[Thm:tensor\], we obtain the conclusion. The right action of $\widehat{H}_{Q}(\beta)$ -------------------------------------------- Summarizing the discussion so far, we have obtained a $(U_{q}(L{\mathfrak{g}}), \widehat{H}_{Q}(\beta))$-bimodule structure on the left $U_{q}(L{\mathfrak{g}})$-module $${\widehat{V}}^{\otimes \beta} := \bigoplus_{{\mathbf{i}}\in I^{\beta}} {\widehat{V}}^{\otimes {\mathbf{i}}}$$ such that the following diagram commutes: $$\xy \xymatrix{ U_{q}(L{\mathfrak{g}}) \ar[r] \ar[d]^-{\widehat{\Phi}_{\beta}} & {\mathop{\mathrm{End}}\nolimits}({\widehat{V}}^{\otimes \beta}) \ar[d]^-{\cong} & {\widehat{H}}_{Q}(\beta)^{\mathrm{op}} \ar[l]_-{\exists \psi} \ar[d]^-{\cong} \\ {\widehat{K}}^{G_{\beta}}({Z^{\bullet}}_{\beta})_{{\Bbbk}} \ar[r] & {\mathop{\mathrm{End}}\nolimits}\left( {\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{\beta})_{{\Bbbk}} \right) & {\widehat{K}}^{G_{\beta}}({\mathcal{Z}}_{\beta})_{{\Bbbk}}^{\mathrm{op}}. \ar[l] } \endxy$$ In this subsection, we describe the right action $ \psi : {\widehat{H}}_{Q}(\beta) \to {\mathop{\mathrm{End}}\nolimits}_{U_q}({\widehat{V}}^{\otimes \beta})^{\mathrm{op}} $ of the quiver Hecke algebra ${\widehat{H}}_{Q}(\beta)$ on the space ${\widehat{V}}^{\otimes \beta}$. For each ${\mathbf{i}}= (i_1, \ldots, i_d) \in I^{\beta}$, we set $$v_{{\mathbf{i}}} := (w_{\lambda_{i_1}} \otimes \cdots \otimes w_{\lambda_{i_d}} )\otimes 1 \in {\widehat{V}}^{\otimes {\mathbf{i}}} = ({\mathbb{W}}(\lambda_{i_1}) \otimes \cdots \otimes {\mathbb{W}}(\lambda_{i_d})) \otimes_{\mathcal{O}_{{\mathbf{i}}}} {\widehat{\mathcal{O}}}_{{\mathbf{i}}}.$$ \[Prop:hwspace\] The highest weight space $ \bigoplus_{{\mathbf{i}}\in I^{\beta}} {\widehat{\mathcal{O}}}_{{\mathbf{i}}} v_{{\mathbf{i}}} \subset {\widehat{V}}^{\otimes \beta} $ of weight $\lambda$ is stable under the right action of ${\widehat{H}}_{Q}(\beta)$. Moreover it is isomorphic to the completed polynomial representation $\widehat{P}_{\beta}$ defined in (\[Eq:Phat\]). Note that the connected component of the graded quiver variety ${\mathfrak{M}^{\bullet}}_{\beta} = {\mathfrak{M}}(\lambda)^{T_{\beta}}$ corresponding to the highest weight space is ${\mathfrak{M}}(0, \lambda)^{T_{\beta}} = \mathrm{pt}$ and hence ${\mathfrak{M}}(0, \lambda)^{T_{\beta}} \times_{E_{\beta}} {\mathcal{F}}_{\beta} = {\mathcal{B}}_{\beta}$. Therefore we have $$\bigoplus_{{\mathbf{i}}\in I^{\beta}} {\widehat{\mathcal{O}}}_{{\mathbf{i}}} v_{{\mathbf{i}}} \cong {\widehat{K}}^{G_{\beta}}({\mathfrak{M}}(0, \lambda)^{T_{\beta}} \times_{E_{\beta}} {\mathcal{F}}_{\beta})_{{\Bbbk}} \cong {\widehat{K}}^{G_{\beta}}({\mathcal{B}}_{\beta})_{{\Bbbk}} \cong \widehat{P}_{\beta}$$ as ${\widehat{H}}_{Q}(\beta)$-module, where the last isomorphism comes from (\[Eq:isomP\]) and (\[Eq:isomRRB\]). Henceforth, we normalize the isomorphism ${\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}})_{{\Bbbk}} \cong {\widehat{V}}^{\otimes {\mathbf{i}}}$ of $U_{q}(L{\mathfrak{g}})$-modules in Theorem \[Thm:left\] by multiplying the element of ${\widehat{\mathcal{O}}}_{{\mathbf{i}}}$ corresponding to the ratio $C_{{\mathbf{i}}}^{-1}$ of Todd classes defined in (\[Eq:const\]) for each ${\mathbf{i}}\in I^{\beta}$ so that the isomorphism $$\bigoplus_{{\mathbf{i}}\in I^{\beta}} {\widehat{\mathcal{O}}}_{{\mathbf{i}}} v_{{\mathbf{i}}} = \bigoplus_{{\mathbf{i}}\in I^{\beta}} {\Bbbk}[\![ X_{1}-a_{i_1}, \ldots, X_{d}-a_{i_d} ]\!] v_{{\mathbf{i}}} \xrightarrow{\cong} \widehat{P}_{\beta} = \bigoplus_{{\mathbf{i}}\in I^{\beta}} {\Bbbk}[\![x_{1}, \ldots, x_{d}]\!]1_{{\mathbf{i}}}$$ in Proposition \[Prop:hwspace\] above sends the element $v_{{\mathbf{i}}}$ to $1_{{\mathbf{i}}}$. Now we recall the normalized $R$-matrices. For any pair $(i_1, i_2) \in I^{2}$, we simplify $z_{k} := z_{\lambda_{i_k}}$ for $k=1,2$. Then it is known (see e.g. [@Kashiwara02 Section 8]) that there is a unique $( U_{q} \otimes {\Bbbk}[z_{1}^{\pm 1}, z_{2}^{\pm 1} ] )$-homomorphism, called the normalized $R$-matrix $$R_{i_1, i_2}^{\mathrm{norm}} : {\mathbb{W}}(\lambda_{i_1}) \otimes {\mathbb{W}}(\lambda_{i_2}) \to {\Bbbk}(z_{2}/z_{1})\otimes_{{\Bbbk}[(z_{2} / z_{1})^{\pm 1}]} \left( {\mathbb{W}}(\lambda_{i_2}) \otimes {\mathbb{W}}(\lambda_{i_1}) \right),$$ such that $R^{\mathrm{norm}}_{i_1, i_2} (w_{\lambda_{i_1}} \otimes w_{\lambda_{i_2}}) = w_{\lambda_{i_2}} \otimes w_{\lambda_{i_1}}$. The denominator of the normalized $R$-matrix $R^{\mathrm{norm}}_{i_1, i_2}$ is defined as the monic polynomial $d_{i_1, i_2}(u) \in {\Bbbk}[u]$ of the smallest degree among polynomials satisfying $${\mathop{\text{Im}}}R^{\mathrm{norm}}_{i_1, i_2} \subset d_{i_1, i_2}(z_{2} / z_{1})^{-1} \otimes \left( {\mathbb{W}}(\lambda_{i_2}) \otimes {\mathbb{W}}(\lambda_{i_1}) \right).$$ By [@Kashiwara02 Proposition 9.3], we have $$\label{Eq:denominator} d_{i_1, i_2}(1) \neq 0 .$$ Let ${\mathbb{K}}_{{\mathbf{i}}}$ be the fraction field of the ring ${\widehat{\mathcal{O}}}_{{\mathbf{i}}}$ for each ${\mathbf{i}}\in I^{\beta}$. It is known that the $U_{q} \otimes {\mathbb{K}}_{{\mathbf{i}}}$-module $${\widehat{V}}^{\otimes {\mathbf{i}}}_{{\mathbb{K}}} := {\mathbb{V}}^{\otimes {\mathbf{i}}}\otimes_{\mathcal{O}_{{\mathbf{i}}}} {\mathbb{K}}_{{\mathbf{i}}} ={\widehat{V}}^{\otimes {\mathbf{i}}} \otimes_{{\widehat{\mathcal{O}}}_{{\mathbf{i}}}} {\mathbb{K}}_{{\mathbf{i}}}$$ is irreducible (see e.g. [@Kashiwara02 Proposition 9.5]). For each $w \in {\mathfrak{S}}_{d}$, the ${\Bbbk}$-algebra isomorphism $$\varphi_{w}: {\widehat{\mathcal{O}}}_{{\mathbf{i}}} \xrightarrow{\cong} {\widehat{\mathcal{O}}}_{{\mathbf{i}}\cdot w}; \quad f(X_{1}, \ldots, X_{d}) \mapsto f^{w}(X_{1}, \ldots, X_{d}) := f(X_{w(1)}, \ldots, X_{w(d)})$$ induces an isomorphism ${\mathbb{K}}_{{\mathbf{i}}} \xrightarrow{\cong} {\mathbb{K}}_{{\mathbf{i}}\cdot w}$ of the fraction fields, which we denote by the same symbol $\varphi_{w}$. The pull-back $ \varphi_{w}^{*} {\widehat{V}}_{{\mathbb{K}}}^{\otimes {\mathbf{i}}\cdot w} $ is an irreducible $U_{q}\otimes {\mathbb{K}}_{{\mathbf{i}}}$-module. For each ${\mathbf{i}}\in I^{\beta}$ and $1 \le k < d$, we define the following non-zero $U_{q}\otimes {\mathbb{K}}_{{\mathbf{i}}}$-homomorphism $$R^{{\mathbf{i}}}_{k} := \left(1^{\otimes (k-1)} \otimes R_{i_k, i_{k+1}}^{\mathrm{norm}} \otimes 1^{\otimes(d-k-1)} \right) \otimes \varphi_{s_{k}} : {\widehat{V}}_{{\mathbb{K}}}^{\otimes {\mathbf{i}}} \to \varphi_{s_{k}}^{*} {\widehat{V}}^{\otimes {\mathbf{i}}\cdot s_{k}}_{{\mathbb{K}}}.$$ By the irreducibility, this is an isomorphism and we have $$\label{Eq:hcKhom} {\mathop{\mathrm{Hom}}\nolimits}_{U_{q}\otimes {\mathbb{K}}_{{\mathbf{i}}}} \left( {\widehat{V}}_{{\mathbb{K}}}^{\otimes {\mathbf{i}}}, \varphi_{s_{k}}^{*} {\widehat{V}}^{\otimes {\mathbf{i}}\cdot s_{k}}_{{\mathbb{K}}} \right) = {\mathbb{K}}_{{\mathbf{i}}} \cdot R^{{\mathbf{i}}}_{k}.$$ Let ${\widehat{V}}^{\otimes \beta}_{{\mathbb{K}}} := \bigoplus_{{\mathbf{i}}\in I^{\beta}} {\widehat{V}}^{\otimes {\mathbf{i}}}_{{\mathbb{K}}}$. We regard ${\widehat{V}}^{\otimes \beta} \subset {\widehat{V}}^{\otimes \beta}_{{\mathbb{K}}}$ naturally. \[Thm:main\] The right action of the quiver Hecke algebra ${\widehat{H}}_{Q}(\beta)$ on the space ${\widehat{V}}^{\otimes \beta}$ is given by the following formulas: $$\begin{aligned} \label{Eq:e} v \cdot e({\mathbf{i}}^{\prime}) & = \delta_{{\mathbf{i}}, {\mathbf{i}}^{\prime}} v \\ \label{Eq:x} v \cdot x_{k} & = \log(a^{-1}_{i_k}X_{k}) v \\ \label{Eq:tau} v \cdot \tau_{k} &= \begin{cases} \displaystyle ( \log(a_{i_k}^{-1}X_{k}) - \log(a_{i_{k+1}}^{-1}X_{k+1}) )^{-1} ( R^{{\mathbf{i}}}_{k}(v)- v ) & \text{if $i_k = i_{k+1}$}, \\ (\log(a_{i_k}^{-1}X_{k+1}) - \log(a_{i_{k+1}}^{-1}X_{k})) R^{{\mathbf{i}}}_{k}(v) & \text{if $i_k \leftarrow i_{k+1}$}, \\ R^{{\mathbf{i}}}_{k}(v) & \text{otherwise}, \end{cases}\end{aligned}$$ where $v \in {\widehat{V}}^{\otimes {\mathbf{i}}}$ with ${\mathbf{i}}= (i_{1}, \ldots, i_{d}) \in I^{\beta}$ and $\log(X) := \sum_{m = 1}^{\infty} (-1)^{m+1}(X -1 )^{m}/m. $ The first formula (\[Eq:e\]) is clear from Theorem \[Thm:VV\] (\[Thm:VV:elem\]) and the construction. To prove the second formula (\[Eq:x\]), we assume that the vector $v \in {\widehat{V}}^{\otimes {\mathbf{i}}}$ corresponds to an element $\zeta \in {\widehat{K}}^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}})_{{\Bbbk}}$ under the isomorphism in Theorem \[Thm:left\]. By Theorem \[Thm:VV\] (\[Thm:VV:elem\]), the right action of $e^{x_{k}} \in {\widehat{H}}_{Q}(\beta)$ on ${\widehat{V}}^{\otimes {\mathbf{i}}}$ corresponds to the convolution with the class $\Delta_{*}[\mathcal{L}_{{\mathbf{i}}}(k)] \in {\widehat{K}}^{G_{\beta}}({\mathcal{F}}_{{\mathbf{i}}} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}})_{{\Bbbk}}$ from the right, where $\mathcal{L}_{{\mathbf{i}}}(k)$ is the line bundle on ${\mathcal{F}}_{{\mathbf{i}}}$ defined in Subsection \[Ssec:VV\] and $\Delta : {\mathcal{F}}_{{\mathbf{i}}} \to {\mathcal{F}}_{{\mathbf{i}}} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}}$ is the diagonal embedding. By [@Nakajima01 Lemma 8.1.1], we have $ \zeta * (\Delta_{*}[\mathcal{L}_{{\mathbf{i}}}(k)]) = \zeta \otimes p_{2}^{*}[\mathcal{L}_{{\mathbf{i}}}(k)], $ where $p_{2} : {\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}} \to {\mathcal{F}}_{{\mathbf{i}}}$ is the second projection. The isomorphism (\[Eq:redK\]) translates the operation $- \otimes p_{2}^{*}[\mathcal{L}_{{\mathbf{i}}}(k)]$ on $K^{G_{\beta}}({\mathfrak{M}^{\bullet}}_{\beta} \times_{E_{\beta}} {\mathcal{F}}_{{\mathbf{i}}})$ into the multiplication of the element $y_{k} 1_{{\mathbf{i}}} \in R(H_{\beta})$ on $K^{H_{\beta}}({\widetilde{\mathfrak{Z}}}(\lambda; w_{{\mathbf{i}}})^{T_{\beta}})$. Thus we have $v \cdot e^{x_{k}} = (a_{i_{k}}^{-1} X_{k})v$ (see (\[Diag:R\])). Let us verify the third formula (\[Eq:tau\]). Let $\psi : {\widehat{H}}_{Q}(\beta) \to {\mathop{\mathrm{End}}\nolimits}_{U_{q}} ( {\widehat{V}}^{\otimes \beta})^{\mathrm{op}}$ be the structure morphism. First, we consider the case $i_k = i_{k+1}$. From the commutation relation between $e({\mathbf{i}}) \tau_k$ and $x_l$ in $H_{Q}(\beta)$, and the formula (\[Eq:x\]) for $\psi(x_{l})$ which we have proved in the previous paragraph, we see that $$(\mathcal{D} \psi(e({\mathbf{i}})\tau_{k}) + 1) f = f^{s_{k}}(\mathcal{D} \psi(e({\mathbf{i}})\tau_{k}) + 1)$$ holds in ${\mathop{\mathrm{End}}\nolimits}_{U_{q}} ({\widehat{V}}^{\otimes {\mathbf{i}}})$ for any $f \in {\widehat{\mathcal{O}}}_{{\mathbf{i}}}$, where we put $\mathcal{D} := \log(a^{-1}_{i_{k}}X_{k}) - \log(a_{i_{k+1}}^{-1}X_{k+1})$. In other words, the operator $\mathcal{D} \psi(e({\mathbf{i}})\tau_{k}) + 1$ belongs to $ {\mathop{\mathrm{Hom}}\nolimits}_{U_{q}\otimes {\widehat{\mathcal{O}}}_{{\mathbf{i}}} } ( {\widehat{V}}^{\otimes {\mathbf{i}}}, \varphi_{s_k}^{*}{\widehat{V}}^{\otimes {\mathbf{i}}}). $ Therefore it extends to an operator on the localizations. Namely, we can regard $$\mathcal{D} \psi(e({\mathbf{i}})\tau_{k}) + 1 \in {\mathop{\mathrm{Hom}}\nolimits}_{U_{q} \otimes {\mathbb{K}}_{{\mathbf{i}}} } \left( {\widehat{V}}^{\otimes {\mathbf{i}}}_{{\mathbb{K}}}, \varphi_{s_k}^{*}{\widehat{V}}^{\otimes {\mathbf{i}}}_{{\mathbb{K}}} \right) \cong {\mathbb{K}}_{{\mathbf{i}}} \cdot R^{{\mathbf{i}}}_{k},$$ where the last isomorphism is (\[Eq:hcKhom\]). By Proposition \[Prop:hwspace\] and the formulas in Theorem \[Thm:KL\], we see that $(\mathcal{D} \psi(e({\mathbf{i}})\tau_{k}) + 1) v_{{\mathbf{i}}} = v_{{\mathbf{i}}} = R^{{\mathbf{i}}}_{k}(v_{{\mathbf{i}}})$. Therefore we obtain $\mathcal{D} \psi(e({\mathbf{i}})\tau_{k}) + 1 = R^{{\mathbf{i}}}_{k}$ as an operator on ${\widehat{V}}^{\otimes {\mathbf{i}}}$. The case $i_{k} \neq i_{k+1}$ is easier. In this case, the commutation relation in $H_{Q}(\beta)$ and the formula (\[Eq:x\]) for $\psi(x_{l})$ show that the operator $\psi(e({\mathbf{i}}) \tau_{k})$ already belongs to ${\mathop{\mathrm{End}}\nolimits}_{U_{q}\otimes {\widehat{\mathcal{O}}}_{{\mathbf{i}}}} ( {\widehat{V}}^{\otimes {\mathbf{i}}}, \varphi_{s_{k}}^{*} {\widehat{V}}^{\otimes {\mathbf{i}}})$. Therefore it extends to an element in ${\mathop{\mathrm{Hom}}\nolimits}_{U_{q} \otimes {\mathbb{K}}_{{\mathbf{i}}}}( {\widehat{V}}^{\otimes {\mathbf{i}}}_{{\mathbb{K}}}, \varphi_{s_k}^{*}{\widehat{V}}^{\otimes {\mathbf{i}}}_{{\mathbb{K}}}).$ Then we proceed just as in the previous paragraph to obtain the desired formula (\[Eq:tau\]), taking Proposition \[Prop:hwspace\], the formulas in Theorem \[Thm:KL\] and (\[Eq:hcKhom\]) into consideration. \[Cor:KKKconj\] For any $i_1, i_2 \in I$, the order of zero of the denominator $d_{i_1, i_2}(u)$ at the point $u=a_{i_2} / a_{i_1}$ is at most one. Since we know (\[Eq:denominator\]), we may assume that $i_1 \neq i_2$. We consider a sequence ${\mathbf{i}}= (i_1, i_2) \in I^{\beta}$ with $\beta = \alpha_{i_1} + \alpha_{i_2}$. When $i_{1} \leftarrow i_{2}$, the formula (\[Eq:tau\]) tells us that the operator $(\log(a_{i_1}^{-1}z_{1}) - \log(a_{i_2}^{-1}z_{2})) R^{{\mathbf{i}}}_{1}$ belongs to ${\mathop{\mathrm{Hom}}\nolimits}_{U_{q}}({\widehat{V}}^{\otimes {\mathbf{i}}}, {\widehat{V}}^{\otimes {\mathbf{i}}\cdot s_{1}} )$, where we put $z_{k} = z_{\lambda_{i_k}}$ for $k=1,2$ as before. Notice that $$\log(a^{-1}_{i_1} z_{1}) - \log(a^{-1}_{i_2} z_{2}) \in ( {z_2}/{z_1} - a_{i_2} / a_{i_1} )\cdot {\widehat{\mathcal{O}}}_{{\mathbf{i}}}^{\times}.$$ Therefore we find that the order of zero of $d_{i_1, i_2}(u)$ at $u = a_{i_2} / a_{i_1}$ is at most one. For the other case $i_k \not \leftarrow i_{k+1}$, by the formula (\[Eq:tau\]), the operator $R^{{\mathbf{i}}}_{1}$ already belongs to ${\mathop{\mathrm{Hom}}\nolimits}_{U_{q}} ({\widehat{V}}^{\otimes {\mathbf{i}}}, {\widehat{V}}^{\otimes {\mathbf{i}}\cdot s_{1}} )$. Therefore the order of zero of $d_{i_1, i_2}(u)$ at $u = a_{i_2} / a_{i_1}$ is zero. \[Rem:KKK\] For each ${\mathbf{i}}\in I^{\beta}$, we define a topological ${\Bbbk}$-algebra automorphism $\sigma_{{\mathbf{i}}}$ of ${\widehat{\mathcal{O}}}_{{\mathbf{i}}}$ by setting $$\sigma_{{\mathbf{i}}}(\log(a_{i_k}^{-1}X_{k})) := a_{i_k}^{-1}X_{k} -1$$ for all $k$. This induces a $U_{q}(L{\mathfrak{g}})$-module automorphism $\sigma := \bigoplus_{{\mathbf{i}}\in I^{\beta}} (1 \otimes \sigma_{{\mathbf{i}}})$ on the module ${\widehat{V}}^{\otimes \beta}$. If we twist our right ${\widehat{H}}_{Q}(\beta)$-action by this automorphism $\sigma$ (i.e. we replace the structure map $\psi$ with $\sigma\psi(-) \sigma^{-1}$), we get a new right ${\widehat{H}}_{Q}(\beta)$-action on ${\widehat{V}}^{\otimes \beta}$ given by the following formulas: $$\begin{aligned} \label{Eq:e2} v \cdot e({\mathbf{i}}^{\prime}) & = \delta_{{\mathbf{i}}, {\mathbf{i}}^{\prime}} v \\ \label{Eq:x2} v \cdot x_{k} & = (a_{i_k}^{-1}X_{k} -1 ) v \\ \label{Eq:tau2} v \cdot \tau_{k} &= \begin{cases} \displaystyle ( a^{-1}_{i_k}X_{k} - a_{i_{k+1}}^{-1}X_{k+1})^{-1} ( R^{{\mathbf{i}}}_{k}(v)- v ) & \text{if $i_k = i_{k+1}$}, \\ (a_{i_k}^{-1}X_{k+1} - a_{i_{k+1}}^{-1}X_{k}) R^{{\mathbf{i}}}_{k}(v) & \text{if $i_k \leftarrow i_{k+1}$}, \\ R^{{\mathbf{i}}}_{k}(v) & \text{otherwise}, \end{cases}\end{aligned}$$ where $v \in {\widehat{V}}^{\otimes {\mathbf{i}}}$ with ${\mathbf{i}}= (i_{1}, \ldots, i_{d}) \in I^{\beta}$. This new action is same as Kang-Kashiwara-Kims action in [@KKK18], [@KKK15]. \[Thm:equiv\] The formulas (\[Eq:e\]), (\[Eq:x\]) and (\[Eq:tau\]) (or the formulas (\[Eq:e2\]), (\[Eq:x2\]) and (\[Eq:tau2\])) define a structure of a $(U_{q}(L{\mathfrak{g}}), {\widehat{H}}_{Q}(\beta))$-bimodule on the left $U_{q}(L{\mathfrak{g}})$-module ${\widehat{V}}^{\otimes \beta}$. The functor $M \mapsto {\widehat{V}}^{\otimes \beta} \otimes_{{\widehat{H}}_{Q}(\beta)} M$ gives an equivalence of categories: $${\widehat{H}}_{Q}(\beta) {\text{-}\mathrm{mod}_{\mathrm{fd}}}\xrightarrow{\simeq} {\mathcal{C}}_{Q, \beta}.$$ This follows from the discussions in this subsection, Theorem \[Thm:mine\] and Theorem \[Thm:Morita\]. [10]{} V. Chari and A. Pressley. Quantum affine algebras and affine [H]{}ecke algebras. , 174(2):295–326, 1996. V. Chari and A. Pressley. Weyl modules for classical and quantum affine algebras. , 5:191–223, 2001. N. Chriss and V. Ginzburg. . Birkhauser Boston, Inc., Boston, MA, 1997. D. Edidin and W. Graham. Riemann-[R]{}och for equivariant [C]{}how groups. , 102(3):567–594, 2000. R. Fujita. Affine highest weight categories and quantum affine [S]{}chur-[W]{}eyl duality of [D]{}ynkin quiver types. preprint. arXiv:1710.11288. V. Ginzburg, N. Reshetikhin, and E. Vasserot. Quantum groups and flag varieties. In [*Mathematical aspects of conformal and topological field theories and quantum groups (South Hadley, MA, 1992)*]{}, number 175 in Contemp. Math., pages 101–130. Amer. Math. Soc., Providence, RI, 1994. D. Happel. , volume 119 of [*London Mathematical Society Lecture Note Series*]{}. Cambridge University Press, Cambridge, 1988. H. Hernandez and B. Leclerc. Quantum [G]{}rothendieck rings and derived [H]{}all algebras. , 701:77–126, 2015. S.-J. Kang, M. Kashiwara, and M. Kim. Symmetric quiver [H]{}ecke algebras and [R]{}-matrices of quantum affine algebras. , 211(2):591–685, 2018. S.-J. Kang, M. Kashiwara, and M. Kim. Symmetric quiver [H]{}ecke algebras and [R]{}-matrices of quantum affine algebras, [II]{}. , 164(8):1549–1602, 2015. S.-J. Kang, M. Kashiwara, M. Kim, and S.-j. Oh. Symmetric quiver [H]{}ecke algebras and [R]{}-matrices of quantum affine algebras, [III]{}. ,111(2):420–444, 2015. S.-J. Kang, M. Kashiwara, M. Kim, and S.-j. Oh. Symmetric quiver [H]{}ecke algebras and [R]{}-matrices of quantum affine algebras, [IV]{}. , 22(4):1987–2015, 2016. M. Kashiwara. Crystal bases of modified quantized enveloping algebra. , 73(2):383–413, 1994. M. Kashiwara. On level-zero representations of quantized affine algebras. , 112(1):117–175, 2002. S. Kato. oincare-[B]{}irkhoff-[W]{}itt bases and [K]{}hovanov-[L]{}auda-[R]{}ouquier algebras. , 163(3):619–663, 2014. M. Khovanov and A. Lauda. A diagrammatic approach to categorification of quantum groups. [I]{}. , 13:309–347, 2009. H. Nakajima. Quiver varieties and finite-dimensional representations of quantum affine algebras. , 14(1):145–238, 2000. H. Nakajima. Quiver varieties and tensor products. , 146(2):399–449, 2001. S.-j. Oh and T. Scrimshaw. Categorical relations between [L]{}anglands dual quantum affine algebras: [E]{}xceptional cases. preprint. arXiv:1802.09253. R. Rouquier. 2-[K]{}ac-[M]{}oody algebras. preprint. arXiv:0812.5023. M. Varagnolo and E. Vasserot. Canonical bases and [KLR]{}-algebras. , 659:67–100, 2011. [^1]: Department of Mathematics, Kyoto University, Oiwake Kita-Shirakawa Sakyo Kyoto 606-8502 JAPAN, `E-mail:[email protected]`
{ "pile_set_name": "ArXiv" }
--- abstract: 'We address the question whether features known from quantum chromodynamics (QCD) can possibly also show up in solid-state physics. It is shown that spinless fermions of charge $e$ on a checkerboard lattice with nearest-neighbor repulsion provide for a simple model of confined fractional charges. After defining a proper vacuum the system supports excitations with charges $\pm e/2$ attached to the ends of strings. There is a constant confining force acting between the fractional charges. It results from a reduction of vacuum fluctuations and a polarization of the vacuum in the vicinity of the connecting strings.' author: - 'F. Pollmann and P. Fulde' title: 'On confined fractional charges: a simple model' --- One of the most fascinating aspects of quantum chromodynamics (QCD) is the existence of particles with fractional charges, i.e., of quarks and their confinement [@Gross73; @Politzer73]. Of particular interest is that the confining force is constant, i.e., independent of the distance of the separated particles. The question may be posed whether this physical feature is restricted to QCD or whether it occurs in a similar form in other areas of physics, e.g., in solid-state theory. There are well known examples where parts of the above mentioned physics show up in solids. We mention here that in heavily doped polyacetylen excitations with fractional charges may exist [@Su81]. In two dimensions the fractional quantum Hall effect is another example for the appearance of fractionally charged excitations [@Laughlin83]. But in both cases there is no constant confining force. On the other hand, confinement is nearly realized, e.g., in an Ising antiferromagnet with a hole put into it. When this hole is moving with hopping matrix element $t$ it generates a string of frustrated spins and therefore a constant confining force - almost, one should add, because there exist so-called Trugman paths [@Trugman88] which enable the hole to avoid generating such string. The probability for those paths is small, though, i.e., the effective hopping matrix element is of order $t^5/J^4$ where $J$ is the coupling constant of interacting neighboring Ising spins. But in this example there are no fractional charges appearing. The aim of the present communication is to point out a simple Hamiltonian for spinless (or fully spin polarized) fermions of charge $e$ defined on a frustrated lattice, here a checkerboard lattice, which leads to excitations with fractional charges $\pm e/2$ and a constant confining force. The advantage of studying such a model system is that it enables us to describe in detail the origin of the confining energy and constant confining force. We consider the following Hamiltonian on a checkerboard lattice [@FuldeP02] $$H = -t \sum_{\langle ij \rangle} \left( c^\dagger_i c_j + h.c. \right) + V \sum_{\langle ij \rangle} n_i n_j \label{eq:HtVij}$$ where $c^\dagger_i(c_i)$ create (destroy) spinless fermions on site $i$ and $n_i = c^\dagger_i c_i$. The notation $\langle ij \rangle$ refers to neighboring sites. The checkerboard lattice can be considered as a projection of a pyrochlore lattice onto a plane. The latter consists of corner sharing tetrahedra and therefore each of its sites has six nearest neighbors. We are interested here in the limit $V \gg |t|$. ---------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- (a) ![(a) Ground-state configuration maximizing the number of flipable hexagons, one of which is explicitely shown. Neighboring occupied sites (dots) are connected by a solid line. (b) Two defects with charge $e/2$ of an arbitrarily chosen allowed configuration are connected by a string consisting of an odd number of particles, shown by the black line. \[fig1\]](squiggle.eps "fig:"){width="5cm"} (b) ![(a) Ground-state configuration maximizing the number of flipable hexagons, one of which is explicitely shown. Neighboring occupied sites (dots) are connected by a solid line. (b) Two defects with charge $e/2$ of an arbitrarily chosen allowed configuration are connected by a string consisting of an odd number of particles, shown by the black line. \[fig1\]](doped.eps "fig:"){width="5cm"} ---------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- First we want to define a vacuum state for our system. For that purpose we consider a lattice of $N$ sites with $N/2$ particles. When $t = 0$ the repulsion energy is minimized if every crisscrossed square is occupied by two particles (”tetrahedron” rule) [@Anderson56]. In that case the (classical) ground state is approximately $(\frac{4}{3})^{\frac{3}{2} N}$ - fold degenerate as known from the ice problem [@Pauling60]. One special of those configurations $|\phi_i^{(0)}\rangle$ is shown in Fig. \[fig1\]a. Here we have connected neighboring occupied sites by a solid line. Assuming periodic boundary conditions those lines form loops which are either contractible or go around the torus. The degeneracy of the ground states is partially lifted by the hopping term [@Runge04]. To lowest order in $t/V$ this occurs through ring hopping as described by the effective Hamiltonian $$\begin{aligned} H_{\mathrm{eff}} & = & \frac{12t^3}{V^2} \sum_{{\pspicture[0.35](0.4,0.2) \psset{linewidth=0.03,linestyle=solid} \pspolygon[](0,0.1)(0.1,0.0)(0.3,0)(0.4,0.1)(0.3,0.2)(0.1,0.2) \endpspicture\;}} c^+_{j_6} c^+_{j_4} c^+_{j_2} c_{j_5} c_{j_3} c_{j_1}~~~. \label{eq:HeffFermions} \end{aligned}$$ It acts on the subspace spanned by the different ground-state configurations and within that space on hexagons with sites $j_1, ... j_6$ which are alternating occupied and empty. One of those hexagons is indicated in Fig. \[fig1\]a. Note that ring exchange is quite common in He$^3$ (see Ref. [@Thouless65]) and in frustrated spin systems [@Diep05]. When matrix elements $\langle \phi_i^{(0)} |H_{\mathrm{eff}}| \phi_j^{(0)}\rangle$ are calculated it matters for the sign of the result whether the site in the center of the hexagon is occupied or empty. For that reason it is advantageous to rewrite $H_{\mathrm{eff}}$ in the form $$\begin{aligned} H_{\mbox{\tiny eff}} & = & -g\sum_{\{{\pspicture[0.35](0.4,0.2) \psset{linewidth=0.03,linestyle=solid} \pspolygon[](0,0.1)(0.1,0.0)(0.3,0)(0.4,0.1)(0.3,0.2)(0.1,0.2) \endpspicture\;}\}}\left(\Big | { \; \pspicture[0.35](1,0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[](0,0.25)(0.25,0)(0.75,0)(1,0.25) (0.75, 0.5) (0.25, 0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.0)(0.75,0)(0.5,0.25) \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.5)(0.75,0.5)(0.5,0.25) \psline[](0.5,0)(0.5,0.5) \psset{linewidth=0.08,linestyle=solid} \psdots[linecolor=black,dotsize=.15](0.25,0)(0.25,0.5) (1,0.25)(0.5,0.25) \endpspicture\;}\Big \rangle \Big \langle { \; \pspicture[0.35](1,0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[](0,0.25)(0.25,0)(0.75,0)(1,0.25) (0.75, 0.5) (0.25, 0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.0)(0.75,0)(0.5,0.25) \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.5)(0.75,0.5)(0.5,0.25) \psline[](0.5,0)(0.5,0.5) \psset{linewidth=0.08,linestyle=solid} \psdots[linecolor=black,dotsize=.15](0,0.25)(0.75,0) (0.75,0.5)(0.5,0.25) \endpspicture\;}\Big | -\Big | { \; \pspicture[0.35](1,0.5) \psdots[linecolor=black,dotsize=.15] (0.25,0)(0.25,0.5) (1,0.25) \psset{linewidth=0.03,linestyle=solid} \pspolygon[](0,0.25)(0.25,0)(0.75,0)(1,0.25) (0.75, 0.5) (0.25, 0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.0)(0.75,0)(0.5,0.25) \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.5)(0.75,0.5)(0.5,0.25) \psline[](0.5,0)(0.5,0.5) \psset{linewidth=0.08,linestyle=solid} \endpspicture\;}\Big \rangle \Big \langle { \; \pspicture[0.35](1,0.5) \psdots[linecolor=black,dotsize=.15](0,0.25)(0.75,0) (0.75,0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[](0,0.25)(0.25,0)(0.75,0)(1,0.25) (0.75, 0.5) (0.25, 0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.0)(0.75,0)(0.5,0.25) \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.5)(0.75,0.5)(0.5,0.25) \psline[](0.5,0)(0.5,0.5) \psset{linewidth=0.08,linestyle=solid} \endpspicture\;}\Big | +\mbox{H.c.}\right) \label{eq:C4_H_eff}\end{aligned}$$ with $g = 12 t^3/V^2$. The sum is over all horizontal and vertical hexagons of the system. The remaining notation is self-explanatory. Note that terms of order $t^2/V$ yield identical contributions for all ground states and therefore do not lift their degeneracy. Exact diagonalization of finite clusters with different geometries shows that within the quantum-mechanical ground state $| \psi_0 \rangle = \sum_i \alpha_i | \phi_i^{(0)} \rangle$ those $| \phi_i^{(0)} \rangle$ will have largest amplitudes $\alpha_i$ which contain the largest numbers of flipable hexagons. An investigation of the configurations $| \phi_i^{(0)} \rangle$ of finite clusters with up to 1000 sites shows that the ones with the largest number of flipable hexagons form squiggles. The latter have been first identified by Penc and Shannon [@Penc06]. Their unit cell has 40 sites. From translation and rotation of the unit cell a corresponding number of different configurations $|\phi_{i}^{(0)}\rangle$ can be generated which maximize the number $N_{\mbox{fl}}$ of flipable hexagons. One of them is shown explicitely in Fig. \[fig1\]a. These configurations correspond to two different charge orderings which are related by a rotation of $90^{\circ}$ around a symmetry axis. This is seen by counting the number of particles along the lines denoted by 1 and 2 in Fig. \[fig3\]a. The charge ratio in the two directions is $\langle n_{x}\rangle:\langle n_{y}\rangle=2:3$ and for the rotated configuration $3:2$. When in a small cluster with periodic boundary conditions the squiggles can not develop, configurations with maximum $N_{\mbox{fl}}$ have a different shape. For quantum mechanical calculations we have used a cluster of size $\sqrt{72}\times\sqrt{72}$ with periodic boundary conditions. From the numerical analysis of such a cluster we have determined the distribution of the weight of configurations with given $N_{\mbox{fl}}$. The result is shown in Fig. \[fig2\]. A single configurations $|\phi_{i}^{(0)}\rangle$ contributes to $|\psi_{0}\rangle$ the more, the more flipable hexagons it has. But due to their large number, the largest total contribution to the ground state of, e.g., a 72 site cluster comes from configurations $|\phi_{i}^{(0)}\rangle$ with 14 flipable hexagons, i.e., $N_{\mbox{fl}}=14.$ Since the squiggle phase cannot be realized in the considered cluster, we find a ratio $\langle n_{x}\rangle:\langle n_{y}\rangle=1:2$ and $2:1$ instead for the configurations which maximize $N_{\mbox{fl}}$. The Hamiltonian (\[eq:C4\_H\_eff\]) conserves that ratio. Since charge ordering is solely caused by the effective kinetic Hamiltonian (\[eq:C4\_H\_eff\]), we have here a nice example of order by disorder. The quantum mechanical ground state is therefore 2-fold degenerate. We define the vacuum by choosing one of the ground states, i.e., we are dealing with a symmetry broken (degenerate) vacuum. ![Average weights of configurations with $N_{\rm fl}$ flipable hexagons in the quantum mechanical ground state of a 72 site checkerboard cluster. \[fig2\]](average_weight.eps){width="7cm"} Next we place an additional particle onto an empty site of the checkerboard lattice. Hereby we select that particular ground state which in the presence of the added particle gives the lowest energy. As a consequence two neighboring tetrahedra have three particles attached to them. When the hopping term of Eq. (\[eq:HtVij\]) is applied the two tetrahedra separate. That does not cost any additional repulsive energy. Each of the two tetrahedra has a charge $e/2$ attached to it, if the charge of the added particle was $e$. Separating the two tetrahedra with three particles generates a string consisting of an odd number of sites which is linked to two loops with an even number of sites each. This is shown in Fig. \[fig1\]b for an arbitrarily chosen configuration. The fractional charges sit at those linking points. Next we determine the changes of the kinetic energy density due to the presence of the two charges $e/2$. This is done by keeping the two tetrahedra with the fractional charges fixed at ${\bf 0}$ and ${\bf r}$ and determining the ground state $\bar{\psi}_0 ({\bf 0},{\bf r})$ and its energy for a cluster of 72 sites. The latter can be decomposed into local contributions by calculating the following expectation value of the energy at site $i$ $$\begin{aligned} \epsilon_i& = & -\frac g 6\sum_{\{{\pspicture[0.35](0.4,0.2) \psset{linewidth=0.03,linestyle=solid} \pspolygon[](0,0.1)(0.1,0.0)(0.3,0)(0.4,0.1)(0.3,0.2)(0.1,0.2) \endpspicture\;}| i \in {\pspicture[0.35](0.4,0.2) \psset{linewidth=0.03,linestyle=solid} \pspolygon[](0,0.1)(0.1,0.0)(0.3,0)(0.4,0.1)(0.3,0.2)(0.1,0.2) \endpspicture\;}\}}\Big\langle \bar{\psi}_0(\mathbf 0,\mathbf r)\Big| \left(\Big | { \; \pspicture[0.35](1,0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[](0,0.25)(0.25,0)(0.75,0)(1,0.25) (0.75, 0.5) (0.25, 0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.0)(0.75,0)(0.5,0.25) \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.5)(0.75,0.5)(0.5,0.25) \psline[](0.5,0)(0.5,0.5) \psset{linewidth=0.08,linestyle=solid} \psdots[linecolor=black,dotsize=.15](0.25,0)(0.25,0.5) (1,0.25)(0.5,0.25) \endpspicture\;}\Big \rangle \Big \langle { \; \pspicture[0.35](1,0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[](0,0.25)(0.25,0)(0.75,0)(1,0.25) (0.75, 0.5) (0.25, 0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.0)(0.75,0)(0.5,0.25) \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.5)(0.75,0.5)(0.5,0.25) \psline[](0.5,0)(0.5,0.5) \psset{linewidth=0.08,linestyle=solid} \psdots[linecolor=black,dotsize=.15](0,0.25)(0.75,0) (0.75,0.5)(0.5,0.25) \endpspicture\;}\Big | -\Big | { \; \pspicture[0.35](1,0.5) \psdots[linecolor=black,dotsize=.15] (0.25,0)(0.25,0.5) (1,0.25) \psset{linewidth=0.03,linestyle=solid} \pspolygon[](0,0.25)(0.25,0)(0.75,0)(1,0.25) (0.75, 0.5) (0.25, 0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.0)(0.75,0)(0.5,0.25) \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.5)(0.75,0.5)(0.5,0.25) \psline[](0.5,0)(0.5,0.5) \psset{linewidth=0.08,linestyle=solid} \endpspicture\;}\Big \rangle \Big \langle { \; \pspicture[0.35](1,0.5) \psdots[linecolor=black,dotsize=.15](0,0.25)(0.75,0) (0.75,0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[](0,0.25)(0.25,0)(0.75,0)(1,0.25) (0.75, 0.5) (0.25, 0.5) \psset{linewidth=0.03,linestyle=solid} \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.0)(0.75,0)(0.5,0.25) \pspolygon[fillcolor=lightgray, fillstyle=solid](0.25,0.5)(0.75,0.5)(0.5,0.25) \psline[](0.5,0)(0.5,0.5) \psset{linewidth=0.08,linestyle=solid} \endpspicture\;}\Big | +\mbox{H.c.}\right)\Big | \bar{\psi}_0(\mathbf 0,\mathbf r) \Big \rangle~~~. \label{eq:epsig6}\end{aligned}$$ The sum is over all hexagons containing site $i$. The result is shown in Fig. \[fig3\]a,b. One notices a decrease in kinetic energy in the region between the two fractional charges, i.e., along the connecting string. The reason is that due to the topological changes caused by the generated string the number of flipable hexagons positioned between the fractional charges is decreased. This energy decrease is proportional to the length of the generated string and implies a constant confining force. The origin of that force is clear. It is due to a reduction of vacuum fluctuations caused by the topological changes in the vacuum when the fractional charges are separated. Consider a configuration with two defects of charge $e/2$ separated along a diagonal (see Fig. \[fig3\]b). Then we find that the maximal number of flipable hexagons is reduced by $2d$, where $d$ is the number of crisscrossed squares between the defects. We have also determined the density changes caused by the separation of the fractional charges (see Fig. \[fig3\]c,d). They are obtained from the ground state in the presence of the two defects $|\bar{\psi}_0({\bf 0},{\bf r})\rangle$ by simply calculating $$\delta n_i = \left< \bar{\psi}_0 ({\bf 0},{\bf r}) \mid n_i \mid \bar{\psi}_0 ({\bf 0},{\bf r}) \right> - \left< \psi_0 \mid n_i \mid \psi_0 \right>~~~. \label{eq:deltani}$$ ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- (a)![(a) and (b): local loss of kinetic energy due to a separation of two fractionally charged defects. Radii of the dots are proportional to the local energy loss. The positions of defects are shown by red colored crisscrossed square. (c) and (d): red and blue dots show an increase (decrease) of the local density (vacuum polarization). \[fig3\]](kiner_qm_0_1_delta_vacuum_x3_1.eps "fig:"){width="50mm"} (b)![(a) and (b): local loss of kinetic energy due to a separation of two fractionally charged defects. Radii of the dots are proportional to the local energy loss. The positions of defects are shown by red colored crisscrossed square. (c) and (d): red and blue dots show an increase (decrease) of the local density (vacuum polarization). \[fig3\]](kiner_qm_2_3_delta_vacuum_x3_1.eps "fig:"){width="50mm"} (c)![(a) and (b): local loss of kinetic energy due to a separation of two fractionally charged defects. Radii of the dots are proportional to the local energy loss. The positions of defects are shown by red colored crisscrossed square. (c) and (d): red and blue dots show an increase (decrease) of the local density (vacuum polarization). \[fig3\]](n_qm_0_1_delta_x08_1.eps "fig:"){width="50mm"} (d)![(a) and (b): local loss of kinetic energy due to a separation of two fractionally charged defects. Radii of the dots are proportional to the local energy loss. The positions of defects are shown by red colored crisscrossed square. (c) and (d): red and blue dots show an increase (decrease) of the local density (vacuum polarization). \[fig3\]](n_qm_2_3_delta_x08_1.eps "fig:"){width="50mm"} ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- While the total changes add up to zero, the vacuum is modified by the breaking up of the charge. It is polarized along the connecting string. The above considerations can be also formulated in terms of a dimer model [@Rokhsar88]. For that purpose one connects the centers of the crisscrossed squares ending up with a square lattice. If the site on a link is occupied a dimer is attached to it. The tetrahedron rule implies that each site of the square lattice is touched by two dimers. The same findings discussed here apply also when a particle is removed from the ground state or when a particle-hole excitation is generated out of the ground state. They remain nearly unmodified for the case of quarter filling. We also expect to find a similar picture for the kagomé lattice, another frustrated lattice in which the classical ground-state manifold can be described by a dimer model on a bipartite lattice. We have not discussed here the three dimensional pyrochlore structure where also fractional charges may occur [@FuldeP02] because the numerics is much more involved. Neither have we discussed possible experimental verification of fractional charges in frustrated lattices [@Pollmann06], e.g., in the rather common spinels (for the latter see, e.g., Ref. [@Tsuda00]). Instead we refer to the literature. Summarizing, we have found that a simple Hamiltonian (\[eq:HtVij\]) when applied to a half-filled frustrated lattice leads in the strong coupling limit to fractional charges $\pm e/2$ with a constant confining force between them. The latter is due to modifications of the symmetry broken vacuum between the charges, i.e., the vacuum fluctuations are diminished and the vacuum is polarized along the string connecting the fractional charges. Our findings suggest that a number of features known from QCD are also expected to occur in solid-state physics. Conversely, one would hope that by studying frustrated lattices or dimer models one might be able to obtain better insight into certain aspects of QCD. A crucial difference between the confinement in the introduced model and in QCD is the fact that we can adjust the confining force by varying the ratio $t/V$. For the considered limit $|t|\ll V$, the fractionally charged particles form bound pairs of large spatial extent [@Pollmann06]. Thus we expect that charge fractionalization has observable effects even if the fractionally charged particles are confined. For instance an exchange of fractionally charged particles might occur at finite doping. We want to thank Dr. Karlo Penc (Budapest) and Dr. Nic Shannon (Bristol) for sharing their own observations on the nature of the undoped ground state. Also discussions with Dr. Joseph Betouras (St. Andrews) on the mechanisms of confinement are gratefully acknowledged. [0]{} D. J. Gross and F. Wilczek, Phys. Rev. Lett. **30**, 1343 (1973) H. D. Politzer, Phys. Rev. Lett. **30**, 1346 (1973) W. P. Su and J. R. Schrieffer, Phys. Rev. Lett. **46**, 738 (1981) R. B. Laughlin, Phys. Rev. Lett. **50**, 1395 (1983) S. A. Trugman, Phys. Rev. B **37**, 1597 (1988) P. Fulde and K. Penc and N. Shannon, Ann. Phys. (Leipzig) **11**, 892 (2002) P. W. Anderson, Phys. Rev.**102**, 1008 (1956) L. Pauling, *The Nature of the Chemical Bond* (Cornell, Ithaca, 1960) E. Runge and P. Fulde, Phys. Rev. B **70**, 245113 (2004) D. J. Thouless, Proc. Phys. Soc. London **86**, 893 (1965) H. T. Diep, *Frustrated Spin Systems* (World Scientific, Singapore, 2005) K. Penc and N. Shannon, private communication and to be published (2006) D. S. Rokhsar and A. A. Kivelson, Phys. Rev. Lett **61**, 2376 (1988) F. Pollmann and P. Fulde and E. Runge, Phys. Rev. B, **73**, 125121 (2006) N. Tsuda and K. Nasu and A. Fujimori and K. Siratori, *Electronic Conduction in Oxides*, vol. 94 (Springer-Verlag, Berlin, 2000), 2nd ed.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We present extensive ultraviolet (UV) and optical photometry, as well as dense optical spectroscopy for type II Plateau (IIP) supernova [SN 2016X]{} that exploded in the nearby ($\sim$ 15 Mpc) spiral galaxy [UGC 08041]{}. The observations span the period from 2 to 180 days after the explosion; in particular, the *Swift* UV data probably captured the signature of shock breakout associated with the explosion of [SN 2016X]{}. It shows very strong UV emission during the first week after explosion, with contribution of $\sim$ 20 – 30% to the bolometric luminosity (versus $\lesssim$ 15% for normal SNe IIP). Moreover, we found that this supernova has an unusually long rise time of about 12.6 $\pm$ 0.5 days in the $R$ band (versus $\sim$ 7.0 days for typical SNe IIP). The optical light curves and spectral evolution are quite similar to the fast-declining type IIP object SN 2013ej, except that [SN 2016X]{} has a relatively brighter tail. Based on the evolution of photospheric temperature as inferred from the $Swift$ data in the early phase, we derive that the progenitor of [SN 2016X]{} has a radius of about 930 $\pm$ 70 . This large-size star is expected to be a red supergiant star with an initial mass of $\gtrsim$ 19 – 20 M$_{\odot}$ based on the mass $--$ radius relation of the Galactic red supergiants, and it represents one of the most largest and massive progenitors found for SNe IIP.' author: - | F. Huang,$^{1,2,3}$[^1] X.-F. Wang,$^{1}$[^2] G. Hosseinzadeh,$^{4,5}$ P. J. Brown,$^{6}$ J. Mo,$^{1}$ J.-J. Zhang,$^{3,7,8}$ K.-C. Zhang,$^{1}$ T.-M. Zhang,$^{9}$ D.-A. Howell,$^{4,5}$ I. Arcavi,$^{4,5,10}$ C. McCully,$^{4,5}$ S. Valenti,$^{11}$ L.-M. Rui,$^{1}$ H. Song,$^{1}$ D.-F. Xiang,$^{1}$ W.-X. Li,$^{1}$ H. Lin,$^{1}$ L.-F. Wang$^{6}$\ $^{1}$Physics Department and Tsinghua Centre for Astrophysics, Tsinghua University, Beijing, 100084, China\ $^{2}$Department of Astronomy, Shanghai Jiao Tong University, Shanghai, 200240, China\ $^{3}$Key Laboratory for the Structure and Evolution of Celestial Objects, Chinese Academy of Sciences, Kunming 650011, China\ $^{4}$Department of Physics, University of California, Santa Barbara, CA 93106-9530, USA\ $^{5}$Las Cumbres Observatory, 6740 Cortona Dr., Suite 102, Goleta, CA 93117-5575, USA\ $^{6}$George P. and Cynthia Woods Mitchell Institute for Fundamental Physics & Astronomy, Texas A&M University, Department of Physics and Astronomy, 4242 TAMU, College Station, TX 77843, USA\ $^{7}$Yunnan Observatories, Chinese Academy of Sciences, Kunming 650011, China\ $^{8}$Centre for Astronomical Mega-Science, Chinese Academy of Sciences, 20A Datun Road, Chaoyang District, Beijing, 100012, China\ $^{9}$Key Laboratory of Optical Astronomy, National Astronomical Observatories, Chinese Academy of Sciences, Beijing 100012, China\ $^{10}$Einstein Fellow\ $^{11}$Department of Physics, University of California, 1 Shields Avenue, Davis, CA 95616-5270, USA\ date: 'Accepted XXX. Received YYY; in original form ZZZ' title: 'SN 2016X: A Type II-P Supernova with A Signature of Shock Breakout from Explosion of A Massive Red Supergiant' --- \[firstpage\] supernovae: general $-$ supernovae: individual: [[SN 2016X]{}]{} $-$ galaxies: individual: [UGC 08041]{} Introduction ============ Type II supernovae (SNe) are the outcome of massive stars (with initial mass $\geq$ 8 ; [[e.g.]{}]{}  ) experiencing gravitational core collapse after energy exhaustion at the end of life. They are characterized by P-cygni profile of Balmer lines in the early optical spectra compared to type I SNe . Based on the behaviors of light curves, SNe II are further divided into two subclasses: those with a prolonged plateau lasting $\sim$ 100 days are called type IIP, while those with a linear decline trend after maximum belong to type IIL . Recently, statistical work with large samples from different surveys tends to favour for a continuum distribution of the observational properties of SNe II ([[e.g.]{}]{} @2014ApJ...786...67A [@2015ApJ...799..208S; @2016MNRAS.459.3939V]). As the most abundant sub-type, SNe IIP occupy about 70% of all observed SNe II in a volume-limited sample [@2011MNRAS.412.1441L]. The observed plateau in the light curve results from the propagation of a cooling and recombination wave through the SN envelope . The presence of prominent hydrogen lines indicates that they retain a significant fraction of hydrogen envelopes before explosion. Analysis of the archive images allow direct detections of the progenitors for a few SNe IIP, which are generally found to be red supergiant (RSG) stars with a mass range of 8.5–16.5 . The observational limit is lower than the prediction from theoretical models, e.g., 8 to 25 . This inconsistency might be somewhat related to the presence of substantial circumstellar dust around the RSGs, which could lead to the underestimate of luminosity and hence the initial mass of the progenitor stars [@2012ApJ...759L..13F; @2012ApJ...756..131V; @2014ApJ...787..139D]. SNe IIP show a large diversity in the observational properties, such as peak luminosity, plateau length, expansion velocity, and synthesized nickel mass [@2003ApJ...582..905H]. These are connected with the explosion mechanism and the physical characteristics of the progenitors such as mass, explosion energy, and initial radius [@2009ApJ...703.2205K; @2011ApJ...741...41P]. Dozens of SNe IIP have been extensively studied from the ultraviolet to the near-infrared wavelength, [[i.e.]{}]{} SN 2005cs [@2009MNRAS.394.2266P], SN 2009N [@2014MNRAS.438..368T], and SN 2013ej [@2014MNRAS.438L.101V; @2015ApJ...807...59H], which helps take a deep look into the observed diversity and the progenitor physics. [SN 2016X]{} provides another opportunity for such kind of study. [SN 2016X]{} (ASASSN-16at) was discovered by All Sky Automated Survey for SuperNovae (ASAS-SN) on 2016 Jan. 20.59 (UT dates are used throughout this paper) in the nearby SBd galaxy [UGC 08041]{} (z=0.004408 from NED) at a $V$-band magnitude of $\sim$15.1 mag. The J2000 coordinates of the SN are $\alpha$ = 12$^h$55$^m$15.50$^s$ and $\delta = +00{\mbox{$^\circ$}}05 \arcmin 59.7 \arcsec$, approximately 60$\arcsec\,$ south and 42$\arcsec\,$ east from the centre of [UGC 08041]{} [@2016ATel.8566....1B]. The last non-detection was reported on Jan. 18.35 with a limit of $V > 18.0$ mag, but it was detected on 2016 Jan. 19.49 at $V$ $\sim$ 16.6 mag and Jan. 19.50 at $V \sim$ 17.0 mag. We therefore adopt 2016 Jan. 18.9 (MJD = 57405.92 $\pm$ 0.57) as the explosion time. An optical spectrum obtained on Jan. 20.75 suggests that it is a young core-collapse SN [@2016ATel.8567....1H], while another spectrum obtained on Jan. 23.88 confirms that it is a type II-P SN [@2016ATel.8584....1Z]. @2016ATel.8588....1G reported the discovery of X-rays from [SN 2016X]{} with *Swift*, which indicates that [SN 2016X]{} may have experienced moderate interaction with circumstellar material or stellar wind at early phase. We therefore triggered an instant follow-up campaign to study the photometric and spectroscopic evolution of this young type II-P supernova. The distance to its host galaxy is estimated to be 15.2 Mpc (distance modulus $\mu$ = 30.91 $\pm$ 0.43 mag) using Tully-Fisher method [@2014MNRAS.444..527S], which is adopted throughout this work. In this paper, we present photometry and spectroscopy of the nearby type IIP [SN 2016X]{}. In Section \[sec:obs\], we describe the observations and data reduction process for photometric and spectroscopic data. In Section \[sec:lc\], we study the photometric behavior of [SN 2016X]{}. The spectroscopic evolution is presented in Section \[sec:spec\]. We discuss the explosion parameters and progenitor properties of [SN 2016X]{}  in Section \[sec:discuss\], and summarize our conclusions in Section \[sec:summ\]. Observations and Data Reduction {#sec:obs} =============================== Photometry {#subsec:phot} ---------- ### Ground-based Observation {#subsubsec:optobs} High-cadence, broad-band photometric data of [SN 2016X]{} was obtained in Johnson $UBV$ and Sloan $gri$ filters with the 1.0 m telescopes of Las Cumbres Observatory (LCO; @2013PASP..125.1031B), spanning from 2016 Jan. 21 to 2016 Jul. 6. We also used the 0.8 m Tsinghua University-NAOC telescope (TNT; @2008ApJ...675..626W [@2012RAA....12.1585H]) at Xinglong Observatory and the Lijiang 2.4 m telescope (LJT; @2015RAA....15..918F) of Yunnan Astronomical Observatories in China to collect photometry in Johnson-Cousin $UBVRI$ filters. The observations began on 2016 Jan. 23 and ended on 2016 Jun. 3. All data were pre-processed with standard <span style="font-variant:small-caps;">IRAF</span>[^3] routines, including the corrections for bias, overscan, flat-field, and cosmic-ray removal. For TNT and LJT data, instrumental magnitudes were determined using the point-spread function (PSF) photometry with the <span style="font-variant:small-caps;">SNOoPy</span> package[^4]. The LCO data were reduced using `lcogtsnpipe` [@2016MNRAS.459.3939V]. The colour terms and extinction coefficients were derived from observations of Landolt stars on photometric nights [@1992AJ....104..340L]. The photometric zeropoints were determined by comparing the magnitudes of 10 field stars (marked in Figure \[fig:finder\]) to the values transformed from the Sloan Digital Sky Survey (SDSS) Data Release 9 catalogue [@2012ApJS..203...21A] using the relation from @2008AJ....135..264C. The coordinates and magnitudes of the reference stars around [SN 2016X]{} are listed in Table \[tab:standstar\], and the final calibrated magnitudes of [SN 2016X]{} are presented in Table \[tab:phot\_tnt\]–\[tab:phot\_LCO\]. ![image](finder.eps){width="50.00000%"} ### Swift UVOT Observations {#subsubsec:uvobs} [SN 2016X]{} was also observed in the ultraviolet and optical bands with the Ultra-Violet/Optical Telescope (UVOT; [@2005SSRv..120...95R]) on board the *Swift* spacecraft [@2004ApJ...611.1005G]. The space-based observations were obtained in the $uvw2, uvm2, uvw1, u, b$, and $v$ filters, covering the period from 2016 Jan. 21 to 2016 Mar. 5, and these data were taken from the *Swift* Optical/Ultraviolet Supernova Archive[^5] (SOUSA; ). The data reduction is based on the method described in @2009AJ....137.4517B, including subtraction of the host galaxy count rates and usage of the revised UV zeropoints and time-dependent sensitivity loss from @2011AIPC.1358..373B. The UVOT magnitudes of [SN 2016X]{} are listed in Table \[tab:uvot\]. Spectroscopy {#subsec:spec} ------------ The spectroscopic observations of [SN 2016X]{} started on 2016 Jan. 20 and continued until 2016 Jun. 9, corresponding to $\sim$2 day to $\sim$140 days after the explosion. A total of 40 low-resolution optical spectra were collected using the LCO 2 m Faulkes Telescope North (FTN; with FLOYDS), the Lijiang 2.4 m telescope (with YFOSC; [@2016PASP..128k5005F]), and the Xinglong 2.16 m telescope (with BFOSC). A journal of spectroscopic observations is given in Table \[tab:spelog\]. The spectroscopic data were reduced in a standard manner under the <span style="font-variant:small-caps;">IRAF</span> environment. After bias and overscan corrections, flat-fielding and cosmic-ray removal, one dimensional spectra were extracted using the optimal extraction method [@1986PASP...98..609H]. The wavelength calibration was done using the Fe/Ar and Hg/Ar lamp spectra, and the fluxes were calibrated using spectrophotometric standards observed on the same night with the same instrumental set-up. FLOYDS spectra were reduced using the `floydsspec` pipeline. Photometric Evolution {#sec:lc} ===================== The light curves of [SN 2016X]{} in UV and optical bands are shown in Figure \[fig:lc\], ranging from 2 to 170 days after explosion. The UV luminosity rises to the peak in a short time, followed by a rapid decline. The optical light curves resemble the evolution of typical SNe IIP but with relatively faster declines during the plateau phase. We present detailed analysis in the following subsections. ![Light curves of [SN 2016X]{} in ultraviolet (UV) and optical bands. The insert is a zoom on the early-time UV light curve, with polynomial fitting to the data around maximum. Prominent UV emission is clearly seen at a few days before the primary UV peaks. The phase is given relative to the estimated explosion date, MJD = 57,405.92.[]{data-label="fig:lc"}](lc.eps){width="50.00000%"} Swift UV Light Curves {#subsec:uvlc} ---------------------- The *Swift* UVOT observations of [SN 2016X]{} were triggered immediately after its discovery. The very early light curves in the $uvw2$ and $uvm2$ bands show an initial decline before rising to the peak at t $\sim$ 5 days after explosion (see the insert panel of Figure \[fig:lc\]). This indicates that [SN 2016X]{} may have another UV peak within 1–2 days from the explosion, which could be due to the breakout of a blast shockwave through the progenitor star’s outer envelope after the core-collapse explosion [@1977ApJS...33..515F; @1978ApJ...223L.109K]. The observed UV trough might thus be associated with the cooling of shock breakout, when the temperature behind the shock is lower than that at the shock front [@2008Sci...321..223S]. Such a UV trough had ever been reported for two SNe IIP at relatively larger distances, [[i.e.]{}]{} SNLS-04D2dc [@2008Sci...321..223S] and SNLS-06D1jd [@2008ApJ...683L.131G]. By adopting a polynomial fitting to the early data, we obtained m$_{uvw2}$(max) = 12.79 mag on 4.53 d, m$_{uvm2}$(max) = 12.61 mag on 4.88 d, and m$_{uvw1}$(max) = 12.68 mag on 5.02 d relative to the explosion date. The rise time for the primary UV peaks is $\sim$ 2 days longer than that of SNLS-04D2dc and SNLS-06D1jd, indicating that [SN 2016X]{} may have a progenitor with a larger initial radius. After the maximum, the SN declines quickly in the *swift* $uvw2, uvm2, uvw1$, and $u$ bands, with a rate of 0.245 $\pm$ 0.012, 0.269 $\pm$ 0.022, 0.208 $\pm$ 0.027, 0.135 $\pm$ 0.011 mag d$^{-1}$, respectively. While the corresponding decay rate is 0.047 $\pm$ 0.011, 0.016 $\pm$ 0.014 mag d$^{-1}$ in *swift* $b$ and $v$ bands. Note that [SN 2016X]{} shows a faster decline in $uvm2$ than in $uvw2$ and $uvw1$, which is against the usual trend that the decay rate steepens at shorter wavelengths. This opposite trend is also seen in other SNe IIP ([[i.e.]{}]{} SN 2005cs), and it might be related to the fact that more Fe III and Fe II lines are concentrated within the $uvm2$ bandpass [@2007ApJ...659.1488B]. Figure \[fig:absuv\] shows *Swift* UVOT absolute light curves of [SN 2016X]{} and some well-observed SNe IIP. Extinction corrections have been applied to all of our objects. As it can be seen, [SN 2016X]{} lies on the bright side of SNe IIP, and it reached the UV maximum 2–3 days later than other objects with UV observations. After t $\approx$ 1 month from the peak, the UV light curves seem to flatten out especially in the uvw1 and uvw2 filters, and this is similarly seen in SN 2012aw, SN 2013ab, and SN 2013ej. At this phase, the UV emission becomes very weak and the photometry can be significantly affected by optical photons leaked out of the red tails of the UV filters [@2016AJ....152..102B]. ![Comparison of the UV absolute light curves of [SN 2016X]{} with a few well-observed SNe IIP. For the comparison SNe, the distance modulus $\mu$ in mag, and extinction $A_V$ in mag are listed in the brackets after the name of each SN sample, and the references are: SN 2005cs (29.26, 0.095), @2007ApJ...659.1488B, @2009MNRAS.394.2266P;SN 2006bp (31.22, 0.08), @2007ApJ...664..435I;SN 2012aw (29.98, 0.08), @2013ApJ...764L..13B; SN 2013ab (31.90, 0.14), @2015MNRAS.450.2373B;SN 2013ej (29.91, 0.19), @2015ApJ...807...59H; SN 2014cx (31.74, 0.31), @2016ApJ...832..139H.[]{data-label="fig:absuv"}](uvmcomp.eps){width="50.00000%"} Optical Light Curves {#subsec:lcs} -------------------- The overall evolution of the optical light curves of [SN 2016X]{} can be divided into four main phases: the rising phase ($\sim$ 15 days), the plateau phase ($\sim$ 90 days), the transitional phase ($\sim$ 100 days), and the nebular phase ($\ge$ 100 days). The densely sampled data obtained immediately after the explosion allow us to catch the rising evolution of [SN 2016X]{} in very early phase. Using polynomial fit to the observed data around the maximum light, we are able to estimate the dates of maximum light and the peak magnitudes in different filters. The results for the phases of maximum and peak magnitudes in different bands are listed in Table \[tab:photpara\]. After the maximum light, the $B$-band magnitude declines by $\sim$ 4.0 mag in 100 days, which is larger than the typical value for SNe IIP ([[i.e.]{}]{} ${\beta}^B_{100} < 3.5$ mag; ). The $V$-band declines by $\sim$ 0.8 mag from the peak brightness in the first 50 days after explosion, which is also larger than normal SNe IIP (i.e., $s$50$_V < 0.5$ mag; @2014MNRAS.445..554F). Moreover, there are a few luminous SNe IIP ([[e.g.]{}]{} SNe 2007od, 2007pk, 2009bw, 2009dd, and 2013ej) that are found to show similar large post-maximum magnitude declines [@2015MNRAS.448.2608V; @2015ApJ...807...59H]. This indicates that a larger V-band decline should be used to make a distinguish between SNe IIP and SNe IIL, or these fast-declining SNe IIP may actually represent a subclass linking normal SNe IIP and SNe IIL. From the end of the plateau phase, the SN starts a transitional phase with a very rapid flux drop. For example, the $V$-band magnitude drops by $\sim$ 2.0 mag during the phase from t$\approx$ +90 days to t$\approx$ +130 days. After t $\approx$ 110 days, the SN enters into the nebular phase powered by the radioactive decay (i.e., to ). The decline rates at this phase are estimated to be 0.79, 1.44, 1.22, and 1.14 mag (100d)$^{-1}$ in $BVRI$ bands, respectively. Rise time --------- The rise time is an important parameter to constrain the properties of progenitor and explosion physics of SNe, which is typically defined as the time between the explosion epoch and the maximum light. Following the definition by , we adopt the maximum-light date as the time when the $r/R$-band magnitude rises by less than 0.01 mag per day. Based on a sample of 20 SNe IIP and IIL, found that SNe II show a diversity of rise time, with an average value of 7.0 $\pm$ 0.3 days for SNe IIP. The rise time is found to depend more sensitively on the progenitor radius than the mass and explosion energy [@2011ApJ...728...63R]. On the other side, recent studies indicate that the rise time of SNe II only shows a weak correlation with their luminosities [@2016MNRAS.459.3939V; @2016ApJ...820...33R]. This is in contrast to previous conclusion that brighter SNe II tend to have longer rise time . Fitting a low-order polynomial to the data around maximum, we find that the $r$-band light curve has a rise time of 12.6 $\pm$ 0.5 days for [SN 2016X]{}, and an absolute peak magnitude of $-$17.00 $\pm$ 0.43 mag. Figure \[fig:trise\] shows the comparison of $r$/$R$-band light curves and rise time between [SN 2016X]{} and some SNe II with early photometry. One can see that [SN 2016X]{} has a longer rise time than typical SNe IIP, while the absolute magnitude at the end of rise follows the brighter-slower trend. The longer rise time of [SN 2016X]{} indicates that its initial radius should be larger than that of normal SNe IIP, as predicted by the fact that photons take longer time to reach the surface of exploding star. ![image](risecom.eps){width="46.00000%"} ![image](trise.eps){width="45.00000%"} Reddening and Colour Curves {#subsec:colo} --------------------------- The Galactic reddening along the line of sight to [SN 2016X]{} is $E(B-V)_{\rm MW}$ = 0.02 mag [@2011ApJ...737..103S]. The host galaxy reddening is estimated using the colour method raised by @2010ApJ...715..833O which assumes that the intrinsic $V-I$ colour is constant ([[i.e.]{}]{}, $(V-I)_0$ = 0.656 mag) toward the end of the plateau phase. Fitting the $V$-band light curve with Equation (4) from @2010ApJ...715..833O, we obtain the middle of the transition phase as $t_{\rm{PT}}$ = 95 d. Using the $V-I$ colour at 65 d and correcting for the Galactic reddening, we obtain ${A_v}(host)$ = 0.05$\pm$0.21 mag. Thus we adopt the extinction $E(B-V)_{\rm tot}$ = 0.04 mag for [SN 2016X]{}. In Figure \[fig:colour\], we show the reddening corrected $(U-B)_0, (B-V)_0, (V-R)_0$, and $(V-I)_0$ colour curves of [SN 2016X]{} together with those of a few comparison SNe IIP. The colour evolution of [SN 2016X]{} shows similar trend with that of other SNe IIP. At early time, the $(U-B)_0$ and $(B-V)_0$ colours are quite blue and they evolve towards redder colours rapidly as a result of faster expansion and cooling of the ejecta. In comparison, the $(V-R)_0$ and $(V-I)_0$ colours evolve more slowly with a rate of $<$ 0.5 mag in 30 days. During the plateau phase ($\sim$30–110) days, the $(U-B)_0$ and $(B-V)_0$ colours become progressively red by $\sim$ 1 mag as the cooling rate decreases, while $(V-R)_0$ and $(V-I)_0$ colours show little change. The $(B-V)_0$ colour shows a peak during the transitional phase around t $\sim$ +110 days, which is also visible in other SNe IIP. In the nebular phase ($>$ 120 days), the $B-V$ colour becomes gradually bluer, similar to that of SN 1999em and SN 2014cx. ![The colour-curve evolution of [SN 2016X]{}, along with that of other well-studied SNe IIP (SNe 1999em, 2004et, 2005cs, and 2014cx). All the colours have been corrected for both the Galactic and host-galaxy reddening.[]{data-label="fig:colour"}](colour.eps){width="50.00000%"} Bolometric Light Curve {#subsec:bolo} ---------------------- Due to the lack of near-infrared observations, we calculated the quasi-bolometric luminosity of [SN 2016X]{} following the same method as described in @2015ApJ...807...59H. After corrections for the line-of-sight extinction, the broadband magnitudes were converted into fluxes at the effective wavelength when $V$-band observations were available. The data in other bands, if not obtained, were estimated by interpolating the observations on adjacent nights. The spectral energy distribution (SED) were integrated, and the observed fluxes were converted to luminosity with the Tully-Fisher distance from [@2014MNRAS.444..527S]. Figure \[fig:bolo\] shows the quasi-bolometric UV+optical ($UBVRI$) light curve of [SN 2016X]{}, compared with that of some representative SNe IIP. The peak luminosity is estimated to be as log $L_{\rm bol}$ = 42.17 . Note that the calculations of the quasi-bolometric light curves still suffer large uncertainties in the distance modulus. The plateau luminosity of [SN 2016X]{} is not constant but shows a monotonic decline up to t $\sim$ 90 days after explosion, which is similar with SN 2004et and SN 2013ej. The decline rate during the plateau phase is faster than other normal SNe IIP but comparable to the fast-declining type IIP SN 2013ej. The tail luminosity is lower than that of comparison SNe IIP except for the sub-luminous SN 2005cs, indicating that a relatively small amount of   was synthesized in the explosion. Using the least-square fitting, the decline rate at the nebular phase is estimated to be 0.6 mag (100d)$^{-1}$. ![The quasi-bolometric light curve of [SN 2016X]{} compared with that of a few well-studied SNe IIP.[]{data-label="fig:bolo"}](bolo.eps){width="50.00000%"} For the bolometric luminosity, the UV flux has a significant contribution in the early time ($\le$ 30 d), as shown in Figure \[fig:uvratio\]. The UV contribution can reach $\gtrsim$30% for [SN 2016X]{}, which is much higher than other comparison SNe IIP (i.e., $\sim$15%). Prominent UV emission is also in agreement with the higher temperature and larger progenitor radius derived for [SN 2016X]{} in Section \[subsec:radi\]. After about one month, the UV contribution becomes marginally important for most SNe IIP when entering into the plateau phase. Note that the above calculations of UV fraction may suffer from the uncertainties in dust extinctions applied for different SNe IIP. ![Temporal evolution of the UV contribution of SN 2016X in the first two months, compared with some well-studied SNe IIP (including SNe 2005cs, 2007od, 2008in, 2012A, 2012aw, 2013ej, and 2014cx). The comparison data are extracted from Swift Optical/Ultraviolet Supernova Archive.[]{data-label="fig:uvratio"}](uvratiocom.eps){width="50.00000%"} Spectroscopic Analysis {#sec:spec} ====================== Evolution of Optical Spectra {#subsec:spe} ---------------------------- A total of 40 optical spectra of [SN 2016X]{} covering the phase from +2 d to +140 d after the explosion are displayed in Figure \[fig:spec\]. The phases marked in the plot are relative to the explosion date estimated in §3.1. All spectra have been corrected for the recession velocity of the host galaxy (1321$\pm$2 )[^6]. The main spectral features are identified in previous studies for SNe IIP [@2002PASP..114...35L; @2004MNRAS.347...74P], and are also marked in Fig \[fig:speccomp\]. The first spectrum, taken at less than 2 days after explosion, shows a featureless blue continuum, consistent with a very young event of core-collapse explosion. The blue continuum indicates that the photosphere has a temperature that is above 10$^4$ K. At t $\approx$ 2.6 d, shallow hydrogen Balmer lines, and He [i]{} 5876 lines with broad P-Cygni profiles become visible. The blue wing of  absorption indicates that the expansion velocity can reach up to $\sim$18,000 . A double P-Cygni absorption of  appears in the t=+8d spectrum (see Fig \[fig:speccomp\](a)), and disappears after one month since explosion. The high-velocity feature is also reported in other SNe IIP, which might be due to Si [ii]{} 6355. After two weeks since explosion (t $\ge$ 15 d), the He [i]{} feature vanishes and is replaced by Na [i]{} line at the similar position. Apart from hydrogen Balmer lines, O [i]{} 7774, Ca [ii]{} H $\&$ K (3934, 3968), Ca [ii]{} NIR triplet (8498, 8542, 8662), and Fe [ii]{} multiplets are also clearly seen in the spectra. During the photospheric phase, the spectra turn progressively redder, and a number of narrow metal lines (Fe [ii]{}, Ti [ii]{}, Sc [ii]{}, Ba [ii]{}, Mg [ii]{}, et al.) emerge in the spectra. These features grow progressively stronger and dominate the spectra over time. After $\sim$90 days, the continuum flattens, and the spectra become dominated by emission lines, meaning that the SN enters into the nebular phase. The  emission profile shows a weak asymmetric feature (also seen in Figure \[fig:speccomp\]c). The asymmetric feature has been commonly observed in a few SNe IIP ([[i.e.]{}]{}, SNe 1999em, 2004dj, and 2013ej), and might result from interaction with circumstellar medium, asymmetry in the line-emitting region [@2002PASP..114...35L], or bipolar  distribution in a spherical envelope [@2006AstL...32..739C]. The subsequent spectra show permitted lines due to metals, when the outer ejecta became optically thin. And the spectra are characterized by the presence of forbidden lines \[O [i]{}\] 6300, 6364 and \[Ca [ii]{}\] 7291, 7324. Comparison with Other SNe IIP ----------------------------- In Figure \[fig:speccomp\], we compare the spectra evolution of [SN 2016X]{} to a few other SNe IIP at similar phases, [[i.e.]{}]{} the early phase at one week, the plateau phase at 2 months, and the nebular phase at 4 months after explosion. [SN 2016X]{} shows similarities with these comparison SNe IIP (especially SN 2013ej and SN 2014cx) in the spectral evolution. In the early phase, the spectrum of [SN 2016X]{} shows weaker and broader profiles of Balmer lines and He [I]{} line compared to SN 1999em and SN 2005cs. During the plateau phase, the spectra of [SN 2016X]{} and the comparison objects are dominated by metal lines, including Fe [ii]{}, Ti [ii]{}, Sc [ii]{}, Ba [ii]{}, and Mg [ii]{} etc. The forbidden lines such as \[O I\] and \[Ca II\] emerge in the spectra when the SNe enter into the nebular phase. In comparison, SN 2005cs shows much narrower absorption features and redder continuum at this phase. ![Spectral comparison of [SN 2016X]{} with those of other well-studied SNe IIP at early (1 week), plateau (80 d), and nebular (130 d) phases.[]{data-label="fig:speccomp"}](speccom.eps){width="50.00000%"} Expansion Velocities {#subsec:vel} -------------------- The measurement of the ejecta velocities and comparison with that of other SNe IIP are presented in this subsection. The expansion velocities of hydrogen and metal lines are measured by using SPLOT in IRAF to locate the absorption minima. The upper panel of Figure \[fig:velo\] shows the line velocities of , , Fe II 5169, 5018 and 4924. During the first week after the explosion, the expansion velocity of hydrogen is above 10,000 and it declines very rapidly. Later on, the velocity then declines in an exponential trend over time. The velocity of Fe II lines, which is a good indicator of photospheric velocity, is always lower than that of hydrogen lines and it decreases below 3,000 after 90 d. This can be explained by that the Fe II lines are formed in the inner-layers with larger optical depths. In the lower panel of Figure \[fig:velo\], we compare the velocity evolution of Fe 5169 between [SN 2016X]{} and other SNe IIP. It is obvious that the velocity of [SN 2016X]{} higher than SN 1999em (by $\sim$ 1,000 ) and SN 2005cs (by $\sim$ 3,000 ), and close to SN 2013ej, SN 2004et, and SN 2014cx. ![Upper: The velocity evolution of  and Fe [ii]{} lines. Lower: Comparison of evolution of photospheric velocity (measured by Fe 5169) of [SN 2016X]{} with some well-studied type II-P SNe such as SN 1999em [@2002PASP..114...35L], SN 2004et [@2006MNRAS.372.1315S], SN 2005cs [@2009MNRAS.394.2266P], SN 2013ej [@2015ApJ...807...59H], and SN 2014cx [@2016ApJ...832..139H].[]{data-label="fig:velo"}](velocity.eps){width="8.5cm"} Discussion {#sec:discuss} ========== Nickel Mass {#subsec:nick} ----------- The amount of  synthesized in the explosion of SNe IIP can be estimated by the luminosity of their late-time light curves. In the nebular phase, the light curve is powered by the radioactive decay of  to  and  to , with $e$-folding time of 8.8 d and 111.26 d, respectively. The mass of  of SN 1987A has been accurately determined to be 0.075 $\pm$ 0.005  [@1996snih.book.....A]. We adopt a linear least square fit to the nebular luminosity during the phase 120–160 d, and obtain $L$([SN 2016X]{})/ $L$(SN 1987A) = 0.43 at 140 d, from which we derive $M(^{56}{\rm Ni}) = 0.032\pm$0.006 . Assuming that the $\gamma$ photons produced from the  to  are fully thermalized, the  mass can be also estimated from the tail luminosity. Using Equation (2) in @2003ApJ...582..905H, we estimate the  mass to be $M( ^{56}{\rm Ni}) = 0.034 \pm$ 0.005 . found a tight correlation between the  mass and a steepness parameter of the $V$-band light curve at the transitional phase. For [SN 2016X]{} we fit the $V$-band light curve and estimate the steepness parameter $S$ as 0.099 mag day$^{-1}$ at the epoch of inflection $t_i = 92$ day. Using the empirical relation ($\log\,M( ^{56}{\rm Ni}) = -6.2295\,S -0.8147$), the mass of  for [SN 2016X]{} is estimated to be 0.037$\pm$0.003 . This value is consistent with that derived from the tail luminosity. Therefore, the average value of  mass is taken as 0.034 $\pm$ 0.006 . Properties of Progenitor {#subsec:radi} ------------------------ For CC SNe, shortly after the shock breakout, the shock-heated stellar envelope cools down due to the outward expansion. The timescale of cooling depends mainly on the initial radius of the progenitor, opacity, and gas composition. And the early light curves of SNe are dominated by the radiation from the expanding envelope. Some simple analytic expressions have been developed to describe the properties of the emitted radiation and are used to constrain the progenitor radius ([[e.g.]{}]{} @2011ApJ...728...63R [@2011ApJ...729L...6C; @2017ApJ...838..130S]). For example, progenitors with larger radius (i.e., RSG with 500–1000 ) stay at higher temperature and cool down at a slower pace than those with smaller radius (i.e., BSG with 50–100 ), as indicated by the expression where $f_{\rho}$ represents density profile, $E_{51}$ is the energy in units of 10$^{51}$ erg, $R_{*,13}$ is the radius in units of $10^{13}$ cm, $\kappa_{0.34}$ is the opacity in units of 0.34 cm$^2$ g$^{-1}$, and $t_5$ is time in units of 10$^5$ s. Thanks to the timely follow-up observations from the *Swift* UVOT, we are able to better construct the spectral energy distribution and estimate the corresponding blackbody temperature (cooling phase of the shock breakout) for [SN 2016X]{} in the early phase. This allows us to constrain its progenitor radius by fitting to the temperature evolution. Adopting an optical opacity of 0.34 ${cm}^2 g^{-1}$ and a RSG density profile $f_{\rho}=0.13$ in the Eq.(13) from @2011ApJ...728...63R, we yield an initial radius of 860–990  for the progenitor of [SN 2016X]{} , as shown in Figure \[fig:radius\]. We also overplot the temperature evolution and the best-fit progenitor radius for SN 1987A, SN 2006bp, SN 2013ej, and SN 2014cx. One can see that [SN 2016X]{} has an apparently large progenitor in comparison with other SNe IIP. Using the SuperNova Explosion Code (SNEC, @2015ApJ...814...63M), @2016ApJ...829..109M find that the early properties of the light curves of SNe IIP depend sensitively on the radius of the progenitor star, with a relationship between the $g$-band rise time and the radius at the time of explosion (i.e., log $R [R_{\odot}]$ = 1.225 log $t_{\rm rise}$ \[day\] + 1.692). We also use this relation to estimate the size of the progenitor star. For [SN 2016X]{} , the $g$-band rise time is estimated as 10.60$\pm$0.40 days, which leads to an estimate of 890$\pm$40  for the progenitor of [SN 2016X]{} . This analysis, together with the result from shock breakout cooling, favours that [SN 2016X]{} has a larger progenitor with a radius up to $\sim$ 900-1000 . ![Radius estimates using the prescription from @2011ApJ...728...63R. The best fit for [SN 2016X]{} is 860 – 990 . Over-plotted are the comparison objects SN 1987A, 2006bp, 2013ej and 2014cx [@2014MNRAS.438L.101V; @2016ApJ...832..139H].[]{data-label="fig:radius"}](radius.eps){width="8.5cm"} Based on the RSG sample in the Milky Way and Magellanic Clouds (MC) [@2005ApJ...628..973L; @2006ApJ...645.1102L], @2015MNRAS.451.2212G found that there is a general tendency that the more massive RSG stars have larger radius sizes. A tight mass–radius relation can be obtained for the RSG stars in the Milky Way, [[i.e.]{}]{} $R/R_{\odot}=1.4(M/M_{\odot})^{2.2}$, as shown in Figure \[fig:mrad\] (see the dashed line). This relation gives a rough estimate of 18.5–19.7  for the progenitor of [SN 2016X]{}. In this plot, we also show the progenitor mass and radius estimated from photospheric cooling/hydrodynamic analysis and HST archive images for a sample of SNe IIP. We notice that mass and radius of these SNe IIP seem to follow that of the Galactic or MC RSGs, except that the hydrodynamic method gives a larger mass and a smaller radius for SN 2012aw. Table \[tab:massnrad\] listed the details of these estimates and the references. For comparison, we overplot the mass–radius relation derived from red supergiants in the MCs (see the dotted line). Given a radius, the star will have a larger mass for lower metallicity. This can be explained with that more metal-poor stars usually lose their mass at a lower efficiency. As the host galaxy of [SN 2016X]{} UGC 08041 is a late-type Sd galaxy and the sn locates at its outskirts, it is possible that the progenitor of [SN 2016X]{} has a relatively lower metallicity. Considering this effect and hence the possible mass loss of the progenitor star before the explosion, the mass range we derived for the progenitor of [SN 2016X]{} should be a lower limit. Along with SN 2012aw and SN 2012ec, the high mass derived for the progenitor of [SN 2016X]{} indicates that RSGs with an initial mass around or above 20.0  could lead to an explosion of type IIP SN. ![The progenitor mass and radius for a sample of SNe IIP. The blue dots represent the estimates using hydrodynamic modeling, while the black dots are results from analysis of the pre-explosion images (see the reference from Table \[tab:massnrad\]). Dashed lines represent the mass–radius relation derived from the Galactic (upper) and Magellanic-cloud (lower) RSGs, respectively [@2015MNRAS.451.2212G].[]{data-label="fig:mrad"}](mrad.eps){width="8.5cm"} Summary {#sec:summ} ======= In this paper, we present the ultraviolet/optical photometry and low resolution spectroscopic observations for the type IIP [SN 2016X]{} up to 180 days after explosion. The high-quality UV/optical data allow us to place interesting constraints on the observational properties of [SN 2016X]{} and its progenitor. A brief summary of our results are listed below. The Swift UVOT data reveals the presence of prominent UV emissions at just only 2 days before the primary UV peaks, which is very likely related to the shock breakout of very massive stars. For [SN 2016X]{} the UV contribution to the total flux can reach $\gtrsim$30% for [SN 2016X]{} in the early phase, while the typical value is $\sim$15%. In particular, this supernova is found to have a very long rise time before reaching the maximum light, i.e., 12.6$\pm$0.5 days in the $R$ band, in contrast to $\sim$7.0 days for normal SNe IIP. The photometric and spectral evolution is overall similar to SN 2013ej. Using the early-time temperature evolution inferred from the *Swift* UV photometry, we derived an initial radius of 860–990  for the progenitor of [SN 2016X]{}. The long $g$-band rise time of [SN 2016X]{} also indicates a large progenitor radius of $\sim$ 890  according to the rise time – radius relation from the SNEC. Based on the mass – radius relation of the Galactic RSG, we also obtain a rough mass estimate of 18.5–19.7  for the progenitor of [SN 2016X]{}, which provides further evidence that massive stars with an initial mass up to 19-20 could also produce an explosion of type IIP supernova. Acknowledgements {#acknowledgements .unnumbered} ================ We acknowledge the support of the staff of the Lijiang 2.4m and Xinglong 2.16m telescope. Funding for the LJT has been provided by Chinese Academy of Sciences and the People’s Government of Yunnan Province. The LJT is jointly operated and administrated by Yunnan Observatories and Centre for Astronomical Mega-Science, CAS. This work is supported by the National Natural Science Foundation of China (NSFC grants 11178003, 11325313, and 11633002). This work was also partially supported by the Collaborating Research Program (OP201702) of the Key Laboratory of the Structure and Evolution of Celestial Objects, Chinese Academy of Sciences. This work makes use of observations from Las Cumbres Observatory. DAH, CM, and GH are supported by the US National Science Foundation grant 1313484. Support for IA was provided by NASA through the Einstein Fellowship Program, grant PF6-170148. The work made use of Swift/UVOT data reduced by P. J. Brown and released in the Swift Optical/Ultraviolet Supernova Archive (SOUSA). SOUSA is supported by NASA’s Astrophysics Data Analysis Program through grant NNX13AF35G. J.-J. Zhang is supported by the NSFC (grants 11403096, 11773067), the Key Research Program of the CAS (Grant NO. KJZD-EW- M06), the Youth Innovation Promotion Association of the CAS, and the CAS “Light of West China” Program. T.-M. Zhang is supported by the NSFC (grants 11203034). [99]{} Ahn, C. P., Alexandroff, R., Allende Prieto, C., et al. 2012, , 203, 21 Anderson, J. P., Gonz[á]{}lez-Gait[á]{}n, S., Hamuy, M., et al. 2014, , 786, 67 Arnett, D. 1996, Supernovae and Nucleosynthesis: An Investigation of the History of Matter, from the Big Bang to the Present, by D. Arnett. Princeton: Princeton University Press, 1996. Barbarino, C., Dall’Ora, M., Botticella, M. T., et al. 2015, , 448, 2312 Barbon, R., Ciatti, F., & Rosino, L. 1979, , 72, 287 Bayless, A. J., Pritchard, T. A., Roming, P. W. A., et al. 2013, , 764, L13 Bock, G., Shappee, B. J., Stanek, K. Z., et al. 2016, The Astronomer’s Telegram, 8566 Bose, S., Valenti, S., Misra, K., et al. 2015, , 450, 2373 Breeveld, A. A., Landsman, W., Holland, S. T., et al. 2011, American Institute of Physics Conference Series, 1358, 373 Brown, P. J., Dessart, L., Holland, S. T., et al. 2007, , 659, 1488 Brown, P. J., Holland, S. T., Immler, S., et al. 2009, , 137, 4517 Brown, T. M., Baliber, N., Bianco, F. B., et al. 2013, , 125, 1031 Brown, P. J., Breeveld, A. A., Holland, S., Kuin, P., & Pritchard, T. 2014, , 354, 89 Brown, P. J., Breeveld, A., Roming, P. W. A., & Siegel, M. 2016, , 152, 102 Chevalier, R. A., & Irwin, C. M. 2011, , 729, L6 Chonis, T. S., & Gaskell, C. M. 2008, , 135, 264 Chugai, N. N. 2006, Astronomy Letters, 32, 739 Dall’Ora, M., Botticella, M. T., Pumo, M. L., et al. 2014, , 787, 139 Elmhamdi, A., Chugai, N. N., & Danziger, I. J. 2003, , 404, 1077 Ekstr[ö]{}m, S., Georgy, C., Eggenberger, P., et al. 2012, , 537, A146 Falk, S. W., & Arnett, W. D. 1977, , 33, 515 Fan, Y.-F., Bai, J.-M., Zhang, J.-J., et al. 2015, Research in Astronomy and Astrophysics, 15, 918 Fan, Z., Wang, H., Jiang, X., et al. 2016, , 128, 115005 Faran, T., Poznanski, D., Filippenko, A. V., et al. 2014, , 445, 554 Filippenko, A. V. 1997, , 35, 309 Fraser, M., Ergon, M., Eldridge, J. J., et al. 2011, , 417, 1417 Fraser, M., Maund, J. R., Smartt, S. J., et al. 2012, , 759, L13 Fraser, M., Maund, J. R., Smartt, S. J., et al. 2014, , 439, L56 Gall, E. E. E., Polshaw, J., Kotak, R., et al. 2015, , 582, A3 Gal-Yam, A., Kasliwal, M. M., Arcavi, I., et al. 2011, , 736, 159 Gehrels, N., Chincarini, G., Giommi, P., et al. 2004, , 611, 1005 Gezari, S., Dessart, L., Basa, S., et al. 2008, , 683, L131 Gonz[á]{}lez-Gait[á]{}n, S., Tominaga, N., Molina, J., et al. 2015, , 451, 2212 Grassberg, E. K., Imshennik, V. S., & Nadyozhin, D. K. 1971, , 10, 28 Grasberg, E. K., & Nadezhin, D. K. 1976, , 44, 409 Grupe, D., Dong, S., Shappee, B. J., et al. 2016, The Astronomer’s Telegram, 8588 Horne, K. 1986, , 98, 609 Hosseinzadeh, G., Arcavi, I., McCully, C., et al. 2016, The Astronomer’s Telegram, 8567 Hamuy, M. 2003, , 582, 905 Huang, F., Li, J.-Z., Wang, X.-F., et al. 2012, Research in Astronomy and Astrophysics, 12, 1585 Huang, F., Wang, X., Zhang, J., et al. 2015, , 807, 59 Huang, F., Wang, X., Zampieri, L., et al. 2016, , 832, 139 Ibeling, D., & Heger, A. 2013, , 765, L43 Immler, S., Brown, P. J., Milne, P., et al. 2007, , 664, 435 Jerkstrand, A., Fransson, C., Maguire, K., et al. 2012, , 546, A28 Kasen, D., & Woosley, S. E. 2009, , 703, 2205 Klein, R. I., & Chevalier, R. A. 1978, , 223, L109 Landolt, A. U. 1992, , 104, 340 Leonard, D. C., Filippenko, A. V., Gates, E. L., et al. 2002, , 114, 35 Levesque, E. M., Massey, P., Olsen, K. A. G., et al. 2005, , 628, 973 Levesque, E. M., Massey, P., Olsen, K. A. G., et al. 2006, , 645, 1102 Li, W., Leaman, J., Chornock, R., et al. 2011, , 412, 1441 Lisakov, S. M., Dessart, L., Hillier, D. J., Waldman, R., & Livne, E. 2017, , 466, 34 Maund, J. R., Smartt, S. J., & Danziger, I. J. 2005, , 364, L33 Maund, J. R., Fraser, M., Smartt, S. J., et al. 2013, , 431, L102 Maund, J. R., Mattila, S., Ramirez-Ruiz, E., & Eldridge, J. J. 2014, , 438, 1577 Morozova, V., Piro, A. L., Renzo, M., et al. 2015, , 814, 63 Morozova, V., Piro, A. L., Renzo, M., & Ott, C. D. 2016, , 829, 109 Nomoto, K. 1984, , 277, 791 Nomoto, K., & Hashimoto, M. 1988, , 163, 13 Olivares E., F., Hamuy, M., Pignata, G., et al. 2010, , 715, 833 Pastorello, A., Zampieri, L., Turatto, M., et al. 2004, , 347, 74 Pastorello, A., Valenti, S., Zampieri, L., et al. 2009, , 394, 2266 Patat, F., Barbon, R., Cappellaro, E., & Turatto, M. 1994, , 282, 731 Pumo, M. L., & Zampieri, L. 2011, , 741, 41 Pumo, M. L., Zampieri, L., Spiro, S., et al. 2017, , 464, 3013 Rabinak, I., & Waxman, E. 2011, , 728, 63 Roming, P. W. A., Kennedy, T. E., Mason, K. O., et al. 2005, , 120, 95 Rubin, A., Gal-Yam, A., De Cia, A., et al. 2016, , 820, 33 Sahu, D. K., Anupama, G. C., Srividya, S., & Muneer, S. 2006, , 372, 1315 Sanders, N. E., Soderberg, A. M., Gezari, S., et al. 2015, , 799, 208 Sapir, N., & Waxman, E. 2017, , 838, 130 Schawinski, K., Justham, S., Wolf, C., et al. 2008, Science, 321, 223 Schlafly, E. F., &Finkbeiner, D. P. 2011, , 737, 103 Smartt, S. J. 2009, , 47, 63 Smartt, S. J. 2015, , 32, e016 Sorce, J. G., Tully, R. B., Courtois, H. M., et al. 2014, , 444, 527 Tak[á]{}ts, K., Pumo, M. L., Elias-Rosa, N., et al. 2014, , 438, 368 Tomasella, L., Cappellaro, E., Fraser, M., et al. 2013, , 434, 1636 Valenti, S., Sand, D., Pastorello, A., et al. 2014, , 438, L101 Valenti, S., Sand, D., Stritzinger, M., et al. 2015, , 448, 2608 Valenti, S., Howell, D. A., Stritzinger, M. D., et al. 2016, , 459, 3939 Van Dyk, S. D., Cenko, S. B., Poznanski, D., et al. 2012, , 756, 131 Wang, X., Li, W., Filippenko, A. V., et al. 2008, , 675, 626 Zheng, X.-M., & Zhang, J.-J. 2016, The Astronomer’s Telegram, 8584 ------ ---------------------- ---------------------- ----------- ----------- ----------- ----------- ----------- Star $\alpha_{\rm J2000}$ $\delta_{\rm J2000}$ $U$ $B$ $V$ $R$ $I$ ID (h m s) (  ) (mag) (mag) (mag) (mag) (mag) 1 12:55:10.458 0:09:19.79 18.58(03) 18.38(04) 17.58(03) 17.11(03) 16.66(06) 2 12:55:22.596 0:08:42.00 14.47(02) 13.91(03) 13.47(02) 13.10(05) 12.39(09) 3 12:55:26.850 0:04:14.02 18.81(04) 17.78(06) 16.62(03) 15.95(04) 15.33(09) 4 12:55:21.901 0:04:33.88 17.92(03) 17.59(05) 16.77(03) 16.29(04) 15.81(07) 5 12:55:11.357 0:04:50.16 18.38(03) 17.24(07) 15.72(04) 14.78(08) 13.72(16) 6 12:55:00.181 0:05:08.91 16.93(02) 17.05(04) 16.48(03) 16.14(03) 15.78(05) 7 12:55:16.953 0:04:05.97 20.37(08) 19.44(07) 17.94(04) 16.92(10) 15.56(21) 8 12:55:26.305 0:03:09.69 16.20(02) 15.98(04) 15.30(03) 14.90(03) 14.49(06) 9 12:55:14.299 0:03:17.25 19.16(04) 18.97(04) 18.27(03) 17.85(03) 17.41(06) 10 12:55:17.333 0:03:16.92 15.72(02) 15.68(04) 15.01(03) 14.60(03) 14.15(06) ------ ---------------------- ---------------------- ----------- ----------- ----------- ----------- ----------- -------------- ----------- ------------- ------------------ ------------------ ------------------ ----------- ------------------ UT Date MJD Phase$^{a}$ $U$ $B$ $V$ $R$ $I$ (yy/mm/dd) (day) (mag) (mag) (mag) (mag) (mag) 2016 Jan. 30 57417.705 11.79 13.28(01) 14.26(03) 14.11(01) 13.95(01) 13.78(04) 2016 Jan. 31 57418.755 12.84 13.19(01) 14.32(02) 14.09(02) 13.89(03) 13.77(02) 2016 Feb. 01 57419.885 13.97 13.46(01) 14.28(04) 14.10(04) 13.89(05) 13.75(05) 2016 Feb. 02 57420.710 14.79 13.61(02) 14.44(03) 14.16(04) 14.06(05) 93.94(04) 2016 Feb. 04 57422.775 16.86 [  $\cdots$  ]{} 14.44(03) 14.21(03) 13.99(03) 13.82(02) 2016 Feb. 05 57423.870 17.95 13.88(02) 14.51(03) 14.26(04) 14.00(02) 13.84(04) 2016 Feb. 14 57432.680 26.76 14.82(01) 14.96(02) 14.41(01) 14.13(03) 13.91(02) 2016 Feb. 15 57433.885 27.97 14.97(02) 15.01(03) 14.46(01) 14.15(04) 13.99(02) 2016 Feb. 16 57434.690 28.77 15.03(02) 15.07(02) 14.49(03) 14.16(03) 13.99(04) 2016 Feb. 17 57435.695 29.78 15.24(02) 15.09(03) 14.49(04) 14.13(05) 13.99(04) 2016 Feb. 19 57437.885 31.97 [  $\cdots$  ]{} 15.19(03) 14.53(03) 14.21(02) 13.98(03) 2016 Feb. 20 57438.675 32.76 15.40(03) 15.20(04) 14.60(04) 14.21(04) 14.04(04) 2016 Feb. 22 57440.705 34.79 [  $\cdots$  ]{} 15.24(01) 14.58(02) 14.20(01) 14.03(01) 2016 Feb. 23 57441.885 35.97 [  $\cdots$  ]{} 15.33(03) 14.54(03) 14.28(04) [  $\cdots$  ]{} 2016 Mar. 01 57448.845 42.93 15.93(04) 15.65(03) 14.74(04) 14.36(06) 14.14(05) 2016 Mar. 02 57449.690 43.77 15.96(05) 15.51(03) 14.72(04) 14.29(04) 14.07(05) 2016 Mar. 03 57450.670 44.75 [  $\cdots$  ]{} 15.52(02) 14.67(03) 14.28(03) 14.09(03) 2016 Mar. 05 57452.625 46.71 15.98(05) 15.56(05) 14.67(05) 14.22(04) 14.03(05) 2016 Mar. 10 57457.635 51.72 16.12(03) 15.61(05) 14.66(05) 14.26(05) 14.03(04) 2016 Mar. 11 57458.645 52.73 16.12(05) 15.60(04) 14.75(02) 14.31(03) 14.08(03) 2016 Mar. 12 57459.630 53.71 16.26(04) 15.70(04) 14.87(04) 14.40(04) 14.18(04) 2016 Mar. 13 57460.625 54.71 16.23(05) 15.70(04) 14.66(07) 14.30(06) 14.05(04) 2016 Mar. 20 57467.715 61.80 16.43(07) 15.85(03) 14.81(04) 14.46(05) 14.17(05) 2016 Mar. 29 57476.685 70.77 17.02(05) 16.13(03) 15.00(04) 14.54(03) 14.30(04) 2016 Mar. 30 57477.695 71.78 17.18(07) 16.11(04) 15.04(03) 14.60(06) 14.32(04) 2016 Apr. 03 57481.750 75.83 17.35(05) 16.24(02) 15.04(01) 14.61(04) 14.28(03) 2016 Apr. 04 57482.750 76.83 17.39(09) 16.22(03) 15.11(03) 14.59(03) 14.33(05) 2016 Apr. 13 57491.750 85.83 17.80(07) 16.55(03) 15.23(03) 14.71(03) 14.48(03) 2016 Apr. 16 57494.750 88.83 18.40(14) 16.75(03) 15.42(04) 14.84(04) 14.58(02) 2016 Apr. 22 57500.750 94.83 [  $\cdots$  ]{} [  $\cdots$  ]{} 15.87(05) 15.25(05) 14.91(03) 2016 Apr. 23 57501.750 95.83 [  $\cdots$  ]{} 17.41(10) 16.18(04) 15.44(03) 15.03(03) 2016 May 06 57514.500 108.58 19.59(34) 18.14(05) 16.77(03) 16.05(04) 15.67(04) 2016 May 07 57515.750 109.83 [  $\cdots$  ]{} [  $\cdots$  ]{} [  $\cdots$  ]{} 16.14(17) 15.76(17) 2016 Jun. 02 57541.500 135.58 [  $\cdots$  ]{} 18.36(05) 17.23(04) 16.53(03) 16.20(03) -------------- ----------- ------------- ------------------ ------------------ ------------------ ----------- ------------------ $^{a}$ Relative to the explosion date, MJD = 57,405.92.\ -------------- ---------- ------------- ------------------ ------------------ ----------- ------------------ ------------------ UT Date MJD Phase$^{a}$ $U$ $B$ $V$ $R$ $I$ (yy/mm/dd) (day) (mag) (mag) (mag) (mag) (mag) 2016 Jan. 23 57410.91 4.99 13.36(02) 14.29(04) 14.38(03) 14.31(04) 14.33(02) 2016 Jan. 28 57415.92 10.00 13.23(02) 14.02(05) 14.04(04) 14.09(04) 13.81(05) 2016 Feb. 02 57420.88 14.96 13.53(06) 14.38(02) 14.16(02) 13.93(03) 13.83(03) 2016 Feb. 04 57422.83 16.91 13.90(17) 14.49(09) 14.36(08) 14.16(09) 13.95(05) 2016 Feb. 13 57431.92 26.00 14.88(05) 14.86(05) 14.37(02) 14.10(01) 13.90(08) 2016 Feb. 16 57434.85 28.93 15.19(02) 15.06(02) 14.48(02) 14.17(02) 14.02(01) 2016 Feb. 18 57436.90 30.98 15.27(03) 15.27(03) 14.69(03) 14.23(03) 14.06(01) 2016 Feb. 23 57441.84 35.92 15.79(03) 15.45(02) 14.67(05) 14.30(07) 14.06(11) 2016 Feb. 28 57446.83 40.91 15.83(07) 15.49(08) 14.62(05) 14.41(09) [  $\cdots$  ]{} 2016 Mar. 02 57449.93 44.01 15.96(02) 15.47(03) 14.72(02) 14.32(03) 14.08(02) 2016 Mar. 03 57450.84 44.92 15.96(04) 15.49(02) 14.69(03) 14.26(03) 14.07(03) 2016 Mar. 11 57458.85 52.93 16.21(03) 15.59(04) 14.75(02) 14.30(06) 14.09(02) 2016 Mar. 16 57460.91 54.99 16.36(13) 15.78(06) 14.86(04) 14.45(03) 14.14(03) 2016 Mar. 18 57465.74 59.82 16.66(12) 15.89(04) 14.85(04) [  $\cdots$  ]{} [  $\cdots$  ]{} 2016 Mar. 20 57467.88 61.96 [  $\cdots$  ]{} 15.86(01) 14.92(02) 14.46(02) 14.21(02) 2016 Apr. 04 57482.75 76.83 [  $\cdots$  ]{} 16.39(03) 15.09(04) 14.63(04) 14.37(02) 2016 Apr. 09 57487.73 81.81 17.56(08) 16.39(04) 15.19(03) 14.68(04) 14.41(01) 2016 Apr. 17 57495.79 89.87 [  $\cdots$  ]{} 17.05(15) 15.45(12) 14.96(03) 14.51(05) 2016 Apr. 26 57504.72 98.80 19.36(16) 18.00(06) 16.58(04) 15.83(04) 15.51(06) 2016 May 02 57510.68 104.76 [  $\cdots$  ]{} [  $\cdots$  ]{} 16.92(03) 16.23(05) 15.85(04) 2016 May 11 57519.69 113.77 19.66(33) 18.03(13) 16.88(06) 16.26(07) 15.96(06) 2016 May 27 57535.70 129.78 [  $\cdots$  ]{} 18.30(06) 17.08(03) 16.40(03) 16.07(02) 2016 Jun. 03 57542.67 136.75 [  $\cdots$  ]{} 18.29(14) 17.17(09) 16.55(05) 16.17(05) -------------- ---------- ------------- ------------------ ------------------ ----------- ------------------ ------------------ $^{a}$ Relative to the explosion date, MJD = 57,405.92.\ -------------- ----------- ----------- ------------------ ------------------ ------------------ ------------------ ------------------ ------------------ UT Date MJD Phase$^a$ $U$ $B$ $V$ $g$ $r$ $i$ (yy/mm/dd) (day) (mag) (mag) (mag) (mag) (mag) (mag) 2016 Jan. 21 57408.340 2.421 13.723(023) 14.711(042) 14.818(035) 14.714(032) 14.903(030) 15.021(020) 2016 Jan. 22 57409.360 3.441 13.718(052) 14.504(042) 14.675(041) 14.473(022) 14.775(036) 14.845(042) 2016 Jan. 25 57412.275 6.356 13.615(012) 14.199(044) 14.186(052) 14.265(023) 14.364(047) 14.328(028) 2016 Jan. 26 57413.315 7.396 13.276(023) 14.128(030) 14.083(032) 14.073(033) 14.370(041) 14.258(033) 2016 Jan. 26 57413.340 7.421 13.259(027) 14.130(027) 14.070(026) 14.055(030) 14.147(022) 14.216(026) 2016 Jan. 28 57415.085 9.166 13.300(044) 14.088(038) 14.162(035) 14.026(036) 14.159(040) 14.274(039) 2016 Jan. 28 57415.235 9.316 13.276(041) 14.162(032) 14.198(034) 14.113(041) 14.095(050) 14.106(048) 2016 Feb. 01 57419.245 13.326 13.509(052) 14.288(041) 14.109(029) 14.150(026) 14.043(024) 14.176(028) 2016 Feb. 03 57421.090 15.171 13.565(016) 14.348(041) 13.991(026) 14.168(044) 14.023(032) 14.161(032) 2016 Feb. 07 57425.095 19.176 14.284(018) 14.515(058) 14.261(050) 14.308(029) 14.189(052) 14.211(039) 2016 Feb. 12 57430.655 24.736 14.460(035) 14.790(056) 14.407(054) 14.513(030) 14.222(038) 14.315(035) 2016 Feb. 15 57433.995 28.076 14.995(073) 15.067(042) [  $\cdots$  ]{} 14.631(043) 14.269(040) 14.387(043) 2016 Feb. 19 57437.720 31.801 [  $\cdots$  ]{} 15.178(053) 14.490(038) 14.870(036) 14.342(034) 14.381(025) 2016 Feb. 23 57441.565 35.646 15.356(037) 15.323(044) 14.682(039) 14.888(045) 14.486(042) 14.467(029) 2016 Feb. 28 57446.035 40.116 15.689(032) 15.426(032) 14.657(039) 14.973(037) 14.533(039) 14.500(038) 2016 Mar. 02 57449.765 43.846 [  $\cdots$  ]{} [  $\cdots$  ]{} [  $\cdots$  ]{} 14.985(036) 14.548(053) 14.516(025) 2016 Mar. 08 57455.340 49.421 [  $\cdots$  ]{} 15.628(062) 14.710(042) 15.285(042) 14.454(035) 14.502(034) 2016 Mar. 11 57458.665 52.746 15.881(057) 15.621(045) 14.787(029) 15.314(052) 14.613(037) 14.487(033) 2016 Mar. 16 57463.075 57.156 16.141(063) 15.769(052) 14.799(017) [  $\cdots$  ]{} [  $\cdots$  ]{} [  $\cdots$  ]{} 2016 Mar. 18 57465.255 59.336 16.195(049) 15.864(031) 14.843(039) 15.345(028) 14.572(030) 14.532(022) 2016 Mar. 21 57468.915 62.996 16.425(100) 15.967(048) 14.914(031) 15.330(043) 14.609(031) 14.629(035) 2016 Mar. 30 57477.110 71.191 [  $\cdots$  ]{} 16.178(046) 15.027(039) 15.503(031) 14.728(035) 14.757(038) 2016 Apr. 02 57480.830 74.911 16.855(065) 16.252(044) 15.045(034) 15.574(036) 14.824(040) 14.783(028) 2016 Apr. 07 57485.535 79.616 17.352(171) 16.305(050) 15.126(038) 15.588(031) 14.780(028) 14.799(032) 2016 Apr. 10 57488.785 82.866 17.925(144) 16.512(047) 15.213(047) 15.719(030) 14.891(044) 14.876(044) 2016 Apr. 15 57493.775 87.856 97.478(250) 16.829(062) 15.361(039) 15.939(032) 15.012(027) 15.000(032) 2016 Apr. 22 57500.465 94.546 [  $\cdots$  ]{} 17.227(058) 15.943(037) 16.401(034) 15.444(041) 15.363(040) 2016 Apr. 26 57504.740 98.821 [  $\cdots$  ]{} 17.973(062) 16.559(040) 17.152(038) 16.073(039) 16.077(043) 2016 May 01 57509.795 103.876 [  $\cdots$  ]{} 18.453(085) 16.666(047) 17.558(042) 16.402(040) [  $\cdots$  ]{} 2016 May 02 57510.720 104.801 [  $\cdots$  ]{} 18.464(085) 16.815(044) 17.580(041) 16.376(041) 16.301(041) 2016 May 02 57510.835 104.916 [  $\cdots$  ]{} 18.445(063) 16.891(044) 17.529(060) 16.396(029) [  $\cdots$  ]{} 2016 May 04 57512.030 106.111 [  $\cdots$  ]{} 18.284(093) 16.776(038) 17.391(035) 16.307(026) 16.240(034) 2016 May 04 57512.740 106.821 [  $\cdots$  ]{} 18.401(059) 16.806(040) 17.518(026) 16.417(040) 16.289(045) 2016 May 05 57513.710 107.791 [  $\cdots$  ]{} 18.292(065) 16.766(049) 17.424(037) 16.322(037) 16.273(043) 2016 May 10 57518.965 113.046 [  $\cdots$  ]{} 18.189(057) 16.871(055) 17.581(060) 16.459(042) 16.376(037) 2016 May 11 57519.935 114.016 [  $\cdots$  ]{} 18.307(069) 16.860(033) 17.364(034) 16.412(027) 16.398(040) 2016 May 12 57520.050 114.131 [  $\cdots$  ]{} 18.338(045) 16.900(041) 17.454(032) 16.431(029) 16.443(037) 2016 May 14 57522.935 117.016 [  $\cdots$  ]{} 18.500(112) 16.940(034) [  $\cdots$  ]{} [  $\cdots$  ]{} [  $\cdots$  ]{} 2016 May 15 57523.123 117.204 [  $\cdots$  ]{} 18.468(115) 16.926(043) 17.485(038) 16.460(037) 16.388(040) 2016 May 20 57528.760 122.841 [  $\cdots$  ]{} 18.225(102) 17.089(061) [  $\cdots$  ]{} 16.470(038) 16.397(033) 2016 May 28 57536.065 130.146 [  $\cdots$  ]{} 18.331(035) 17.056(037) [  $\cdots$  ]{} [  $\cdots$  ]{} [  $\cdots$  ]{} 2016 Jun. 04 57543.760 137.841 [  $\cdots$  ]{} [  $\cdots$  ]{} 17.416(188) 17.841(029) 16.780(032) 16.824(024) 2016 Jun. 05 57544.763 138.844 [  $\cdots$  ]{} 18.729(040) 17.372(042) 17.934(034) 16.859(028) 16.853(026) 2016 Jun. 06 57545.885 139.966 [  $\cdots$  ]{} 18.682(051) 17.466(033) 17.927(041) 16.883(031) 16.851(026) 2016 Jun. 07 57546.745 140.826 [  $\cdots$  ]{} 18.608(066) 17.489(038) [  $\cdots$  ]{} [  $\cdots$  ]{} [  $\cdots$  ]{} 2016 Jun. 08 57547.715 141.796 [  $\cdots$  ]{} 18.734(054) 17.494(038) 17.931(031) 17.007(029) 16.870(041) 2016 Jun. 23 57562.705 156.786 [  $\cdots$  ]{} 18.484(060) 17.416(043) 17.855(041) 16.812(034) [  $\cdots$  ]{} 2016 Jun. 24 57563.705 157.786 [  $\cdots$  ]{} 18.612(159) 17.516(101) [  $\cdots$  ]{} [  $\cdots$  ]{} [  $\cdots$  ]{} 2016 July 06 57575.745 169.826 [  $\cdots$  ]{} 18.634(045) 17.668(045) 18.069(037) 16.871(034) 17.084(028) -------------- ----------- ----------- ------------------ ------------------ ------------------ ------------------ ------------------ ------------------ $^{a}$ Relative to the explosion date, MJD = 57,405.92.\ -------------- ---------- ----------- ----------- ------------------ ------------------ ------------------ ------------------ ------------------ UT Date MJD Phase$^a$ $uvw$2 $uvm$2 $uvw$1 $U$ $B$ $V$ (yy/mm/dd) (day) (mag) (mag) (mag) (mag) (mag) (mag) 2016 Jan. 21 57408.07 2.15 12.74(03) 12.72(03) 12.91(03) 13.49(03) 14.83(04) 14.91(06) 2016 Jan. 22 57409.45 3.53 12.89(04) 12.75(04) 12.87(04) 13.25(04) 14.55(04) 14.74(06) 2016 Jan. 22 57409.79 3.87 12.82(04) 12.66(04) 12.80(04) 13.20(04) 14.49(04) 14.57(06) 2016 Jan. 23 57410.19 4.27 12.79(04) 12.62(04) 12.73(04) 13.11(04) 14.36(04) 14.47(06) 2016 Jan. 24 57411.82 5.90 13.00(04) 12.72(04) 12.72(04) 12.95(04) 14.24(04) 14.22(05) 2016 Jan. 25 57412.50 6.58 13.17(03) 12.86(03) 12.82(03) 13.00(03) 14.21(03) 14.21(05) 2016 Jan. 27 57414.25 8.33 13.55(04) [  $\cdots$  ]{} [  $\cdots$  ]{} [  $\cdots$  ]{} [  $\cdots$  ]{} [  $\cdots$  ]{} 2016 Feb. 01 57419.03 13.11 14.57(05) [  $\cdots$  ]{} 13.87(05) 13.28(04) 14.29(04) 14.13(05) 2016 Feb. 01 57419.76 13.84 14.76(06) [  $\cdots$  ]{} 14.03(05) 13.36(04) 14.30(04) 14.18(05) 2016 Feb. 06 57424.69 18.77 16.11(20) [  $\cdots$  ]{} 15.38(07) 14.05(05) 14.49(05) [  $\cdots$  ]{} 2016 Feb. 07 57425.10 19.18 16.25(08) 16.43(08) 15.33(06) 14.06(04) 14.53(04) 14.28(05) 2016 Feb. 08 57426.63 20.71 16.60(09) 16.83(10) 15.67(07) 14.25(05) 14.69(05) 14.31(06) 2016 Feb. 09 57427.96 22.04 16.92(09) 17.12(09) 15.89(07) 14.55(05) 14.68(05) 14.27(05) 2016 Feb. 10 57428.39 22.47 16.89(10) 17.16(11) 15.92(08) 14.54(05) 14.72(05) 14.36(06) 2016 Feb. 17 57435.68 29.76 18.24(19) 18.66(26) 17.02(10) 15.59(07) 15.14(05) 14.48(06) 2016 Feb. 21 57439.80 33.88 18.54(18) 19.39(34) 17.21(11) 15.83(08) 15.26(05) 14.62(06) 2016 Feb. 23 57441.59 35.67 18.40(18) 18.91(26) 17.36(10) 16.10(08) 15.34(05) 14.71(07) 2016 Mar. 1 57448.77 42.85 18.75(22) [  $\cdots$  ]{} 17.64(14) 16.29(09) 15.52(06) 14.81(07) 2016 Mar. 5 57452.46 46.54 18.95(22) [  $\cdots$  ]{} 17.73(13) 16.39(09) 15.61(06) 14.80(06) -------------- ---------- ----------- ----------- ------------------ ------------------ ------------------ ------------------ ------------------ $^{a}$ Relative to the explosion date, MJD = 57,405.92.\ -------------- ---------- ----------- ------------- ---------- ------------------------------------ UT Date MJD Phase$^a$ Range Exposure Telescope + Instrument (days) (Å) (s) 2016 Jan. 20 57407.74 1.82 3300–9,000 900 LCO 2.0 m Telescope South + FLOYDS 2016 Jan. 21 57408.52 2.60 3300–10,000 900 LCO 2.0 m Telescope North + FLOYDS 2016 Jan. 23 57410.48 4.56 3250–10,000 900 LCO 2.0 m Telescope North + FLOYDS 2016 Jan. 23 57410.68 4.76 3250–10,000 900 LCO 2.0 m Telescope South + FLOYDS 2016 Jan. 23 57410.89 4.97 3500–9000 1500 Lijiang 2.4 m + YFOSC 2016 Jan. 25 57412.47 6.55 3250–10,000 900 LCO 2.0 m Telescope North + FLOYDS 2016 Jan. 25 57412.67 6.75 3250–10,000 900 LCO 2.0 m Telescope South + FLOYDS 2016 Jan. 27 57414.52 8.60 3250–10,000 900 LCO 2.0 m Telescope North + FLOYDS 2016 Jan. 28 57415.54 9.62 3250–10,000 1200 LCO 2.0 m Telescope North + FLOYDS 2016 Jan. 28 57415.93 10.01 3500–9100 1200 Lijiang 2.4 m + YFOSC 2016 Jan. 31 57418.73 12.81 3350–10,000 1200 LCO 2.0 m Telescope South + FLOYDS 2016 Feb. 2 57420.88 14.96 3500–9100 1800 Lijiang 2.4 m + YFOSC 2016 Feb. 3 57421.46 15.54 3250–10,000 1200 LCO 2.0 m Telescope North + FLOYDS 2016 Feb. 6 57424.71 18.79 3400–10,000 1200 LCO 2.0 m Telescope South + FLOYDS 2016 Feb. 10 57428.68 22.76 3400–10,000 1200 LCO 2.0 m Telescope South + FLOYDS 2016 Feb. 13 57431.90 25.98 3500–9100 1500 Lijiang 2.4 m + YFOSC 2016 Feb. 16 57434.69 28.77 3550–10,000 1200 LCO 2.0 m Telescope South + FLOYDS 2016 Feb. 16 57434.86 28.94 3500–9100 1500 Lijiang 2.4 m + YFOSC 2016 Feb. 18 57436.87 30.95 3500–9100 1500 Lijiang 2.4 m + YFOSC 2016 Feb. 22 57440.47 34.55 3550–10,000 1200 LCO 2.0 m Telescope North + FLOYDS 2016 Feb. 23 57441.84 35.92 3500–9100 1500 Lijiang 2.4 m + YFOSC 2016 Feb. 28 57446.54 40.62 3500–10,000 1200 LCO 2.0 m Telescope North + FLOYDS 2016 Feb. 28 57446.84 40.92 3500–9100 1500 Lijiang 2.4 m + YFOSC 2016 Mar. 6 57453.64 47.72 3500–10,000 1200 LCO 2.0 m Telescope North + FLOYDS 2016 Mar. 11 57458.86 52.94 3500–9100 1800 Lijiang 2.4 m + YFOSC 2016 Mar. 12 57459.69 53.77 3700–10,000 1200 LCO 2.0 m Telescope South + FLOYDS 2016 Mar. 18 57465.51 59.59 3600–10,000 1200 LCO 2.0 m Telescope North + FLOYDS 2016 Mar. 18 57465.75 59.83 3500–9100 1800 Lijiang 2.4 m + YFOSC 2016 Mar. 21 57468.74 62.82 3950–10,000 1200 LCO 2.0 m Telescope South + FLOYDS 2016 Mar. 27 57474.71 68.79 3700–9150 2850 Lijiang 2.4 m + YFOSC 2016 Apr. 4 57482.72 76.80 3650–9150 2100 Lijiang 2.4 m + YFOSC 2016 Apr. 7 57485.51 79.59 3800–10,000 1200 LCO 2.0 m Telescope South + FLOYDS 2016 Apr. 13 57491.69 85.77 3900–10,000 1200 LCO 2.0 m Telescope South + FLOYDS 2016 Apr. 17 57495.69 89.77 3900–8780 2100 Xinglong 2.16 m + BFOSC 2016 Apr. 17 57495.80 89.88 3600–9100 2100 Lijiang 2.4 m + YFOSC 2016 Apr. 26 57504.72 98.80 3600–9100 2100 Lijiang 2.4 m + YFOSC 2016 Apr. 29 57507.26 101.34 3500–10,000 1200 LCO 2.0 m Telescope North + FLOYDS 2016 May 2 57510.72 104.80 3500–9170 2100 Lijiang 2.4 m + YFOSC 2016 May 21 57529.27 123.35 4500–9300 1800 LCO 2.0 m Telescope North + FLOYDS 2016 Jun. 9 57548.34 142.42 3500–10,000 3600 LCO 2.0 m Telescope North + FLOYDS -------------- ---------- ----------- ------------- ---------- ------------------------------------ $^{a}$ Relative to the explosion date, MJD = 57,405.92. ------------------------ ------- ------- ------- ------- ------- ------- ------- ------- $U$ $B$ $g$ $V$ $r$ $R$ $i$ $I$ (mag) (mag) (mag) (mag) (mag) (mag) (mag) (mag) Peak magnitude 13.25 14.14 14.04 14.05 13.99 13.91 14.07 13.77 Phase of maximum$^{a}$ 9.26 9.60 10.60 11.26 13.70 13.76 13.55 14.11 Plateau magnitude – – 15.31 14.67 14.60 14.46 14.50 14.07 Decay rate (mag/100 d) – 0.58 0.99 1.35 1.25 1.22 1.05 1.14 ------------------------ ------- ------- ------- ------- ------- ------- ------- ------- $^{a}$ Relative to the explosion date, MJD = 57,405.92.\ ----------- ---------------------- ------------------ ---------------- -------------- -------------------------------------------------------------------- mass radius mass radius () () () () SN 2005cs $9_{-2}^{+3}$ [  $\cdots$  ]{} 11 360 $\pm$ 70 [@2005MNRAS.364L..33M; @2017MNRAS.464.3013P] SN 2008bk $12.9_{-1.8}^{+1.6}$ 470 $\pm$ 16 12 502 [@2014MNRAS.438.1577M; @2017MNRAS.466...34L] SN 2009N [  $\cdots$  ]{} [  $\cdots$  ]{} 13 $\pm$ 2 287$\pm$ 43 [@2014MNRAS.438..368T] SN 2009md $8.5_{-1.5}^{+6.5}$ [  $\cdots$  ]{} 10 288 [@2011MNRAS.417.1417F; @2017MNRAS.464.3013P] SN 2012A $10.5_{-2}^{+4.5}$ [  $\cdots$  ]{} 14 $\pm$ 2 260 $\pm$ 40 [@2013MNRAS.434.1636T] SN 2012aw 14 – 26 1040 $\pm$ 100 22 – 24 290 – 580 [@2012ApJ...756..131V; @2012ApJ...759L..13F; @2014ApJ...787..139D] SN 2012ec 14 – 22 1030 $\pm$ 180 14.0 – 14.6 230 $\pm$ 70 [@2013MNRAS.431L.102M; @2015MNRAS.448.2312B] SN 2013ej 8 – 15.6 [  $\cdots$  ]{} 12.5 $\pm$ 1.9 415 $\pm$ 62 [@2014MNRAS.439L..56F; @2015ApJ...807...59H] SN 2016X [  $\cdots$  ]{} [  $\cdots$  ]{} 18.5 – 19.7 925 $\pm$ 65 this work ----------- ---------------------- ------------------ ---------------- -------------- -------------------------------------------------------------------- \[lastpage\] [^1]: E-mail: [email protected] [^2]: E-mail: wang\[email protected] [^3]: IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation (NSF). [^4]: http://sngroup.oapd.inaf.it/snoopy.html [^5]: http://swift.gsfc.nasa.gov/docs/swift/sne/swift\_sn.html [^6]: http://leda.univ-lyon1.fr/
{ "pile_set_name": "ArXiv" }
--- author: - | Ali Punjani\ Department of Computer Science\ University of Toronto\ `[email protected]`\ Marcus A. Brubaker\ Department of Computer Science\ University of Toronto\ `[email protected]` bibliography: - 'refs.bib' title: 'Microscopic Advances with Large-Scale Learning: Stochastic Optimization for Cryo-EM' ---
{ "pile_set_name": "ArXiv" }
--- abstract: 'Deterministic light-induced spin squeezing in an atomic gas is limited by photon shot noise or, equivalently, by atomic state information escaping with the light field mediating the effective atom-atom interaction. We show theoretically that the performance of cavity spin squeezing \[M.H. Schleier-Smith, I.D. Leroux, and V. Vuletić, Phys. Rev. A **81**, 021804(R) (2010)\] can be substantially improved by erasing the light-atom entanglement, and propose several methods for doing so. Accounting for light scattering into free space, quantum erasure improves the scaling of cavity squeezing from $S^{-1/2}$ to $S^{-2/3}$, where $S$ is the total atomic spin.' author: - 'Ian D. Leroux' - 'Monika H. Schleier-Smith' - Hao Zhang - Vladan Vuletić bibliography: - 'shotnoise-cancellation.bib' title: Unitary Cavity Spin Squeezing by Quantum Erasure --- Introduction ============ Squeezed spin states [@Kitagawa1993] are among the simplest many-body entangled states to describe and characterize, since their quantum correlations appear as an improved signal-to-noise ratio in certain measurements. This improvement makes squeezed spin states potentially useful for precision metrology beyond the standard quantum limit (SQL) that is set by the quantum fluctuations of independent particles. In particular, spin squeezing might increase the stability of atomic clocks, magnetometers, and other measurements based on atom interferometry [@Wineland1992; @Wineland1994; @Louchet-Chauvet2010; @Leroux2010:clock]. Many approaches to spin squeezing have been proposed [@Wineland1992; @Kuzmich1998; @Sorensen1999; @Sorensen2001; @Andre2002; @Bouchoule2002; @Zhang2003; @Stockton2004; @Hammerer2004; @Madsen2004; @Takeuchi2005; @Genes2006; @Meiser2008; @Saffman2009; @Schleier-Smith2010:dynsq; @Schleier-Smith2011:err; @Tasgin2011] and a number of them have been demonstrated experimentally, including atomic absorption of squeezed light [@Hald1999], entangling gates in an ion trap [@Meyer2001], projection by quantum non-demolition (QND) measurement [@Appel2009; @Schleier-Smith2010:msmt; @Chen2011], atom-atom collisions [@Esteve2008; @Gross2010; @Riedel2010], and light-mediated interaction between distant atoms in an optical resonator [@Leroux2010:dynsq; @Leroux2011:err]. This last method, cavity squeezing [@Schleier-Smith2010:dynsq; @Schleier-Smith2011:err], has generated the strongest spin squeezing to date, with 5.6 dB observed (no noise subtracted) and 10 dB inferred (detection noise subtracted) [@Leroux2010:dynsq]. Cavity squeezing relies on the off-resonant interaction between an ensemble of atoms and a light field circulating in an optical resonator cavity [@Schleier-Smith2010:dynsq]. The atoms’ state-dependent index of refraction modifies the cavity resonance frequency. If the cavity is driven by a probe laser, the atom-induced resonance frequency shift changes the optical power circulating in the cavity, modifies the AC Stark shift it imparts to the atoms, and thus affects the phase of the atomic state. This phase shift, which depends on the states of all atoms in the ensemble, introduces correlations between them and produces a squeezed state of the total atomic spin. In the cavity squeezing scheme analyzed in Ref. [@Schleier-Smith2010:dynsq], the atomic spin is entangled with the outgoing light field so that an outside observer can gain information about the atomic state. While that information can be used to reduce the variance of a transverse spin component [@Kuzmich1998; @Kuzmich2000; @Takano2009] and hence to perform conditional spin squeezing by QND measurement [@Appel2009; @Schleier-Smith2010:msmt; @Chen2011], it is ignored in unconditional cavity squeezing and thus causes non-unitary evolution in the spin subspace. Equivalently, the AC Stark shift fluctuations associated with the probe photon shot noise cause an undesirable growth of the spin uncertainty region, reducing the achievable cavity squeezing. ![(Color online) Setup for cavity squeezing. (a) Atoms with two pseudospin states $\upk,\downk$ optically coupled to an excited state $\exck$. (b) The atoms are uniformly coupled to an optical cavity, either symmetric (i) or one-sided (ii), whose resonance frequency $\ocav$ lies between the two optical transition frequencies. (c) The cavity is driven by a pulse detuned from its resonance frequency, so that the intracavity intensity depends on the atom-induced cavity frequency shift. (d) The resulting AC Stark shift imparts an -dependent phase shift to the atoms, shearing the uncertainty distribution of a coherent spin state on the Bloch sphere of total atomic spin $\vc{S}$.[]{data-label="fig:setup"}](setup){width="\linewidth"} In this paper we propose schemes that suppress the effects of photon shot noise, disentangle the outgoing light field and the atomic spin, and ideally result in unitary evolution in the spin subspace. We show that the photon shot noise suppression can be achieved either by creating effective photon number states using high-efficiency photodetectors and classical feedback or by a spin echo technique for canceling quantum noise. Practically, this proposal shows that the performance of unconditional light-induced squeezing can be improved well beyond the limit set by photon shot noise or light-atom entanglement. Fundamentally, it shows that one can engineer light-mediated interactions between distant atoms without leaving a trace of the atomic state in the light field. A related idea has recently been proposed [@Trail2010] in a modification of the free-space scheme put forward by Takeuchi *et al.* for spin squeezing by two-fold light-atom interaction [@Takeuchi2005]. The modified scheme of Ref. [@Trail2010] can produce exponential squeezing for short times, but is ultimately limited by light scattering into unobserved free-space modes. The near-unity single-atom cooperativity readily available in optical resonators [@Tanji-Suzuki2011] reduces the rate of this scattering relative to the squeezing, and allows the scheme we present here to achieve substantially greater spin squeezing for a given atom number than any free-space technique. In Sec. \[sec:ideal\] we analyze cavity squeezing in a one-sided cavity where the coupling of the intracavity light field to the continuum of external light modes is treated exactly in the input-output formalism [@Collett1984; @Gardiner1985]. For an input light field prepared in a near-monochromatic photon number state no information about the atomic state is contained in the outgoing light field and the spin dynamics are therefore unitary. We show that the evolution of the collective atomic spin is then described by the one-axis twisting Hamiltonian introduced by Kitagawa and Ueda [@Kitagawa1993]. In Sec. \[sec:lightnoise\] we introduce the effects of non-ideal input light states, and in Sec. \[sec:schemes\] we use the results to evaluate several practical implementation strategies: one using a spin-echo technique to erase the phase information from the outgoing light field, one using a squeezed input light field, and one using high-efficiency photodetectors to generate an effective photon number state of the light field. In Sec. \[sec:scatt\] we consider the unavoidable effects of light scattering into free space by the atomic ensemble, deriving the associated limit on cavity squeezing with quantum erasure. Unitary Cavity Squeezing Using an Input Photon Fock State {#sec:ideal} ========================================================= In Ref. [@Schleier-Smith2010:dynsq] the authors analyze cavity squeezing for atoms in a symmetric Fabry-Pérot cavity driven by a coherent field tuned to the slope of the resonance \[Fig. \[fig:setup\](b)(i)\]. The performance of the scheme is limited by the entanglement of the atomic spin variables with the light field leaving the resonator. If the information in the light field is not used, then it must be traced over, leading to decoherence of the atomic state. This adds spin uncertainty which ultimately limits the attainable amount of spin squeezing. That this decoherence can be eliminated is most easily seen in a setup with a one-sided cavity \[Fig. \[fig:setup\](b)(ii)\]. Provided that the cavity is lossless, all incident light is reflected, and nothing about the intracavity dynamics can be inferred from the reflected power. If the near-monochromatic input light is in a photon number state rather than a coherent state, then any phase shift imparted to the reflected light cannot be measured either. With no way for an outside observer to learn the state of the atoms inside the cavity, their dynamical evolution must be unitary, as we show explicitly below. Briefly, we model the system with an effective Hamiltonian that dispersively couples the atoms’ pseudospin degrees of freedom to the optical resonator mode. After diagonalizing this Hamiltonian, we use it to analyze the evolution of an arbitrary initial pseudospin state under the action of an incident light pulse with definite photon number. For a sufficiently monochromatic pulse, the system evolves into a product state of the field and pseudospin degrees of freedom, where the transformation of the pseudospin state is generated by the one-axis twisting Hamiltonian [@Kitagawa1993]. The system we consider consists of $N$ identical atoms uniformly coupled to an optical resonator that, in contrast to the system considered in Ref. [@Schleier-Smith2010:dynsq], is one-sided (Fig. \[fig:setup\]). The atoms have two stable states separated in energy by $\hbar \oat$, such as hyperfine or magnetic sublevels, which we label as the pseudospin states $\upk_i$ and $\downk_i$ for the $i$-th atom. We define total pseudospin operators $2\Sz = \sum_{i=1}^N \left( \upk_i\upb_i - \downk_i\downb_i \right)$, the population difference between the stable states, and $\Splus = \Sx+\ii\Sy = \sum_{i=1}^N\upk_i\downb_i$, $\Sminus = \adj\Splus$. For simplicity, we consider the case where the two optical transitions $\upk\leftrightarrow\exck$ and $\downk\leftrightarrow\exck$ connecting the stable states to the optically active excited state $\exck$ couple to the resonator with the same single-photon Rabi frequency $2g$. The cavity resonance frequency $\ocav$ is chosen halfway between the two optical transition frequencies, so that their single-photon detunings $\pm \Delta = \pm\oat/2$ are of equal magnitude and opposite sign. Within the rotating-wave approximation and neglecting, for now, the scattering of photons into free space by the atoms, the Hamiltonian for the intracavity system is $$\begin{aligned} \frac{H_\text{cav}}{\hbar} &= \oat \Sz + \sum_{i=1}^{N} \ocav \exck_i \excb_i + \ocav \nc +\notag \\ &\qquad + \sum_{i=1}^{N} \left[g \left(\upk_i \excb_i + \downk_i \excb_i\right)\cdag + H.c.\right].\end{aligned}$$ In this equation the first two terms describe the energy of the atoms, the third describes the energy of the cavity mode with photon annihilation operator , and the terms proportional to $g$ describe the coupling of the atoms to the light field. We are interested in the linear, dispersive regime of atom-field interactions where $\Delta$ is large compared to $g$, to the cavity linewidth $\kappa$, and to the excited-state linewidth $\Gamma$. Assuming that the intracavity photon number remains low enough to keep the excited-state population negligible (i.e. $\avg{\nc} g^2 / \Delta^2 \ll 1$), we adiabatically eliminate the excited state to replace the full atom-cavity interaction by an AC Stark shift of the two pseudospin states. In so doing we also neglect stimulated Raman processes which, under the same set of assumptions, are too far off resonance to be significant. Using the input-output formalism [@Collett1984; @Gardiner1985] to describe the coupling with the field outside the cavity, we obtain the following effective Hamiltonian for the dynamics of the pseudospin and of the light field: $$\begin{aligned} \label{eq:EffHam} \frac{H}{\hbar} &= \oat \Sz + \ocav \nc + \int\dd\omega \omega \nb{\omega} \notag \\ &\qquad + \Omega \nc \Sz + \ii \sqrt{\frac{\kappa}{2 \pi}} \int\dd\omega [\bdag{\omega} \cfield - \cdag \bfield{\omega}].\end{aligned}$$ The first three terms of the Hamiltonian account for the energy of the atoms, of the cavity field, and of the field outside the cavity with its continuous spectrum of creation operators . The fourth term represents the dispersive coupling between the cavity field and the atoms, and may be interpreted either as an AC Stark shift of the atomic levels by the light field or as a modification of the cavity resonance frequency by the atomic index of refraction [@Schleier-Smith2010:dynsq]. The coefficient $\Omega = 2 g^2 / \Delta$ is both the cavity frequency shift per atomic spin flip and the atomic transition frequency shift per intracavity photon. The final integral is the coupling between the cavity and the external field through the cavity’s partially transmissive input mirror, leading to a damping of the energy stored in the cavity field at rate $\kappa$. The Hamiltonian $H$ may be exactly diagonalized. First, note that both $\Sz$ and the total photon number $\nc + \int\dd\omega \nb{\omega}$ are conserved. Thus, all product states $\ket{m} \otimes \ket{0}$ of an atomic eigenstate $\ket{m}$ of with the electromagnetic vacuum $\ket{0}$ are eigenstates of the full Hamiltonian. The eigenstates with non-zero photon number require additional labels to specify the spectral distribution of the photons. To find those eigenstates, we follow Fano’s procedure for diagonalizing a discrete state (the cavity mode) coupled to a continuum (the external field) [@Fano1961], which yields an operator that annihilates a photon in an eigenstate of the total field: $$\begin{aligned} \afield{\omega} &= \frac{1}{\sqrt{(\omega - \ocav - \Omega \Sz)^2 + \kappa^2 / 4}} \left[\ii \sqrt{\frac{\kappa}{2 \pi}} \cfield\right. \notag \\ &\qquad \left.+ \frac{\kappa}{2 \pi} \pv\int\dd\omega^\prime \frac{\bfield{\omega^\prime}}{\omega - \omega^\prime} + (\omega - \ocav - \Omega \Sz) \bfield{\omega}\right].\end{aligned}$$ is a reminder that the Cauchy principal value of the integral must be taken over the pole at $\omega = \omega^\prime$. This field operator has the usual commutation relation $\comm{\afield{\omega}}{\adag{\omega^\prime}} = \delta(\omega^\prime - \omega)$ and allows us to rewrite the Hamiltonian (\[eq:EffHam\]) into the much simpler form $$\frac{H}{\hbar} = \oat \Sz + \int\dd\omega \omega \na{\omega}$$ in which the first term is just the bare energy of the atoms and the second describes, for a given atomic eigenstate $\ket{m}$, a set of decoupled harmonic oscillators. The excitations created by the $\adag{\omega}$ operators are photons with amplitudes to be either in the cavity ($\cdag$) or in the outside continuum ($\bdag{\omega^\prime}$), with the intracavity component resonantly enhanced near the atom-shifted cavity resonance frequency $\omega \approx \ocav + \Omega \Sz$. A one-photon eigenstate can now be generated by acting with the total field raising operator $\adag{\omega}$ on any of the vacuum states, $$H \adag{\omega} \ket{m} \otimes \ket{0} = \hbar(\oat m + \omega) \adag{\omega} \ket{m} \otimes \ket{0},$$ and repeated applications of can yield arbitrary $n$-photon states. We are now equipped to calculate the evolution of the atomic spin state under the action of input light pulses. In particular, we consider an incident pulse containing exactly $n$ photons incident on the cavity, as described by the initial state $$\ket{\Psi_{(-t_0)}} = \ket{\psi_a} \otimes \frac{1}{\sqrt{n!}} \left(\int\dd\omega \ee^{\ii \omega t_0} B(\omega) \bdag{\omega}\right)^n \ket{0}$$ where is an arbitrary state of the atoms and $B(\omega)$ is the pulse amplitude spectrum. We have written this initial state explicitly as a product state of the atoms and field, emphasizing that the creation operator acts only on the field outside the cavity. We take the initial time $-t_0$ to be far in the past, before the pulse arrives at the resonator (specifically $t_0 \gg \max \abs{\dd B(\omega) / \dd\omega}^{2/3}$). Re-expressed in terms of field eigenstates, the initial state is $$\begin{aligned} \ket{\Psi_{(-t_0)}} &= \sqrt{\frac{(-\ii)^n}{n!}} \times \\ &\qquad \times \left(\int\dd\omega \ee^{\ii (\omega t_0 - \Phi_\omega)} B(\omega) \adag{\omega}\right)^n \ket{\psi_a} \otimes \ket{0} \notag\end{aligned}$$ where we have introduced an atom-dependent operator $$\Phi_\omega = \arctan \left(\frac{\kappa / 2}{\omega - \ocav - \Omega \Sz}\right) - \frac{\pi}{4}$$ which represents the phase lag of the intracavity field’s response to an external drive at frequency $\omega$. In this form, all components of evolve with known frequency, so it is straightforward to find the final state at $+t_0$, long after the light pulse has reflected from the cavity. Writing this final state in terms of operators that act separately on the field () and on the atoms ($\Sz$), we see that it is, in general, entangled: $$\begin{aligned} \ket{\Psi_{(+t_0)}} &= \frac{(-\ii)^n \ee^{-2 \ii \oat t_0 \Sz}}{\sqrt{n!}} \times \\ &\qquad \times \left(\int\dd\omega \ee^{-\ii (\omega t_0 + 2 \Phi_\omega)} B(\omega) \bdag{\omega}\right)^n \ket{\psi_a} \otimes \ket{0}.\notag\end{aligned}$$ However, for a near-monochromatic input pulse centered on a frequency with a bandwidth much less than $\kappa$, the atomic operator $\Phi_\omega$ is approximately $\Phi_\oprobe$ over the spectrum of the pulse and the state factorizes into $$\begin{aligned} \ket{\Psi_{(+t_0)}} &= \ee^{-2 \ii (n \Phi_\oprobe + \oat t_0 \Sz)} \ket{\psi_a} \otimes \\ &\qquad \otimes \frac{(-\ii)^n}{\sqrt{n!}} \left(\int\dd\omega B(\omega) \ee^{-\ii \omega t_0} \bdag{\omega}\right)^n \ket{0}.\notag\end{aligned}$$ In this limit of an incident monochromatic $n$-photon Fock state the final field state is independent of the atomic state (the pulse has simply been reflected by the one-sided cavity), while the atomic state has been transformed by the unitary operator $$U_n=\ee^{-2 \ii (n \Phi_\oprobe + \oat t_0 \Sz)}.$$ The dynamics of interest in the spin subspace are encoded in the nonlinear operator $\Phi_\oprobe$. When the pulse frequency is tuned to the slope of the Lorentzian cavity resonance, $\oprobe = \ocav + \kappa / 2$, and provided that $N \Omega \ll \kappa$ so that the atoms do not shift the cavity resonance frequency by a large fraction of the cavity linewidth, we can expand $\Phi_\oprobe$ to second order in the spin rotation angle $\pphase = 2 \Omega / \kappa$ imparted by a single incident photon. We find $$U_n = \ee^{-\ii \left(2 \oat t_0 \Sz + n \pphase \Sz + \frac{n}{2} \pphase^2 \Sz^2\right)},$$ where the terms linear in generate a precession of the atomic pseudospin and the term quadratic in generates a shearing or -dependent rotation. To isolate the shearing term it is convenient to apply this transformation twice, separated by a $\pi$ pulse on the atoms which inverts . The overall effect is the two-pulse transformation $$\label{eq:Urhomu} U_{\rho,\mu} = \ee^{-\ii \left(\rho \Sz + \frac{1}{2} \mu \Sz^2 \right)}$$ obtained by the sequence $U_{n_1}$–$\pi$–$U_{n_2}$ with photon numbers $n_1$ and $n_2$. Here $\rho = (n_2 - n_1) \pphase$ is the phase rotation angle and $\mu = (n_2 + n_1) \pphase^2$ is the strength of the shearing action expressed as an atomic phase shift per unit change in . In the ideal case where the incident photon number is identical in the two pulses, $n_1=n_2$, we find the one-axis twisting transformation $U_{0,\mu} = \ee^{-\ii \mu \Sz^2 / 2}$ considered by Kitagawa and Ueda [@Kitagawa1993]. Their results for the spin squeezing achievable by this transformation are reviewed in Appendix \[app:oat\]. Thus we find that the two-fold interaction of an $n$-photon pulse with the atom-cavity system can realize the $\Sz^2$ one-axis twisting Hamiltonian, at least when scattering into free space is ignored (see Sec. \[sec:scatt\]). The unitary evolution in the spin subspace can be understood as a consequence of the absence of photon shot noise, or equivalently, of the absence of phase information in the reflected light that would reveal the atomic state. While the same transformation can be accomplished with a single input pulse followed by a photon-number-dependent rotation of the atomic spin, it is easier to implement using the double-pulse sequence, as we shall see below. Effects of Uncertain Photon Number {#sec:lightnoise} ================================== For input pulses with uncertain photon number, we can find the expectation value of an atomic observable by averaging over the possible photon numbers of the input state. In general, such an average will depend on all moments of the photon number distribution. However, an important special case arises when the phase per photon $\pphase$ is small and the photon numbers are large such that the distributions for $n_1$ and $n_2$ approach Gaussians. In the common case where the fluctuations in the photon number difference $(n_2 - n_1)$ have a variance which scales with the total photon number, the rotation angle variance can be expressed as $\var\rho = \psnfrac \avg\mu$, with a proportionality constant which is 1 for the case of independent photon shot noise fluctuations on $n_1$ and $n_2$. The variance of the shearing $\var{\mu}$, meanwhile, is higher-order in and vanishes in the limit we are considering, so we will not distinguish between $\mu$ and its expectation value $\avg\mu$ hereafter. In this regime the squeezing is determined only by the total number of atoms $2S$, the shearing $\mu$ set by the mean total photon number $\avg{n_1+n_2}$, and the ratio of rotation uncertainty to shearing set by the variance of the photon number difference $n_1-n_2$ between the two pulses. For large $S$, an initial atomic pseudospin polarized along $\unitx$, $\mu \ll 1$, and $\avg{\rho}=0$, the final spin expectation values are approximated as $$\begin{aligned} \avg\Sx &= S \ee^{-\frac{1}{2} \pvar}, \\ \avg\Sy &= \avg\Sz = 0, \\ \avg{\Sy^2} &= \frac{S}{2} \Bigl[1 + S \bigl(1 - \ee^{-2 \pvar}\bigr)\Bigr], \\ \avg{\Sz^2} &= \frac{S}{2}, \\ \avg{\Sy \Sz + \Sz \Sy} &= 2 \avg{\Sz^2} \mu \avg{\Sx} = S^2 \mu \ee^{-\frac{1}{2} \pvar} \label{eq:SySz},\end{aligned}$$ where $\pvar = \psnfrac \mu + \mu^2 \avg{\Sz^2}$ is the characteristic phase variance resulting from both the uncertainty on the photon number and from the shearing. The mean spin length is reduced as the phase broadening wraps the uncertainty region around the Bloch sphere. The transverse variance is initially the projection noise $S/2$, then grows as the phase variance scaled by the length of the Bloch vector $S^2\pvar$, before saturating near $S^2/2$ when the uncertainty distribution has completely wrapped around the Bloch sphere. The cross correlation $\avg{\Sy \Sz + \Sz \Sy}$ is the product of the variance of , the phase change $\mu$ per unit , and the change in per unit phase. Using Eq. (\[eq:Smin\]) to compute the minimum transverse spin variance $\var{\Smin}$ and comparing this variance to the reduced mean spin length gives the metrological squeezing parameter [@Wineland1992; @Wineland1994] $$\sq = \frac{2 S \var{\Smin}}{\avg{\Sx}^2}.$$ $\sq^{-1}$ is the squared signal-to-noise ratio, relative to the standard quantum limit, of a measurement of the total pseudospin’s direction. ![(Color online) Cavity squeezing for $S=10^4$ and varying degrees of photon shot noise suppression: none ($\psnfrac = 1$, dotted red), 90% ($\psnfrac = 0.1$, dashed blue), 99% ($\psnfrac = 0.01$, chain dotted green), and complete ($\psnfrac = 0$, solid black). In this figure, noise due to photon scattering into free space is ignored, corresponding to a cavity with single-atom cooperativity parameter $\eta \gg 1$ (see Sec. \[sec:scatt\]).[]{data-label="fig:gaussian-squeezing"}](squeezing-vs-shotnoise.mps) Figure \[fig:gaussian-squeezing\] shows the squeezing parameter calculated as a function of shearing strength $\mu$ for $S=10^4$ and residual fractions of photon shot noise of 1, 0.1, 0.01, and 0. In general the squeezing parameter first decreases as $\sq \approx 2 \psnfrac / (S \mu)$ under the action of the shearing before rising again as $\sq \approx S^2 \mu^4 / 24$ as the curvature of the Bloch sphere deforms the uncertainty region. In the limit $\psnfrac \rightarrow 0$ we recover the ideal squeezing from perfect one-axis twisting as considered in Ref. [@Kitagawa1993] and Appendix \[app:oat\]: $\sq \approx (S \mu)^{-2} + S^2 \mu^4 / 24$, leading to an optimum squeezing that scales as $\sq \approx 12^{2/3} / (8 S^{2/3})$. Even if the input field contains a definite photon number, imperfections in the resonator can introduce optical loss. Since the loss of photons is a random process, it reintroduces noise on the number of photons which interact with the atoms, and thus decoherence. Such optical loss processes can quite generally be understood as coupling the cavity to a second continuum of modes into which the photons are scattered. For example, we might imagine a second continuum of modes $\bfield\omega^\prime$ behind the right-hand cavity mirror of Fig. \[fig:setup\](b)(i) and then allow this mirror to be partly transparent. Since the cavity field couples to a weighted sum of the incident field mode and the various independent unobserved modes $\bfield\omega^\prime$ we can, following Fano, identify a single linear combination of all the field modes at a given frequency that couples maximally to the cavity mode and one or more orthogonal combinations (as many as there are unobserved fields) that do not couple to the cavity mode at all [@Fano1961] (see Appendix \[sec:sagnac\] for a practical consequence of this separation into coupled and uncoupled modes). If the average fraction lost of the input pulse at one half-linewidth detuning is $\loss \ll 1$, then the probability of an incident photon being in one of these uncoupled modes and failing to interact with the cavity mode is $\frac{1}{2} (1 - \sqrt{1 - 2 \loss}) \approx \loss / 2$, yielding a binomial distribution for the cavity-coupled photons whose variance is $\psnfrac \approx \loss / 2$ times photon shot noise. In other words, the noise in the photon number difference $n_2 - n_1$ to which the squeezing is sensitive is just the shot noise of the lost photons. Practical Photon Shot Noise Suppression Schemes {#sec:schemes} =============================================== In this section we show how to achieve reduced photon shot noise $\psnfrac<1$ using several schemes of practical interest. Spin Echo Quantum Eraser for Coherent Input Pulses -------------------------------------------------- If the input consists of two independent coherent pulses (Fig. \[fig:double-pulses\], top), $\var{(n_2 - n_1)} = \avg{n_2} + \avg{n_1}$ and $\psnfrac = 1$. In this case we recover the scaling obtained by Takeuchi *et al.* [@Takeuchi2005] for their polarization-feedback scheme in free space: $\sq \approx 2 / (S \mu) + S^2 \mu^4 / 24$, with the best squeezing, obtained for a shearing parameter $\mu \approx 12^{1/5} / S^{3/5}$, scaling as $\sq \approx 5 / (384^{1/5} S^{2/5})$. Note that these results do not take into account free-space scattering, which can be ignored only if a cavity with cooperativity $\eta>1$ is used (see Sec. \[sec:scatt\]). ![(Color online) Cavity squeezing with coherent pulses. Previous experimental demonstrations of cavity squeezing [@Leroux2010:dynsq] used two separate coherent pulses with statistically independent photon numbers $n_1$ and $n_2$ (top). If, instead, a single coherent pulse is used for both interactions with the cavity (bottom), the two photon numbers $n_1$ and $n_1^\prime$ differ only by the optical losses between the two interactions and the effect of their correlated fluctuations can be canceled out by a spin-echo sequence.[]{data-label="fig:double-pulses"}](double-pulses.mps) However, if the same coherent pulse is reused for both interactions by storing it in an optical delay line while the $\pi$ pulse is applied to the atoms (Fig. \[fig:double-pulses\], bottom), then ideally $n_2 = n_1$ and $\rho=\psnfrac = 0$. Information about is encoded in the optical phase shift of the coherent pulse after its first reflection from the cavity, but this is erased during the second reflection, which applies the opposite phase shift since has changed sign in the interim. Thus the light is disentangled from the atoms at the end of the sequence and the overall evolution of the atoms is unitary. In a real implementation there will be losses in the delay line used to store the pulse between its two interactions with the cavity, such that $n_2$ will not be precisely equal to $n_1$ and $\rho$ will not exactly vanish. If a fraction $\loss$ of the photons is lost between the two pulses, then $\var{(n_2 - n_1)} = \avg{n_1} \loss \approx (\avg{n_2} + \avg{n_1}) \loss / 2$, where the second approximation holds for small losses, giving $\psnfrac \approx \loss/2$. Again, the squeezing is limited by the shot noise of the lost photons. Squeezed Input Pulses --------------------- One demonstrated approach to improving optical atom detection for a given number of photons is to use a squeezed state of the input light field [@Wolfgramm2010]. We therefore consider the effect of incident pulses whose fluctuations in-phase with the coherent amplitude $\alpha$ are squeezed so as to reduce intensity noise in the cavity. The photon number variance of such pulses can be parametrized as [@Yuen1976; @Loudon1987] $\var{n} = \abs\alpha^2 \ee^{-2 s} + 2 \sinh^2 s \cosh^2 s$ where $\alpha$ is the coherent amplitude of the pulse. The usual optical squeezing parameter $s$ is related to the shot noise suppression factor by $\psnfrac = \ee^{-2 s} + (2 \sinh^2 s \cosh^2 s) / \avg{n_2 + n_1}$. Note that $\psnfrac$ does not decrease monotonically with $s$ but reaches a minimum which improves with photon number as $\psnfrac \sim (n_2 + n_1)^{-1/3}$, because squeezing reduces the fluctuations in a quadrature of the incident electric field rather than in its magnitude. Delivering such squeezed light to the cavity in the presence of optical losses remains experimentally challenging, but sources providing -squeezed light to gravitational observatories have been demonstrated [@Vahlbruch2010], so that suppression factors of order $\psnfrac\sim10^{-1}$ may be attainable by this approach. Generation of Effective Fock States by Measurement {#sec:Fock} -------------------------------------------------- Another approach to suppressing photon shot noise relies on the conservation of total photon number in this scheme: every photon sent onto the cavity must leave it and, for a single-ended cavity with negligible loss, every photon leaves in the same reflected mode. This identity between photon numbers of the input and output fields allows an effective input Fock state to be produced using high-efficiency photon counters and classical feedback. Such feedback-generated Fock states were studied in the early days of light squeezing research [@Machida1986] but have not seen wide use because, in the absence of a QND photodetector, the photon Fock state is destroyed in the very detection process that generates it. However, since the spin-squeezing setup does not change the photon number we may place it inside the feedback loop, before the photodetector (Fig. \[fig:fock-projection\]), thus sidestepping this difficulty. ![(Color online) By counting photons after their interaction with the cavity, the squeezing light pulse can be projected onto a definite photon number state. With fast classical feedback it can even be steered to a predetermined number state for unconditional photon shot noise suppression.[]{data-label="fig:fock-projection"}](fock-projection.mps) In the simplest scheme, a coherent laser pulse is sent onto the cavity and the reflected light is collected by a photon counter. If a perfect photon counter detects $n$ photons in the light pulse reflected from the cavity, then the system is projected into the state obtained for an incident $n$-photon Fock state. The photon-counting measurement has removed the uncertainty on the energy of the incident light pulse and has destroyed the complementary information in the phase of the light field, which would have revealed the atomic state via the cavity frequency shift. Conditioned on the reflected photon number, the atomic dynamics are unitary: although the evolution of the spin state depends *a priori* on the uncertain photon number in the incident pulse, *a posteriori* the experimenter knows exactly which unitary operation was performed by the light pulse whose photon number was measured. Note that no useful information can be obtained from the photon arrival times at the counter. Since the spectrum of the incident light must be much narrower than the cavity linewidth, the arrival time of the photons has a Fourier-limited uncertainty much larger than the cavity lifetime and one cannot determine whether a photon entered the resonator or merely bounced off the input mirror. Rather than contenting oneself with conditional unitary evolution, one can deterministically generate the equivalent of an input Fock state by applying direct feedback to the incident light. One must merely count reflected photons and switch the light source off once some target photon number $n$ has been reached. Finite photodetector quantum efficiency will introduce an uncertainty on the number of input photons for a given detected photon count. The residual photon shot noise obtained by this technique will therefore be at best $\psnfrac = 1-Q$, where $Q$ is the quantum efficiency of the photodetector. Note that a transmission-based QND measurement of the atomic spin as used in Ref. [@Schleier-Smith2010:msmt] can squeeze initially as $\sq \approx 2 / (S \mu Q)$, which is slower than the $\sq \approx 2 \gamma / (S \mu)$ of cavity squeezing for any finite quantum efficiency. Even a perfect photon-shot-noise-limited phase measurement of the light reflected from a one-sided cavity could squeeze only as $\sq \approx 1 / (4 S \mu Q)$, so that for $Q\gtrsim\unit[85]{\%}$ conditional squeezing by measurement still proceeds more slowly than cavity-feedback squeezing using the same photodetector to suppress photon shot noise fluctuations. Effects of Scattering Into Free Space {#sec:scatt} ===================================== So far, we have neglected scattering of photons into free space. Like atom loss in collisional squeezing of Bose-Einstein condensates [@Sinatra2011], such photon loss degrades the performance of light-induced spin-squeezing schemes [@Hammerer2004; @Madsen2004] unless a suitable level scheme is used to avoid its effects [@Saffman2009]. In the simple and symmetric model we consider, the average number of photons scattered into free space per atom in the ensemble is given by $$2\halfscatt = (n_1 + n_2) \frac{\pphase}{2} \frac{\Gamma}{\Delta} = \frac{\mu}{2 \eta}$$ where the numbers of photons Raman- and Rayleigh-scattered are equal to each other and to $\halfscatt$. The scattering depends only on the shearing $\mu$ and the cavity cooperativity $\eta = 4 g^2 / (\kappa \Gamma)$, so that for any finite single-atom cooperativity $\eta$ scattering into free space is inescapable at any detuning $\Delta$. Scattering leads directly to atomic decoherence by revealing the state of certain atoms to a hypothetical observer outside the cavity and, in the case of Raman scattering, by randomly flipping some spins in the ensemble. Here we apply the treatment of these effects given in Ref. [@Schleier-Smith2010:dynsq] to the case of unitary cavity feedback. All photons scattered into free space reveal the internal state of the scattering atom to a hypothetical observer. For Raman scattering, the state is encoded in the frequency of the scattered light. For Rayleigh scattering, it is encoded in the phase of the scattered field, because the laser detuning from resonance has opposite sign for the two spin states [@Uys2010]. Any atom which scatters a photon into free space therefore acquires an unknown phase, entangled with the information lost in the scattered field and uncorrelated with that of the other atoms in the ensemble. The mean length of the Bloch vector is thus reduced from $S$ by a factor of $\contrast = \ee^{-2 \halfscatt}$ corresponding to the fraction of atoms which have scattered no photons. Raman scattering, in addition, modifies the relative population of $\upk$ and $\downk$, and forces us to distinguish between the spin component $\Sz$ found at the end of the squeezing and its time-averaged value $\Szbar$ during the squeezing process. The distribution of $\Sz$ is unaltered by the Raman scattering, since it already corresponded to the sum of independent random $\pm1/2$ contributions from the uncorrelated atoms in the initial state. But since spins which flip partway through the squeezing pulse contribute less to the time average, $\avgSzbarsq$ is reduced to $(1 - 2 \halfscatt / 3) S / 2$, to leading order in $\halfscatt$. Since it is $\Szbar$ which sets the average atom-induced shift of the light intensity inside the cavity, and thus the phase shift imparted to the atoms by the Stark effect, the phase variance in turn is reduced to $\pvar^\prime \approx \psnfrac \mu + \mu^2 \avgSzbarsq$. Similarly, the factor of $\avg{\Sz^2}$ in the $\Sy$–$\Sz$ correlation \[Eq. (\[eq:SySz\])\]—which expressed the correlation between the $\Sz$ found at the end of the squeezing and the atom-induced change to the light shift that modifies $\Sy$—becomes $\avg{\Sz \Szbar} \approx (1 - \halfscatt) S / 2$, again to leading order in $\halfscatt$. Combining these effects, we find an adjusted set of spin moments $$\begin{aligned} \avg\Sx &= S \contrast \ee^{-\frac{1}{2} \pvar^\prime}, \\ \avg\Sy &= \avg\Sz = 0, \\ \avg{\Sy^2} &= \frac{S}{2}, \Bigl[1 + S \contrast^2 \bigl(1 - \ee^{-2 \pvar^\prime}\bigr)\Bigr], \\ \avg{\Sz^2} &= \frac{S}{2}, \\ \avg{\Sy \Sz + \Sz \Sy} &= S^2 (1 - \halfscatt) \contrast \mu \ee^{-\frac{1}{2} \pvar^\prime}.\end{aligned}$$ Note that photon number fluctuations due to atomic absorption can be neglected because, while the fraction of atoms which scatter a photon is fixed by $\mu$ and $\eta$, the fraction of photons scattered vanishes in the large-photon-number limit considered here. The resulting squeezing is plotted in Fig. \[fig:squeeze-w-scatt\] for different cavity cooperativities $\eta$ and for perfect photon shot noise suppression ($\psnfrac=0$). ![(Color online) Cavity squeezing for $S=10^4$ with input pulses of definite photon number for a perfect cavity (solid black), and for finite single-atom cooperativities $\eta=1$ (chaindotted green), $\eta=0.1$ (dashed blue) and $\eta=0.01$ (dotted red). For reference, the gray curve shows squeezing for a perfect cavity ($\eta\rightarrow\infty$) without photon shot noise suppression ($\psnfrac=1$).[]{data-label="fig:squeeze-w-scatt"}](squeezing-vs-scattering.mps) For the weak- and moderate-coupling regimes where scattering is the dominant limitation on squeezing ($\eta < 1$), we find $\sq \approx (S \mu)^{-2} + \mu / (3 \eta)$. As in the ideal case, the squeezed variance is initially suppressed by the square of the squeezing parameter, but the noise from scattering into free space adds a variance which scales linearly with the shearing. This leads to an optimum squeezing $\sq \approx 6^{1/3} / 2 (S \eta)^{2/3}$ for a shearing parameter $\mu \approx (6 \eta)^{1/3} / S^{2/3}$. Note that $4 S \eta$ corresponds to the resonant optical depth of the atomic ensemble probed through the cavity, and that the achievable squeezing therefore scales as optical depth to the $-2/3$. This is the same scaling reported by Trail *et al.* for their analogous polarization-based spin-squeezing scheme [@Trail2010]. The dashed curve of Fig. \[fig:squeeze-w-scatt\] shows the squeezing achievable with photon shot noise suppression in a setup otherwise similar to that used in Ref. [@Leroux2010:dynsq], using $2S=2\times 10^4$ atoms of $\Rb$ ($\Gamma / \abs\Delta = 1.8\times 10^{-3}$) in a resonator with a single-atom cooperativity for the relevant transitions of $\eta=0.1$ so that $\pphase = \eta\Gamma / \abs\Delta = 1.8\times 10^{-4}$. For a shearing parameter of $\mu=1.8\times 10^{-3}$ corresponding to a photon number of $2.7\times 10^4$ in each incident pulse, the squeezing reaches , a substantial improvement over the achievable in the same system without photon shot noise suppression. For the strong-coupling regime $\eta \gg 1$ the curvature becomes significant before scattering can decohere the ensemble, and the ideal squeezing behavior of Sec. \[sec:ideal\] is restored. Note that the scaling of the achievable squeezing with atom number (as $S^{-2/3}$) is the same for finite $\eta$ as it is in this ideal limit of $\eta \rightarrow \infty$. Once the effect of photon shot noise has been suppressed, the scattering into free space costs only a constant factor in squeezing performance. Conclusion ========== In this paper we have shown how to improve cavity squeezing performance by disentangling the atomic variables from the light field which mediates the interatomic interaction. We have suggested several ways of doing this, including a spin-echo sequence that erases the phase information in a coherent light pulse, and the use of photodetectors and classical feedback to generate effective Fock states of the input field. Once the entanglement between atoms and outgoing light field is eliminated, even a moderate cavity cooperativity $\eta \sim 1$ suffices to obtain squeezing performance close to the limit set by the curvature of the Bloch sphere. This limit, in turn, could be overcome by two-axis counter-twisting [@Kitagawa1993], which can be realized by alternating periods of one-axis twisting with rotations of the atomic spin [@Liu2011]. Furthermore, since unitary $\Sz^2$ evolution in combination with rotations suffices, in principle, to implement any unitary map on the Bloch sphere [@Chaudhury2007], the techniques we have presented could enable the production of non-Gaussian entangled states of ensembles comprising tens of thousands of atoms. Further studies are needed to determine which states are attainable given realistic experimental imperfections. I.D.L. acknowledges support from NSERC; M.H.S.-S. acknowledges support from the Hertz Foundation and the NSF. This work was supported by the NSF, DARPA, and ARO. Squeezing by One-Axis Twisting {#app:oat} ============================== This appendix summarizes the results for squeezing by one-axis twisting first obtained by Kitagawa and Ueda [@Kitagawa1993]. In order to prepare a squeezed state, we begin with the atoms in a totally symmetric but unentangled coherent spin state (CSS) along the axis of pseudospin $$\begin{aligned} \ket{\psi_a} &= \left(\frac{\upk + \downk}{\sqrt{2}}\right)^{\otimes N} \\ &= \sum_{m = -S}^S\sqrt{\frac{1}{2^{2 S}}\binom{2 S}{S + m}}\ket{m}.\end{aligned}$$ The second form explicitly shows the CSS’s binomial distribution of eigenvalues. In this state $\avg{\Sx} = S = N / 2$, $\avg{\Sy} = \avg{\Sz} = 0$, $\avg{\Sy \Sz + \Sz \Sy} = 0$ and $\avg{\Sy^2} = \avg{\Sz^2} = S / 2$. Afer the transformation $U_{\rho,\mu}$ defined in Eq. (\[eq:Urhomu\]), the distribution is unmodified but the phases between levels have acquired a quadratic dependence on $m$. The expectation values become $$\begin{aligned} \avg\Sx &= S \cos (\rho) \cos^{2 S - 1} \left(\frac{\mu}{2}\right), \\ \avg\Sy &= S \sin (\rho) \cos^{2 S - 1} \left(\frac{\mu}{2}\right), \\ \avg{\Sy^2} &= \frac{S}{2} \Bigl[1 + \Bigl(S - \frac{1}{2}\Bigr), \bigl(1 - \cos (2 \rho) \cos^{2 S - 2} (\mu)\bigr)\Bigr],\end{aligned}$$ and $$\begin{aligned} \avg{\Sy \Sz + \Sz \Sy} &= S (2 S - 1) \cos (\rho) \times \notag \\ &\qquad \times \sin \left(\frac{\mu}{2}\right) \cos^{2 S - 2} \left(\frac{\mu}{2}\right).\end{aligned}$$ For $\rho = 0$ the mean spin remains aligned along and the minimum variance transverse to this direction is given by $$\var{\Smin} = \frac{1}{2} \left(u_+ - \sqrt{u_-^2 + \avg{\Sy \Sz + \Sz \Sy}^2}\right) \label{eq:Smin}$$ with $u_\pm = \avg{\Sy^2} \pm \avg{\Sz^2}$. For large $S$ and in the region of significant squeezing $S^{-1} \ll \mu \ll S^{-1/2}$ the squeezing parameter is approximately $\sq \approx (S \mu)^{-2} + S^2 \mu^4 / 24$, decreasing under the action of the shearing until the curvature of the Bloch sphere deforms the uncertainty region. The best squeezing obtained is $\sq \approx 12^{2/3} / 8 S^{2/3}$ for a shearing parameter $\mu \approx 12^{1/6} / S^{2/3}$. Simulating a One-Sided Cavity {#sec:sagnac} ============================= Throughout this paper we have considered a single-ended cavity. Many cavity-QED experiments find it convenient to use two-sided Fabry-Pérot resonators. Such two-sided cavities would mix a Fock state input from one end with vacuum fluctuations admitted through the other end of the cavity, restoring much of the photon shot noise we wish to suppress. Equivalently, information on the cavity detuning (and hence the atomic state) is available in the ratio of transmitted to reflected photon numbers, and this information leak entails atomic decoherence. Fortunately, it is possible to convert a symmetric cavity into an effective one-sided cavity using only external optics. ![(Color online) Equivalence of one-sided (top left) and two-sided (top right) cavities: the left and right ports of a symmetric Fabry-Pérot resonator can be combined on a beam splitter to isolate the linear combination of the two fields that couples to the intracavity field (bottom right). Conversely, the single input of a one-sided cavity can be mixed with an auxiliary mode to yield an effective two-sided resonator (bottom left).[]{data-label="fig:cavity-conversion"}](one-two-sided-cavities){width="\linewidth"} Figure \[fig:cavity-conversion\] illustrates the principle of this conversion. For any given frequency, transverse mode, and polarization, the symmetric cavity couples to two spatially separated input fields (right and left). Since there is only one cavity mode near the given frequency with the given transverse mode and polarization, it must couple only to some linear combination, labeled here as $b^\dagger$, of the two input fields. The orthogonal combination $d^\dagger$ does not couple to the cavity mode at all. Classically, $b^\dagger$ ($d^\dagger$) corresponds to simultaneous illumination from left and right with phases chosen so as to give constructive (destructive) interference within the resonator. Enclosing the Fabry-Pérot in a Sagnac interferometer overlaps the left and right fields at the input, so that with appropriately chosen path lengths the maximally-coupled superposition $b^\dagger$ is isolated from the uncoupled mode $d^\dagger$. To an observer looking into the $b^\dagger$ port of the input beam splitter, the apparatus appears to be a single-ended cavity which couples to no other field modes. The $d^\dagger$ input, uncoupled from the cavity, remains available for interferometer stabilization.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We predict a novel temperature-driven phase transition of DNA below the melting transition. The additional, intermediate phase exists for repetitive sequences, when the two strands have different lengths. In this phase, the excess bases of the longer strand are completely absorbed as bulge loops inside the helical region. When the temperature is lowered, the excess bases desorb into overhanging ends, resulting in a contour length change. This continuous transition is in many aspects analogous to Bose Einstein condensation. Weak sequence disorder renders the transition discontinuous.' author: - 'Richard A. Neher' - Ulrich Gerland title: An intermediate phase in DNA melting --- The base-pairing interaction between the two strands of DNA is not only pivotal to its biological function [@Alberts:02], but also leads to intriguing applications in nanotechnology [@Seeman:03]. One approach to probe this interaction is to monitor the DNA conformation as a function of temperature. Experimentally, one can observe the number of basepairs formed (using UV absorption [@Wartell:85; @SantaLucia:04]), as well as changes of intra-molecular distances on the nanometer scale (using modern single-molecule techniques [@Zhuang:03]). On the theoretical side, the temperature dependence of DNA conformations has been studied for almost fifty years, using models of various degrees of complexity [@Zimm:59; @Hill:59; @Poland:66b; @Peyrard:89; @Cule:97; @Kafri:00; @Theodorakopoulos:00; @Garel:04]. Particular attention has been paid to the characteristics of the melting transition, where the two strands separate completely. Whereas early models yielded only a crossover [@Zimm:59], the Poland-Scheraga (PS) model [@Poland:66b] was the first to display a phase transition, albeit a continuous one, which appeared to be at variance with the experimentally observed sharp jump in the fraction of bound basepairs [@Wartell:85]. Only recently have mechanisms been proposed [@Kafri:00; @Theodorakopoulos:00] which yield an abrupt, first order transition. So far, however, most analyses of DNA melting have incorporated only native interactions, i.e. base pairs that occur in the ground state of the molecule (see [@Hill:59; @Garel:04] for notable exceptions). It is our aim here to show that such non-native interactions can introduce an intermediate phase in the melting behavior of DNA, associated with an additional conformational transition before strand separation. Non-native interactions are particularly relevant for repetitive DNA sequences, which are common in genomes [@Lovett:04]. Periodic DNA, with e.g. a single base repeat such as `TTT…` or a higher order repeat such as `CAGCAG…`, can take on basepairing patterns with asymmetric loops and the two complementary strands can be shifted relative to each other. Here, we consider the general situation where the two strands can have arbitrary lengths $N$, $M$. We describe the DNA using a generalized PS model [@Garel:04] and calculate its equilibrium behavior analytically. We find that for $N\neq M$, the bound phase splits into two separate phases. The low temperature phase is characterized by an extensive length of the unbound end on the longer strand, whereas in the new intermediate phase these overhanging bases are absorbed into the helical region. Mathematically, and also conceptually, many aspects of this transition are analogous to Bose-Einstein condensation (BEC), as “particles” (bases) condense into a single “state” (the overhanging end), which thereby acquires macroscopic “occupation” (length). Obviously, the analogy extends only to the behavior of the partition function, as there is no quantum coherence in the DNA problem. Effectively, the transition amounts to a temperature sensitive change in the contour length of the DNA molecule, which should be observable with optical or single molecule methods. While the transition is continuous for perfectly periodic sequences, we find that it becomes a first order transition once (weak) sequence disorder is introduced. We also show that the non-native interactions can change the order of the melting transition, as has been conjectured previously [@Kafri:02b]. ![\[fig:melting\_setup\] A possible configuration of two complementary DNA strands with a repetitive sequence (a bead represents one repeat unit). Note that repetitive sequences can form base pairing patterns with asymmetric loops. In general we allow for different strand lengths $N$, $M$. The last repeat units (squares) are permanently bound. ](fig1){width="\columnwidth"} [*DNA model.—*]{} We consider two DNA strands with lengths $N$ and $M\geq N$, respectively, and describe their interaction with a generalized PS model [@Garel:04; @Neher:04]. Specifically, a base $i\le N$ of the lower strand can form a base pair $(i,j)$ with every complementary base $j\le M$ of the upper strand, whereas the formation of base pairs within a strand can be neglected (since we are interested only in sequences with a high degree of complementarity and a low degree of self-complementarity). Due to geometrical constraints, we may neglect the ‘crossing’ of base pairs, e.g. two base pairs $(i_1,j_1)$ and $(i_2,j_2)$ with $i_1<i_2$ but $j_1>j_2$. The basepairing pattern ${{\cal S}}$, i.e. the set of all formed base pairs, then creates a DNA conformation consisting of bound segments alternating with (possibly asymmetric) loops, see [Fig. \[fig:melting\_setup\]]{}. To simplify the discussion, we enforce the base pair $(N,M)$ at the right end, so that we need to consider only one overhanging end. Experimentally, this boundary condition would be realized e.g. by a few particularly strong basepairs at one end. To each basepairing pattern ${{\cal S}}$, we assign a statistical weight ${{\cal Q}}({{\cal S}})$, which takes the form of a product with factors of four different types: (i) a Boltzmann factor $q=e^{{{\varepsilon_{\rm b}}}/k_BT}$ for every basepair with binding energy $-{{\varepsilon_{\rm b}}}<0$, (ii) a Boltzmann factor $g^2=e^{-{{\varepsilon_{\ell}}}/k_BT}$ for every loop with loop initiation cost ${{\varepsilon_{\ell}}}>0$, (iii) an entropic factor ${{B_{\ell}}}(m)=s^m m^{-c}$ for each loop, which is the increase in the number of polymer configurations when $m$ bases form a (floppy) loop instead of being in a (rigid) double helical conformation, (iv) and a similar entropic factor ${A}(n)=s^n n^{-\bar{c}}$ for a single-stranded end of $n$ bases. Here, the exponents $c,\bar{c}$ in the entropic factors are universal in that they are independent of the detailed polymer properties, but are sensitive to excluded volume interactions. For interacting self-avoiding loops one has $c\approx2.15$, while $\bar{c}\approx 0.1$ [@Kafri:02b]. Whereas the value of $c$ determines the critical behavior at the melting transition [@Kafri:00], the non-universal constant $s$ has no qualitative effect on the melting behavior (we use $s=10$ in all numerical examples). In the following, we apply the DNA model to perfectly periodic sequences, where each repeat unit can be treated as an effective base with renormalized parameters (we use ${{\varepsilon_{\rm b}}}=6$ and ${{\varepsilon_{\ell}}}=3$ in temperature units, $k_B=1$). We emphasize that our simplistic model for the involved energies and entropies is meant to illustrate the physical phenomena in a transparent way, but leads to an unrealistic temperature scale. With a more detailed description [@SantaLucia:04], we find that all of the interesting behavior happens at accessible temperatures [@future]. [*Free energy of periodic DNA.—*]{} To obtain the equilibrium properties of the DNA model, we calculate the partition sum over all basepairing patterns, $Z_N^M=\sum_{{{\cal S}}}{{\cal Q}}({{\cal S}})$. By separating the single stranded ends from the double stranded part, see [Fig. \[fig:melting\_setup\]]{}, we write $Z_N^M$ as $$\label{eq:completeZ} Z_N^M=\sum_{i=0}^{N-1}\sum_{j=0}^{M-1}{A}(i){A}(j)\,W_{N-i}^{M-j}\;.$$ Here, $W_{r}^t$ is the partition function of two complementary and periodic strands of length $r$ and $t$ with the first and last base pair formed. $W_{r}^t$ obeys the recursion relation $$\label{eq:recursion} W_{r+1}^{t+1}=qW_r^t+g^2q\sum_{k+m>0}^{k<r,m<t}{{B_{\ell}}}(k\!+\!m)W_{r-k}^{t-m},$$ with the initial conditions $W_{1}^1=q$ and $W_{1}^{i}=W_{i}^1=0$ for $i>1$ [@Garel:04; @Neher:04]. Eqs. (\[eq:completeZ\]) and (\[eq:recursion\]) can be used to calculate $Z_N^M$ for finite lengths $N$, $M$. To extract the thermodynamic behavior in the limit of long strands, we take the $z$-transform $\hat{Z}(x,y)=\sum_{N,M=0}^\infty Z_N^M\,x^Ny^M$. Compared to the related case of a single self-complementary RNA strand folding back onto itself [@Bundschuh:99; @deGennes:68], we need two instead of one transformation variables here, due to the second strand of DNA. One obtains $$\label{eq:Z_melting} {\hat{Z}}(x,y)=\frac{{\hat{A}}(x){\hat{A}}(y)qxy}{1-qxy+\frac{qg^2xy}{x-y}(y{\hat{B}}(y)-x{\hat{B}}(x))}\;,$$ where the transforms of the entropic factors are given by ${\hat{A}}(z)=\phi_{\bar{c}}(sz)+1$ and ${\hat{B}}(z)=\phi_c(sz)$, with the polylogarithm $\phi_c(z)=\sum_{n=1}^\infty z^n n^{-c}$. The $z$-transformation carried out above amounts to a change from the canonical to the grand canonical ensemble. The transformation variables $x,y$ play the role of fugacities for bases in the lower and upper strand, respectively. However, for the ensuing discussion, it is advantageous to keep the length $N$ of the shorter strand fixed as a reference. Hence, we perform the inverse transformation for the lower strand by contour integration in $x$, see [Fig. \[fig:contourint\]]{}, to obtain the partition sum $Z_N(y_0)$ for $N$ bases on the lower strand and the upper strand coupled to a “nucleotide reservoir” with fixed fugacity $y_0$. Whenever both strands are bound, ${\hat{Z}}(x,y)$ has a singularity at ${{x^*}}(y_0)<s^{-1}$, see [Fig. \[fig:contourint\]]{}. For large $N$, the contour integration is dominated by the residue at ${{x^*}}(y_0)$, leading to $Z_N(y_0)={\hat{A}}(y_0){{x^*}}(y_0)^{-N}$. Hence, the free energy of the bound phase is given by $N f_b(y_0)-{T}\ln{\hat{A}}(y_0)$, where the first term is the contribution of the helical region with a free energy per length $f_b(y_0)={T}\ln{{x^*}}(y_0)$, and the second term is the contribution from the unbound end of the longer strand. ![\[fig:contourint\] The contour integration required for the inverse transformation of ${\hat{Z}}(x,y_0)$ is given by the sum of the integral along the branchcut $[s^{-1},\infty]\subset\mathbb{R}$ and the integral encircling the singularity of ${\hat{W}}(x,y_0)$ at $x={{x^*}}(y_0)$. This isolated singularity only exists below the melting temperature. ](fig2){width="0.55\columnwidth"} The free energy for given $N$ and $M$ is then obtained by saddle point integration, $$\label{eq:freeenergy} \frac{F(T,N,M)}{{T}}=-\ln{\hat{A}}(y_0)+ N \frac{f_b(y_0)}{{T}}+M\ln(y_0),$$ where the fugacity $y_0$ is determined by $$\label{eq:meanlength} \begin{split} M=\langle M \rangle_{y_0}&= y_0 \frac{\partial \ln{\hat{A}}(y_0)}{\partial y_0}-N\frac{y_0}{{T}}\frac{\partial f_b(y_0)}{\partial y_0}\;. \end{split}$$ ![\[fig:condensate\] Top: The length of the unbound end, normalized by the total length $N$ of the shorter strand. For finite systems ($N=1000$, dashed line), the unbound end shrinks to a minimal value and increases again, as the melting temperature is approached. Monte Carlo simulation data (circles) agrees well with the analytical result. In the $N\to\infty$ limit, the overhang length diverges below $T_c$ and is of order 1 for $T>T_c$. Bottom: The fraction of bound basepairs $\theta$ as a function of temperature. For periodic sequences with $c=2.15$, $\theta$ vanishes with zero slope, whereas a random sequence shows a first order phase transition. When increasing $c$ to 3.15, the periodic sequence displays a similar first order transition. ](fig3){width="0.9\columnwidth"} [*Phase diagram.—*]{} To extract the physical behavior of the DNA model from Eqs. (\[eq:freeenergy\]) and (\[eq:meanlength\]), we focus on two observables, the total number of base pairs, $N\theta$, and the length of the single-stranded overhang. The fraction $\theta$ of bound base pairs is calculated from the free energy per length of the helical region as $$\label{eq:theta} \theta= -\frac{q}{{T}} \frac{\partial f_b(y_0)}{\partial q}\;.$$ To obtain the overhang length, we note that the right hand side of (\[eq:meanlength\]) decomposes the total length $M$ of the upper strand into two contributions, where the first term is the expected overhang length and the second term corresponds to the number of bases in the helical region. The dashed line in [Fig. \[fig:condensate\]]{} (top) shows the overhang length as a function of temperature, for $N=1000$ and $M=1150$. At low temperatures, the two DNA strands are completely aligned, so that all $M-N$ excess bases of the longer strand form an overhanging end. However, we observe that the overhang length decreases with increasing temperature, dropping almost to zero before it rises again sharply at even higher temperature. We see in [Fig. \[fig:condensate\]]{} (bottom) that this drop occurs in a temperature range where almost all possible base pairs are formed, and the rise occurs when the two strands separate. These observations suggest that a temperature-driven conformational transition occurs before the melting transition. This transition is in fact completely analogous to BEC, as [Eq. (\[eq:meanlength\])]{} parallels the behavior of the equation of state for an ideal Bose gas: If we divide [Eq. (\[eq:meanlength\])]{} by our system size $N$ and introduce the “particle density” $\alpha=M/N$, we obtain $$\label{eq:alpha} \alpha=\frac{1}{N}\,\frac{\phi_{\bar{c}-1}(s\,y_0)}{\phi_{\bar{c}}(s\,y_0)+1}+{\bar{\alpha}}(y_0) \;,$$ where ${\bar{\alpha}}(y_0)=-\frac{y_0}{{T}}\frac{\partial f_b(y_0)}{\partial y_0}\geq 1$ is the density inside the helical region. In [Eq. (\[eq:alpha\])]{}, the first term on the right hand side corresponds to the occupation of the ground state of an ideal Bose gas, whereas ${\bar{\alpha}}(y_0)$ is analogous to the occupation of the excited states. The fugacities of a Bose gas and our DNA are bounded: for the former, by the energy of the ground state, and for the DNA by the weight of an unbound monomer, i.e. $y_0\le s^{-1}$. The population of the excited states increases monotonically with the fugacity, and attains a finite maximal value, in our case ${{\bar{\alpha}_{max}}}={\bar{\alpha}}(s^{-1})$ (provided the loop exponent $c>2$). If the density $\alpha$ exceeds this maximal value the length of the unbound end has to diverge in order to accommodate the remaining bases. In other words, the length of the unbound end becomes extensive. In an analogous way, the ground state of a Bose gas is macroscopically populated at low temperatures. In this “condensate” phase, the fugacity is locked to the value $s^{-1}$ in the thermodynamic limit ($N,M\rightarrow\infty$, ${\alpha}=const.$). The deviation for finite systems scales as $s^{-1}-y_0\sim 1/N$, see [Fig. \[fig:bec\]]{} (left). In the opposite case, where $\alpha<{{\bar{\alpha}_{max}}}$, there is a solution to [Eq. (\[eq:alpha\])]{} with $y_0<s^{-1}$ and the unbound end remains finite for all system sizes. ![\[fig:bec\] Left: The fugacity $y_0$ vs. $T$ for different system sizes $N$. In the thermodynamic limit, $y_0=s^{-1}$ for $T<T_c$. As for BEC, $y_0$ approaches its limiting value as $s^{-1}-y_0\sim 1/N$. Right: Phase diagram of periodic DNA. At low temperatures, both strands are completely aligned and excess bases of the longer strand form an unbound end. In the intermediate phase, all excess bases are absorbed into the helical region. ](fig4a "fig:"){width="0.5\columnwidth"} ![\[fig:bec\] Left: The fugacity $y_0$ vs. $T$ for different system sizes $N$. In the thermodynamic limit, $y_0=s^{-1}$ for $T<T_c$. As for BEC, $y_0$ approaches its limiting value as $s^{-1}-y_0\sim 1/N$. Right: Phase diagram of periodic DNA. At low temperatures, both strands are completely aligned and excess bases of the longer strand form an unbound end. In the intermediate phase, all excess bases are absorbed into the helical region. ](fig4b "fig:"){width="0.44\columnwidth"} It is easily shown that ${{\bar{\alpha}_{max}}}$ approaches 1 at low temperatures, and consequently all excess bases of the longer strand are condensed in the overhang, as illustrated in [Fig. \[fig:bec\]]{} (right). As $T$ increases, more and more bases are absorbed in the helical region (${{\bar{\alpha}_{max}}}$ increases), and the system enters the intermediate phase at $T=T_c$, where ${{\bar{\alpha}_{max}}}={\alpha}$. At $T_c$ the condensate fraction vanishes, as the solid line shows in [Fig. \[fig:condensate\]]{} (top). If $T$ is raised to the melting temperature $T_m$, which is independent of ${\alpha}$, the strands separate and $\theta$ vanishes (denatured phase). Note that the intermediate phase exists only when ${\alpha}$ is not too large. It has been previously predicted [@Kafri:02b] that the loop exponent $c$ is effectively reduced by one for periodic sequences compared to the standard PS-model with native base pairs only. This prediction is explicitly confirmed by our exact calculation of the free energy. We find [@future], that there is no melting transition if $c\leq 2$, that the transition is continuous if $2< c \leq3$ and of first order if $c>3$. For $2< c\leq 3$, we obtain $\theta\sim |T-T_m|^{\frac{3-c}{c-2}}$, using the same method as [@Fisher:66] for the standard PS-model. To illustrate this, we plot $\theta$ for periodic sequences and for the standard PS-model in [Fig. \[fig:condensate\]]{} (bottom). Whereas for the latter $\theta$ drops discontinuously to zero for $c=2.15$, $\theta$ of periodic DNA vanishes with zero slope. Only after increasing $c$ artificially to $3.15$ does periodic DNA exhibits a similar first order transition. ![\[fig:mutations\] Mutations in the sequence renders the transition to the intermediate phase discontinuous. The plot shows Monte Carlo data for the length of the unbound end for evenly spaced mutations every 100 and 200 bases. This length drops discontinuously at a critical temperature ${\tilde{T}_c}$, which depends on the mutation density and the binding strength of mutated basepairs.](fig5a "fig:"){width="0.47\columnwidth"} ![\[fig:mutations\] Mutations in the sequence renders the transition to the intermediate phase discontinuous. The plot shows Monte Carlo data for the length of the unbound end for evenly spaced mutations every 100 and 200 bases. This length drops discontinuously at a critical temperature ${\tilde{T}_c}$, which depends on the mutation density and the binding strength of mutated basepairs.](fig5b "fig:"){width="0.47\columnwidth"} [*Weak sequence disorder.—*]{} Is the intermediate phase identified above robust against sequence disorder? To address this question, we replace a small fraction of base pairs by bases that can pair with each other, but not with other bases in the sequence. Fig. 5 shows the average length of the overhanging end, obtained by Monte Carlo simulation, for evenly spaced mutations with densities $0.005$ and $0.01$ and mutation strength $\bar{\varepsilon}_b=2$ . The plot suggests that in the presence of weak sequence disorder the transition described above remains, but is of first order instead of being continuous. The unbound end keeps its ground state length up to certain temperature, and then shortens rapidly. This is readily understood, when comparing the energy barriers for forming bulgeloops with and without mutations. The formation of a bulgeloop on the longer strand of a perfectly periodic molecule requires only the initiation energy ${{\varepsilon_{\ell}}}$. In the presence of mutation, however, shifting both strands breaks mutated basepairs. Hence, to form a bulge loop, all mutations to the left of the loop have to be broken and the energy barrier for loop formation grows with the distance from the end. Due to this extensive energy barrier for loop formation, mutated basepairs stay bound in a finite temperature range. For a sufficiently low density of mutations, there is a temperature ${\tilde{T}_c}$, at which the entropy gained by distributing excess bases in loops along the molecule outweighs the energetic costs [@Neher:05]. Below ${\tilde{T}_c}$ all mutations are bound, if $T>{\tilde{T}_c}$ as many mutations open, as are necessary to absorb all excess bases. On increasing the mutation density, ${\tilde{T}_c}$ approaches the melting temperature and the intermediate phase vanishes. [*Discussion.—*]{} We have identified a BEC-like conformational transition in periodic and nearly periodic DNA, which occurs below the melting transition. This transition leads to a change in the contour length of the DNA molecule, which is roughly proportional to $M-N$. We also expect an effect on the persistence length of the helical region due to the increased density of bulgeloops. The hallmark of the transition, i.e. the shortening of the unbound end, could be directly observed by resonant energy transfer between fluorescent dyes located at the ends of the two strands. We expect the existence of the intermediate phase to be independent of the details of our model. Furthermore, we have shown that the additional conformations possible for repetitive sequences change the critical behavior at the melting transition. We are grateful for important comments by E. Frey and H. Wagner. We acknowledge financial support by the [*Deutsche Forschungsgemeinschaft*]{} through the Emmy Noether Program. [21]{} 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: 'First measurement of sub-threshold $\Sigma$(1385) production is presented. Experimental data are presented for Al+Al reactions at 1.9$A$ GeV measured with the FOPI detector at SIS/GSI. The $\Sigma$(1385)/$\Lambda$ ratio is found to be in good agreement with the transport and statistical model predictions. The results allow for a better understanding of sub-threshold strangeness production and strangeness exchange reaction which is the dominant process for $K^-$ production below and close-to threshold.' author: - 'X. Lopez' - 'N. Herrmann' - 'P. Crochet' - 'A. Andronic' - 'V. Barret' - 'Z. Basrak' - 'N. Bastid' - 'M.L. Benabderrahmane' - 'P. Buehler' - 'M. Cargnelli' - 'R. Čaplar' - 'E. Cordier' - 'P. Dupieux' - 'M. Dželalija' - 'L. Fabbietti' - 'Z. Fodor' - 'I. Gašparić' - 'Y. Grishkin' - 'O.N. Hartmann' - 'K.D. Hildenbrand' - 'B. Hong' - 'T.I. Kang' - 'J. Kecskemeti' - 'M. Kirejczyk' - 'Y.J. Kim' - 'M. Kiš' - 'P. Koczon' - 'M. Korolija' - 'R. Kotte' - 'A. Lebedev' - 'Y. Leifels' - 'V. Manko' - 'J. Marton' - 'T. Matulewicz' - 'M. Merschmeyer' - 'W. Neubert' - 'D. Pelte' - 'M. Petrovici' - 'F. Rami' - 'W. Reisdorf' - 'M.S. Ryu' - 'P. Schmidt' - 'A. Schüttauf' - 'Z. Seres' - 'B. Sikora' - 'K.S. Sim' - 'V. Simion' - 'K. Siwek-Wilczyńska' - 'V. Smolyankin' - 'G. Stoicea' - 'K. Suzuki' - 'Z. Tyminski' - 'P. Wagner' - 'E. Widmann' - 'K. Wiśniewski' - 'D. Wohlfarth' - 'Z.G. Xiao' - 'I. Yushmanov' - 'X.Y. Zhang' - 'A. Zhilin' - 'J. Zmeskal' - 'P. Kienle' - 'T. Yamazaki' title: | Sub-threshold production of $\Sigma$(1385) baryons\ in Al+Al collisions at 1.9$A$ GeV --- Relativistic heavy ion collisions at SIS energies provide interesting opportunities for studying hot and dense nuclear matter. It allows to address fundamental aspects of nuclear physics such as the nuclear-matter equation-of-state [@Aichelin:1986ss; @Hartnack:2005tr; @Sturm:2000dm; @Fuchs:2000kp] and the question whether hadron properties undergo modifications in such an environment [@chiral]. Indications for in-medium modifications of charged kaon production and propagation have been experimentally observed by the KaoS [@lastkaos] and the FOPI [@Crochet:2000fz; @Wisniewski:2001dk] collaborations at SIS energies. This beam energy range (1 - 2$A$ GeV) is particularly well suited for studying in-medium properties of strange particles since, as they are produced below or close-to threshold, their production process is sensitive to nuclear in-medium effects. While sub-threshold $K^+$ are mostly produced via multi-step processes, transport models indicate that sub-threshold $K^-$ production results from strangeness exchange reactions $\pi + {\rm Y} \leftrightarrow K^- + {\rm B}$ whose rate is intimately linked to the hyperon (${\rm Y} = \Lambda, \Sigma$) yield [@Fuchs:2005zg; @Hartnack:2001zs]. In addition, recent calculations based on the chiral theory predict an important coupling of the $K^-$ to the medium via $\Sigma(1385)$, $\Lambda(1405)$ and $\Lambda(1520)$ resonances [@Cassing:2003vz; @Lutz:2001dq; @Schaffner-Bielich:1999cp; @Tolos:2000fj] and reveal, in particular, the importance of the $\Sigma(1385)$ hyperon in sub-threshold $K^-$ production [@Lutz:2003id]. Theoretical predictions on the in-medium cross-section of this strangeness exchange reaction differ widely [@Cassing:2003vz; @Lutz:2001dq; @Schaffner-Bielich:1999cp; @Tolos:2000fj] and, on the other hand, experimental data on $\Sigma(1385)$ are scarce.\ We report in this letter on the first measurement of charged $\Sigma(1385)$ (called $\Sigma^*$ in the following) production in Al+Al collisions at a beam kinetic energy of 1.9$A$ GeV. This measurement corresponds to sub-threshold production since the threshold beam kinetic energy for $\Sigma^*$ production in elementary reaction is $E_{\rm thr.} = 2.33~{\rm GeV}$. The measured yield is compared to the predictions of statistical and transport models. The only other available measurement of $\Sigma^*$ production yield in heavy ion collisions has been reported recently by the STAR collaboration for the Au+Au system at the top RHIC energy ($\sqrt{s_{NN}} = 200~{\rm GeV}$) [@Adams:2006yu; @Salur:2004ar].\ The experiment has been performed with the FOPI detector at the SIS Heavy-Ion Synchrotron of GSI-Darmstadt by using an Al beam of kinetic energy of 1.9$A$ GeV on an Al target. The beam intensity was chosen to be $8\cdot 10^5$ ions/s and the target thickness was 567 mg/cm$^2$. The FOPI detector is an azimuthally symmetric apparatus made of several sub-detectors which provide charge and/or mass identification over nearly the full solid angle. The central part of the detector set-up, placed in a super-conducting solenoid, consists of a Central Drift Chamber (CDC) surrounded by a barrel of plastic scintillators. The forward part is composed of two walls of plastic scintillators and a second drift chamber. For the present analysis, $\Lambda$ and $\Sigma^*$ were reconstructed from their decay products measured in the CDC ($23^\circ~<~\theta_{lab}~<~114^\circ$). These are identified by their mass, which is determined from the correlation between magnetic rigidity and specific energy loss. More details about the configuration and performances of the different components of the FOPI apparatus can be found in [@Gobbi:1992hw; @Ritman:1995td; @Andronic:2000cx]. The events were centrality selected by imposing conditions on the multiplicity of charged particles detected in the forward part of the detector. The results presented in the following correspond to the most central collisions ($\sigma_{geo}\le315$ mb) representing about 20$\%$ of the total geometrical cross section.\ In order to assess the P($\Sigma^{*-}+\Sigma^{*+}$)/P($\Lambda+\Sigma^{0}$) ratio, yields of $\Sigma^{*-}$, $\Sigma^{*+}$ as well as primary $\Lambda$ have to be determined. Note that the reconstructed $\Lambda$ yield includes decay $\Lambda$ from $\Sigma^0$ which cannot be isolated with the FOPI detector because $\Sigma^0$ decays into a $\Lambda$ and a photon. While primary $\Lambda$ are reconstructed from the invariant mass of $(p,\pi^-)$ pair candidates (see [@Ritman:1995tn; @mmxl] for more details about $\Lambda$ reconstruction in FOPI), $\Sigma^*$ are reconstructed from a topological analysis of their double two body decay: $$\begin{aligned} \begin{array}{lllllll} \Sigma^{*\pm}~\rightarrow~&\Lambda&\!\!\!\!\!+~\,\pi^\pm & ({\rm BR} = 88\%, c\tau = 5~{\rm fm}) \nonumber \\ &{\makebox[0mm][l]{\rule{0.33em}{0mm}\rule[0.55ex]{0.044em}{1.55ex}}\rightarrow}&p\,\,+~\pi^- & ({\rm BR} = 64\%, c\tau = 7.89~{\rm cm}). \end{array}\end{aligned}$$ Since the short life-time of the $\Sigma^*$ does not allow to disantengle its decay vertex from the collision vertex, its reconstruction consists in correlating $\Lambda$ with primary pions. Therefore, in order to minimize the combinatorial background in the $(\Lambda,\pi^\pm)$ pair candidates, a high purity sample of $\Lambda$ is extracted from the data. These $\Lambda$ are reconstructed from identified pions and protons with the following conditions: i\) a cut on momentum ($p_{lab} \ge 0.1$ GeV/$c$) is applied in order to reject particles which are spiraling inside the CDC; ii\) the distance of closest approach (DCA) between tracks and the primary vertex in the transverse plane (DCA$_{p}$ $>$ 0.8 cm, DCA$_{\pi^-}$ $>$ 1.9 cm) is used to enhance the fraction of protons and pions coming from a secondary vertex; iii\) a cut on transverse flight distance ($d_{t}<20$ cm) and on the pointing angle of the $\Lambda$ ($\Delta\phi<4^{o}$), the latter being the difference between the geometrical and kinematical azimuthal angles; iv\) a cut on transverse momentum of the $\Lambda$ in the laboratory ($p_{t}\ge0.3$ GeV/$c$) is used to increase the purity of the signal. The invariant mass of the particle pair is calculated from its four-momenta at the intersection point of the two tracks. The corresponding distribution is shown in the inset of Fig. \[fig:minv1\]. A total of about 10$^5$ $\Lambda$ are reconstructed with a signal-to-background ratio close to 10 from the analysis of 290 million events. $\Sigma^{*\pm}$ are then reconstructed by calculating the invariant mass of these $\Lambda$ with charged pions after applying a two-sigma mass selection cut around the $\Lambda$ nominal mass (shown by the dashed lines in the inset of Fig. \[fig:minv1\]). ![\[fig:minv1\] Invariant mass spectra of $p\pi^-$ pairs (inset, upper panel) and $\Lambda \pi^{\pm}$ pairs. The solid histogram and crosses denote the data and the scaled mixed-event background, respectively (upper panel). The lower panel shows the signal after background subtraction. The following characteristics of the signal are shown: number of counts in the signal (S), signal-to-background ratio (S/B) and significance (SIGNIF). The parameters extracted from the fit to the data (mean mass value (MEAN) and the width ($\Gamma$)) are also reported.](fig1_new.eps){width="6.5cm"} The reconstructed $\Sigma^{*-}$ and $\Sigma^{*+}$ are shown in Fig. \[fig:minv1\]. The combinatorial background is obtained with the event-mixing method [@Berger:1976eb]. For that purpose, the two decay particles ($\Lambda$ and $\pi^{\pm}$) are taken from two different events which present the same particle multiplicity in the CDC. In addition, the two events are aligned to the reaction plane in order to have the same reference system for both particles. The reaction plane is estimated event by event utilizing the transverse momentum procedure detailed in [@Danielewicz:1985hn]. In the invariant mass range 1.35 to 2 GeV/$c^2$, 16 excited states of $\Sigma$’s have already been observed [@rev]. Therefore, the normalization of the mixed background was done in the mass range 1.25 - 1.32 GeV/$c^2$ (grey zone in the upper panel of Fig. \[fig:minv1\]) in order not to bias the reproduction of the combinatorial background. The shape of the resulting mixed-event background describes the combinatorial background and is indicated by crosses in Fig. \[fig:minv1\] (upper panel). The vertical bars of the crosses correspond to the statistical errors. After background subtraction (Fig. \[fig:minv1\], lower panel), the remaining peak in the mass spectrum is fitted with a Breit-Wigner function. Within the width interval ($\Gamma$), about 3100 $\Sigma^{*\pm}$ are found for the applied set of cuts in the analysis. The mean mass value and width extracted from the fit are in a good agreement, within statistical errors, with the values reported by the Particle Data Group [@rev]. The width of the $\Sigma^*$ is expected to increase with baryonic density [@oset] but the accuracy of our measurement doesn’t permit to draw any conclusion about a possible broadening. On the right side of the $\Sigma^{*\pm}$ signal, some excited states are visible as it has been observed in the invariant mass distribution presented by the STAR collaboration [@Adams:2006yu; @Salur:2004ar]. The losses due to decay, acceptance and reconstruction efficiency have to be corrected in order to extract the particle yields. These corrections are determined by means of extensive GEANT simulations modeling the full detector response. The IQMD model [@Hartnack:1997ez] is used as generator for the underlying heavy-ion event. The $\Sigma^{*\pm}$ and $\Lambda$ resonances are embedded in those events (one signal per event) and they are generated separately with a momentum distribution according to the Siemens-Rasmussen formula [@sie] which describes an expanding system with a temperature $T$ and a radial expansion velocity $\beta$. The choice of the values of these parameters ($T=90~{\rm MeV}$, $\beta=0$ and $0.3$) is imposed by previous measurements [@Hong:1997mr; @mm]. Afterwards, a full simulation of the detector, including resolutions in energy deposition and spatial position, Front-End-Electronics processing and tracking, is performed and the resulting output is subject to the same reconstruction procedure as for the experimental data. Simulated and experimental spectra of all relevant quantities were carefully compared. Since no significant differences were found, we conclude that the apparatus is properly described by the simulation [@xl; @mm1]. Finally, the reconstruction efficiency is determined by computing the ratio of reconstructed particles in the simulation to those initially embedded into the background events. The systematic error on the $\Lambda$ and $\Sigma^{*\pm}$ yields was evaluated in two steps. First the reconstruction efficiency was estimated from simulated $\Lambda$ and $\Sigma^{*\pm}$ with two values of the radial flow velocity ($\beta=0$ and $\beta=0.3$). Then the $\Lambda$ and $\Sigma^{*\pm}$ signals were reconstructed under different sets of conditions on the relevant quantities previously discussed ($p_{lab}$, DCA, $d_{t}$, $\Delta\phi$ and $p_t$). In order to minimize the statistical errors, the yields of $\Sigma^{*+}$ and $\Sigma^{*-}$ are summed and the following ratio can be extracted: $$\label{yield} \frac{\mathrm{P}(\Sigma^{*-}+\Sigma^{*+})}{\mathrm{P}(\Lambda+\Sigma^{0})} = 0.125\pm 0.026(\mathrm{stat.})\pm 0.033(\mathrm{syst.}).$$ Note that using such a ratio has the additional advantage that the influence of the chosen momentum distribution for the simulated signal is cancelled out to some extent. It is worth to point out that the measured ratio vary, within errors, by less than a factor two from SIS to RHIC despite the different processes involved in the $\Sigma^*$ production [@Adams:2006yu; @Salur:2004ar]. Our experimental ratio is compared to a statistical model predictions for Al+Al collisions  [@Andronic:2005yp]. The results are shown in Fig. \[fig:comp\] where other particle species such as proton, pion, $\Lambda$ and $K^0$ are included. ![\[fig:comp\] Measured hadron yield ratios compared to the thermal model calculations for Al+Al collisions at 1.9$A$ GeV. The symbols are data and the lines are the model calculations. ](fig2_new.eps){width="6.5cm"} The statistical model, based on the canonical ensemble, reproduces the ratios with a temperature of 76 MeV and a baryonic chemical potential of $\mu_{B}=816$ MeV. This chemical temperature is lower than the kinetical temperature of 95 $\pm$ 10 MeV derived from previous measurements [@Hong:1997mr; @mm]. The same model, used to extract thermal parameters from Au+Au collisions at $\sqrt{s_{NN}}=2.7$ GeV [@Andronic:2005yp], reproduced particle ratios with T = 64 MeV and $\mu_{B}=760$ MeV. It is surprising to notice that at the same energy, a lighter system (Al+Al) brings out higher temperature and baryonic chemical potential, while the volume represents 10% of the one of the Au+Au system. Finally, a higher temperature appears better suited to describe particle ratios when including a baryonic resonance and further studies on other strange particles including mesonic resonances as $\phi$ and $K$(892) will be useful to constrain the temperature. As it has been reported in [@becattini], an additional parameter as $\gamma_S$ which takes into account a non fully chemical equilibrium of strange particles could also be used to describe all the particle ratios. The measurement is also compared to the prediction from the UrQMD transport model [@Bleicher:2002rx] for Al+Al collisions [@vogel]. It is worth to point out that this model is the only transport code predicting $\Sigma^*$ production at SIS energies. The results of comparisons with statistical and transport models are summarized in Tab. \[tab-1\]. Yield ratio Data Therm. Model UrQMD ------------------------------------------------------------------------ ------------------- -------------- ------- $\frac{{\rm P}(\Sigma^{*-}+\Sigma^{*+})}{{\rm P}(\Lambda+\Sigma^{0})}$ 0.125 $\pm$ 0.042 0.097 0.177 : \[tab-1\] Experimental yield ratio and predictions from statistical and transport models. The error of the experimental result corresponds to the quadratic sum of statistical and systematic errors. The transport model prediction for Al+Al collisions is in relatively good agreement with the measurement when taking into account the upper limit of the error bars. The dominant process involved in the UrQMD model in order to produce $\Sigma^*$ is the fusion of $\Lambda$ and pions (76%) with an average cross section of about 37 mb [@vogel]. The other processes used in the transport code to produce $\Sigma^*$ are the interactions between $\Sigma$ and pions (12%) and, $N^*$ and $\Delta$ resonances with baryons (12%). The time dependence of these reactions exhibits a maximum around 8 fm/$c$ which is in agreement with the predictions of the IQMD transport model [@Hartnack:2001zs] on the time production of strange particles such as kaons [@hart21]. During the time span between the $\Sigma^*$ production and the chemical freeze-out, about 8% of pions coming from the decay of the $\Sigma^*$ are lost in inelastic interactions. This marginal loss could be attributed to the small size of the system created in Al+Al collisions.\ In summary, we have presented new results about the P($\Sigma^{*-}+\Sigma^{*+}$)/P($\Lambda+\Sigma^{0}$) ratio in Al+Al collisions at 1.9$A$ GeV. For the first time, the $\Sigma^{*\pm}$ resonances were measured 400 MeV below their production threshold. The experimental result was compared to the predictions of a statistical model and a transport model. The measurement is in good agreement with the transport model prediction where the dominant reaction to produce the $\Sigma^*$ is the fusion of $\Lambda$ and pions. On the other hand, the comparison with the thermal model suggests that a chemical freeze-out temperature of about 80 MeV is needed to describe the measured ratio. The reported measurement should be used as an input for other transport codes for testing calculations on sub-threshold strangeness production. In addition, a measurement of the system size dependence of the $\Sigma^*$ production, as well as other strange resonances, will give the opportunity to assess effects of a partial restoration of the chiral symmetry at high baryon density that can be reached at SIS energies.\ We are grateful to M. Bleicher and S. Vogel for providing us model calculations and for intensive discussions. This work was supported by the German BMBF under Contract No. 06HD154, by the Korea Science and Engineering Foundation (KOSEF) under grant No. F01-2006-000-10035-0, by the mutual agreement between GSI and IN2P3/CEA, by the Hungarian OTKA under grant No. 47168, within the Framework of the WTZ program (Project RUS 02/021), by DAAD (PPP D/03/44611) and by DFG (Projekt 446-KOR-113/76/04). We have also received support by the European Commission under the 6th Framework Program under the Integrated Infrastructure on: Strongly Interacting Matter (Hadron Physics), Contract No. RII3-CT-2004-506078. [29]{} J. Aichelin and C. M. Ko, Phys. Rev. Lett.  [**55**]{}, 2661 (1985). C. Hartnack, H. Oeschler and J. Aichelin, Phys. Rev. Lett.  [**96**]{}, 012302 (2006). C. Sturm [*et al.*]{} \[KAOS Collaboration\], Phys. Rev. Lett.  [**86**]{}, 39 (2001). C. Fuchs, A. Faessler, E. Zabrodin and Y. M. Zheng, Phys. Rev. Lett.  [**86**]{}, 1974 (2001). C. M. Ko, V. Koch, G. Li, Ann. Rev. Nucl. Part. Sci. [**47**]{}, 505 (1997). A. Förster [*et al.*]{} \[KaoS collaboration\], Phys. Rev. C [**75**]{}, 024906 (2007). P. Crochet [*et al.*]{} \[FOPI collaboration\], Phys. Lett. B [**486**]{}, 6 (2000). K. Wisniewski [*et al.*]{} \[FOPI Collaboration\], Eur. Phys. J. A [**9**]{}, 515 (2000). C. Fuchs, Prog. Part. Nucl. Phys.  [**56**]{}, 1 (2006). C. Hartnack, H. Oeschler and J. Aichelin, Phys. Rev. Lett.  [**90**]{}, 102302 (2003). W. Cassing, L. Tolos, E. L. Bratkovskaya and A. Ramos, Nucl. Phys. A [**727**]{}, 59 (2003). M. F. M. Lutz and C. L. Korpa, Nucl. Phys. A [**700**]{}, 309 (2002). J. Schaffner-Bielich, V. Koch and M. Effenberger, Nucl. Phys. A [**669**]{}, 153 (2000). L. Tolos, A. Ramos, A. Polls and T. T. S. Kuo, Nucl. Phys. A [**690**]{}, 547 (2001). M. F. M. Lutz, Prog. Part. Nucl. Phys.  [**53**]{}, 125 (2004). J. Adams [*et al.*]{} \[STAR Collaboration\], Phys. Rev. Lett. [**97**]{}, 132301 (2006). S. Salur \[STAR Collaboration\], J. Phys. G [**31**]{}, S179 (2005). A. Gobbi [*et al.*]{} \[FOPI Collaboration\], Nucl. Instrum. Meth. A [**324**]{}, 156 (1993). J. Ritman \[FOPI Collaboration\], Nucl. Phys. Proc. Suppl.  [**44**]{}, 708 (1995). A. Andronic [*et al.*]{} \[FOPI Collaboration\], Nucl. Phys. A [**679**]{}, 765 (2001). J. L. Ritman [*et al.*]{} \[FOPI Collaboration\], Z. Phys. A [**352**]{}, 355 (1995). M. Merschmeyer, X. Lopez, N. Bastid, P. Crochet, N. Herrmann [*et al.*]{} \[FOPI Collaboration\], Phys. Rev. C [**76**]{}, 024906 (2007). E. L. Berger, R. Singer, G. H. Thomas and T. Kafka, Phys. Rev. D [**15**]{}, 206 (1977). P. Danielewicz and G. Odyniec, Phys. Lett. B [**157**]{}, 146 (1985). W.-M. Yao [*et al.*]{}, J. Phys. G [**33**]{}, 1 (2006). Murat M. Kaskulov and E. Oset, Phys. Rev. C [**73**]{}, 045213 (2006). C. Hartnack, R. K. Puri, J. Aichelin, J. Konopka, S. A. Bass, H. Stoecker and W. Greiner, Eur. Phys. J. A [**1**]{}, 151 (1998). P.J. Siemens and J.O. Rasmussen, Phys. Rev. Lett. [**42**]{}, 880 (1979). B. Hong [*et al.*]{} \[FOPI Collaboration\], Phys. Rev. C [**57**]{}, 244 (1998) \[Phys. Rev. C [**58**]{}, 603 (1998)\]. M. Merschmeyer et al., GSI Scientific 2005, 208 [**2006-1**]{}. X. Lopez, PhD thesis, University of Clermont-Ferrand (2004). M. Merschmeyer, PhD thesis, University of Heidelberg (2004). A. Andronic, P. Braun-Munzinger and J. Stachel, Nucl. Phys. A [**772**]{}, 167 (2006). F. Becattini and G. Pettini, Phys. Rev. C [**67**]{}, 015205 (2003). M. Bleicher, Nucl. Phys. A [**715**]{}, 85 (2003). S. Vogel (private communication). C. Hartnack, \[arXiv:nucl-th/0507002\].
{ "pile_set_name": "ArXiv" }
--- abstract: 'A generalized Kummer surface $X$ obtained as the quotient of an abelian surface by a symplectic automorphism of order 3 contains a $9{\bf A}_{2}$-configuration of $(-2)$-curves. Such a configuration plays the role of the $16{\bf A}_{1}$-configurations for usual Kummer surfaces. In this paper we construct $9$ other such $9{\bf A}_{2}$-configurations on the generalized Kummer surface associated to the double cover of the plane branched over the sextic dual curve of a cubic curve. The new $9{\bf A}_{2}$-configurations are obtained by taking the pullback of a certain configuration of $12$ conics which are in special position with respect to the branch curve, plus some singular quartic curves. We then construct some automorphisms of the K3 surface sending one configuration to another. We also give various models of $X$ and of the generic fiber of its natural elliptic pencil.' author: - 'David Kohel, Xavier Roulleau, Alessandra Sarti' title: A special configuration of 12 conics and generalized Kummer surfaces --- ÅŁłø¶§\#1[[\#1]{}]{} Introduction ============ A Kummer surface $\Km(A)$ is the minimal desingularization of the quotient of an abelian surface $A$ by the involution $[-1]$. It is a K3 surface containing $16$ disjoint $\cu$-curves, which lie over the $16$ singularities of $A/\langle[-1]\rangle$. We call such set of curves a $16{\bf A}_{1}$-configuration. A well-known result of Nikulin [@Nikulin] gives the converse: if a K3 surface contains a $16{\bf A}_{1}$-configuration, then it is the Kummer surface of an abelian surface $A$, such that the $16$ $\cu$-curves lie over the singularities of $A/\langle[-1]\rangle$. Shioda [@ShiodaRemarks] then asked the following question: if two abelian surfaces $A$ and $B$ satisfy $\Km(A)\simeq\Km(B)$, is it true that $A\simeq B$ ? Gritsenko and Hulek [@GriH] gave a negative answer to that question in general. In [@RS1], [@RS2], we studied and constructed examples of two $16{\bf A}_{1}$-configurations on the same Kummer surface such that their associated abelian surfaces are not isomorphic. Kummer surfaces have natural generalizations to quotients of an abelian surface $A$ by other symplectic groups $G\subseteq\Aut(A)$. If $G\cong\ZZ/3\ZZ$, then the quotient surface $A/G$ has $9$ cusp singularities, in bijection with the fixed points of $G$. Its minimal desingularization, denoted by $\Km_{3}(A)$, is a K3 surface which contains $9$ disjoint $\mathbf{A}_{2}$-configurations, i.e. pairs $(C,C')$ of $\cu$-curves such that $CC'=1$. It is then natural to ask if an isomorphism $\Km_{3}(A)\simeq\Km_{3}(B)$ between two generalized Kummer surfaces implies that $A$ and $B$ are isomorphic. With this question in mind, in the present paper we construct geometrically several $9\mathbf{A}_{2}$-configurations on some generalized Kummer surfaces previously studied by Birkenhake and Lange [@BL]. Their construction is as follows. For $\l$ generic, the dual of a cubic curve $E_{\l}:x^{3}+y^{3}+z^{3}-3\l xyz=0$ is a sextic curve $C_{\l}$ with a set $\mathcal{P}_{9}$ of $9{\bf A}_{2}$ singularities corresponding to the nine inflection points on $E$. The minimal desingularization $X_{\l}$ of the double cover of $\PP^{2}$ branched over $C_{\l}$ is a generalized Kummer surface with a natural $9\mathbf{A}_{2}$-configuration $\mathcal{A}_{0}$. The surface $X_{\l}$ has a natural elliptic fibration $\varphi:X_{\l}\to\PP^{1}$ for which the $18$ $\cu$-curves in the $9\mathbf{A}_{2}$-configuration are sections, and the reduced strict transform of $C_{\l}$ is a fiber. In order to find other $\cu$-curves on $X_{\l}$ we study the set $\mathcal{C}_{12}$ of conics that contain at least $6$ points in $\mathcal{P}_{9}$. One has The set $\mathcal{C}_{12}$ has cardinality $12$. Each conic in $\mathcal{C}_{12}$ contains exactly $6$ points in $\mathcal{P}_{9}$ and through each point in $\mathcal{P}_{9}$ there are $8$ conics. The sets $(\mathcal{P}_{9},\mathcal{C}_{12}$) form therefore a $(9_{8},12_{6})$-configuration. The configuration $(\cP_{9},\cC_{12})$ has interesting symmetries, e.g. there are $8$ conics among the $12$ passing through a fixed point $q$ in $\mathcal{P}_{9}$ and the $8$ points in $\cP_{9}\setminus\{q\}$, which form a $8_{5}$ point-conic configuration. The freeness of the arrangement of curves $\cC_{12}$ is studied in [@PS], where we learned that this configuration has been also independently discovered in [@DLPU]. The irreducible components of the curves in the K3 surface $X_{\l}$ above the $12$ conics are $24$ $\cu$-curves. This set of $\cu$-curves contains nine $8{\bf A}_{2}$ sub-configurations which are the strict transform of the nine $8_{5}$ sub-configurations of conics. Using the pullback to $X_{\l}$ of some $9$ special (singular) quartic curves, we are able to complete each of these $8{\bf A}_{2}$-configurations into a new $9{\bf A}_{2}$-configuration $\mathcal{A}_{k},\,(k\in\{1,\dots,9\})$. According to [@Barth], to a $9{\bf {A}_{2}}$-configuration corresponds an Abelian surface $A$ and an order $3$ symplectic group $G$ such that $X_{\l}\simeq\Km_{3}(A)$ and the $9{\bf {A}_{2}}$-configuration is the exceptional divisor of the minimal desingularisation $\Km_{3}(A)\to A/G$ (this is the analog of Nikulin’s result for $16{\bf {A}_{1}}$-configurations). A Kummer structure on K3 surface $X$ is an isomorphism class of Abelian surfaces $A$ such that $X\simeq\Km(A)$. Kummer structures are in one-to-one correspondence with the orbits under $\aut(X)$ of Nikulin’s $16{\bf {A}_{1}}$-configurations (see e.g. [@RS1 Proposition 21]). Similarly, one can define a generalized Kummer structure on a K3 surface $X$ as an isomorphism class of pairs $(A,G$) of abelian surfaces $A$ and order $3$ symplectic group $G$ such that $X\simeq\Km_{3}(A)$. We show that there is again a one-to-one correspondence between generalized Kummer structures and the orbits of the $9{\bf {A}_{2}}$-configurations. Using the $9{\bf {A}_{2}}$-configurations we constructed, we obtain: The $9{\bf {A}_{2}}$-configurations $\mathcal{A}_{1},\dots,\mathcal{A}_{9}$ we obtained on the K3 surface $X_{\l}$ are contained in the $\aut(X_{\lambda})$-orbit of $\mathcal{A}_{0}$. For the proof we construct automorphisms sending one configuration to another. Some of these automorphisms are obtained by using translations by torsion sections of the natural elliptic fibration $\varphi:X_{\l}\to\PP^{1}$, and some other by using the Torelli Theorem for K3 surfaces. We then continue our study of the surface $X_{\l}$ by obtaining various models in projective space, in particular as a degree $8$ non-complete intersection in $\PP^{5}$. We construct a model of the generic fiber $E_{K3}$ of the fibration $\varphi$. The fibration $\varphi:X_{\l}\to\PP^{1}$ has $8$ singular fibers of type $\tilde{{\bf A}}_{2}$; the $24$ $\cu$-curves above the $12$ conics in $\mathcal{C}_{12}$ are contained in these fibers. A Hessian model of the generic fiber of $\varphi$ is $$E_{K3}:x^{3}+y^{3}+z^{3}+\frac{\l^{3}(t^{2}+3)-4t^{2}}{\l^{2}(t^{2}-1)}xyz=0.$$ We also get a Weierstrass model of $E_{K3}$. It turns out that the Mordell-Weil group of the fibration $\varphi$ has rank $1$ and torsion $(\ZZ/3\ZZ)^{2}$; we compute its generators. Using the translation maps constructed from the model $E_{K3}$, we can acquire other $9\mathbf{A}_{2}$-configurations from the previously known one. We also obtain another construction of the K3 surface $X_{\l}$ as a double plane: The surface $X_{\l}$ is the minimal desingularization of the double cover of $\PP^{2}$ branched over the sextic curve which is the union of the elliptic curves $E_{\l}$ and its Hessian $${\mathrm{He}}(\l):\,\,\,\,x^{3}+y^{3}+z^{3}+\frac{(\l^{3}-4)}{\l^{2}}xyz=0,$$ The strict transform on $X_{\l}$ of the $12$ lines of the Hesse arrangement in $\PP^{2}$ are the $24$ $\cu$-curves above the $12$ conics. **Acknowledgements** The authors wish to thank Cédric Bonnafé, Igor Dolgachev, Antonio Laface, Ulf Persson, Piotr Pokora, Giancarlo Urzúa, and also Carlos Rito for sharing his program `LinSys`. Part of the computations were done using Magma software [@Magma]. The second author thanks the Max-Planck Institute for Mathematics of Bonn for its hospitality and support. Preliminaries ============= Notations and conventions\[subsec:Notations-and-conventions\] ------------------------------------------------------------- Let $\eta:Y\to Z$ be a dominant map between two surfaces and let $C\hookrightarrow Z$ be a curve. In this paper, the reduced pullback of $C$ minus the irreducible components contracted by $\eta$ is called the strict transform of $C$ on $Y$. We say that two $\cu$-curves $E,E'$ on a K3 surface form an $\mathbf{A}_{2}$-configuration if their intersection matrix is: $\left(\begin{array}{@{}c@{\,}c@{}} -2 & \ 1\\ \ 1 & -2 \end{array}\right)\cdot$ We say that the $\cu$-curves $E_{1},E_{1}',\dots,E_{n},E_{n}'$ form an $n{\bf A}_{2}$-configuration if the curves $E_{j},E_{j}',\,j\in\{1,\dots,n\}$ form $n$ disjoint ${\bf A}_{2}$-configurations. We recall (see [@Dolgachev]) that a $(v_{r},b_{k})$-configuration is the data of two sets $\mathcal{P}_{v},\mathcal{L}_{b}$ of respective cardinality $v$ and $b$, plus a subset $I\subset\mathcal{P}_{v}\times\mathcal{L}_{b}$, such that for each element $p$ in $\mathcal{P}_{v}$, there are $r$ elements $l$ in $\mathcal{L}_{b}$ such that $(p,l)\in I$, and for each element $l$ in $\mathcal{L}_{b}$, there are $k$ elements $p$ in $\mathcal{P}_{v}$ such that $(p,l)\in I$. One has $vr=bk$. When $v=b$ and $r=k$, we call such a configuration a $v_{r}$-configuration. Generalized Kummer structures\[subsec:Generalized-Kummer-structures\] --------------------------------------------------------------------- Let $X$ be a generalized Kummer surface. By the results of Barth [@Barth] if $\mathcal{C}=A_{1},A_{1}',\dots,A_{9},A_{9}'$ is a $9{\bf A}_{2}$-configuration on $X$, there exists an abelian surface $B$ and a symplectic group $G$ of order $3$ such that $X\simeq\Km_{3}(B)$, such that the curves of the configuration $\mathcal{C}$ are the exceptional curves of the minimal desingularization $X\to B/G$. The abelian surface $B$ is obtained by taking the blow-up $\tilde{X}\to X$ at the $9$ intersection points $A_{k}A_{k}'$ ($k=1,...,9$). The strict transform of $\mathcal{C}$ on $\tilde{X}$ is the union of $18$ disjoint $(-3)$-curves, the result of Barth gives that there exists a unique triple cover map $\tilde{B}\to\tilde{X}$ branched over the $18$ $(-3)$-curves; the reduced pull-back of these curves are $(-1)$-curves and the pull-back of the exceptional curves on $\tilde{X}$ are $9$ $(-3)$-curves. The minimal model of $\tilde{B}$ is the abelian surface $B$. A generalized Kummer structure (of order $3$) on a K3 surface $X$ is an isomorphism class of pairs $(A,G)$ of abelian surfaces equipped with an order $3$ symplectic automorphism subgroup $G\subset\aut(A)$, such that $X\simeq\Km_{3}(A)$, where $\Km_{3}(A)$ is the minimal desingularization of $A/G$. There is a one-to-one correspondence between Kummer structures on $X$ and $\aut(X)$-orbits of $9{\bf A}_{2}$-configurations. Let $\mu:X\to X$ be an automorphism of $X$ sending the configuration $\mathcal{C}$ to the configuration $\mathcal{C}'$. Let $B,B'$ be the abelian surfaces and let $G\subset\aut(B),\,G'\subset\aut(B')$ be the order $3$ automorphism groups such that $X$ is the minimal desingularization of $B/G$ and $B'/G'$, and the exceptional curves of the minimal desingularization are respectively in $\mathcal{C}$, $\mathcal{C}'$. Let us prove that there exists an isomorphism $\tau$ between the abelian surfaces $B,B'$ such that $G'=\tau G\tau^{-1}$.\ As above let $\tilde{X}$, $\tilde{X}'$ be the blow-up of $X$ at the $9$ singular points of $\mathcal{C}$ and $\mathcal{C}'$ respectively. The automorphism $\mu$ extends to an isomorphism $\tilde{\mu}:\tilde{X}\to\tilde{X}'$. The map $\tilde{B}\to\tilde{X}\stackrel{\tilde{\mu}}{\to}\tilde{X}'$ is branched with order $3$ over the strict transform of $\mathcal{C}'$ in $\tilde{X}'$. By uniqueness of the triple cover, there is an isomorphism $\tilde{\tau}:\tilde{B}\to\tilde{B}',$ (which gives an isomorphism to $\tau:B\to B'$ between the minimal models). Moreover, the order $3$ automorphisms group $\tilde{G}$ of transformations of the triple cover $\tilde{B}\to\tilde{X}$ is sent by transport of structures to the order $3$ automorphism group $\tilde{G}'$ of transformations of the triple cover $\tilde{B}'\to\tilde{X}'$, i.e. $\tilde{G}'=\tilde{\tau}\tilde{G}\tilde{\tau}^{-1}$. This property is preserved when taking the minimal models: $G'=\tau G\tau^{-1}$. For the converse, let $(B,G)$ and $(B',G')$ be equipped abelian surfaces such that $\Km_{3}(B)=X=\Km_{3}(B')$ and there is an isomorphism $\tau:B\to B'$ with $G=\tau^{-1}G'\tau$. The map $B\to B'\to B'/G'$ is $3$ to $1$ and $G$-invariant, thus it induces an isomorphism $B/G\simeq B'/G'$. That isomorphism extends to the desingularization: $X=\Km_{3}(B)\simeq\Km_{3}(B')=X$ and sends the first $9{\bf A}_{2}$-configuration to the second. The Néron-Severi and transcendental lattices\[subsec:The-N=00003D0000E9ron-Severi-group\] ----------------------------------------------------------------------------------------- ### The Néron-Severi lattice Let $A$ be an abelian surface with a symplectic action of a group $G:=\ZZ/3\ZZ$. The quotient $A/G$ has $9{\bf A}_{2}$ singularities and the minimal resolution is a generalized Kummer surface $X:=\Km_{3}(A)$ which carries a $9{\bf A}_{2}$-configuration $A_{1},A_{1}',\ldots,A_{9},A_{9}'$ of $\cu$-curves. Observe that the abelian surface has Picard number at least $3$, see [@barth2 Proposition on p. 10] and the K3 surface $X$ has generically Picard number $\rho(X)$ equal to $19$. Let $\calK_{3}$ denote the minimal primitive (rank $18$) sub-lattice of the K3 lattice $H^{0}(X,\ZZ)$ containing the $9$ configurations ${\bf A}_{2}$. The lattice $\calK_{3}$ is described as follows. By [@bertin Proof of Proposition 1.3], it is generated by the classes $A_{1},A_{1}',\ldots,A_{9},A_{9}'$ and the following three classes $$\begin{array}{c} v_{1}=\frac{1}{3}\sum_{i=1}^{9}(A_{i}-A_{i}')\\ v_{2}=\frac{1}{3}((A_{2}-A{}_{2}')+2(A_{3}-A_{3}')+A_{6}-A_{6}'+2(A_{7}-A_{7}')+A_{8}-A_{8}'+2(A_{9}-A_{9}'))\\ v_{3}=\frac{1}{3}(A_{4}-A_{4}'+2(A_{5}-A_{5}')+A_{6}-A_{6}'+2(A_{7}-A_{7}')+2(A_{8}-A_{8}')+A_{9}-A_{9}'). \end{array}$$ Then the discriminant group $\calK_{3}^{\vee}/\calK_{3}$ is generated by the classes $$\begin{array}{c} w_{1}=\frac{1}{3}(A_{5}-A_{5}'+A_{7}-A_{7}'+A_{8}-A_{8}')\\ w_{2}=\frac{1}{3}(2(A_{4}-A_{4}')+A_{6}-A_{6}'+2(A_{7}-A_{7}')+A_{8}-A_{8}')\\ w_{3}=\frac{1}{3}(A_{3}-A_{3}'+A_{5}-A_{5}'+A_{6}-A_{6}') \end{array}$$ with intersection matrix: $$\left(\begin{array}{ccc} -2 & -2 & -\frac{2}{3}\\ -2 & -\frac{20}{3} & -\frac{2}{3}\\ -\frac{2}{3} & -\frac{2}{3} & -2 \end{array}\right).$$ Assume $\rho(X)=19$ (the minimal possible) and that $D_{2}$, a generator of $\calK_{3}^{\perp}\subset\NS X)$, has square $D_{2}^{2}=2$. Then \[prop:The-NS-latt-Discri54\]The Néron-Severi lattice is $\NS X)=\ZZ D_{2}\oplus\calK_{3}$ and its discriminant group has order $54$. Denote by $\mathcal{L}_{2}:=\ZZ D_{2}\oplus\calK_{3}$ and assume there exists a non-zero class $\bar{w}\in\mathcal{L}_{2}^{\vee}/\mathcal{L}_{2}$ such that $w\in\NS X)$. By the previous description of $\calK_{3}$ and the fact that $D_{2}^{2}=2$, we get $w=\tfrac{aD_{2}}{2}+\tfrac{v}{3}$ with $\frac{v}{3}\in\calK_{3}^{\vee}/\calK_{3}$ and integer $a\in\ZZ$ which must be odd since $\bar{w}\neq0$. We get $$w^{2}=\tfrac{a^{2}}{2}+\tfrac{v^{2}}{2}=\tfrac{9a^{2}+2v^{2}}{18}\in\ZZ$$ This gives $9a^{2}+2v^{2}\in18\ZZ$, but $a$ is odd so this is not possible. This shows that such an element as $\bar{w}$ does not exist and $\NS X)=\mathcal{L}_{2}$. The discriminant group of $\NS X)$ is the direct sum of the discriminant groups of $\ZZ D_{2}$ and $\mathcal{K}_{3}$, which from the above description are isomorphic to respectively $\ZZ/2\ZZ$ and $(\ZZ/3\ZZ)^{3}$, thus the claim on the order. ### The transcendental lattice Since $\NS X)=\ZZ D_{2}\oplus\mathcal{K}_{3}$, we know that the discriminant group is $$\tfrac{\NS X)^{\vee}}{\NS X)}=\tfrac{(\ZZ D_{2})^{\vee}}{\ZZ D_{2}}\oplus\tfrac{\mathcal{K}_{3}^{\vee}}{\mathcal{K}_{3}}.$$ The intersection matrix of the generators $w_{1},w_{2},w_{3}-w_{1}-w_{2}+\frac{1}{2}D_{2}$ is $$\begin{aligned} \mathcal{N}:=\left(\begin{array}{ccc} 0&0&-\frac{2}{3}\\ 0&-\frac{2}{3}&0\\ \frac{2}{3}&0&\frac{1}{2}\\ \end{array} \right).\end{aligned}$$ Recall that the quadratic form on the discriminant group has values in $\QQ/2\ZZ$ and the corresponding bilinear form has values in $\QQ/\ZZ$, see e.g. [@morrison Section 2]. The transcendental lattice $T_{X}$ of $X$ has rank $3$ signature $(2,1)$, and since $H^{2}(X,\ZZ)$ is unimodular $\det\,T_{X}=-2\cdot3^{3}$ with discriminant form the same as the discriminant form of $\NS X)$ scaled by $-1$. As in [@sarti_tran Definition 2.1], we define We call the discriminant $d$ of an indefinite rank $3$ lattice [*small*]{} if $4\cdot d$ is not divisible by $k^{3}$ for any non square natural number $k$ congruent to $0$ or $1$ modulo $4$. The lattice $T_{X}$ is small, thus by [@conway Theorem 21, p. 395] and [@conway Corollary 1.9.4] it is uniquely determined by its signature and its discriminant form. Consider now the rank three lattice $T$ of signature $(2,1)$ with Gram matrix$$\begin{aligned} \left(\begin{array}{ccc} 0&3&0\\ 3&6&-3\\ 0&-3&6\\ \end{array} \right).\end{aligned}$$It has determinant $-2\cdot3^{3}$ and if one calls $u_{1}$, $u_{2}$, $u_{3}$ its generators, one computes that the discriminant group is generated by $$\tfrac{2u_{1}}{3},\,\,\tfrac{u_{3}}{3},\,\,\tfrac{3u_{1}+2u_{2}+u_{3}}{6}$$ and the discriminant form is the same as $-\mathcal{N}$ so that by unicity: The transcendental lattice $T_{X}$ is isometric to $T$. The Hesse, dual Hesse, and related configurations\[subsec:The-Hesse,-dual\] --------------------------------------------------------------------------- Let us recall the construction and properties of the Hesse configuration before introducing a configuration with analogous properties in the next section. Let $E\hookrightarrow\PP^{2}$ be a smooth cubic curve and $\mathcal{T}_{9}$ its set of $9$ inflection points. Fixing an inflection point as the group identity, the set $\mathcal{T}_{9}$ is the $3$-torsion subgroup $E[3]$. The *Hesse configuration* is defined to be the pair $(\mathcal{T}_{9},\cL_{12})$, where $\cL_{12}$ is the set of $12$ lines through pairs of points in $\mathcal{T}_{9}$. Since each line meets $\mathcal{T}_{9}$ at $3$ points and each point of $\mathcal{T}_{9}$ is contained in $4$ lines, the Hesse configuration is a $(9_{4},12_{3})$-configuration. Removing one point of the Hesse configuration and its $4$ incident lines, one obtains a symmetric $8_{3}$-configuration of $8$ points and $8$ lines. We note that the construction is not unique to the curve $E$. If $\cP$ is the set of inflection points of any nonsingular cubic plane curve $E$, then there exists a pencil of elliptic curves (called the Hesse pencil) whose set of inflection points is $\cP$. By taking the $12$ points and $9$ lines in the dual space $(\PP^{2})^{*}$ corresponding to the $12$ lines and $9$ points dual to the Hesse configuration, one obtains a dual $(12_{3},9_{4})$-configuration called the *dual Hesse configuration*. As above, removing one line and the 4 points it meets gives a dual symmetric $8_{3}$-configuration. As abstract configurations, the above configurations can be realized from the symmetric $13_{4}$-configuration of points and lines in the the plane $\PP^{2}(\mathbb{F}_{3})$. Removing a fixed point and the set of $4$ lines which pass through it, gives the $(12_{3},9_{4})$-configuration, and removing a fixed line and the 4 points it contains gives the $(9_{4},12_{3})$-configuration. The $(9_{4},12_{3})$-configuration is naturally identified with $\AA^{2}(\FF_{3})=\PP^{2}(\FF_{3})\setminus\PP^{1}(\FF_{3})$ equipped with its system of affine lines, consistent with the identification $\mathcal{T}_{9}=E[3]\cong\FF_{3}^{2}$ in the Hesse configuration. \[sec:Nine-new-\]Nine new $9\mathbf{A}_{2}$-configurations ========================================================== \[subsec:confOfConics\]$(9_{8},12_{6})$ and $8_{5}$-configurations of conics ---------------------------------------------------------------------------- Let us fix $\l\notin\{1,\omega,\omega^{2}\}$, for $\omega$ such that $\omega^{2}+\omega+1=0$. The dual $C_{\l}$ of the elliptic curve $$E_{\l}:\,\,x^{3}+y^{3}+z^{3}-3\l xyz=0,$$ is a $9$-cuspidal sextic curve, (i.e. a sextic curve with $9$ cusps) and conversely any $9$-cuspidal sextic curve is obtained in that way. The curve $C_{\l}$ is: $$\begin{array}{c} C_{\l}:\,\,(x^{6}+y^{6}+z^{6})+2(2\l^{3}-1)(x^{3}y^{3}+x^{3}z^{3}+y^{3}z^{3})\\ \hfill-6\l^{2}xyz(x^{3}+y^{3}+z^{3})-3\l(\l^{3}-4)x^{2}y^{2}z^{2}=0. \end{array}$$ The images by the dual map of the $9$ inflection points of $E_{\l}$ are the $9$ cusps of $C_{\l}$. The set $\mathcal{P}_{9}$ of the $9$ cusps $p_{1},\dots,p_{9}$ is $$\begin{array}{ccc} p_{1}=(\l:1:1), & p_{4}=(\l:\omega:\omega^{2}), & p_{7}=(\l:\omega^{2}:\omega)\\ p_{2}=(1:\l:1), & p_{5}=(\omega^{2}:\l:\omega), & p_{8}=(\omega:\l:\omega^{2})\\ p_{3}=(1:1:\l), & p_{6}=(\omega:\omega^{2}:\l), & p_{9}=(\omega^{2}:\omega:\l). \end{array}$$ When $\l$ varies, the closure of the set of points $p_{j}$ is a line, denoted by $L_{j}$; we obtain in that way a set $\mathcal{L}_{9}$ of $9$ lines. Dually, the points on $L_{j}$ correspond to the pencil of lines meeting in the inflection point (corresponding to $p_{j}$) of the elliptic curve $E_{\l}$. One can check moreover that the line $L_{j}$ is the tangent line to the cusp $p_{j}$. Let $\mathcal{P}_{12}$ be the intersection points set of the lines in $\mathcal{L}_{9}$. \[thm:12Conics9Points\] The set $\mathcal{P}_{12}$ has cardinality $12$. The pair $(\mathcal{P}_{12},\mathcal{L}_{9})$ forms a $(12_{3},9_{4})$-configuration which is the dual Hesse configuration. The set $\cC_{12}$ of conics containing $6$ points in $\cP_{9}$ has cardinality $12$; each conic of $\mathcal{C}_{12}$ is smooth. The pair of sets $(\cP_{9},\cC_{12})$ of points and conics form a $(9_{8},12_{6})$-configuration.\ More precisely, the union of the pairwise intersections of the $12$ conics in $\mathcal{C}_{12}$ is $\mathcal{P}_{9}\cup\mathcal{P}_{12}$. The intersections between the conics are transverse. Moreover, each point of $\mathcal{P}_{12}$ is contained in exactly two conics. A conic $C$ in $\mathcal{C}_{12}$ meets $9$ conics of $\mathcal{C}_{12}$ in $4$ points of $\mathcal{P}_{9}$ and the two remaining conics in $3$ points of $\mathcal{P}_{9}$ and one point of $\mathcal{P}_{12}$. By a computer search, the $12$ conics are: $$\begin{array}{c} C_{1,2,3,4,5,6}:\,\,x^{2}+(\l+1)(\omega xy+\omega^{2}xz+yz)+\omega^{2}y^{2}+\omega z^{2}=0,\hfill\\ C_{1,2,3,7,8,9}:\,\,x^{2}+(\l+1)(\omega^{2}xy+\omega xz+yz)+\omega y^{2}+\omega^{2}z^{2}=0,\hfill\\ C_{1,2,4,5,7,8}:\,\,xy-\l z^{2}=0,\hfill\\ C_{1,2,4,6,8,9}:\,\,x^{2}+(\omega\l+1)(xy+\omega xz+\omega yz)+y^{2}+\omega^{2}z^{2}=0,\hfill\\ C_{1,2,5,6,7,9}:\,\,x^{2}+(\omega^{2}\l+1)(xy+\omega^{2}xz+\omega^{2}yz)+y^{2}+\omega z^{2}=0,\hfill\\ C_{1,3,4,5,8,9}:\,\,x^{2}+(\omega\l+\omega^{2})(xy+yz+\omega xz)+\omega y^{2}+z^{2}=0,\hfill\\ C_{1,3,4,6,7,9}:\,\,-\l y^{2}+xz=0,\hfill\\ C_{1,3,5,6,7,8}:\,\,x^{2}+(\omega^{2}\l+\omega)(xy+yz+\omega^{2}xz)+\omega^{2}y^{2}+z^{2}=0,\hfill\\ C_{2,3,4,5,7,9}:\,\,x^{2}+(\l+\omega^{2})(xy+xz+\omega^{2}yz)+\omega y^{2}+\omega z^{2}=0,\hfill\\ C_{2,3,4,6,7,8}:\,\,x^{2}+(\l+\omega)(xy+xz+\omega yz)+\omega^{2}(y^{2}+z^{2})=0,\hfill\\ C_{2,3,5,6,8,9}:\,\,\l x^{2}-yz=0,\hfill\\ C_{4,5,6,7,8,9}:\,\,x^{2}+(\l+1)(xy+xz+yz)+y^{2}+z^{2}=0,\hfill \end{array}$$ where the index $i,j,\dots,n$ of the conic $C_{i,j,\dots,n}$ means that this conic contains the $6$ points $p_{s},\,s\in\{i,j,\dots,n\}$. It is easy to see that the points in $\mathcal{P}_{9}$ are in general position: no line contains $3$ cusps, thus the conics are smooth. From the data of the conics and the knowledge of the points in $\mathcal{P}_{9}$ they contain, one can check the assertions about the configuration of the $12$ conics and the $9$ points. If one renumbers the $12$ conics by their order $C_{1},\dots,C_{12}$ from the top to bottom of the above list, one obtains that the pair of (indexes of) conics which have an intersection point not in $\mathcal{P}_{9}$ are $$\begin{array}{c} (1,2),\,(1,12),\,(2,12),\,(3,7),\,(3,11),\,(4,8),\\ (4,9),\,(5,6),\,(5,10),\,(6,10),\,(7,11),\,(8,9), \end{array}$$ and correspondingly, the $12$ points are $$\begin{array}{c} (1:1:1),\,(\omega:\omega^{2}:1),\,(\omega^{2}:\omega:1),\,(1:0:0),\,(0:1:0),\,(\omega^{2}:1:1),\\ (1:\omega^{2}:1),\,(\omega:1:1),\,(1:\omega:1),\,(\omega^{2}:\omega^{2}:1),\,(0:0:1),\,(\omega:\omega:1). \end{array}$$ respectively. One can check easily that these $12$ points in $\mathcal{P}_{12}$ are the intersection points of the lines in $\mathcal{L}_{9}$, which lines form the dual Hesse arrangement (see Section \[subsec:The-Hesse,-dual\]). By Bézout’s Theorem, the intersections between the conics are transverse. Let $q\in\mathcal{P}_{9}$ and define $\mathcal{P}_{q}=\mathcal{P}_{9}\setminus\{q\}$. \[thm:8\_5Config\] The set $\mathcal{C}_{q}$ of conics containing the point $q$ has cardinality $8$.\ The set of points $\mathcal{P}_{q}$ and the set of conics $\mathcal{C}_{q}$ form an $8_{5}$-configuration: each point is on $5$ conics and each conic contains $5$ of the points in $\mathcal{P}_{q}$.\ For each conic $C$ in $\mathcal{C}_{q}$ there exists a unique conic $C'\in\mathcal{C}_{q}$ such that there is a unique point in the intersection of $C$ and $C'$ which is not in $\mathcal{P}_{q}$. That can be checked directly from the datas in the proof of Proposition \[thm:12Conics9Points\]. Let $X_{\l}$ be the minimal desingularization of the double cover branched over the sextic curve $C_{\l}$ with $9$ cusps. We denote by $\eta:X_{\l}\to\PP^{2}$ the natural map and we denote by $A_{j},A_{j}'$ the two $\cu$-curves in $X$ above the point $p_{j}$ in $\mathcal{P}_{9}$ (so that the curves $A_{j},A_{j}'$, $j\in\{1,\dots,9\}$ form a $9\mathbf{A}_{2}$-configuration). We have \[lem:2A2config\]The strict transform by the map $\eta:X_{\l}\to\PP^{2}$ of a conic $C\in\mathcal{C}_{12}$ is the union of two disjoint $\cu$-curves $\t_{C},\t_{C}'$.\ Let $C,D$ be two conics in $\mathcal{C}_{12}$. Suppose that $C$ and $D$ meet in $4$ points in $\mathcal{P}_{9}$. Then the $\cu$-curves $\t_{C},\t_{C}',\t_{D},\t_{D}'$ are disjoint.\ Suppose that $C$ and $D$ meet in $3$ points in $\mathcal{P}_{9}$. Then, up to exchanging $\t_{D}$ and $\t_{D}'$, the curves $\t_{C},\t_{D},\t_{C}',\t_{D}'$ form a $2\mathbf{A}_{2}$-configuration. Rather than performing a double cover and taking the resolution of surface singularities, we perform three blow-ups at each cusp $q\in\mathcal{P}_{9}$ of $C_{\l}$, so that the branch locus is smooth and near $q$ it is the union of the strict transform of $C_{\l}$ and a $\cu$-curve $E_{2}$ (see also Figure \[figniceDiag\]). The three exceptional curves are $E_{2}$ and $E_{1},E_{3}$ where $E_{j}^{2}=-j$ and $E_{1}E_{2}=E_{1}E_{3}=1$, $E_{2}E_{3}=0$. On the double cover, the reduced image inverse of the curves $E_{1},E_{2},E_{3}$ are respectively a $\cu$-curve, a $(-1)$-curve and two disjoint $(-3)$-curves. Contracting the $(-1)$-curve and then the image of the $(-2)$-curve, we get the K3 surface $X_{\l}$.\ By that local computation, we see that for $C\in\mathcal{C}_{12}$, the curves $\t_{C},\t_{C'}$ are disjoint (the strict transform ${\mkern 1.5mu\overline{\mkern-1.5mu{C}\mkern-1.5mu}\mkern 1.5mu}$ of $C$ under the blow-up map do not meet the branch locus, and the two curves above ${\mkern 1.5mu\overline{\mkern-1.5mu{C}\mkern-1.5mu}\mkern 1.5mu}$ remains disjoint after contracting the $9$ $(-1)$-curves on the double cover, and then contracting the images of the $9$ $(-2)$-curves).\ Suppose $C$ and $D$ meet in $4$ points in $\mathcal{P}_{9}$. The intersection being transverse, the strict transforms ${\mkern 1.5mu\overline{\mkern-1.5mu{C}\mkern-1.5mu}\mkern 1.5mu},{\mkern 1.5mu\overline{\mkern-1.5mu{D}\mkern-1.5mu}\mkern 1.5mu}$ of the curves $C,D$ under the $3$ blow-ups at each cusps are two disjoint curves not meeting the branch curve. As above the $4$ curves above the remain disjoint in $X_{\l}$ after contracting the $(-1)$-curves.\ If $C$ and $D$ meet in $3$ points in $\mathcal{P}_{9}$ then they meet transversely at a unique point not in $\mathcal{P}_{9}$. Then taking the above notations, we have this time ${\mkern 1.5mu\overline{\mkern-1.5mu{C}\mkern-1.5mu}\mkern 1.5mu}{\mkern 1.5mu\overline{\mkern-1.5mu{D}\mkern-1.5mu}\mkern 1.5mu}=1$, so that the last assertion holds. Let $\mathcal{P}_{q}$ and $\mathcal{C}_{q}$ as above. Using Theorem \[thm:8\_5Config\] and Lemma \[lem:2A2config\], we get: The strict transform by $\eta$ of the $8$ conics in $\mathcal{C}_{q}$ forms an $8\mathbf{A}_{2}$-configuration. For each point $q=p_{j},$ $j\in\{1,\dots,9\}$, we denote by $\mathcal{A}_{j}'$ the corresponding $8\mathbf{A}_{2}$-configuration on $X_{\l}$. In order to obtain new $9{\bf A}_{2}$-configurations, one needs to find other $\mathbf{A}_{2}$-configurations. This will be done in the next section by using singular quartics instead of conics. Using a computer, we found eight $8{\bf A}_{2}$-configurations in the set of $32$ $\cu$-curves which is the union of the two $8\mathbf{A}_{2}$-configurations $\mathcal{A}_{j}'$ and $\{A_{1},A_{1}',\dots,A_{9},A_{9}'\}\setminus\{A_{j},A_{j}'\}$. However one can compute that the orthogonal complement of $6$ of them are lattices with no $\cu$-classes, thus one cannot complete these $6$ configurations into $9\mathbf{A}_{2}$-configurations. The $32$ $\cu$-curves can be realized as lines in a projective model of $X_{\l}$, see Proposition \[prop:degree8Model\]. \[subsec:-9new-9A2-configurations\]Nine new $9\mathbf{A}_{2}$-configurations ----------------------------------------------------------------------------- Let $p_{j}\in\mathcal{P}_{9}$ be one of the $9$ cusp singularities of the sextic $C_{\l}$. There exists a unique quartic curve $Q_{j}$ containing $\mathcal{P}_{9}$ with a unique singularity at the point $p_{j}$, which is of multiplicity $3$. The singularity has type ${\bf D}_{5}$: it has two tangents, one branch is smooth while the other branch is a cuspidal singularity. The tangent at the cusp of $Q_{j}$ is also the tangent to the cuspidal singularity of the sextic $C_{\l}$ at $p_{j}$.\ The curve $Q_{j}$ has geometric genus $0$. Its strict transform by $\eta$ on $X_{\l}$ is the union of two $\cu$-curves $\g_{j},\g_{j}'$ which form an $\mathbf{A}_{2}$-configuration. The curves $\g_{j},\g_{j}'$ and the $16$ curves in $\mathcal{A}_{j}'$ form a $9\mathbf{A}_{2}$-configuration $\mathcal{A}_{j}$. We give in the Appendix the equations of the $9$ curves $Q_{j},$ $j\in\{1,\dots,9\}$. These curves have been constructed using the `LinSys` program by C. Rito which enables to find curves of given degree with prescribed singularities and given tangencies at a set of points in the plane. Conversely, one can check that the singularity of $Q_{j}$ at $p_{j}$ has multiplicity $3$, is resolved by one blow-up, with the exceptional divisor meeting the strict transform in two points, one of multiplicity $2$.\ The curve $Q_{j}$ has genus $0$, (see e.g. [@GH Chapter 4, Section 2]). By Bézout’s Theorem, the intersections of the quartic $Q_{j}$ with the $8$ conics in $\mathcal{C}_{12}$ that contain $p_{j}$ are transverse, so that the curves in $\mathcal{A}_{j}'$ are disjoint from $\g_{j},\g_{j}'$ and we thus get a $9\mathbf{A}_{2}$-configuration. See Figure \[subsec:confOfConics\] for the behavior of the quartic curve $Q_{j}$ under the double cover and its desingularization. \[figniceDiag\] (4,-5) node [The horizontal arrows are blow-up maps]{}; (0,-2.5) node [$X_\lambda$]{}; (0,1.7) node [$\mathbb{P}^2$]{}; plot (, \^3); plot (, -\^3); plot (, 2\*\^3); plot (, -2\*\^3); (1,1) arc (0:180:0.5) ; (0,-1) – (0,1); (0.75,-1) node [$Q_j$]{}; (-0.48,-1) node [$C_\lambda$]{}; (1.5,0) – (2.2,0); plot (/2+3, sqrt abs ); plot (/2+3, -sqrt abs ); plot (/2+3, sqrt abs ); plot (/2+3, -sqrt abs ); (3,-1.3) – (3,1.9); (3.5,1) arc (-35:145:0.5) ; (3.9,0) – (4.6,0); (4.8,0) – (6,0); (5.4,1.8) – (5.4,-1.2); (4.8,1.2) – (6,-1.2); (4.8,-1.2) – (5.8,1.2\*2/3); (5.8,1.2\*2/3) arc (-20:85:0.6) ; (6.3,0) – (7,0); (7.3,0) – (8.8,0); (7.5,1.3) – (7.5,-1.3); (7.9,1.3) – (7.9,-0.9); (8.2,0.5) – (8.2,-1.3); (8.5,1.3) – (8.5,-0.9); (8.2,0.5) arc (180:90:0.5) ; (7.85,-1.1) node [$-2$]{}; (8.5,-1.1) node [$-3$]{}; (9.1,0) node [$-1$]{}; (8,-1.4) – (8,-2.1); (9.1,-1.7) node [$2:1$ cover]{}; (7.3-0.3,-3.5) – (9,-3.5); (7.5-0.3,1.3-3.5) – (7.5-0.3,-1-3.5); (7.9-0.3,1.3-3.5) – (7.9-0.3,-1-3.5+0.3); (8.2-0.3,0.5-3.5) – (8.2-0.3,-1-3.5); (8.5-0.3,1.3-3.5) – (8.5-0.3,-1-3.5+0.3); (9-0.3,1.3-3.5) – (9-0.3,-1-3.5+0.3); (8.2-0.3,0.5-3.5) arc (180:90:0.5) ; (8.7-0.3,0.5-3.5) arc (180:90:0.5) ; (8.7-0.3,0.5-3.5) – (8.7-0.3,-1-3.5); (9.3,-3.5) node [$-2$]{}; (8.5-0.35,-1-3.36) node [$-3$]{}; (9-0.3,-1-3.36) node [$-3$]{}; (7.9-0.32,-1-3.36) node [$-1$]{}; (6.15,-3.5) – (6.85,-3.5); (7.3-0.3-3.2,-3.5) – (9-3.2,-3.5); (7.5-0.3-3.2,1.3-3.5) – (7.5-0.3-3.2,-1-3.5); (8.2-0.3-3.2,0.5-3.5) – (8.2-0.3-3.2,-1-3.5); (8.5-0.3-3.2,1.3-3.5) – (8.5-0.3-3.2,-1-3.5+0.3); (9-0.3-3.2,1.3-3.5) – (9-0.3-3.2,-1-3.5+0.3); (8.2-0.3-3.2,0.5-3.5) arc (180:90:0.5) ; (8.7-0.3-3.2,0.5-3.5) arc (180:90:0.5) ; (8.7-0.3-3.2,0.5-3.5) – (8.7-0.3-3.2,-1-3.5); (8.5-0.35-3.2,-1-3.36) node [$-3$]{}; (9-0.3-3.2,-1-3.36) node [$-3$]{}; (9.3-3.45,-3.35) node [$-1$]{}; (2.9,-3.5) – (3.6,-3.5); (0.8,-2.2) – (0.8,-4.5); (0.2,-3.8) – (2.8,-2.5); (0.2,-3.2) – (2.8,-4.5); (0.4,-3.1) – (1.6,-4.3); (0.4,-3.9) – (1.6,-2.7); (1.6,-2.7) arc (135:18:0.65) ; (1.6,-4.3) arc (360-135:360-20:0.65) ; (2.9,-2.4) node [$-2$]{}; (2.9,-4.6) node [$-2$]{}; (2.7,-3.9) node [$\gamma_j$]{}; (2.7,-3.2) node [$\gamma_j'$]{}; In Section \[subsec:A-Hessian-model\], we study some automorphisms of $X_{\l}$. We obtain that the nine above $9{\bf A}_{2}$-configurations are in the same orbit under the action by $3$-torsion of the Mordell-Weil group of a fibration of $X_{\l}$. Projective, Hessian and Weierstrass models of $X_{\protect\l}$ ============================================================== \[subsec:NaturalEllFib\]The natural elliptic fibration of $X_{\protect\l}$ -------------------------------------------------------------------------- Let $D_{2}$ be the big and nef divisor on $X_{\l}$ which is the pull back of a line in $\PP^{2}$. For $\l$ generic, the divisors $D_{2},A_{1},A_{1}'\dots,A_{9},A_{9}'$ form a $\QQ$-base $\mathcal{B}$ of $\NS X_{\l})_{/\QQ}$; they generate an index $3^{6}$ sub-lattice of $\NS X_{\l})$ (see also Proposition \[prop:The-NS-latt-Discri54\]). Let $\mu:Y_{\l}\to\PP^{2}$ be the blow-up of the plane at the $9$ cusps of the sextic curve $C_{\l}$ and let $E_{1},\dots,E_{9}$ be the exceptional curves. The strict transform by $\mu$ of the curve $C_{\l}$ is the smooth genus $1$ curve $${\mkern 1.5mu\overline{\mkern-1.5mu{C}\mkern-1.5mu}\mkern 1.5mu}_{\l}=\mu^{*}C_{\l}-2{\textstyle \sum_{j=1}^{9}}E_{i},\text{ such that }{\mkern 1.5mu\overline{\mkern-1.5mu{C}\mkern-1.5mu}\mkern 1.5mu}_{\l}^{2}=0,$$ where $E_{1},\dots,E_{9}$ are the exceptional curves over $p_{1},\dots,p_{9}$. The surface $X_{\l}$ is the double cover of $Y_{\l}$ branched over $\bar{C_{\l}}$; we denote by $$\eta':X_{\l}\to Y_{\l}$$ the double cover morphism (so that $\eta'^{*}E_{j}=A_{j}+A_{j}'$) and by $F_{o}$ the ramification locus, so that $2F_{o}=\eta^{'*}\bar{C}_{\l}$. Since $2F_{o}=\eta'^{*}\bar{C_{\l}}\equiv6D_{2}-2\sum_{i=1}^{9}A_{j}+A_{j}'$, we get $$F_{o}\equiv3D_{2}-\left({\textstyle \sum_{j=1}^{9}}A_{j}+A_{j}'\right).$$ Let $L\hookrightarrow\PP^{2}$ be a line; the curve $C_{\l}$ belongs to the linear system $$\d=|6L-2{\textstyle \sum_{j=1}^{9}}p_{j}|$$ of sextic curves with a double point at points in $\mathcal{P}_{9}$. One computes that this linear system is $1$ dimensional. Moreover there exists a unique cubic curve ${\mathrm{Ca}}(\l)$ (called the Cayleyan curve, see [@AD]) that contains the $9$ points in $\mathcal{P}_{9}$, which is $${\mathrm{Ca}}(\l):\,\,x^{3}+y^{3}+z^{3}-\tfrac{(\l^{3}+2)}{\l}xyz=0,\label{eq:cayleysian}$$ so that $2{\mathrm{Ca}}(\l)\in\d$. The linear system $\d$ lifts to a base point free linear system $\d'$ on $Y_{\l}$ with $\bar{C}_{\l}\in\d'.$ The linear system $\d'$ defines a morphism $\varphi':Y_{\l}\to\PP^{1}$ and induces an elliptic fibration $$\varphi:X_{\l}\to\PP^{1}$$ for which $F_{o}$ is a fiber. Let $p,q$ be the images by $\varphi'$ of the strict transforms in $X_{\l}$ of ${\mathrm{Ca}}(\l)$ and $C_{\l}$. In fact the surface $X_{\l}$ is the fiber product of the fibration $\varphi'$ and the quadratic transformation $\PP^{1}\to\PP^{1}$ branched at $p,q$. Indeed both maps $X_{\l}\to Y_{\l}$ and $Y_{\l}\times_{\PP^{1}}\PP^{1}$ has the same branch locus in the rational surface $Y_{\l}$.\ The curves $A_{1},A_{1}',\dots,A_{9},A_{9}'$ are sections of $\varphi$, and one can check that the curves $\t_{1},\t_{1}',\dots,\t_{9},\t'_{9}$ are also sections (see Figure \[subsec:confOfConics\]). \[Thm:24InTheFibers\] The fibration $\varphi$ contracts the $24$ $\cu$-curves $\varTheta_{j}$, $j\in\{1,\dots,24\}$ which are above the $12$ conics $\mathcal{C}_{12}$. The singular fibers of $\varphi$ are $8$ fibers of type $\tilde{\mathbf{A}}_{2}$, each singular fiber is the union of $3$ curves $\varTheta_{j}$. For $\l$ generic, the fibration $\varphi$ has fibers with non-constant moduli and the Mordell-Weil group of the fibration $\varphi$ is $\ZZ\times(\ZZ/3\ZZ)^{2}$. The following four sextic curves $$\begin{array}{c} C_{123456}+C_{123789}+C_{456789},\,C_{124578}+C_{134679}+C_{235689},\\ C_{124689}+C_{135678}+C_{234579},\,C_{125679}+C_{134589}+C_{234678}, \end{array}$$ belong to the linear system $\d$ of sextic curves that have multiplicity $2$ at the points in $\mathcal{P}_{9}$; actually their singularities are nodes. By the results in the proof of Theorem \[thm:12Conics9Points\], the strict transform to $Y_{\l}$ of the above $4$ sextic form $4$ fibers of type $\tilde{\mathbf{A}}_{2}$, which lies in the étale locus of $\eta$. Their strict transform on $X_{\l}$ is therefore the union of eight fibers of type $\tilde{\mathbf{A}}_{2}$. A fiber of type $\tilde{\mathbf{A}}_{2}$ contributes to $3$ in the Euler characteristic of $X_{\l}$, which is equal to $24$. Since there are $8\tilde{\mathbf{A}}_{2}$ singular fibers, the fibration has no other singular fibers. The $24$ curves $\varTheta_{j}$ above the $12$ conics are in the fibers, thus are contracted by $\varphi$. The strict transform ${\mathrm{He}}(\l)$ on $X_{\l}$ of ${\mathrm{Ca}}(\l)$ is smooth, of genus $1$ (we will see that this is the Hessian of the curve $E_{\l}$, thus the notation; see also Remark \[rem:doublecoverCayleysian\]). Since ${\mathrm{He}}(\l)\cdot F_{o}=0$, we have that ${\mathrm{He}}(\l)\equiv F_{o}$. The curve $F_{o}$ is isomorphic to $E_{\l}$. For generic $\l$, the curves ${\mathrm{Ca}}(\l)$ and $E_{\l}$ have distinct $j$-invariants, thus the fibers of $\varphi$ have a non-constant moduli. Since the fibration is not isotrivial, results of Shioda (see [@Shioda Corollary 1.5]) apply and tell that the Mordell-Weil group of sections of $\varphi:X_{\l}\to\PP^{1}$ has rank $1=19-(2+8(3-1))$. In fact, elliptic fibrations of K3 surfaces are classified by Shimada in [@Shimada]. A table with the $3278$ possible cases is available in [@Shimada1]. Our fibration is case number $2373$ in that table, where one can find moreover that the torsion part of its Mordell-Weil group is isomorphic to $(\ZZ/3\ZZ)^{2}$. \[subsec:Two-polarizations-and\]Two polarizations and a degree $8$ projective model ------------------------------------------------------------------------------------ The divisor $$D_{14}=4D_{2}-\left({\textstyle \sum_{j=1}^{9}}A_{j}+A_{j}'\right)$$ is linearly equivalent to $D_{2}+F$ ($F$ a fiber of $\varphi$) and is effective. Let us define $D_{8}=D_{14}-(A_{1}+A_{1}')$. \[prop:degree8Model\]The divisors $D_{8}$ and $D_{14}$ are ample of square $D_{8}^{2}=8,$ $D_{14}^{2}=14$. The linear system $|D_{8}|$ is base point free, non-hyperelliptic, and defines an embedding $X_{\l}\hookrightarrow\PP^{5}$ as a degree $8$ surface. For $d\in\NN^{*}$, let $n_{d}$ be the number of $\cu$-curves of degree $d$ for $D_{8}$. The series $\sum n_{d}T^{d}$ begins with $$32T+20T^{2}+334T^{4}+576T^{5}+880T^{6}+8640T^{7}+17784T^{8}...,$$ in particular $X_{\l}$ contains $32$ lines and $20$ conics. Let $B$ be a $\cu$-curve such that $D_{14}B\leq0$. Since $D_{2}$ is effective and $D_{2}^{2}>0$, one has $D_{2}B\geq0$, moreover since $F$ is a fiber, $FB\geq0$ and we must have $D_{2}B=0=FB$. That implies that $B$ is an irreducible component of a singular fiber, ie $B=\varTheta_{j}$ for some $j\in\{1,\dot{,24\}}$. But since $D_{2}\varTheta_{j}=2$ for all $j$, such a curve $B$ cannot exist, thus $D_{14}$ is ample. Let us prove that $D_{8}$ is ample. We have $$\g_{1}+\g_{1}'\equiv4D_{2}-2(A_{1}+A_{1}')-{\textstyle \sum_{j=1}^{9}}(A_{j}+A_{j}'),$$ thus $D_{8}\equiv A_{1}+A_{1}'+\g_{1}+\g'_{1}$ and the divisor $D_{8}$ is effective. We check that $D_{8}A_{1}=D_{8}A_{1}'=$$D_{8}\g_{1}=D_{8}\g'_{1}=1$ and $D_{8}^{2}=8$, therefore $D_{8}$ is nef and big. Suppose that there is a $\cu$-curve $B$ on $X_{\l}$ such that $D_{8}B=0$. Then by the above expression of $D_{8}$, one has $A_{1}B=A_{1}'B=0$. Let $L\hookrightarrow\PP^{2}$ be a line. For $j\in\{2,\dots,9\}$, let us consider the linear system $$\d_{j}=|4L-(p_{1}+p_{j}+{\textstyle \sum_{j=1}^{9}}p_{k})|$$ of the quartic curves that go through the points in $\mathcal{P}_{9}$ and with multiplicity $2$ at $p_{1}$ and $p_{j}$. Using `LinSys`, one can compute that for each $j>1$, the linear system $\d_{j}$ is a pencil of curves and the base points set is $\mathcal{P}_{9}$. Moreover, the generic element $\vartheta_{j}$ of $\d_{j}$ is an irreducible curve of geometric genus $1$ which cuts $C_{\l}$ in $\mathcal{P}_{9}$ and two more points. Thus we obtain that for each $j>1$, the strict transform of $\vartheta_{j}$ is an irreducible curve $\G_{j}$ such that $$D_{8}\equiv\G_{j}+A_{j}+A_{j}'\text{ and }\G_{j}^{2}=2.$$ Since $D_{8}B=0$, we obtain $A_{j}B=A_{j}'B=0$ for all $j\in\{1,\dots,9\}$. Since the orthogonal of the classes $A_{j},A_{j}'$, $j\in\{1,\dots,9\}$ (on which $B$ belongs) is generated by $D_{2}$, the class of $B$ must be a multiple of $D_{2}$ and have positive square, which is absurd. Therefore $D_{8}$ is ample. Suppose that there is a fiber $F'$ such that $D_{8}F'\in\{1,2\}$. Observe that by using the expression for $D_{8}$, we get that $F'\Gamma_{j}=0,1,2$. If $F'\Gamma_{j}=0$, then $\Gamma_{j}$ is contained in a fiber of the fibration determined by $F'$, but this is not possible since $\Gamma_{j}^{2}=2$. If $F'\Gamma_{j}=1$, then $\Gamma_{j}$ is a section of the fibration so is a rational curve, but again this is not possible. If $F'\Gamma_{j}=2$ (we can assume that this holds for all $j$, otherwise we are in a previous case), then $F'$ is in the orthogonal complement of the $A_{j},A_{j}'$ but this is not possible since this is generated by $L$, which is of square $2$. Therefore there are no such fiber $F'$ and using [@SaintDonat], we obtain that the linear system $|D_{8}|$ is base-point free and gives an embedding of $X_{\l}$. With respect to the divisor $D_{8}$, the degrees of the curves $A_{1},A_{1}',\g_{1},\g_{1}'$ equal $2$ and the degrees of curves $A_{i},A_{i}',\,i\geq2$ is $1$. For the assertions on the number of rational curves of degree $d\leq8$ we used an algorithm (see e.g. [@Roulleau]), which computes the classes of $\cu$-curves in $\NS X_{\l})$ of given degrees with respect to a fixed ample class. Proceeding in a similar way as in the proof of Proposition \[prop:degree8Model\], we obtain: Let $i,j\in\{1,\dots,9\}$, $i\neq j$. The divisor $$D_{i,j}=D_{14}-(A_{i}+A_{i}'+A_{j}+A_{j}')$$ is nef of square $2$ and the linear system $|D_{i,j}|$ is base point free. One can compute that the intersection with $D_{ij}$ is $0$ for the $10$ curves $\t_{ijklmn},\t'_{ijklmn}$ (where $\{k,l,m,n\}\subset\{1,\dots,9\}$ is a set of $4$ elements such that the conic $C_{ijklmn}$ exists), and for the $\cu$-curve which is the strict transform on $X_{\l}$ of the line through cusps $p_{i},p_{j}$. \[subsec:A-Hessian-model\]A Hessian model of the natural fibration of $X_{\protect\l}$ -------------------------------------------------------------------------------------- ### The generic fiber of the elliptic fibration $\varphi$ and $18$ rational points Let $f_{\l}$ be the equation of the $9$ cuspidal sextic $C_{\l}$ which is the dual of $E_{\l}$, and let $c_{\l}$ be the equation of the Cayleyan elliptic curve ${\mathrm{Ca}}(\l)$ (see equation ), the unique cubic that goes through the $9$ cusps of the sextic curve $C_{\l}$. We recall (see Section \[subsec:NaturalEllFib\]) that $Y_{\l}$ is the blow-up of the plane at the $9$ points in $\mathcal{P}_{9}$; it has a natural elliptic fibration $\varphi'$, coming from the pencil of sextic curves which have double points at the $9$ points in $\mathcal{P}_{9}$, pencil which is generated by $C_{\l}$ and $2{\mathrm{Ca}}(\l)$. A singular model of $Y_{\l}$ is therefore obtained as the surface in $\PP^{1}\times\PP^{2}$ with equation $uf_{\l}-vc_{\l}^{2}=0$, where $u,v$ are the coordinates of $\PP^{1}$. The projection onto $\PP^{1}$ induces the fibration $\varphi':Y_{\l}\to\PP^{1}$. A singular model of the $K3$ surface $X_{\l}$ is the surface $X_{\l}^{sing}$ in $\PP^{1}\times\PP^{2}$ with equation $u^{2}f_{\l}-v^{2}c_{\l}^{2}=0$; again the projection onto $\PP^{1}$ induces the natural fibration $X_{\l}^{sing}\to\PP^{1}$, where the generic fibers are $9$-nodal sextic curves. In order to obtain a smooth model of $X_{\l}$, let us consider the linear system $L_{4}(\mathcal{P}_{9})$ of quartics that contain the $9$ cusps. The linear system $L_{4}(\mathcal{P}_{9})$ has (projective) dimension $5$ and defines a rational map $\phi:\PP^{2}\dashrightarrow\PP^{5}$ not defined on $\mathcal{P}_{9}$. One computes that the image of $X_{\l}^{sing}$ by the rational map $$(i_{d},\phi):\PP^{1}\times\PP^{2}\dashrightarrow\PP^{1}\times\PP^{5}$$ is a smooth model of $X_{\l}$; from Section \[subsec:NaturalEllFib\], the images of the cusps are the $18$ $\cu$-curves on $X_{\l}$ forming a $9{\bf A}_{2}$-configuration. Taking the generic point over $\PP^{1}$, one get a smooth genus $1$ curve in $\PP_{/\QQ(t)}^{5}$ (where $t=\frac{u}{v}$). That curve $E_{K3}$ has naturally $18$ rational points, corresponding to the $18$ $\cu$-curves. Using Magma, we computed a Hessian model $E_{K3}\hookrightarrow\PP_{/\QQ(t)}^{2}$, which is \[thm:AHessian-model-of\]A model of the generic fiber of the fibration $X_{\l}\to\PP^{1}$ is $$E_{K3}\,\,\,\,\,x^{3}+y^{3}+z^{3}+\frac{\l^{3}(t^{2}+3)-4t^{2}}{\l^{2}(t^{2}-1)}xyz=0.\label{eq:ofEK3}$$ The elliptic curve $E_{K3}$ contains the $9$ obvious $3$-torsion points $$\begin{array}{c} Q_{1}=(0:-1:1),\,\,Q_{2}=(-1:0:1),\,\,Q_{3}=(-1:1:0),\hfill\\ Q_{4}=(0:-\o:1),\,\,Q_{5}=(\o+1:0:1),\,\,Q_{6}=(-\o:1:0),\hfill\\ Q_{7}=(0:\o+1:1),\,\,Q_{8}=(-\o:0:1),\,\,Q_{9}=(\o+1:1:0). \end{array}$$ (where $\o^{2}+\o+1=0$; we take $Q_{1}$ as the neutral element) and the following $9$ points $$\begin{array}{c} P_{1}=(-2t:\l(t+1):\l t+\l),\hfill\\ P_{2}=(\l(t-1):-2t:\l t+\l),\hfill\\ P_{3}=(-\l t-\l:-\l t+\l:2t),\hfill\\ P_{4}=((2\o+2)t:\l(\o t+\o):\l t+\l)\hfill\\ P_{5}=(\l(\o+1)(-t+1):-2\o t:\l t+\l),\hfill\\ P_{6}=((\o+1)\l(t+1):\o\l(-t+1):2t),\hfill\\ P_{7}=(-2\o t:-\l(\o+1)(t+1):\l t-\l),\hfill\\ P_{8}=(\l\o(t-1):(2\o+2)t:\l t+\l),\hfill\\ P_{9}=(-\l\o(t-1):(\o+1)\l(t-1):2t).\hfill \end{array}$$ Together, these $18$ points are the above-mentioned rational points of $E_{K3/\QQ(\o,t)}$ corresponding to the $18$ sections of the fibration of $X_{\l}$. For the neutral element of $E_{K3}$, let us choose $Q_{1}$. One can check that point $P_{k}$ is the translate of $P_{1}$ by the $9$ torsion point $Q_{k}$ ($k\in\{1,\dots,9\}$), i.e. $P_{k}=P_{1}+Q_{k}$. ### A smooth model of $X_{\protect\l}$ in $\protect\PP^{1}\times\protect\PP^{2}$ By taking the homogenization of the generic fiber $E_{K3}$ in , we get a natural model of the K3 surface $X_{\l}$ as $$\l^{2}(u^{2}-v^{2})(x^{3}+y^{3}+z^{3})+(\l^{3}(u^{2}+3v^{2})-4u^{2})xyz=0.\label{eq:HessianK3}$$ in the space $\PP^{1}\times\PP^{2}$ (with coordinates $u,v;x,y,z$, where $t=\frac{u}{v}$). That model is smooth, and the generic fibers are smooth cubic curves, by contrast with the previous model $X_{\l}^{sing}$. We denote by $(P)$ the section in $X_{\l}$ corresponding to the points $P\in E_{K3}$. Using Magma, it is then possible to obtain the equations of the $\cu$-curves (also sections) $A_{j}=(Q_{j})$, resp. $A_{j}'=(P_{i})$, which are on $X_{\l}\hookrightarrow\PP^{1}\times\PP^{2}$. We can check that: The $9$ curves $A_{j}+A_{j}'$ ($j\in\{1,\dots,9\}$) form a $9\mathbf{A}_{2}$-configuration. We use the equations of the $\cu$-curves $A_{j},A_{j}'$ in the model $X_{\l}\subset\PP^{1}\times\PP^{2}$ to check that $A_{j}A_{j}'=1$ and $A_{j}A_{k}'=A_{j}A_{k}=A_{j}'A_{k}'=0$ for $k\neq j$. In fact, one already knows that $3$-torsion sections are disjoint by [@Miranda VII, Proposition 3.2] (thus the sections $A_{j}'$, being translated of the group of $3$-torsion sections, are also disjoint). One can check moreover that the $9$ intersection points of $A_{j}$ with $A_{j}'$ for $i=1,\dots,9$ are on the fiber over $0$ of the fibration $\varphi$, fiber which is isomorphic to $E_{\l}$. Using the addition law on the elliptic curve $E_{K3}$, one can find other points of $E_{K3}$, and therefore sections of $\varphi$. By example the following points $$\begin{array}{c} R_{1}=(-2t:\l(t-1):\l t+\l),\hfill\\ R_{2}=(\l(t+1):-2t:\l t-\l),\hfill\\ R_{3}=(-\l t+\l:-\l t-\l:2t),\hfill\\ R_{4}=((2\o+2)t:\l\o(t-1):\l t+\l),\hfill\\ R_{5}=(-\l(\o+1)(t+1):-2\o t:\l t-\l),\hfill\\ R_{6}=((\o+1)\l(t-1):-\o\l(t+1):2t),\hfill\\ R_{7}=(-2\o t:\l(\o+1)(-t+1):\l t+\l),\hfill\\ R_{8}=(\l\o(t+1):(2\o+2)t:\l t-\l),\hfill\\ R_{9}=(-\o\l t+\o\l:(\o+1)(\l t+\l):2t),\hfill \end{array}$$ are the points $R_{i}=-P_{1}+Q_{i},$ $i\in\{1,\dots,9\}$. We have The $9$ curves $A_{i},\,(R_{i})$, $i=1,...,9$, form a $9{\bf A}_{2}$-configurations. The curves $A_{i}$ (resp. $(R_{i})$) are images of the curves $A_{i}'$ (resp. $A_{i}$) by the translation by $-A_{1}'$ and we already know that the curves $A_{i},A_{i}'$ form a $9{\bf A}_{2}$-configuration. We already know that the fiber at $0$ of the elliptic K3 surface $\varphi:X_{\l}\to\PP^{1}$ is isomorphic to $E_{\l}$, moreover: \[rem:doublecoverCayleysian\]From equation \[eq:HessianK3\], the fiber at $\infty$ of $X_{\l}\to\PP^{1}$ is the elliptic curve $${\mathrm{He}}(\l):\,\,\,\,x^{3}+y^{3}+z^{3}+\tfrac{(\l^{3}-4)}{\l^{2}}xyz=0,$$ which is in fact the Hessian of the curve $E_{\l}$. The $j$-invariants of ${\mathrm{Ca}}(\l)$ and ${\mathrm{He}}(\l)$ are distinct, in particular these curves are not isomorphic. The Cayleyan curve $\text{Ca}(\l)$ is the quotient of Hessian ${\mathrm{He}}(\l)$ of $E_{\l}$ by a $2$-torsion point; in particular the two curves are $2$-isogeneous, see [@AD]. ### \[subsec:A-CompletInterModelInP5\]A degree 8 non-complete intersection model in $\protect\PP^{5}$ One can check that the map from $X_{\l}\subset\PP^{1}\times\PP^{2}$ to $\PP^{5}$ obtained as the product of the identity map of $\PP^{1}$ with the Segre embedding composed with the projection to $\PP^{5}$, is an embedding with image a degree $8$ K3 surface in $\PP^{5}$ defined by the following $5$ equations: $$\begin{array}{c} -U_{2}U_{4}+U_{1}U_{5},\,\,\,\,-U_{2}U_{3}+U_{0}U_{5},\,\,\,\,-U_{1}U_{3}+U_{0}U_{4},\hfill\\ \lambda^{2}(U_{0}^{2}U_{3}-U_{3}^{3}+U_{1}^{2}U_{4}-U_{4}^{3})+(\lambda^{3}-4)U_{0}U_{1}U_{5}\\ \hfill+\lambda^{2}(U_{2}^{2}U_{5}+3\lambda U_{3}U_{4}U_{5}-U_{5}^{3}),\\ \lambda^{2}(U_{0}^{3}+U_{1}^{3}+U_{2}^{3}-U_{0}U_{3}^{2})+(\lambda^{3}-4)U_{0}U_{1}U_{2}\\ \hfill+\lambda^{2}(-U_{1}U_{4}^{2}+3\lambda U_{0}U_{4}U_{5}-U_{2}U_{5}^{2}), \end{array}$$ in particular this is not a complete intersection (in fact, by using [@Beauville Chapter VIII, Exercice 11], a K3 surface has a degree $8$ smooth model which is not a complete intersection if and only if it has a smooth model in $\PP^{1}\times\PP^{2}$ of bi-degree $(2,3)$). Using the images of the known $\cu$-curves on $X_{\l}$, one finds that this surface in $\PP^{5}$ contains at least $33$ lines. The involution $\s'$ defined by $u\to(-u_{0}:-u_{1}:-u_{2}:u_{3}:u_{4}:u_{5})$ acts on $X_{\l}\hookrightarrow\PP^{5}$. Also one can check that the order $3$ automorphisms $$\begin{array}{c} \a_{1}:u\to(u_{1}:u_{2}:u_{0}:u_{4}:u_{5}:u_{3}),\hfill\\ \a_{2}:u\to(u_{0}:\o u_{1}:\o^{2}u_{2}:u_{3}:\o u_{4}:\o^{2}u_{5}) \end{array}$$ act on $X_{\l}$ and so does the involution $$\b:u\to(u_{0}:u_{2}:u_{1}:u_{3}:u_{5}:u_{4}).$$ The fixed point set of $\a_{1}$ is the union of $6$ points, the fixed point set of $\beta$ is a genus $2$ curve. The involution $\s'\b$ is symplectic. Using the above equations of $X_{\l}$ and the equations of curves $A_{1},\dots,A_{9}'$ in $\PP^{5}$, one can check that the automorphism group $G_{\mathcal{C}}$ that preserve globally the $9{\bf A}_{2}$-configuration $A_{1},\dots,A_{9}'$ contains the group $G_{18}$ isomorphic to $\ZZ_{3}\rtimes S_{3}$ generated by $\a_{1},\a_{2},\s'\b$. We prove in Section \[subsec:stabilizerGroupFirstConf\] that $G_{18}$ has index $2$ in the group $G_{\mathcal{C}}$ preserving the configuration $A_{1},\dots,A_{9}'$. \[subsec:A-Weierstrass-equation\]A Weierstrass equation ------------------------------------------------------- We recall that $\o$ is such that $\o^{2}+\o+1=0$. In the Appendix, we define three polynomials $A,B,D$ in $\QQ(\o)(t)$ of respective degree $8$, $12$ and $8$. Let us prove the following result \[thm:TheWeiModel\]The following elliptic curve $$E_{1/\QQ(\o,t)}:\,\,\,y^{2}=x^{3}-\frac{1}{48}Ax+\frac{1}{864}B$$ gives a minimal Weierstrass model of the elliptic K3 surface $X_{\l}$. The $8$ singular fibers $\tilde{\mathbf{A}_{2}}$ of $X_{\l}$ are over the $8$ zeros of $D$. One computes that the $j$-invariant of the elliptic curve $E_{K3}$ is $$j=-\frac{A^{3}}{(\l^{2}(\l^{3}-1)D)^{3}}.$$ For any $j\notin\{0,1728\}$, the elliptic curve $$E_{0}(j)\,\,\,\,y^{2}=x^{3}-\frac{1}{48}\frac{j}{j-1728}x+\frac{1}{864}\frac{j}{j-1728}$$ has $j$-invariant equal to $j$. In our case, we compute that we have $$\frac{j}{j-1728}=\frac{A^{3}}{B^{2}},$$ where $A$ and $B$ are defined in the Appendix. By taking the change of variables $$x'=u^{2}x,\,y'=u^{3}y$$ with $u=(B/A)^{1/2}$ in the equation of $E_{0}(j)$, we obtain the elliptic curve $E_{1}$. The curve $E_{1}$ has also its $j$-invariant equals to $j$, is also in Weierstrass form, but its coefficients are coprime degree $8$ and $12$ polynomials in $t$. The discriminant of the equation of $E_{1}$ is $$\Delta=-(\l^{2}(\l^{3}-1)D)^{3},$$ where $D$ is a product of $8$ degree $1$ polynomials in $t$ (see appendix). According to [@Miranda Table IV.3.1], the associated elliptic surface is a K3 surface with $8$ singular fibers of type $\tilde{{\bf A}}_{2}$. Using Magma, we finally obtain an isomorphism defined over $\QQ(\o,t)$ between the Hesse model $E_{K3}$ and the Weierstrass model $E_{1}$. On automorphisms of the K3 surface $X_{\protect\l}$ =================================================== \[subsec:On-the-Mordell-Weil\]On the Mordell-Weil lattice of the elliptic fibration $\varphi$ --------------------------------------------------------------------------------------------- For a point $P\in E_{K3}(\QQ(\o,t))$, let us denote by $(P)\hookrightarrow X_{\l}$ the corresponding section of $\varphi:X_{\l}\to\PP^{1}$. We denote by $\tau\in\aut(X)$ the automorphism which is the translation by $A_{1}'$. We have Modulo torsion, the section $A_{1}'=(P_{1})$ generates the Mordell-Weil lattice $\text{MWL}(X_{\l})$ of sections. One can compute the classes in $\NS X{}_{\l})$ of the curves $(R_{i})$ (which are the translate by $(-P_{1})$ of the curves $A_{i}$); we give these classes in the Appendix. Using that knowledge, we get the matrix representation on $\NS X_{\l})$ of the action of the automorphism $\tau$. The characteristic polynomial of $\tau$ is $(T-1)^{3}(T^{2}+T+1)^{8}.$ Let $O=A_{1}$ be the zero section, and let $F$ be a fiber of $\varphi:X_{\l}\to\PP^{1}$. Using the knowledge of the action of the automorphism $\tau$ (which is the translation by $(P_{1})$) on $\NS X)$, we get that $$(6P_{1})-2(3P_{1})+O\equiv6F$$ in $\NS X_{\l})$, thus (see e.g. [@Silverman Chapter III, Theorem 9.5]) $\left\langle 3P_{1},3P_{1}\right\rangle =6$ and $\left\langle P_{1},P_{1}\right\rangle =\frac{2}{3}$, where $\left\langle \cdot,\cdot\right\rangle $ is the bilinear pairing on $\text{MWL}(X_{\l})$ associated to the canonical height. Let $\text{Triv}(X_{\l})$ be the lattice generated the zero section and the fibers components of the fibration. The determinant formula [@ShiodaSchutt Corollary 6.39] is $$|\det\,\NS X_{\l})|=|\det\,\text{Triv}(X_{\l})|\cdot\det\,\text{MWL}(X_{\l})/|\text{MWL}(X_{\l})|^{2}.$$ By Proposition \[prop:The-NS-latt-Discri54\], we know that $|\det\,\NS X_{\l})|=54$. We have moreover $\det\,\text{Triv}(X_{\l})=-3^{8}$ and $|\text{MWL}(X_{\l})|^{2}=3^{4}$, thus we obtain that $$\det\,\text{MWL}(X_{\l})=\tfrac{2}{3}.$$ By Theorem \[Thm:24InTheFibers\], the group $\text{MWL}(X_{\l})$ has rank $1$; since $\left\langle P_{1},P_{1}\right\rangle =\frac{2}{3}$, we conclude that $P_{1}$ generates $\text{MWL}(X_{\l})$ modulo torsion. Using the action of $\tau$ and its powers, we can obtain more classes in $\NS X_{\l})$ of the sections on the K3 surface $X_{\l}\to\PP^{1}$. We searched the $9{\bf A}_{2}$-configurations among a set of $45$ sections, but we obtained only the expected ones, i.e. the $9{\bf A}_{2}$-configuration that are translate of the configuration $A_{i},A_{i}'$, $i\in\{1,\dots,9\}$. Since these configurations are images of one configuration by an automorphism (the translation by $A_{1}'$ and its multiples), these $9{\bf A}_{2}$-configurations give the same generalized Kummer structure. \[subsec:More-auto\]More elements of the automorphism group and another double plane model ------------------------------------------------------------------------------------------ The K3 surface $X_{\l}$ is constructed as the minimal desingularization of the double cover of the plane branched over the sextic curve $C_{\l}$. Let $\s\in\aut(X)$ be the corresponding involution. By construction $\s(A_{j})=A_{j}'$, $\s(A_{j}')=A_{j}$ and $\sigma$ preserves the fiber of the fibration $X_{\l}\to\PP^{1}$. From that we know the action of $\s$ on the Néron-Severi group, since we know also the action of $\tau$, one can compute that We have $\s=\tau\s\tau.$ Let us recall that we denoted by $(R_{i})$ the sections corresponding to the points $R_{i}=-P_{1}+Q_{i}$ in $E_{K3}$. We also have a model $X_{\l}\subset\PP^{1}\times\PP^{2}$. From the equations of the curves involved, we obtain that:\ a) the natural map $\pi_{2}:X_{\l}\to\PP^{2}$ induced by the projection $\PP^{1}\times\PP^{2}\to\PP^{2}$ is a $2$ to $1$ map ,\ b) it contracts the $\cu$-curves $A_{1},\dots,A_{9}$ to the $9$ torsion base points $\mathcal{T}_{9}$ of the Hesse pencil $x^{3}+y^{3}+z^{3}-3\mu xyz=0,$ $\mu\in\PP^{1}$.\ c) for $j\in\{1,\dots,9\}$, the curves $A_{j}'$ and $(R_{j})$ are mapped to a line $L_{j}'$ that contains exactly one point of $\mathcal{T}_{9}$, that line is therefore tangent to the sextic curve at its other intersection points. The $9$ lines are in general position. We can therefore compute the action of the involution $\s'\in\aut(X)$ corresponding to the double cover $\pi_{2}$ on the Néron-Severi group (the curves $A_{j}'$ and $(R_{j})$ are exchanged and one can check that the fiber is preserved; the classes of $(R_{j})$ in base $F,A_{j},A_{j}'$ are in the appendix). The action of $\s'$ on $X_{\l}\hookrightarrow\PP^{5}$ is given in sub-section \[subsec:A-CompletInterModelInP5\]. We get: We have $\tau=\s\sigma'$. We compute moreover that the pull-back by $\pi_{2}$ of the $9$ points of $\mathcal{T}_{9}$ are the irreducible curves $A_{1},\dots,A_{9}$, in fact we have The surface $X_{\l}$ is the minimal desingularization of the double cover of the plane branched over the sextic curve $C_{6}$ which is the union of the elliptic curves $$E_{\l}:\,\,\,\,x^{3}+y^{3}+z^{3}-3\l xyz=0$$ and the Hessian of $E_{\l}$: $${\mathrm{He}}(\l):\,\,\,\,x^{3}+y^{3}+z^{3}+\tfrac{(\l^{3}-4)}{\l^{2}}xyz=0.$$ Let $F_{0}$ and $F_{\infty}$ be the two fibers of the fibration $X_{\l}\to\PP^{1}$ at $0$ and infinity. One computes that $\pi_{2}(F_{0})=E_{\l}$, $\pi_{2}(F_{\infty})={\mathrm{He}}(\l)$, and moreover $$\pi_{2}^{*}(E_{\l})=2F_{0}+{\textstyle \sum_{j=1}^{9}}A_{j},\,\,\pi_{2}^{*}({\mathrm{He}}(\l))=2F_{\infty}+{\textstyle \sum_{j=1}^{9}}A_{j},$$ thus $E_{\l}+{\mathrm{He}}(\l)$ is the branch locus of the map $\pi_{2}$. By Bézout Theorem, the singularities of $E_{\l}+{\mathrm{He}}(\l)$ are nodal since the curves $E_{\l}$ and ${\mathrm{He}}(\l)$ meet at at $\mathcal{T}_{9}$. For each $j=1,\dots,9$, the images by $\pi_{2}$ of the two sections $(-2P_{1}+Q_{j}),\,(2P_{1}+Q_{j})$ is the same quartic curve, which is nodal with $3$ nodes. The images of $(-P_{1}+Q_{j}),\,(P_{1}+Q_{j})$ are the $9$ lines, their coordinates in the dual plane with base $x,y,z$ are the same as the points in $\mathcal{P}_{9}$. These $9$ lines are the inflection lines of the curve $E_{\l}$. We recall that the Hesse configuration is the point-line configuration $(9_{4},12_{3})$ of the $9$ points in $\mathcal{T}_{9}$ and the $12$ lines $\mathcal{L}_{12}$ such that each line contains $3$ points in $\mathcal{T}_{9}$ and each point is on $4$ lines. \[prop:The12linesHesse\]The images by $\pi_{2}$ in $\PP^{2}$ of the $24$ irreducible components of the singular fibers of the fibration $\varphi:X_{\l}\to\PP^{1}$ are the $12$ lines of the Hesse configuration. We give in Theorem \[thm:TheWeiModel\] the $8$ points $p\in\PP^{1}$ such that the fiber $F_{p}$ over $p$ is singular. We are then able to compute these singular fibers in $X_{\l}\subset\PP^{1}\times\PP^{2}$ and their images in $\PP^{2}$. Using the elliptic curve $E_{K3}$, we obtain that the sub-group $\text{Tor}_{3}$ of order $3$ elements in the Mordell-Weil lattice $\text{MWL}(X_{\l})$ is generated by two order $3$ elements $t_{1},\,t_{2}$ which acts on $\NS X_{\l})$ via $$t_{1}(A_{j})=A_{\s j},t_{1}(A_{j}')=A_{\s j}',\,\,\,t_{2}(A_{j})=A_{\mu j},t_{2}(A_{j}')=A_{\mu j}'$$ where $\sigma=(1,2,3)(4,5,6)(7,8,9)$ and $\mu=(1,4,7)(2,5,8)(3,6,9)$. The elements of $\text{Tor}_{3}$ commute with $\sigma$ (and of course with $\tau$). The action of $\text{Tor}_{3}$ is transitive on the nine $9{\bf A}_{2}$-configurations we found in section \[subsec:-9new-9A2-configurations\]. \[subsec:stabilizerGroupFirstConf\]On the stabilizer group of natural $9{\bf A}_{2}$-configuration -------------------------------------------------------------------------------------------------- Recall that $X_{\l}$ is the minimal desingularization of the double cover of $\PP^{2}$ ramified over the $9$-cuspidal sextic $C_{\l}$. The strict transform on $X_{\l}$ of $C_{\l}$ is a smooth elliptic curve isomorphic to $E_{\l}$. The linear system defined by that elliptic curve defines an elliptic fibration which we denote $\varphi:X_{\l}\to\PP^{1}$, and which we call the *natural* fibration. The curves $A_{j},A_{j}'$ (above the cusps) are sections of $\varphi$, the $24$ $\cu$-curves $\t_{J},\t_{J}'$ (for some set $J\subset\{1,\dots,9\}$ of order $6$) above the $12$ conics are contained in the fibers (see Proposition \[Thm:24InTheFibers\]), their classes are given in the Appendix, under a simpler labelling $\varTheta_{j},\,j\in\{1,\dots,24\}$. We have \[prop:TheNS\] An entire base $\mathcal{B}_{1}$ of the Néron-Severi group is $$\begin{array}{c} F,A_{1},A_{1}',A_{2},A_{2}',A_{3},A_{3}',A_{4},A_{4}',A_{5},A_{5}',A_{6},A_{7},A_{7}',\\ \varTheta_{5},\varTheta_{14},\varTheta_{22},\varTheta_{23},\varTheta_{20}, \end{array}$$ where $F$ is a fiber of the fibration $\varphi$. The discriminant group $A_{\NS X)}\simeq\ZZ/2\ZZ\times(\ZZ/3\ZZ)^{3}$ is generated by $$\begin{array}{c} w_{0}=\frac{1}{2}(0,1,1,0,0,0,0,1,1,0,0,0,1,1,1,1,1,0,0)_{\mathcal{B}_{1}},\hfill\\ v_{1}'=\frac{1}{3}(A_{3}+2A_{3}'+A_{5}+2A_{5}'+A_{7}+2A_{7}'),\hfill\\ v_{2}'=\frac{1}{3}(A_{2}+2A_{2}'+2A_{4}+A_{4}'+2A_{5}+A_{5}'+A_{7}+2A_{7}'),\\ v_{3}'=\frac{1}{3}(A_{1}+2A_{1}'+A_{4}+2A_{5}'+A_{7}+2A_{7}').\hfill \end{array}$$ By Proposition \[prop:The-NS-latt-Discri54\], the discriminant group has order $54$. The lattice generated by the elements in $\mathcal{B}$ has rank $19$ and discriminant equal to $54$, thus these elements generate $\NS X)$. By taking the inverse of the intersection matrix of the vectors in $\mathcal{B}$, we get the generators of the discriminant group. The non-immediate part of the Gram matrix of the base $\mathcal{B}_{1}$ is the intersection matrix of the $5$ curves $\varTheta_{j},\,j=5,14,22,23,20$ with the curves in $\mathcal{B}_{1}$, which matrix is: Let $G_{\mathcal{C}}$ be the automorphism sub-group of $\aut(X_{\l})$ preserving the configuration $\mathcal{C}=A_{1},A_{1}',\dots,A_{9},A_{9}'$. The proof of the following Proposition will also serve as a preliminary for the proof of Theorem \[MAINthm\] in the next section: \[pro:The-group-Aut-preserveConf\]The group $G_{\mathcal{C}}$ is isomorphic to $\ZZ_{2}\times(\ZZ_{3}\rtimes S_{3})$, where $\ZZ_{3}\rtimes S_{3}$ acts symplectically on $X_{\l}$. The center of $G_{\mathcal{C}}$ is generated by the non-symplectic involution $\s$ associated to the double cover map $X_{\l}\to\PP^{2}$ branched over the cuspidal sextic curve $C_{\l}$. Let $\phi$ be an element of $G_{\mathcal{C}}$. Since it preserves globally the configuration $\mathcal{C}$, it must map its orthogonal complement (generated by $D_{2}$, the pull-back of a line) to itself. There are $2^{9}9!$ bijective maps $$\mu:\{A_{1},A_{1}',\dots,A_{9},A_{9}'\}\to\{A_{1},A_{1}',\dots,A_{9},A_{9}'\}$$ which preserves the incidence relations of $\mathcal{C}$. Since a linear map is defined by the images of the vectors of a base, each map $\mu$ extends to a linear automorphism $\phi_{\mu}:\,\NS X_{\l})\otimes\QQ\to\NS X_{\l})\otimes\QQ$ sending $D_{2}$ to $D_{2}$ and the configuration $\mathcal{C}$ to itself. The action of $\phi$ on $\NS X)$ must be one of these maps. Using the entire basis $\mathcal{B}_{1}$ of Proposition \[prop:TheNS\], we obtain that among all the possibilities, only $864$ matrices in base $\mathcal{B}_{1}$ of such $\phi_{\mu}$ are in $GL_{19}(\ZZ)$. These $864$ matrices are the elements of a group $G_{864}$ isomorphic to the product of $\ZZ/2\ZZ$ with $AGL_{2}(\FF_{3})$, the affine linear group of the space $\FF_{3}^{2}$. The center of $G_{864}$ has order $2$ and is generated by the matrix of the non-symplectic involution $\sigma$ defined in section \[subsec:More-auto\]. In the appendix, we give two generators $g_{1},g_{2}$ (of respective order $8$ and $6$) of the group $G_{432}\subset G_{864}$ isomorphic to $AGL_{2}(\FF_{3})$. Their action on the $2$-torsion part of the discriminant group $A_{\NS X)}$ is trivial, and one computes that their action on the $3$-torsion part of the discriminant group $A_{\NS X)}$ is by the matrices $$\bar{g}_{1}=\left(\begin{array}{ccc} 1 & 1 & 1\\ 2 & 1 & 0\\ 1 & 0 & 0 \end{array}\right),\,\bar{g}_{2}=\left(\begin{array}{ccc} 1 & 0 & 0\\ 0 & 2 & 0\\ 0 & 0 & 1 \end{array}\right)$$ in base $v_{1},v_{2},v_{3}$. The group $G_{\text{Disc}}$ generated by $\bar{g}_{1},\,\bar{g}_{2}$ is isomorphic to the symmetric group $S_{4}$. The kernel of the map $G_{432}\to G_{\text{Disc}}$ is a group $G_{18}$ isomorphic to $\ZZ_{3}\rtimes S_{3}$. Let $\text{Tran}(X)$ be the orthogonal complement of $\NS X)$ in $H^{2}(X,\ZZ)$; it is a rank $3$ signature $(2,1)$ lattice. There exists an isomorphism $$\g:A_{\text{Tran}(X)}\to A_{\NS X)}$$ between the discriminant groups, such that the quadratic forms of the groups satisfy $q_{\text{Tran}(X)}\circ\g=-q_{\NS X)}$. Since an element $\phi\in G_{18}$ acts trivially on $A_{\NS X)}$, we can extend it to an isometry $\G$ of $H^{2}(X_{\l},\ZZ)$ by gluing it with the identity map on $\text{Tran}(X)$ (see e.g. [@ShiodaRemarks Theorem 12] and references therein for an example of such construction). Since $\G$ is the identity on the space $\text{Tran}(X)\otimes\CC$ containing the period, this is a Hodge isometry. Since $\G$ naturally preserves the polarization $D_{14}=4D_{2}-\sum_{j}(A_{j}+A_{j}')$ (studied in Proposition \[prop:degree8Model\]), it is effective. Therefore we can apply the Torelli Theorem for K3 surfaces (see [@BHPVdV Chap. VIII, Theorem 11.1]) and conclude that there exists a unique $g\in\aut(X)$ such that $g^{*}=\G$ on $\NS X)$. By [@Huybrecht Corollary 3.3.5], since the Picard number of $X_{\l}$ is odd, the only Hodge isometry on $\text{Tran}(X)$ is $\pm I_{d}$. Suppose that an element $g$ of $G_{432}$ not contained in $G_{18}$ comes from an automorphism of $X_{\l}$. From the definition of $g$, its action $\bar{g}$ on $A_{\NS X)}$ is non trivial, thus its action on $\text{Tran}(X)$ is also non-trivial. The automorphism $g$ is therefore non symplectic and it acts by $-Id$ on $\text{Tran}(X)$. Thus $\bar{g}\in G_{\text{Disc}}$ acts on the discriminant group $A_{\text{Tran}(X)}$ by $-Id$. However $-Id$ is not contained in $G_{\text{Disc}}\simeq S_{4}$: a contradiction and $g$ cannot come from an automorphism of $X_{\l}$. Since $\sigma$ commutes with the elements of $G_{432}$, the automorphism group of the configuration is the direct product of $G_{18}$ and $\left\langle \sigma\right\rangle \simeq\ZZ_{2}$. The group isomorphic to $G_{\mathcal{C}}\simeq\ZZ_{2}\times(\ZZ_{3}\rtimes S_{3})$ contains the $9$ translations by $3$-torsion automorphisms, and the involution $\sigma$. We give generators of the $\ZZ_{3}\rtimes S_{3}$ part of the group $G_{\mathcal{C}}$ in sub-section \[subsec:A-CompletInterModelInP5\]. \[subsec:sendingAConfToAnother\]An automorphism sending a $9{\bf A}_{2}$-configuration to another ------------------------------------------------------------------------------------------------- Let $$\mathcal{C}'=B_{1},B_{1}',\dots,B_{9},B_{9}'$$ be one of the nine $9{\bf A}_{2}$-configurations found in section \[subsec:-9new-9A2-configurations\] (so that $B_{j}B_{j}'=1$, $B_{j}B_{k}=0$ for $j\neq k$), namely we choose the $9{\bf A}_{2}$-configuration $$\mathcal{C}'=\varTheta_{12},\varTheta_{9},\varTheta_{14},\varTheta_{5},\varTheta_{16},\varTheta_{7},\varTheta_{2},\varTheta_{3},\g_{1},\g_{1}',\varTheta_{4},\varTheta_{1},\varTheta_{8},\varTheta_{15},\varTheta_{6},\varTheta_{13},\varTheta_{10},\varTheta_{11},$$ where the classes of the $\cu$-curves $\varTheta_{j}$ (which are the curves $\t_{i,j,k,m,n,o}$ more conveniently labelled and numbered) and $\g_{1},\g_{1}'$ are in the Appendix. The eight ${\bf A}_{2}$-configurations $B_{j},B_{j}',\,j\in\{1,\dots,4,6,\dots,9\}$, are contained in the fibers of $\varphi$ and the ${\bf A}_{2}$-configuration $B_{5},B_{5}'$ is the strict transform under the double cover map of a the quartic curve $Q_{1}$. \[MAINthm\]There exists an automorphism $f$ of $X_{\l}$ sending the curve $A_{j}$ (resp. $A_{j}'$) to the curve $B_{j}$ (resp. $B_{j}'$) for $j\in\{1,\dots,9\}$. In particular, the configuration $\mathcal{C}=A_{1},A_{1}',\dots,A_{9},A_{9}'$ is sent by $f$ to the configuration $\mathcal{C}'$, and therefore the configuration $\mathcal{C}'$ is on the $\aut(X)$-orbit of $\mathcal{C}$. For finding the automorphism $f$ we proceeded as in Proposition \[pro:The-group-Aut-preserveConf\]:\ Let $\phi\in\aut(X_{\l})$ be an automorphism sending $\mathcal{C}$ to $\mathcal{C}'$. The orthogonal complement of $\mathcal{C}$ is generated by $D_{2}$, the pull-back of a line in $\PP^{2}$ and the orthogonal complements of $\mathcal{C}'$ is generated by the divisor $$D_{2}'=7D_{2}-2{\textstyle \sum_{j=1}^{j=9}}(A_{j}+A_{j}')-2(A_{1}+A_{1}'),$$ of square $2$. Since $\phi$ must preserve the Néron-Severi lattice, we have $\phi(D_{2})=D_{2}'$. There are $2^{9}9!$ bijective maps $$\mu:\{A_{1},A_{1}',\dots,A_{9},A_{9}'\}\to\{B_{1},B_{1}',\dots,B_{9},B_{9}'\}$$ which preserves the incidence relations of $\mathcal{C}$ and $\mathcal{C}'$ and which extend uniquely to isometries $\tilde{\phi}_{\mu}$ of $\NS X_{\l})\otimes\QQ$. If two of these maps $\tilde{\phi}_{1},\tilde{\phi}_{2}$ preserve the lattice $\NS X_{\l})$, then $\tilde{\phi}_{2}^{-1}\tilde{\phi}_{1}$ also preserves that lattice, moreover it also preserves the configuration $\mathcal{C}$. It is therefore an element of the group $G_{864}$ defined in the proof of Proposition \[pro:The-group-Aut-preserveConf\]. Thus the set of maps $\tilde{\phi}$ (among the maps $\tilde{\phi}_{\mu}$) that preserve $\NS X_{\l})$ is the orbit of $\tilde{\phi}_{1}$ under the action of $G_{864}$ on the left. The element $\tilde{\phi}_{1}$ defined by sending $A_{j}$ to $B_{j}$ and $A_{j}'$ to $B_{j}'$ preserves the lattice $\NS X_{\l})$ (its matrix $f_{\mathcal{B}_{1}}$ in base $\mathcal{B}_{1}$ is given in the Appendix). Its action on the discriminant group is trivial. Then, as in the proof of Proposition \[pro:The-group-Aut-preserveConf\], we can extend $\tilde{\phi}_{1}$ to an Hodge isometry of $H^{2}(X,\ZZ)$. Let us prove that this isometry is effective. The polarization (see Proposition \[prop:degree8Model\]) $D_{14}=4D_{2}-\sum(A_{j}+A_{j}')$ is sent to $D_{14}'=4D_{2}'-\sum(B_{j}+B_{j}')=D_{2}'+F_{1}$ where $F_{1}$ is (the class of) a fiber of a fibration $f_{1}$ by Remark \[rem:ExtraFib\]. Using the classes of the curves in $\NS X)$, it is easy to check that the curves $B_{j},B_{j}'$ are $18$ sections of $f_{1}$. Using `LinSys`, one finds a pencil of degree $7$ generically irreducible curves in $\PP^{2}$ with singularities of multiplicity $4$ at the point $p_{1}$ in $\mathcal{P}_{9}$, and with multiplicity $2$ at the eight remaining points. The geometric genus of such curve $Q$ is $1$ and $Q$ meet the branch locus at two other points, so that its strict transform on $X_{\l}$ is smooth of genus $2$ in the linear system $|D_{2}'|$. Thus $D_{2}'$ is effective, nef and $|D_{2}'|$ is base point free. Let $C$ be an irreducible $\cu$-curve on $X_{\l}$. If $CF'=CD_{2}'=0$, then $C$ is contained in a fiber of $f_{1}$ and is contracted by $|D_{2}'|$. The second point implies that $C$ is one of the curves $B_{j},B_{j}'$ (otherwise the Picard number of $X_{\l}$ would be $>19$), but then one has $CF'=1$, which is a contradiction. Thus $D_{14}'$ is ample and the isometry is effective. We then conclude as in the proof of Proposition \[pro:The-group-Aut-preserveConf\] that there exists an automorphism of $X_{\l}$ such that its action on the curves $A_{j},A_{j}'$ is as described. We recall that for $k=1,...,9$, the point $p_{k}\in\mathcal{P}_{9}$ is the point over which $A_{k}+A_{k}'$ is contracted by the double cover map $\eta:X_{\l}\to\PP^{2}$. \[rem:ExtraFib\]For any point $p_{k}$, the pencil of lines through $p_{k}$ induces an elliptic fibration $f_{k}:X_{\l}\to\PP^{1}$ such that the curves $A_{k},A_{k}'$ are sections and the curves $A_{j},A_{j}'$ for $j\neq k$ are contained in the fibers. The singular fibers of $f_{k}$ are also $8\tilde{{\bf A}_{2}}$; from the known intersection numbers of these fibers with the elements of base $D_{2},A_{1},\dots,A_{9}'$, we obtain that the class of a fiber is $$F_{k}=D_{2}-(A_{k}+A_{k}')=\tfrac{1}{3}(F+{\textstyle \sum_{j=1}^{9}}(A_{j}+A_{j}'))-(A_{k}+A_{k}')=3D_{2}'-{\textstyle \sum_{j=1}^{9}}(B_{j}+B_{j}')$$ where $F$ a fiber of $\varphi$ and the $B_{j},B_{j}'$ are the curve in the $9{\bf A}_{9}$-configuration found in Section \[subsec:-9new-9A2-configurations\], $D_{2}'$ being the generator of the orthogonal complement of the $B_{j},B_{j}'$’s. The curves $B_{j},B_{j}'$ are sections of $f_{k}$. So the geometric situation for the configuration $A_{1},A_{1}',\dots,A_{9},A_{9}'$ and fibration $f_{p}$ is very similar to the situation for $\mathcal{C}'$ and fibration $\varphi$ for which the $8$ of the ${\bf A}_{2}$-configurations in $B_{1},B_{1}',\dots,B_{9},B_{9}'$ are in the $8$ singular fibers and the remaining one are sections. ### Aligned singularities of the union of the $12$ conics The fibration $\varphi$ has the remarkable property that the $18$ sections $A_{j},A_{j}'$ meet on the same fiber, which is isomorphic to $E_{\l}$. It can be instructive to understand how the similar result holds for the fibration $f_{1}$ (see Remark \[rem:ExtraFib\]) and the curves $B_{j},B_{j}'$, this is the aim of this subsection, which also gives an explanation why each line of the dual Hesse configuration contains $4$ double points of the curve $\sum_{C\in\mathcal{C}_{12}}C$ and these double points form the set $\mathcal{P}_{12}$. We recall that the $12$ conics in $\mathcal{C}_{12}$ meet in either $4$ or $3$ point in $\mathcal{P}_{9}$. If two conics meet in $3$ points in $\mathcal{P}_{9}$ then the fourth intersection point is an ordinary singularity of the union of the $12$ conics. Above the $8$ conics that contain $p_{1}$ are the $16$ $\cu$-curves that gives a $8{\bf A}_{8}$ configuration, which one can complete to a $9{\bf A}_{2}$-configuration according to section \[subsec:-9new-9A2-configurations\]. The $8$ conics containing $p_{1}$ have the property that they meet by pairs into $4$ points $q_{1},\dots,q_{4}$ not in $\mathcal{P}_{9}$ and that these $4$ points are on a line containing $p_{1}$. That line $L_{1}$ is the tangent to the cusp $p_{1}$ (it is one of the lines of the dual Hesse configuration defined in Section \[subsec:The-Hesse,-dual\]). It meets the cuspidal sextic in $p_{1}$ (with multiplicity $3$) and at points $r_{1},r_{2},r_{3}$. One can check that the fiber $F_{0}$ of $f_{1}$ which is the strict transform on $X_{\l}$ of $L_{1}$ is isomorphic to $E_{\l}$ (since we know the branch locus $F_{0}\to L_{1}$). The $8$ points in $X_{\l}$ above $q_{1},\dots,q_{4}$ are the meeting points of the curves in the $8{\bf A}_{2}$ configuration obtained by taking the strict transform of the $8$ conics trough $p_{1}$. Following the Figure \[figniceDiag\], the $9^{th}$ ${\bf A}_{2}$-configuration also has its intersection point on that fiber $F_{0}$. In fact, the fiber $F_{0}$ is the image by an automorphism of $X_{\l}$ of the fiber over $0$ of $\varphi$. The union of the dual Hesse configuration and the $12$ conics in $\mathcal{C}_{12}$ is a line-conic arrangement with $9$ points of multiplicity $9$, $12$ points of multiplicity $5$ and $72$ double points, and with other remarkable properties studied in [@PS]. \[sec:Appendix\] Appendix ========================= Let $D_{2}$ be the pull-back of a line by the double cover map $\eta:X_{\l}\to\PP^{1}$. Let us define the following classes in the $\QQ$-base $\mathcal{B}_{0}=(D_{2},A_{1},A_{1}',\dots,A_{9},A_{9}')$: $$B_{1}=2D_{2}-\tfrac{1}{3}\left({\textstyle \sum_{j=1}^{9}}2A_{j}+A_{j}'\right),\,\,B_{2}=2D_{2}-\tfrac{1}{3}\left({\textstyle \sum_{j=1}^{9}}A_{j}+2A_{j}'\right).$$ We remark that $B_{1}^{2}=B_{2}^{2}=2$, $B_{1}B_{2}=5$ and $B_{1}+B_{2}=D_{14}$. We have $D_{14}A_{j}=D_{14}A_{j}'=1$, therefore $B_{i}A_{j}\in\{0,1\}$, $B_{i}A_{j}'\in\{0,1\}$. Using algorithms described in [@Roulleau], we find that for $j\in\{1,\dots,9\}$, the classes of the curves $\gamma_{j},\gamma_{j}'$ above the quartic $Q_{j}$ are $$\gamma_{j}=B_{1}-(A_{j}+A_{j}'),\,\gamma_{j}'=B_{2}-(A_{j}+A_{j}').$$ It is easy to check that $\gamma_{j}^{2}=\gamma_{j}'^{2}=-2$, $\gamma_{j}\gamma_{j}'=1$, and for $1\leq i\neq j\leq9$, we have $\gamma_{i}\gamma_{j}=\gamma_{i}'\gamma_{j}'=0$ and $\gamma_{i}\gamma_{j}'=3$. In fact, using that the image in $\PP^{2}$ of $\gamma_{j},\gamma_{j}'$ is a quartic curve that goes through the points in $\mathcal{P}_{9}$ with a multiplicity $3$ at $p_{j}$, one gets $$4D_{2}\equiv\gamma_{j}+\gamma_{j}'+2(A_{j}+A_{j}')+{\textstyle \sum_{j=1}^{9}}(A_{j}+A_{j}').$$ The classes in the $\QQ$-base $\mathcal{B}_{0}=(L,A_{1},A_{1}',\dots,A_{9},A_{9}')$ of the $24$ $\cu$-curves $\t_{i,\dots,n},\t'_{i,\dots,n}$ above the $12$ conics $C_{i,\dots,n}$ in $\mathcal{C}_{12}$ are [ $$\begin{array}{c} \t_{123456}=\frac{1}{3}(3,-2,-1,-2,-1,-2,-1,-1,-2,-1,-2,-1,-2,0,0,0,0,0,0),\\ \t'_{123456}=\frac{1}{3}(3,-1,-2,-1,-2,-1,-2,-2,-1,-2,-1,-2,-1,0,0,0,0,0,0),\\ \t_{123789}=\frac{1}{3}(3,-2,-1,-2,-1,-2,-1,0,0,0,0,0,0,-1,-2,-1,-2,-1,-2),\\ \t'_{123789}=\frac{1}{3}(3,-1,-2,-1,-2,-1,-2,0,0,0,0,0,0,-2,-1,-2,-1,-2,-1),\\ \t_{124578}=\frac{1}{3}(3,-2,-1,-1,-2,0,0,-2,-1,-1,-2,0,0,-2,-1,-1,-2,0,0),\\ \t'_{124578}=\frac{1}{3}(3,-1,-2,-2,-1,0,0,-1,-2,-2,-1,0,0,-1,-2,-2,-1,0,0),\\ \t_{124689}=\frac{1}{3}(3,-2,-1,-1,-2,0,0,-1,-2,0,0,-2,-1,0,0,-2,-1,-1,-2),\\ \t'_{124689}=\frac{1}{3}(3,-1,-2,-2,-1,0,0,-2,-1,0,0,-1,-2,0,0,-1,-2,-2,-1),\\ \t_{125679}=\frac{1}{3}(3,-2,-1,-1,-2,0,0,0,0,-2,-1,-1,-2,-1,-2,0,0,-2,-1),\\ \t'_{125679}=\frac{1}{3}(3,-1,-2,-2,-1,0,0,0,0,-1,-2,-2,-1,-2,-1,0,0,-1,-2),\\ \t_{134589}=\frac{1}{3}(3,-2,-1,0,0,-1,-2,-1,-2,-2,-1,0,0,0,0,-1,-2,-2,-1),\\ \t'_{134589}=\frac{1}{3}(3,-1,-2,0,0,-2,-1,-2,-1,-1,-2,0,0,0,0,-2,-1,-1,-2),\\ \t_{134679}=\frac{1}{3}(3,-2,-1,0,0,-1,-2,-2,-1,0,0,-1,-2,-2,-1,0,0,-1,-2),\\ \t'_{134679}=\frac{1}{3}(3,-1,-2,0,0,-2,-1,-1,-2,0,0,-2,-1,-1,-2,0,0,-2,-1),\\ \t_{135678}=\frac{1}{3}(3,-2,-1,0,0,-1,-2,0,0,-1,-2,-2,-1,-1,-2,-2,-1,0,0),\\ \t'_{135678}=\frac{1}{3}(3,-1,-2,0,0,-2,-1,0,0,-2,-1,-1,-2,-2,-1,-1,-2,0,0),\\ \t{}_{234579}=\frac{1}{3}(3,0,0,-2,-1,-1,-2,-2,-1,-1,-2,0,0,-1,-2,0,0,-2,-1),\\ \t'_{234579}=\frac{1}{3}(3,0,0,-1,-2,-2,-1,-1,-2,-2,-1,0,0,-2,-1,0,0,-1,-2),\\ \t{}_{234678}=\frac{1}{3}(3,0,0,-2,-1,-1,-2,-1,-2,0,0,-2,-1,-2,-1,-1,-2,0,0),\\ \t'_{234678}=\frac{1}{3}(3,0,0,-1,-2,-2,-1,-2,-1,0,0,-1,-2,-1,-2,-2,-1,0,0),\\ \t{}_{235689}=\frac{1}{3}(3,0,0,-2,-1,-1,-2,0,0,-2,-1,-1,-2,0,0,-2,-1,-1,-2),\\ \t'_{235689}=\frac{1}{3}(3,0,0,-1,-2,-2,-1,0,0,-1,-2,-2,-1,0,0,-1,-2,-2,-1),\\ \t_{456789}=\frac{1}{3}(3,0,0,0,0,0,0,-1,-2,-1,-2,-1,-2,-2,-1,-2,-1,-2,-1),\\ \t'_{456789}=\frac{1}{3}(3,0,0,0,0,0,0,-2,-1,-2,-1,-2,-1,-1,-2,-1,-2,-1,-2). \end{array}$$ ]{} We also denote by $\varTheta_{j},\,j=1,\dots,24$ these curves in the order of the above list. For $k=1,...,9$, the quartic curves $Q_{k}$ through $\mathcal{P}_{9}$ that have a multiplicity $3$ singular point at $p_{k}$ are: [ $$\begin{array}{c} Q_{1}:\,\,x^{4}-2\l x^{3}y+3\l^{2}x^{2}y^{2}-(\l^{3}+1)xy^{3}+\l y^{4}-2\l x^{3}z+(-\l^{3}+1)xy^{2}z-2\l y^{3}z\hfill\\ +3\l^{2}x^{2}z^{2}+(-\l^{3}+1)xyz^{2}+(\l^{4}+2\l)y^{2}z^{2}-(\l^{3}+1)xz^{3}-2\l yz^{3}+\l z^{4}=0,\\ Q_{2}:\,\,x^{4}-(\l^{3}+1)/\l x^{3}y+3\l x^{2}y^{2}-2xy^{3}+1/\l y^{4}-2x^{3}z+(-\l^{3}+1)/\l x^{2}yz-2y^{3}z\\ +(\l^{3}+2)x^{2}z^{2}+(1-\l^{3})/\l xyz^{2}+3\l y^{2}z^{2}-2xz^{3}-(\l^{3}+1)/\l yz^{3}+z^{4}=0,\\ Q_{3}:\,\,x^{4}-2x^{3}y+(\l^{3}+2)x^{2}y^{2}-2xy^{3}+y^{4}-(\l^{3}+1)/\l x^{3}z+(1-\l^{3})/\l x^{2}yz\hfill\\ +(1-\l^{3})/\l xy^{2}z-(\l^{3}+1)/\l y^{3}z+3\l x^{2}z^{2}+3\l y^{2}z^{2}-2xz^{3}-2yz^{3}+1/\l z^{4}=0,\\ Q_{4}:\,\,x^{4}+(2\omega+2)\l x^{3}y+3\omega\l^{2}x^{2}y^{2}-(\l^{3}+1)xy^{3}-(\omega+1)\l y^{4}-2\omega\l x^{3}z\hfill\\ -(\omega^{2}\l^{3}+\omega+1)xy^{2}z-2\omega\l y^{3}z-(3\omega+3)\l^{2}x^{2}z^{2}+(\omega-\omega\l^{3})xyz^{2}\\ +(\l^{4}+2\l)y^{2}z^{2}-(\l^{3}+1)xz^{3}+(2\omega+2)\l yz^{3}+\omega\l z^{4}=0,\\ Q_{5}:\,\,x^{4}-(\omega^{2}\l^{3}+\omega^{2})/\l x^{3}y+3\omega\l x^{2}y^{2}-2xy^{3}-(\omega+1)/\l y^{4}-2\omega x^{3}z\hfill\\ +(1-\l^{3})/\l x^{2}yz-2\omega y^{3}z-((\omega+1)\l^{3}+2\omega+2)x^{2}z^{2}+(\omega-\omega\l^{3})/\l xyz^{2}\\ +3\l y^{2}z^{2}-2xz^{3}-\omega^{2}(\l^{3}+1)/\l yz^{3}+\omega z^{4}=0,\\ Q_{6}:\,\,x^{4}+(2\omega+2)x^{3}y+(\omega\l^{3}+2\omega)x^{2}y^{2}-2xy^{3}+\omega^{2}y^{4}-(\omega\l^{3}+\omega)/\l x^{3}z\hfill\\ +(1-\l^{3})/\l x^{2}yz-(\omega^{2}\l^{3}-\omega^{2})/\l xy^{2}z-(\omega\l^{3}+\omega)/\l y^{3}z\\ -(3\omega+3)\l x^{2}z^{2}+3\l y^{2}z^{2}-2xz^{3}+(2\omega+2)yz^{3}+\omega/\l z^{4}=0,\\ Q_{7}:\,\,x^{4}-2\omega\l x^{3}y-(3\omega+3)\l^{2}x^{2}y^{2}-(\l^{3}+1)xy^{3}+\omega\l y^{4}+(2\omega+2)\l x^{3}z\hfill\\ +(\omega-\omega\l^{3})xy^{2}z+(2\omega+2)\l y^{3}z+3\omega\l^{2}x^{2}z^{2}-(\omega^{2}\l^{3}-\omega^{2})xyz^{2}\\ +(\l^{4}+2\l)y^{2}z^{2}-(\l^{3}+1)xz^{3}-2\omega\l yz^{3}+\omega^{2}\l z^{4}=0,\\ Q_{8}:\,\,x^{4}-(\omega\l^{3}+\omega)/\l x^{3}y-(3\omega+3)\l x^{2}y^{2}-2xy^{3}+\omega/\l y^{4}+(2\omega+2)x^{3}z\hfill\\ +(1-\l^{3})/\l x^{2}yz+(2\omega+2)y^{3}z+(\omega\l^{3}+2\omega)x^{2}z^{2}-(\omega^{2}\l^{3}-\omega^{2})/\l xyz^{2}\\ +3\l y^{2}z^{2}-2xz^{3}-(\omega\l^{3}+\omega)/\l yz^{3}+\omega^{2}z^{4}=0,\\ Q_{9}:\,\,x^{4}-2\omega x^{3}y+(\omega^{2}\l^{3}-2\omega-2)x^{2}y^{2}-2xy^{3}+\omega y^{4}-(\omega^{2}\l^{3}+\omega^{2})/\l x^{3}z\hfill\\ +(1-\l^{3})/\l x^{2}yz+(-\omega\l^{3}+\omega)/\l xy^{2}z-(\omega^{2}\l^{3}+\omega^{2})/\l y^{3}z\\ +3\omega\l x^{2}z^{2}+3\l y^{2}z^{2}-2xz^{3}-2\omega yz^{3}+\omega^{2}/\l z^{4}=0, \end{array}$$ ]{}where $\o^{2}+\o+1=0$. The matrices in base $\mathcal{B}_{1}$ of the generators of the group $G_{432}\simeq AGL_{2}(\FF_{3})$ preserving the natural $9{\bf A}_{2}$-configuration $A_{1},A_{1}',\dots,A_{9},A_{9}'$ are $$\mbox{\footnotesize\ensuremath{g_{1}=\left(\begin{array}{ccccccccccccccccccc} 1 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & -2 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 1 & 0\\ 0 & 0 & 0 & 0 & 0 & -1 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & -1 & 0\\ 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 & 0\\ 0 & 1 & 0 & 0 & 0 & 1 & -1 & 1 & -1 & 0 & 0 & 0 & 0 & 0 & -1 & 0 & 0 & 1 & 0\\ 0 & 0 & 1 & 0 & 0 & 2 & -2 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & -1 & -1 & 0 & 1 & 0\\ 0 & 0 & 0 & 0 & 0 & 4 & -2 & 1 & 0 & 1 & 0 & 0 & 0 & 0 & -1 & -1 & -1 & 2 & 0\\ 0 & 0 & 0 & 0 & 0 & 2 & -1 & 1 & -1 & 0 & 1 & 0 & 0 & 0 & -1 & 0 & -1 & 1 & 0\\ 0 & 0 & 0 & 0 & 0 & -1 & 1 & 0 & 1 & 0 & 0 & 0 & 1 & 0 & 0 & 0 & 1 & -1 & 0\\ 0 & 0 & 0 & 0 & 0 & -2 & 2 & -1 & 1 & 0 & 0 & 0 & 0 & 1 & 1 & 0 & 1 & -1 & 0\\ 0 & 0 & 0 & 1 & 0 & -1 & 0 & 0 & -1 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 & 1 & -2 & 0 & -1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & -1 & 0\\ 0 & 0 & 0 & 0 & 0 & 2 & -1 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 & -1 & -1 & 1 & 0\\ 0 & 0 & 0 & 0 & 0 & -3 & 2 & -1 & 1 & 0 & 0 & 0 & 0 & 0 & 1 & 1 & 1 & -2 & 0\\ 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & -2 & 1 & -1 & 2 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 1 & -2 & 0\\ 0 & 0 & 0 & 0 & 0 & 1 & 1 & -1 & 2 & 0 & 0 & 0 & 0 & 0 & 1 & -1 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & 3 & -3 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & -1 & -1 & -1 & 1 & 0\\ 0 & 0 & 0 & 0 & 0 & -4 & 2 & -2 & 1 & 0 & 0 & 0 & 0 & 0 & 1 & 1 & 1 & -2 & 1\\ 0 & 0 & 0 & 0 & 0 & 2 & -1 & 1 & 1 & 0 & 0 & 0 & 0 & 0 & -1 & -1 & 0 & 1 & 0 \end{array}\right),} }$$ $$\mbox{\footnotesize\ensuremath{g_{2}=\left(\begin{array}{ccccccccccccccccccc} 1 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & -2 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 1 & 0\\ 0 & 0 & 0 & 0 & 0 & -1 & 1 & 0 & 0 & 1 & 0 & 0 & 0 & 2 & 0 & 0 & 1 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 & 1 & 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & 1 & -1 & 1 & -1 & 0 & 0 & 0 & 0 & -1 & 0 & 0 & -1 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & 2 & -2 & 1 & 0 & 0 & 0 & 0 & 0 & -2 & 0 & 0 & -1 & 0 & 0\\ 0 & 1 & 0 & 0 & 0 & 4 & -2 & 1 & 0 & 0 & 0 & 0 & 0 & -4 & 0 & 0 & -2 & 1 & 0\\ 0 & 0 & 1 & 0 & 0 & 2 & -1 & 1 & -1 & 0 & 0 & 0 & 0 & -2 & 0 & 0 & -1 & 1 & 0\\ 0 & 0 & 0 & 1 & 0 & -1 & 1 & 0 & 1 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 1 & 0 & 0\\ 0 & 0 & 0 & 0 & 1 & -2 & 2 & -1 & 1 & 0 & 0 & 0 & 0 & 2 & 0 & 0 & 1 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & -1 & 0 & 0 & -1 & 0 & 0 & 1 & 0 & 1 & 0 & 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & -2 & 0 & -1 & 0 & 0 & 0 & 0 & 0 & 2 & 0 & 0 & 1 & -1 & 0\\ 0 & 0 & 0 & 0 & 0 & 2 & -1 & 0 & 0 & 0 & 0 & 0 & 1 & -2 & 0 & 0 & -1 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & -3 & 2 & -1 & 1 & 0 & 0 & 0 & 0 & 3 & 0 & 0 & 2 & -1 & 0\\ 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & -2 & 1 & -1 & 2 & 0 & 0 & 0 & 0 & 3 & 0 & -1 & 2 & -1 & 0\\ 0 & 0 & 0 & 0 & 0 & 1 & 1 & -1 & 2 & 0 & 0 & 0 & 0 & 0 & 1 & -1 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & 3 & -3 & 0 & 0 & 0 & 0 & 0 & 0 & -3 & 0 & 0 & -2 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & -4 & 2 & -2 & 1 & 0 & 0 & 0 & 0 & 3 & 0 & 0 & 2 & -1 & 1\\ 0 & 0 & 0 & 0 & 0 & 2 & -1 & 1 & 1 & 0 & 0 & 0 & 0 & -3 & 0 & 0 & -1 & 0 & 0 \end{array}\right),} }$$ The automorphism $f$ of Theorem \[MAINthm\] acts on the Néron-Severi group by $$\mbox{\footnotesize\ensuremath{f_{\mathcal{B}_{1}}=\left(\begin{array}{ccccccccccccccccccc} 0 & 1 & 0 & 0 & 0 & 0 & 1 & 0 & 1 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1\\ 0 & 1 & -1 & 0 & 0 & -1 & 0 & 0 & 0 & -1 & 0 & 1 & -1 & 0 & 0 & 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & -1 & 0 & 0 & 0 & -1 & 0 & 0 & -1 & 0 & 0 & 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 & 1 & -1 & -1 & 0 & 0 & 0 & 0 & 0 & 1 & 0\\ 0 & -1 & 0 & 0 & 0 & 2 & -1 & 0 & 0 & 1 & -1 & -2 & 1 & 0 & 0 & 0 & 0 & 0 & -1\\ 0 & -3 & 2 & 0 & 0 & 2 & -1 & 1 & -1 & 2 & -2 & -3 & 2 & -1 & 0 & 0 & 0 & 0 & -1\\ 0 & -1 & 1 & 0 & 0 & 1 & 0 & 0 & 0 & 1 & -1 & -2 & 1 & -1 & 0 & 0 & 0 & 0 & 0\\ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 1 & -1 & 1 & 1 & 0 & 0 & 0 & 0\\ 1 & 1 & 0 & 0 & 0 & -1 & 0 & 0 & 0 & -1 & 2 & 2 & -1 & 1 & 0 & 0 & 0 & 0 & 1\\ 0 & 1 & -1 & 0 & 0 & -1 & 1 & -1 & 1 & 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 1\\ 0 & 1 & -1 & 0 & 0 & -1 & 1 & -1 & 1 & -1 & 1 & 2 & 0 & 0 & 0 & 0 & 0 & 0 & 1\\ 0 & -1 & 1 & 0 & 0 & 1 & -1 & 0 & 0 & 1 & -1 & -1 & 1 & -1 & 0 & 0 & 0 & 0 & 0\\ 1 & 2 & -1 & 0 & 0 & -2 & 1 & 0 & 0 & -1 & 2 & 2 & -1 & 1 & 0 & 1 & 1 & 0 & 1\\ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & -1 & 0 & 1 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0\\ 1 & 1 & -1 & 0 & 1 & -1 & 0 & 0 & 0 & -1 & 2 & 2 & -1 & 1 & 0 & 1 & 0 & 0 & 0\\ 1 & -1 & 1 & 1 & 0 & 0 & -1 & 1 & -1 & 0 & 1 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0\\ -1 & -2 & 1 & 0 & 0 & 2 & -1 & 0 & 0 & 1 & -2 & -2 & 2 & -1 & 0 & -1 & 0 & 0 & -1\\ 0 & 2 & -1 & 0 & 0 & -2 & 1 & -1 & 0 & -2 & 2 & 3 & -1 & 1 & 0 & 0 & 0 & 0 & 1\\ 0 & -2 & 1 & 0 & 0 & 2 & -1 & 1 & -1 & 1 & -1 & -2 & 1 & 0 & 0 & 0 & 0 & 0 & -2 \end{array}\right).} }$$ Let us define $$S={\textstyle \sum_{j=1}^{9}}A_{i},\,\,S'={\textstyle \sum_{j=1}^{9}}A_{i}'$$ The classes of the $\cu$-curves $(R_{i})$ defined in Section \[subsec:A-Hessian-model\] are $$(R_{i})=2L-\tfrac{1}{3}(S+2S'+3A_{i}+3A'_{i}),$$ where $L$ is the pullback of a line by the double cover map $X_{\l}\to\PP^{2}$. The translation automorphism $\tau$ defined in Section \[subsec:A-Hessian-model\] sends $L$ to the class $$L'=7L-\tfrac{4}{3}(S+2S').$$ The three polynomials $A,B,D$ in $\QQ(\o)(t)$ of Section \[subsec:A-Weierstrass-equation\] are defined as follows: $$\begin{array}{c} A=(\l^{3}t^{2}+3\l^{3}-4t^{2})(\l^{3}t^{2}+3\l^{3}+(6\o+6)\l^{2}t^{2}+(-6\o-6)\l^{2}-4t^{2})\\ \cdot(\l^{3}t^{2}+3\l^{3}-6\l^{2}t^{2}+6\l^{2}-4t^{2})(\l^{3}t^{2}+3\l^{3}-6\o\l^{2}t^{2}+6\o\l^{2}-4t^{2}), \end{array}$$ $$\begin{array}{c} B=(\l^{6}t^{4}+6\l^{6}t^{2}+9\l^{6}+6\l^{5}t^{4}+12\l^{5}t^{2}-18\l^{5}-18\l^{4}t^{4}+36\l^{4}t^{2}-18\l^{4}\hfill\\ -8\l^{3}t^{4}-24\l^{3}t^{2}-24\l^{2}t^{4}+24\l^{2}t^{2}+16t^{4})\hfill\\ \cdot(\l^{6}t^{4}+6\l^{6}t^{2}+9\l^{6}+(-6\o-6)\l^{5}t^{4}+(-12\o-12)\l^{5}t^{2}+(18\o+18)\l^{5}-18\o\l^{4}t^{4}\\ +36\o\l^{4}t^{2}-18\o\l^{4}-8\l^{3}t^{4}-24\l^{3}t^{2}+(24\o+24)\l^{2}t^{4}+(-24\o-24)\l^{2}t^{2}+16t^{4})\\ \cdot(\l^{6}t^{4}+6\l^{6}t^{2}+9\l^{6}+6\o\l^{5}t^{4}+12\o\l^{5}t^{2}-18\o\l^{5}+(18\o+18)\l^{4}t^{4}\hfill\\ +(-36\o-36)\l^{4}t^{2}+(18\o+18)\l^{4}-8\l^{3}t^{4}-24\l^{3}t^{2}-24\o\l^{2}t^{4}+24\o\l^{2}t^{2}+16t^{4}), \end{array}$$ $$\begin{array}{c} D=((\l+2)t-(2\o+1)\l)((\l-2\o-2)t-(2\o+1)\l)((\l+2\o)t-(2\o+1)\l)\\ \cdot(t^{2}-1)((\l+2)t+(2\o+1)\l)((\l-2\o-2)t+(2\o+1)\l)((\l+2\o)t+(2\o+1)\l). \end{array}$$ [10]{} Artebani M., Dolgachev I., The Hesse pencil of plane cubic curves, Ens. Math. (2) 55 (2009), no. 3-4, 235–273. Beauville A., Complex algebraic surfaces, 2nd ed. Lond. Math. Soc., 34. Cambridge University Press, 1996. x+132 pp. J. Bertin, Réseaux de Kummer et surfaces K3, Invent. Math. 93 (1988), no. 2, 267–284. Barth W., K3 surfaces with nine cusps. Geom. Dedicata 72 (1998), no. 2, 171–178. Barth W., On the classification of K3 surfaces with nine cusps, Complex analysis and algebraic geometry, de Gruyter, Berlin, 2000, pp. 41–59. MR 1760871. Barth W., Hulek K., Peters C.A.M., Van de Ven A., Compact complex surfaces. Sd ed. Erg Math Grenz. 3. Folge., 4. Springer-Verlag, Berlin, 2004. xii+436 pp. Birkenhake C., Lange H., A family of Abelian surfaces and curves of genus four, Manus. Math. 85, (1994), 393–407. Bosma W,. Cannon J., Playoust C., The Magma algebra system. I. The user language, Computational algebra and number theory (London, 1993), J. Symbolic Comput. 24, 1997, 3-4, 235--265. Conway J. H., Sloane N. J. A., Sphere packings, lattices and groups, Sde. ed., Grund. der Math. Wiss 290. Springer-Verlag, New York, 1993. xliv+679 pp. Dolgachev I., Abstract configurations in algebraic geometry, The Fano Conference, 423–462, Univ. Torino, Turin, 2004. Dolgachev I., Laface A., Persson U., Urzúa G, Chilean configuration of conics, lines and points, preprint. Griffiths P., Harris J., Principles of Algebraic Geometry, Wiley-Interscience, New York, 1978. xii+813 pp. Gritsenko V., Hulek K., Minimal Siegel modular threefolds, Math. Proc. Cambridge Philos. Soc. 123 (1998), 461–485. Huybrecht D., Lectures on K3 surfaces, Camb. Stud. Adv. Math., 158, Cambridge University Press, 2016. xi+485 pp. Miranda R., The Basic Theory of Elliptic Surfaces, Dottorato di Ricerca in Matematica, ETS Editrice, Pisa, 1989. vi+108 pp. Morrison D. R., On K3 surfaces with large Picard number. Invent. Math. 75 (1984), no. 1, 105–121. Nikulin V.V., On Kummer surfaces, Izv. Akad. Nauk SSSR Ser. Mat. 39 (1975), 278–293. English translation: Math. USSR. Izv, 9 (1975), 261–275. Nikulin V.V., Integer symmetric bilinear forms and some of their geometric applications, Izv. Akad. Nauk SSSR Ser. Mat., 43, 1979, no 1, 111–177. Pokora P., Szemberg T., Conic-line arrangements in the complex projective plane, preprint, ArXiv 2002.01760 Roulleau X., On the geometry of K3 surfaces with finite automorphism group: the compact case, preprint, ArXiv 1909.01909 Roulleau X., Sarti A., Construction of Nikulin configurations on some Kummer surfaces and applications, Math. Annalen 373 (2019), no 11, 7651–7668. Roulleau X., Sarti A., Explicit Nikulin configurations on Kummer surfaces, preprint, ArXiv 1907.12215 Sarti A., Transcendental lattices of some K3-surfaces, Math. Nachr. 281 (2008), no. 7, 1031–1046. Saint Donat B., Projective models of K-3 surfaces. Amer. J. Math. 96 (1974), 602–639. Schütt M., Shioda T., Mordell-Weil lattices, Erg. der Math. 3. Folge. A Series of Modern Surveys in Mathematics 70. Springer, 2019. xvi+431 pp. Silverman J., Advanced topics in the arithmetic of elliptic curves, Graduate Texts in Mathematics, 151. Springer-Verlag, New York, 1994. xiv+525 pp. Shimada I., On elliptic K3 surfaces, Michigan Math. J. 47 (2000), no. 3, 423–446. Shimada I., On elliptic K3 surfaces, arXiv:math/0505140 Shioda T., On elliptic modular surfaces. J. Math. Soc. Japan 24 (1972), 20–59. Shioda T., Some remarks on abelian varieties, J. Fac. Sci. Univ. Tokyo Sect. IA 24 (1977),11–21. David Kohel, Xavier Roulleau,\ Aix-Marseille Université, CNRS, Centrale Marseille,\ I2M UMR 7373,\ 13453 Marseille, France\ `[email protected]`\ `[email protected]`\ Alessandra Sarti,\ Laboratoire de Mathématiques et Applications, UMR CNRS 7348,\ Université de Poitiers, Téléport 2,\ Boulevard Marie et Pierre Curie,\ 86962 Futuroscope Chasseneuil, France\ `[email protected]`\ http://www-math.sp2mi.univ-poitiers.fr/ sarti/
{ "pile_set_name": "ArXiv" }
--- abstract: | We establish strict upper limits for the Casimir interaction between multilayered structures of arbitrary dielectric or diamagnetic materials. We discuss the appearance of different power laws due to frequency-dependent material constants. Simple analytical expressions are in good agreement with numerical calculations based on Lifshitz theory. We discuss the improvements required for current (meta) materials to achieve a repulsive Casimir force.\ Dated: 20 Oct 2005, *Europhysics Letters*, in press.\ PACS. 42.50.Pq – Cavity quantum electrodynamics; 42.50.Lc – Quantum fluctuations, quantum noise; [78.67.-n]{} – [Optical properties of low-dimensional, mesoscopic, and nanoscale materials and structures]{} author: - | C. Henkel$^1$ and K. Joulain$^2$\ $^1$ Institut für Physik, Universität Potsdam, Germany\ $^2$ Laboratoire d’Etudes Thermiques, Ecole Nationale Supérieure\ de Mécanique Aéronautique, Poitiers, France date: 20 October 2005 title: | Casimir force between designed materials:\ what is possible and what not --- Introduction ============ The optical properties of materials that show both a dielectric and magnetic response, have recently attracted much attention (see [@Ramakrishna05] for a review). A number of striking phenomena like perfect lensing and a reversed Doppler effect have been predicted, and experimenters have begun to explore the large parameter space of structural units that can be assembled into artificial materials. Breakthroughs have been reported on the way towards designed susceptibitilies in the near-infrared and visible spectral range [@Pendry04b; @Wegener04]. Quantum electrodynamics in meta materials has recently been explored with particular emphasis on left-handed or negative-index materials [@Klimov02d; @Fleischhauer05a]. We discuss here to what extent the Casimir interaction between two meta material plates can be manipulated by engineering their magneto-dielectric response. Strict limits for the Casimir interaction are proven that apply to all causal and linear materials, including both bulk and multilayer structures. We illustrate these results by computations of the Casimir pressure, considering materials with frequency-dispersive response functions like those encountered in effective medium theories. We derive power law exponents and prefactors and find that a strongly modified Casimir interaction is possible in a range of distances around the resonance wavelength of the response functions. We give estimates for the required temperature range and structure size: it is not unreasonable to expect that improvements in fabrication and detection will allow for experimental observations. One of the most striking changes to the Casimir interaction is a cross over to repulsion. This has been predicted previously for idealized magnetodielectric materials [@Boyer74; @Kenneth02; @Boyer03] and objects suspended in a liquid [@Hartmann91; @Israelachvili]. In the latter case, repulsion has been observed experimentally with colloidal particles [@Sigmund01] and is also used in a recent proposal for measuring Casimir torques [@Capasso05]. Casimir repulsion between mirrors separated by vacuum requires a strong magnetic response [@Kenneth02; @Tomas05] that hardly occurs in conventional ferromagnets [@Camley98; @Kenneth02c-2]. Indeed, to manipulate the Casimir force in the micrometer range and below, where it can be conveniently measured, the key challenge is to achieve a magnetic susceptibility at high frequencies, approaching the visible range. Now, there is a well-known argument due to Landau, Lifshitz, and Pitaevskii that $\mu( \omega ) = 1$ in the visible [@Landau10]. This objection, however, only applies to materials whose magnetization is of atomic origin, where the magnetic susceptibility is $\chi_{\rm m} \sim (v/c)^2 \ll 1$. An array of split ring resonators with sub-wavelength size typically gives, on the contrary, $\chi_{\rm m} \sim (\omega/\vartheta)^2 f \sim 1$, where $\vartheta$ is the resonance frequency and $f$ the filling factor [@Pendry99d; @Smith04a]. As we illustrate below, artificial materials that are structured on the sub-micron scale are promising candidates for a strongly modified Casimir interaction. Lifshitz theory =============== For two perfectly conducting plates held at zero temperature and separated by a distance $d$, Casimir derived a force per unit area given by $ F_{C} = \pi^2 \hbar c /( 240 \, d^4 ) $ [@Casimir48b]. We use the convention that $F_{C} > 0$ corresponds to attraction. For linear media with complex, frequency-dependent material parameters, the force can be computed from Lifshitz theory [@Lifshitz56]. This expression has been re-derived, for plates of arbitrary material and for multilayer mirrors, using different methods . At finite temperature, it can be written in the form $$\begin{aligned} F_{L} &=& 2 k_{B} T {\kern 4.5ex\raisebox{0.25ex}{$'$}\kern -4.5ex}\sum\limits_{n=0}^{\infty} \int\limits_{\xi_{n}/c}^{\infty} \! \frac{ {\rm d}\kappa }{ 2 \pi } \, \kappa^2 \sum_{\lambda} \left( \frac{ {\rm e}^{2 \kappa d } }{ r_{\lambda 1} r_{\lambda 2} } - 1 \right)^{-1}\!\! , \label{eq:Lifshitz-formula}\end{aligned}$$ where the sum is over the imaginary Matsubara frequencies $\omega_{n} = {\rm i} \xi_{n} \equiv 2\pi {\rm i} n k_{B} T / \hbar$ (the $n=0$ term being multiplied by $1/2$), and $\kappa$ is related to the wave vector component perpendicular to the mirrors, $k_{z} = (\omega^2_{n}/c^2 - k_{x}^2 - k_{y}^2)^{1/2} \equiv {\rm i}\, \kappa$. The $r_{\lambda \alpha}$ ($\lambda = {\rm TE}, \, {\rm TM}$, $\alpha = 1, \, 2$) are the reflection coefficients at mirror $\alpha$ for electromagnetic waves with polarization $\lambda$ [@Parsegian70b; @Ninham71]. For homogeneous, thick plates, they are given by $$\begin{aligned} r_{\rm TM} &=& \frac{ \varepsilon ( {\rm i} \xi_{n} ) c \kappa - \sqrt{ \xi_{n}^2 (\varepsilon ( {\rm i} \xi_{n} ) \mu ( {\rm i} \xi_{n} ) - 1) + \kappa^2 c^2 } }{ \varepsilon ( {\rm i} \xi_{n} ) c \kappa + \sqrt{ \xi_{n}^2 (\varepsilon ( {\rm i} \xi_{n} ) \mu ( {\rm i} \xi_{n} ) - 1) + \kappa^2 c^2 } } \label{eq:Fresnel-r}\end{aligned}$$ (exchange $\varepsilon$ and $\mu$ for $r_{\rm TE}$). The zeros of $D_\lambda \equiv {\rm e}^{2 \kappa d } / ( r_{\lambda 1} r_{\lambda 2} ) - 1$ at real frequencies define the eigenmodes of the cavity formed by the two mirrors. Strict limits ============= To derive upper and lower limits for $F_{L}$, we use that the Kramers-Kronig relations [@Landau10] imply real and positive material functions at imaginary frequencies, $\varepsilon( {\rm i}\xi ) \ge 1$, provided the material is passive (non-negative absorption ${\rm Im}\,\varepsilon( \omega ) \ge 0$). As a consequence, the Fresnel formulas (\[eq:Fresnel-r\]) imply $-1 \le r_{\lambda \alpha} \le 1$, and we find $$- \frac{ 1 }{ {\rm e}^{2\kappa d} + 1 } \le \frac{ 1 }{ D_\lambda } \le \frac{ 1 }{ {\rm e}^{2\kappa d} - 1 } \label{eq:D-limits}$$ with the stronger inequalities $0 \le 1/D_{\lambda} \le 1/({\rm e}^{2\kappa d} - 1)$ holding for identical plates. In the latter case, the Casimir force is hence necessarily attractive. The inequalities (\[eq:D-limits\]) saturate for a perfectly conducting mirror facing a perfectly permeable one ($\varepsilon_{1} = \infty$, $\mu_{2} = \infty$, say), and for identical, perfectly reflecting mirrors, respectively. The resulting forces at zero temperature are [@Lifshitz56; @Boyer74] $$T = 0: \quad - \frac{7}{8} F_C \le F_L \le F_C. \label{eq:F-limits-zero-T}$$ In the high-temperature limit, we get similarly [@Tort99] $ -\frac{3}{4} F_T \le F_L \le F_T \equiv \zeta(3) k_B T / (8\pi d^3)$ by keeping in Eq.(\[eq:Lifshitz-formula\]) only the $n=0$ term in the sum. Consider now a mirror made from layers of arbitrary passive materials. Reflection coefficients for such a system can be obtained recursively. For a layer ‘$b$’ separating a medium ‘$a$’ from a substrate ‘$c$’, for example, $$r_{abc} = \frac{ r_{ab} + r_{bc} \,{\rm e}^{ 2 {\rm i} k_{b} w } }{ 1 + r_{ab} r_{bc} \,{\rm e}^{ 2 {\rm i} k_{b} w } } \label{eq:multiple-layer-r}$$ where $r_{ab}$ ($r_{bc}$) describes the reflection from the interface $ab$ ($bc$), respectively, and $w$ is the layer thickness [@BornWolf; @Yeh]. If the substrate $c$ is a multilayer system itself, $r_{bc}$ is the corresponding reflection coefficient. For the imaginary frequencies occurring in the Lifshitz expression (\[eq:Lifshitz-formula\]), the wavevector in the layer is purely imaginary, $k_{b} = {\rm i} \, \kappa_b$, and single-interface coefficients are real \[Eq.(\[eq:Fresnel-r\])\]. From Eq.(\[eq:multiple-layer-r\]), they remain real for multilayer mirrors. In addition, the mapping $r_{ab} \mapsto r_{abc}$ is a conformal one, and if $r_{bc} \,{\rm e}^{ - 2 \kappa_b w }$ is real and $\in[-1,1]$, the interval $[-1,1]$ is mapped onto itself. For multilayer mirrors, we thus obtain again the inequalities $-1 \le r_{\lambda} \le 1$. This generalizes the limits of Refs. [@Lambrecht97; @Genet03c] that are obtained only for layered dielectric mirrors, using transfer matrices. Casimir interaction between metamaterials ========================================= To illustrate these generally valid results, we focus on meta materials described by effective medium theory [@Ramakrishna05; @Pendry99d; @Smith04a]. We adopt Lorentz-Drude formulas for $\varepsilon$ and $\mu$ $$\varepsilon_{\alpha}( {\rm i}\,\xi ) = 1 + \frac{ \Omega_{\alpha}^2 }{ \omega_{\alpha}^2 + \xi^2}, \qquad \mu_{\alpha}( {\rm i}\,\xi ) = 1 + \frac{ \Theta_{\alpha}^2 }{ \vartheta_{\alpha}^2 + \xi^2 } . \label{eq:Lorentz-Drude-1}$$ Regarding the permeability, we have taken the limit of weak absorption and computed $\mu( {\rm i}\,\xi )$ in terms of ${\rm Im}\,\mu( \omega )$ using the Kramers-Kronig relations. This is necessary to ensure high-frequency transparency of the medium. We denote in the following by $\Omega$ a typical resonance or plasma frequency occurring in Eqs.(\[eq:Lorentz-Drude-1\]). The corresponding wavelength, $\Lambda = 2\pi c / \Omega$, provides a convenient distance scale. Note that a (magnetic) resonance wavelength as short as $\sim 3\,\mu{\rm m}$ has already been achieved with material nanofabrication [@Wegener04]. The key advantage of meta materials is that their electric and magnetic ‘plasma frequencies’ $\Omega_\alpha$ and $\Theta_\alpha$ are fairly large as well: a value of $\Theta_\alpha \approx \vartheta_\alpha \sqrt{f} \le (c/a) \sqrt{f}$ is typical for a split-ring resonator array with period $a$ and filling factor $f$ [@Pendry99d]. This property is also necessary, of course, to achieve a left-handed medium ($\varepsilon_\alpha( \omega )$, $\mu_\alpha( \omega ) < 0$ for some real frequencies). The magnetic plasma frequencies occurring in conventional ferromagnets are much smaller [@Camley98], and the impact on the Casimir interaction is weak, as reported recently [@Tomas05]. In the plots shown below, the Casimir pressure is normalized to $\hbar \Omega / d^3$ \[see after Eq.(\[eq:c3-attraction\])\]. In order of magnitude, this corresponds to $10^4\, {\rm pN}\, {\rm mm}^{-2}/(\Lambda / \mu{\rm m})^4$ at a distance $d = \Lambda/2$. This can be measured with sensitive torsion balances [@Lamoreaux97a; @Chan01] or cantilevers [@Mohideen98; @Onofrio02]. We plot in Fig.\[fig:attr-rep\] the result of a numerical integration of Eq.(\[eq:Lifshitz-formula\]), the curves corresponding to different material pairings. One sees that in all cases, the force satisfies the limits (\[eq:F-limits-zero-T\]) that exclude the shaded areas. We observe that materials with negative index of refraction around $\Omega$ show a strongly reduced attraction (Fig.\[fig:attr-rep\](b)). This can be attributed to the reduced mirror reflectivity due to impedance matching. Casimir repulsion is achieved for some distances between mirrors made from different materials (Fig.\[fig:attr-rep\](c,d)). At short distance, i.e. $d \ll \Lambda / 2\pi$, even these pairings show attraction with a power law $1/d^3$. Coating one mirror with a magnetic layer (Fig.\[fig:attr-rep\](c)), there is a sign change around the layer thickness $w$: for $\Lambda/2\pi \ll d \ll w$, the layer behaves like a thick plate, and its material parameters lead to repulsion. The layer can be ignored for $w \ll d$, and one recovers the attraction between the (identical) substrates. This is consistent with asymptotic analysis based on the reflection coefficient (\[eq:multiple-layer-r\]), as we outline below. Detailed calculations show that a large resonance frequency is not sufficient to achieve repulsion, the oscillator strength of the resonances (proportional to $\Omega_{1}$ and $\Theta_{2}$) must be large enough so that $\hbar \Omega_{1}, \hbar \Theta_{2} \gg \max( k_{B} T, \hbar c / d )$. As the temperature is raised, the distance range where repulsion is observed disappears, see Fig.\[fig:finite-T\]. One then finds a $1/d^3$ power law at large distance as well. The different regimes of Fig.\[fig:attr-rep\] can be understood from an asymptotic analysis of Eq.(\[eq:Lifshitz-formula\]). At short distance ($d \ll \Lambda/2\pi$), the integral is dominated by a region in the $\kappa$-$\xi$-plane where the Fresnel coefficients (\[eq:Fresnel-r\]) take the nonretarded forms $r_{{\rm TM}} \to R( \varepsilon) \equiv (\varepsilon - 1)/( \varepsilon + 1 ) > 0$ assuming that $\varepsilon > 1$ and similarly $r_{{\rm TE}} \to R( \mu ) > 0$ unless $\mu = 1$. Proceeding like Lifshitz [@Lifshitz56; @Henkel04a], yields to leading order a power law $F_{L} = c_{3} / d^3$ with a *positive* Hamaker constant given by $$c_{3} = \frac{ k_{B} T }{ 4\pi } {\kern 4.5ex\raisebox{0.25ex}{$'$}\kern -4.5ex}\sum\limits_{n=0}^{\infty} \left\{ {\rm Li}_{3}[ R( \varepsilon_1(\xi_n) ) R( \varepsilon_2(\xi_n) ) ] + {\rm Li}_{3}[ R( \mu_1(\xi_n) ) R( \mu_2(\xi_n) ) ] \right\} , \label{eq:c3-attraction}$$ where ${\rm Li}_{n}( z ) \equiv \sum_{k=1}^\infty z^k / k^n$. It must be noted that for the special case of homogeneous plates, this asymptotic expression actually provides another, much stricter, upper limit to the Casimir force, since $r_{{\rm TM}\alpha} \le R( \varepsilon_{\alpha} )$, $r_{{\rm TE}\alpha} \le R( \mu_{\alpha} )$ and ${\rm Li}_{3}( z )$ is a monotonous function (see Fig.\[fig:attr-rep\]). In order of magnitude, $c_{3} \sim \hbar\Omega$ at low temperatures ($k_{B} T \ll \hbar \Omega$). Compared to ideal mirrors, dispersive plates thus show a much weaker Casimir interaction that is in general attractive (Fig.\[fig:attr-rep\] and Ref.[@Lambrecht97]). At larger distances, $\Lambda/2\pi \ll d \ll \Lambda_{T} \equiv \hbar c / k_{B} T$, the Casimir force follows a $1/d^4$ power law, and repulsion is found provided one of the materials is dominantly magnetic. Here, the non-dispersive results of Ref.[@Kenneth02] are recovered. Finally, for $d \gg \Lambda_{T}$, the leading order force is the term $n=0$ in the sum (\[eq:Lifshitz-formula\]), again an attractive $1/d^3$ law, with a coefficient given by an expression similar to (\[eq:c3-attraction\]), but involving the static material constants, see [@Kenneth02]. The impact of temperature is illustrated in Fig.\[fig:finite-T\]: at high temperature, $k_{B}T \gg \hbar\Omega$, the second Matsubara frequency $\xi_{1}$ falls already into the mirrors’ transparency zone, and the $1/d^3$ power law is valid at all distances. As $T \to 0$, the intermediate repulsive zone appears in the range $\Lambda/2\pi \ll d \ll \Lambda_T/2\pi$. A good agreement with the analytical $1/d^3$ asymptotics is found outside this zone, as shown by the dashed lines. For the resonance wavelength $\Lambda = 3\,\mu{\rm m}$ mentioned above, cooling to a temperature $T \approx 0.1 \,\hbar\Omega/ k_B \sim 50\,{\rm K}$ is required to ‘open up’ the repulsive window. This temperature increases, of course, with materials whose response extends to higher frequencies. Finally, we would like to illustrate the kind of peculiar asymptotics that becomes possible with carefully matched material parameters. This follows Ref.[@Parsegian69] that computes the Van der Waals force on a water film coated on both sides by lipid membranes, finding a weak dependence on the ultraviolet frequency range because both materials have a similar electron density. Consider thus a liquid-filled gap with a similar electron density as medium 2 so that $\varepsilon_{0} = \varepsilon_{2}$, and a permeability $\mu_{0} = \mu_{1} \equiv 1$ matched to medium 1. For simplicity, we assume that these equalities hold at all frequencies. In this case, we can show that the force is repulsive at all distances, even at finite temperature. Indeed, both contributions in Eq.(\[eq:c3-attraction\]) vanish, and the leading order term for high temperatures also vanishes. The high-temperature limit is given by the $n=1$ term in (\[eq:Lifshitz-formula\]). This gives a distance dependence proportional to $\exp( - 4 \pi d / \Lambda_{T})$ similar to what has been observed in some experiments with colloids (mentioned in [@Ackler96]). As the temperature is lowered, this exponential regime still applies for $d \gg \Lambda_{T}$. If $k_{B} T \ll \hbar \Omega$, the $1/d^4$ regime of Ref. [@Kenneth02] exists at intermediate distances $\Lambda/2\pi \ll d \ll \Lambda_{T}/2\pi$. The short distance regime sets in for $d \ll \Lambda$, and an analysis similar to the one leading to Eq.(\[eq:c3-attraction\]) gives ($T = 0$) $$F_{L} = \frac{ \hbar }{ \pi } \sum\limits_{n=1}^{\infty} \left( 2 n d \right)^{2n-3} \int\limits_0^\infty\!\frac{ {\rm d}\xi }{ 2\pi } \Gamma( 3 - 2n, 2 n \xi \, d / c ) \left( - \frac{ \varepsilon_{1} - \varepsilon_{0} }{ \varepsilon_{1} + \varepsilon_{0} } \frac{ (\mu_{2} - \mu_{1}) \xi^2 }{ 4 c^2 } \right)^n \label{eq:c1-repulsion}$$ with $\Gamma( k, z ) \equiv \int_{z}^{\infty}\!{\rm d}t \,t^{k-1} {\rm e}^{-t}$. At short distance, the sum is dominated by the first term, so that to leading order, we get a repulsive power law $F_{L} = - c_{1} / d$ (Fig.\[fig:mismatch\](a)). In order of magnitude, $c_{1} \sim \hbar \Omega^3 / c^2$ and therefore again $-F_{L} \ll F_{C}$, with a cross over occurring around $\Lambda/2\pi$ (see Fig.\[fig:mismatch\](a)). Due to our assumption of a perfect matching $\varepsilon_2 = \varepsilon_0$, this kind of behaviour seems quite remote from experimental reality. As shown in Fig.\[fig:mismatch\](b–d), a slight mismatch between the dielectric functions of liquid and plate leads back to an attractive force, first at short distances, then suppressing the repulsive window altogether. Conclusion ========== We have generalized strict upper and lower limits for the Casimir force. We have shown that a strongly modified Casimir force can occur between dispersive and absorbing mirrors with a sufficiently large magnetic susceptibility, extending results restricted to non-dispersive materials [@Kenneth02]. The most promising way to achieve this repulsion seems the use of meta materials engineered at scales between the nanometer and the micron because they provide a fairly large magnetic oscillator strength. Our results are intrinsically limited to distances $d \gg a$ by our use of effective medium theory. Sufficiently small structures and sufficiently low temperatures then ensure that in the range $a \ll d \ll \Lambda_T/2\pi$, the Casimir interaction can be strongly altered: even if repulsion cannot be achieved in a first step, we expect a significant reduction of the Casimir attraction at distances of a few microns (Fig.\[fig:attr-rep\]). Our thanks for comments and discussion goes to J.-J. Greffet, J.-P. Mulet, M. Wilkens, M. Tomaš, C. Genet, A. Lambrecht, S. Reynaud, and L. Pitaevskii. We thank anonymous referees for constructive criticism. We acknowledge financial support from the bilateral French-German programme “Procope” under project numbers 03199RH and D/0205739. [10]{} S. A. Ramakrishna, Rep. Prog. Phys. [**68**]{}, 449 (2005). T. J. Yen [*et al.*]{}, Science [**303**]{}, 1494 (2004). S. Linden, C. Enkrich, M. Wegener, J. Zhou, T. Koschny, and C. M. Soukoulis, Science [**306**]{}, 1351 (2004). V. V. Klimov, Opt. Commun. [**211**]{}, 183 (2002). J. Kästel and M. Fleischhauer, Phys. Rev. A [**71**]{}, 011804 (2005). T. H. Boyer, Phys. Rev. A [**9**]{}, 2078 (1974). O. Kenneth, I. Klich, A. Mann, and M. Revzen, Phys. Rev. Lett. [**89**]{}, 033001 (2002). T. H. Boyer, American Journal of Physics [**71**]{}, 990 (2003). U. Hartmann, Phys. Rev. B [**43**]{}, 2404 (1991). J. N. Israelachvili, [*Intermolecular and surface forces*]{}, 2nd ed. (Academic Press, San Diego, 2000). S.-W. Lee and W. M. Sigmund, J. Coll. Int. Sci. [**243**]{}, 365 (2001). J. N. Munday, D. Iannuzzi, Y. Barash, and F. Capasso, Phys. Rev. A [**71**]{}, 042102 (2005). M. S. Tomaš, Phys. Lett. A [**342**]{}, 381 (2005). R. E. Camley, M. R. F. Jensen, S. A. Feiven, and T. J. Parker, J. Appl. Phys. [**83**]{}, 6280 (1998). D. Iannuzzi and F. Capasso, Phys. Rev. Lett. [**91**]{}, 029101 (2003); O. Kenneth, I. Klich, A. Mann, and M. Revzen, *ibid.* [**91**]{}, 029102 (2003). L. D. Landau, E. M. Lifshitz, and L. P. Pitaevskii, [*Electrodynamics of continuous media*]{}, 2nd ed. (Pergamon, Oxford, 1984). J. B. Pendry, A. J. Holden, D. J. Robbins, and W. J. Stewart, IEEE Trans. Microwave Theory Tech. [**47**]{}, 2075 (1999). D. R. Smith, P. Rye, D. C. Vier, A. F. Starr, J. J. Mock, and T. Perram, IEICE Trans. Electron. [**E87-C**]{}, 359 (2004). H. B. G. Casimir, Proc. Kon. Ned. Akad. Wet. [**51**]{}, 793 (1948). E. M. Lifshitz, Soviet Phys. JETP [**2**]{}, 73 (1956), \[[*J. Exper. Theoret. Phys. USSR*]{} [**29**]{}, 94 (1955)\]. V. A. Parsegian and B. W. Ninham, Nature [**224**]{}, 1197 (1969). B. W. Ninham and V. A. Parsegian, J. Chem. Phys. [**53**]{}, 3398 (1970). P. Richmond and B. W. Ninham, J. Phys. C: Solid State Phys. [ **4**]{}, 1988 (1971). F. C. Santos, A. Tenório, and A. C. Tort, Phys. Rev. D [**60**]{}, 105022 (1999). K. Schram, Phys. Lett. [**43A**]{}, 282 (1973). J. Schwinger, L. L. DeRaad, Jr., and K. A. Milton, Ann. Phys. (N.Y.) [**115**]{}, 1 (1978). D. Kupiszewska, Phys. Rev. A [**46**]{}, 2286 (1992). F. Zhou and L. Spruch, Phys. Rev. A [**52**]{}, 297 (1995). G. L. Klimchitskaya, U. Mohideen, and V. M. Mostepanenko, Phys. Rev. A [ **61**]{}, 062107 (2000). R. Esquivel-Sirvent, C. Villarreal, and G. H. Cocoletzi, Phys. Rev. A [**64**]{}, 052108 (2001). M. S. Tomaš, Phys. Rev. A [**66**]{}, 052103 (2002). C. Genet, A. Lambrecht, and S. Reynaud, Phys. Rev. A [**67**]{}, 043811 (2003). M. Born and E. Wolf, [*Principles of Optics*]{}, 6th ed. (Pergamon Press, Oxford, 1959). P. Yeh, [*Optical Waves in Layered Media*]{} (John Wiley & Sons, New York, 1988). A. Lambrecht, M.-T. Jaekel, and S. Reynaud, Phys. Lett. A [**225**]{}, 188 (1997). S. K. Lamoreaux, Phys. Rev. Lett. [**78**]{}, 5 (1997), erratum: [**81**]{} (1998) 5475. H. B. Chan, V. A. Aksyuk, R. N. Kleiman, D. J. Bishop, and F. Capasso, Science [**291**]{}, 1941 (2001). U. Mohideen and A. Roy, Phys. Rev. Lett. [**81**]{}, 4549 (1998). G. Bressi, G. Carugno, R. Onofrio, and G. Ruoso, Phys. Rev. Lett. [**88**]{}, 041804 (2002). C. Henkel, K. Joulain, J.-P. Mulet, and J.-J. Greffet, Phys. Rev. A [**69**]{}, 023808 (2004). H. D. Ackler, R. H. French, and Y.-M. Chiang, J. Coll. Int. Sci. [**179**]{}, 460 (1996).
{ "pile_set_name": "ArXiv" }
--- abstract: 'Elastic crystalline membranes exhibit a buckling transition from sphere to polyhedron. However, their morphologies are restricted to convex polyhedra and are difficult to externally control. Here, we study morphological changes of closed crystalline membrane of super-paramagnetic particles. The competition of magnetic dipole-dipole interactions with the elasticity of this magnetoelastic membrane leads to concave morphologies. Interestingly, as the magnetic field strength increases, the symmetry of the buckled membrane decreases from 5-fold to 3-fold, to 2-fold and, finally, to 1-fold rotational symmetry. This gives the ability to switch the membrane morphology between convex and concave shapes with specific symmetry and provides promising applications for membrane shape control in the design of actuatable micro-containers for targeted delivery systems.' author: - Hang Yuan - Monica Olvera de la Cruz bibliography: - 'Literatures.bib' title: Crystalline membrane morphology beyond polyhedra --- =v = = == Polyhedra are of great interest of scientists, mathematicians and engineers. They emerge spontaneously in many fields of science. For example, single crystals take various polyhedra shapes, fullerenes adopt beautiful truncated icosahedron shapes[@Kroto1985], and bacterial micro-compartments are observed in multiple regular and irregular polyhedral shapes[@Fan2010].\ Self-assembled crystalline membranes, like the shells of viruses, generally possess icosahedral symmetry[@Zandi2004]. These icosahedral membrane shapes have been explained by homogeneous elasticity theory[@PhysRevA.38.1005; @PhysRevE.68.051910]. Furthermore, membranes with heterogeneous elasticity have been demonstrated to form various regular and irregular polyhedral shapes[@Vernizzi4292]. Such polyhedral morphologies are formed by the competition between stretching energy and bending energy. Although it is possible to engineer membrane morphologies by arranging defects in closed membrane topologies[@PhysRevLett.111.177801], these morphologies cannot go beyond polyhedra.\ Here, we explore the possibility to create new closed shell morphologies, other than polyhedra, in a controllable manner. For this purpose, we consider elastic membranes of super-paramagnetic particles because of the exceptional penetration of magnetic fields and bio-compatibility, which provide opportunities for biotechnology applications. Magnetoelastic materials form rich morphologies[@Kim2018; @LumE6007] and can accomplish multimodal locomotion[@Hu2018] as well as deformations that generate forces between surfaces[@Brisbois2019] when directed by magnetic fields. The versatility of magnetoelastic filaments, which consist of super-paramagnetic particles connected by elastic linkers, has also been demonstrated experimentally[@Dreyfus2005; @Wang2011; @Wei2016] and numerically[@PhysRevE.95.052606; @Vazquez-Montejo2017].\ Comparing with magnetoelastic filaments and open membranes, closed magnetoelastic membranes, which have additional topological constraints, are found here to generate specific symmetries due to the interplay between nonlinear elasticity and magnetic dipole-dipole interactions. By using molecular dynamics simulations, we find the minimum energy configurations of magnetoelastic membranes, which can be directly controlled by external magnetic fields.\ As dictated by Euler’s polyhedron formula, we start by triangulating a spherical shell with twelve isolated 5-fold disclinations. The disclinations are positioned on the vertices of an inscribed icosahedron (Fig. \[mesh\]) to minimize the interactions between them[@PhysRevB.62.8738], as proposed by Caspar and Klug[@Caspar01011962].\ ![Mesh configuration of the spherical shell according to Caspar and Klug construction, which is characterized by two integers $h$ and $k$[@Caspar01011962]. Above figure shows the example of $(6,6)$ strucutre and it has 1082 vertices, 3240 edges and 2160 faces. Blue vertices correspond to the locations of 5-fold disclinations and there are 12 disclinations in total which are located on vertices of an inscribed icosahedron.[]{data-label="mesh"}](mesh.png){width="45.00000%"} The elastic component of the Hamiltonian of a magnetoelastic membrane, following the discretization scheme of Nelson et al[@PhysRevA.38.1005], is written as $$H_e=\sum_{e\in {\ensuremath{\mathbf{E}}}}\frac{1}{2}k\left({\left| {\ensuremath{\mathbf{r}}}_1^e - {\ensuremath{\mathbf{r}}}_2^e \right|} - l_0\right)^2 + \sum_{e\in {\ensuremath{\mathbf{E}}}} \frac{1}{2}\tilde \kappa {\left| {\ensuremath{\mathbf{n}}}_1^e - {\ensuremath{\mathbf{n}}}_2^e \right|}^2$$ where $k$ is microscopic stretching constant and $\tilde \kappa$ is microscopic bending rigidity. The sum is over all $e$ elements of ${\ensuremath{\mathbf{E}}}$, which is the set of all edges; ${\ensuremath{\mathbf{r}}}_1^e$ and ${\ensuremath{\mathbf{r}}}_2^e$ are two vertices of the edge $e$; and ${\ensuremath{\mathbf{n}}}_1^e$ and ${\ensuremath{\mathbf{n}}}_2^e$ are normal vectors of the two adjacent triangles of the edge $e$; and $l_0$ is the equilibrium length. Note that the corresponding continuum limit of the above discretized Hamiltonian is mesh dependent[@Gompper1996]. With the above described triangulation of a spherical shell, it has been shown that in the continuum limit[@PhysRevA.38.1005; @Nelson2004] Young’s modulus $Y=\frac{2k}{\sqrt 3}$, Poisson’s ratio $\nu =\frac{1}{3}$ and bending rigidity $\kappa=\frac{\tilde \kappa}{\sqrt 3}$.\ Incompressible membranes ($\nu=1/3$) of radius $R$ can be described by two parameters $Y$ and $\kappa$. Then, a single dimensionless parameter, $\gamma=\frac{YR^2}{\kappa}$, called the F$\mathrm{\ddot{o}}$ppl-von Kármán parameter[@Libai1998], completely determines the buckling transition of the system. Nelson et al[@PhysRevE.68.051910] have shown that homogeneous elastic membranes undergo a spontaneous buckling transition from sphere to icosahedron when $\gamma > \gamma^\ast=154$, where $154$ is the value of $\gamma^\ast$ for a flat disk.\ In our study, we place a small super-paramagnetic particle at each vertex. An external magnetic field induces a magnetic dipole on each vertex. Therefore, an additional term from magnetic dipole-dipole interactions is added into the Hamiltonian of the system: $$H_m=-\frac{\mu_0}{4\pi}\sum_{{\ensuremath{\mathbf{r}}}_i,{\ensuremath{\mathbf{r}}}_j\in {\ensuremath{\mathbf{V}}}}\frac{1}{{\left| {\ensuremath{\mathbf{r}}}_{ij} \right|}^3}\left[3\left({\ensuremath{\mbox{\boldmath$ \mu $}}}_i \cdot {\ensuremath{\mathbf{\hat{r}}}}_{ij}\right)\left({\ensuremath{\mbox{\boldmath$ \mu $}}}_j\cdot {\ensuremath{\mathbf{\hat{r}}}}_{ij}\right)-{\ensuremath{\mbox{\boldmath$ \mu $}}}_i\cdot {\ensuremath{\mbox{\boldmath$ \mu $}}}_j\right]$$ where $\mu_0$ is the magnetic permeability in vacuum, ${\ensuremath{\mbox{\boldmath$ \mu $}}}_i$ is the magnetic dipole moment at vertex $i$, ${\ensuremath{\mathbf{V}}}$ is the set of all vertices, ${\ensuremath{\mathbf{r}}}_i$ is the position vector of vertex $i$, ${\ensuremath{\mathbf{r}}}_{ij}={\ensuremath{\mathbf{r}}}_j - {\ensuremath{\mathbf{r}}}_i$ and ${\ensuremath{\mathbf{\hat{r}}}}_{ij}={\ensuremath{\mathbf{r}}}_{ij}/{\left| {\ensuremath{\mathbf{r}}}_{ij} \right|}$ and the sum is over $i\neq j$.\ The magnetic dipole-dipole interaction is long range and anisotropic. A useful simplified form which considers only nearest neighbor interactions in the inextensible limit[^1] is derived and yields: $$H_m\approx \left(\sum_{{\ensuremath{\mathbf{r}}}_i \in {\ensuremath{\mathbf{V}}}^{hex}}6 + \sum_{{\ensuremath{\mathbf{r}}}_i \in {\ensuremath{\mathbf{V}}}^{pen}}5\right)\left({n_z^{i}}^2 - \frac{1}{3}\right)\tilde M$$ where $\tilde M=\frac{1}{4}\frac{\mu_0}{4\pi}\frac{(3\mu)^2}{l_0^3}\frac{2}{3}$, $\mu$ is the induced magnetic dipole moment which assumes only one type of super-paramagnetic particles, ${\ensuremath{\mathbf{V}}}^{hex}$ is the set of vertices with 6 neighbors, ${\ensuremath{\mathbf{V}}}^{pen}$ is the set of vertices with 5 neighbors and $n_z^{i}$ is the z component of normal vector at vertex $i$.\ $\tilde M$ gives the characteristic energy scale for each nearest neighbor pair of magnetic dipole-dipole interactions in the discretization limit. Similar to the case of elastic membranes, a magnetic modulus can be defined in the continuum limit as $M=8\sqrt 3 \frac{\tilde M}{l_0^2}$, and a dimensionless parameter, $\Gamma = \frac{MR^2}{\kappa}$, called magnetoelastic parameter[@PhysRevE.98.032603], can be similarly defined. The magnetoelastic parameter $\Gamma$ characterizes the relative strength between magnetic energy and bending energy(see SI).\ Therefore, the magnetoelastic membrane has one additional energy competition from magnetic dipole-dipole interactions, which is tunable via an external magnetic field. The total magnetoelastic energy of the membrane $H_{em}$ is the sum of elastic and magnetic energies, which divided by $\kappa$ gives the dimensionless form: $$\begin{aligned} \tilde{H}_{em}\left[\{{\ensuremath{\mathbf{r}}}_i\};\gamma,\Gamma\right]&=\frac{H_{em}\left[\{{\ensuremath{\mathbf{r}}}_i\}\right]}{\kappa}\\ &=\tilde{H}_e\left[{\{{\ensuremath{\mathbf{r}}}_i\}};\gamma\right]+\tilde{H}_m\left[{\{{\ensuremath{\mathbf{r}}}_i\}};\Gamma\right] \end{aligned}$$ where tilde indicates dimensionless quantites.\ Besides magnetic and elastic contributions, a volume constraint is also imposed on the membranes to account for internal pressure. This internal pressure is necessary when the membrane is not penetrable, which is modeled as $$H_v=\Lambda\left(\sum_k \Omega_k - V_{ref}\right)^2$$ where $\Omega_k$ is the signed volume of the tetrahedron extended by k-th triangle on the membrane, $V_{ref}$ is the reference volume of the membrane and $\Lambda$ is the Lagrange multiplier which serves role of pressure. $V_{ref}$ is set as volume of the icosahedron after buckling and $\Lambda$ is set to a large enough value such that the membrane has additional rigidity from the volume constraint. The volume constraint is used to better capture effect from the environment surrounding the magnetoelastic membrane and eliminate possible crumpled states[@RevModPhys.79.643]. Corresponding cases without the volume constraint are also explored, and their morphologies generally do not differ significantly from the cases with the volume constraint. Some crumpled states and collapsed states are observed in high field strength limit for the cases without the volume constraint(see SI for more discussions).\ In the simulation, a shifted Lennard-Jones potential is also included for each pair of vertices to account for the exclude volume effect. Each vertex is assigned a point magnetic dipole moment. Stretching and bending are treated with a harmonic bond interaction and a harmonic dihedral interaction, respectively. Magnetic dipole-dipole interactions are calculated without a cutoff. The connectivity of the membrane is preserved during simulations. The external magnetic field is static along z direction. We assume super-paramagnetic particles respond to an external magnetic field instantaneously and ignore rotational degrees of freedom of each vertex because super-paramagnetic particles do not have spontaneous magnetization, which decouples magnetics and elasticity. Kinetic energy is also assigned to each vertex to give a fictitious temperature of the system. The simulations start at high temperature and are gradually annealed to find the minimum energy configuration of the system. This annealing process is repeated several times to ensure that the system is not trapped in local minima.\ A collection of possible morphologies of magnetoelastic membranes obtained by systematically varying the two dimensionless parameters, the F$\mathrm{\ddot{o}}$ppl-von Kármán parameter $\gamma$ and the magnetoelastic parameter $\Gamma$, are shown in Fig. \[morphologies\]. Without magnetic dipole-dipole interactions ($\Gamma$=0), when $\gamma<\gamma^\ast$, the homogeneous elastic membrane tends to stay spherical (Fig. \[morphologies\]e) and when $\gamma>\gamma^\ast$ it buckles into an icosahedron (Fig. \[morphologies\]f) as expected in the conventional homogeneous elastic membranes[@PhysRevE.68.051910].\ At moderate strengths of the magnetic dipole-dipole interaction, as shown in second row of Fig. \[morphologies\], the structures deform since the magnetic dipoles prefer to line up and stay closer to each other to minimize the magnetic energy. When the membrane is relatively soft ($\gamma < \gamma^\ast$), the membrane tends to elongate along the direction of the external magnetic field. However, this is opposed by elastic interactions since elasticity prefers the membrane to stay spherical, resulting in an ellipsoid like membrane morphology as shown in Fig. \[morphologies\]c.\ ![A collection of representative minimum energy morphologies of closed magnetoelastic membranes with different parameters pair $(\gamma,\Gamma)$: F$\mathrm{\ddot{o}}$ppl-von Kármán parameter $\gamma$ and magnetoelastic parameter $\Gamma$; $\gamma$ increases from left to right and $\Gamma$ increases from bottom to top. (a) cylindrical shape (100,50); (b) star shape with four ridges (1000,200); (c) ellipsoid shape (100,25); (d) star shape with six ridges (1000,100); (e) spherical shape (100,0); (f) icosahedral shape (1000,0). Note that first column is shown from y-direction and second column is shown from z-direction to give better illustration of morphologies. Arrows indicate the direction of the external magnetic field.[]{data-label="morphologies"}](morphologies-volume-constraint.png){width="50.00000%"} When the membrane is relatively stiff ($\gamma > \gamma^\ast$), the membrane undergoes an elastically driven buckling transition. The interplay between nonlinear elasticity and magnetic dipole-dipole interactions distorts the icosahedron. The flat regions of the icosahedron bend inward to reduce the distance between magnetic dipoles and disclinations pair up, resulting in a star-like morphology with six ridges as shown in Fig. \[morphologies\]d. Unlike the conventional convex polyhedral morphologies of the purely elastic membranes, the magnetoelastic membranes develop concave regions.\ ![image](curvatures-volume-constraint.png){width="\textwidth"} Then, consider the case of strong magnetic dipole-dipole interactions, as shown in first row of Fig. \[morphologies\]. The elastic energy only becomes comparable with the magnetic energy until the membrane is highly deformed. In this regime, the competition between magnetic energy and elastic energy results in another new family of morphologies.\ When the membrane is easily deformed ($\gamma < \gamma^\ast$), magnetic dipole-dipole interactions tend to elongate the membrane furthermore along the direction of external magnetic field in this high field strength regime. However, the elastic energy can no longer hold the membrane in a spherical or ellipsoidal shape. The membrane forms a cylindrical shape, as shown in Fig. \[morphologies\]a, to minimize the magnetic energy. Although the bending energy is high along edges of two end caps of the cylinder, the total energy decreases by lining up vertices on the side surface of the cylinder.\ When the membrane is relatively rigid ($\gamma > \gamma^\ast$), the elastic energy tries to preserve the total surface area of the membrane since stretching is much more expensive than bending in this case. Meanwhile, the magnetic dipole-dipole interaction tries to reduce the total volume of the membrane to minimize the magnetic energy. This competition, combined with the nonlinearity introduced by the twelve disclinations, results in a star-like morphology with four ridges as shown in Fig. \[morphologies\]b. Note that the membrane in this case is highly bent inward, which reduces its total volume significantly and opens some possible applications as discussed later.\ Among all these mentioned morphologies of the magnetoelastic membrane, the $\gamma\sim 1000$ cases are particularly interesting because this regime corresponds to a typical F$\mathrm{\ddot{o}}$ppl-von Kármán parameter of viral shells[@PhysRevE.68.051910]. In this regime, where both nonlinear elasticity and magnetic dipole-dipole interactions are significant, we find that the magnetoelastic membrane tends to choose configurations that decrease symmetry with increasing external magnetic field strength. This point is illustrated by plotting the mean curvature and energy distribution of the membrane in spherical coordinates as shown in Fig. \[curvatures\].\ In the weak field strength limit, the membrane forms an icosahedron (Fig. \[curvatures\]a) with five-fold rotational symmetry around the z-axis as shown in Fig. \[curvatures\]b and \[curvatures\]c. In this limit, the elastic energy dominates and the magnetic energy is negligible (Fig. \[curvatures\]d). With a moderate external magnetic field strength, the membrane morphology has six ridges(Fig. \[curvatures\]e). However, the 12 isolated disclinations prefer to pair up and form ridges connecting each pair of disclinations[@Lobkovsky1995]. These disclinations pairs are arranged alternatively to maximize the mutual distance in order to reduce interactions between disclinations[@PhysRevB.62.8738] and ridges[@PhysRevE.55.1577], as shown in Fig. \[curvatures\]f and \[curvatures\]g. Because of this alternative arrangement, the membrane with six ridges has only three-fold rotational symmetry around the z-axis, which is also reflected by magnetic energy distribution (Fig. \[curvatures\]h). If the field strength is further increased, the membrane starts to form morphologies with four ridges (Fig. \[curvatures\]i). In this regime, two pairs of disclinations break and there is a single disclination near each of the four concave regions as show in Fig. \[curvatures\]j and \[curvatures\]k. Therefore, the symmetry of the membrane reduces to two-fold rotational symmetry (Fig. \[curvatures\]j, \[curvatures\]k and \[curvatures\]l) around the z-axis. In the extremely high field strengths regime, the magnetic energy completely dominates and the membrane collapses and takes one-fold rotational symmetry(the collapsed state is not shown in Fig. \[curvatures\]).\ A natural question is to ask why the four-fold rotational symmetry is missing among all these above mentioned morphologies. A qualitatively argument suggests that if the system was to form a morphology with four-fold rotational symmetry, the 12 disclinations must be divided into groups of 3 disclinations, which means pairs of disclinations must break up. A direct visualization of the four-fold rotational symmetry case is to image the two-fold rotational symmetry case without alternative arrangement of the disclinations. Since weak field strengths cannot break up disclinations pairs and strong field strengths prefer alternative arrangement of disclinations to lower the system energy, the four-fold rotational symmetry is missed in these morphologies.\ When the volume constraint is removed, we find similar morphologies to those discussed above except in high magnetic field strengths, where they take crumpled or collapsed morphologies with one-fold rotational symmetry(see Fig.2 in SI).\ In summary, crystalline magnetoelastic membranes exhibit concave morphologies beyond the conventional polyhedral shapes found in elastic membranes. Magnetic dipole-dipole interactions give an additional control parameter which is characterized by the magnetoelastic parameter $\Gamma$. Combining with the F$\mathrm{\ddot{o}}$ppl-von Kármán parameter $\gamma$ in the elastic membranes, these two dimensionless parameters provide guidelines for analyzing properties of crystalline magnetoelastic membranes. Importantly, since $\gamma$ is hard to change once a membrane is assembled, the magnetoelastic parameter, which can be easily manipulated by an external magnetic field, provides a way to tune the membrane morphology between convex shapes and concave shapes with specific symmetry.\ Exciting applications, including reversible membrane shape control, design of micro-container, and targeted drug delivery, are expected for the closed crystalline magnetoelastic membranes. For example, since the volume to surface ratio of magnetoelastic membranes can be highly reduced by imposing an external magnetic field, the concentration inside can be much higher than that in the outside environment. This morphological change induced by the external magnetic field can facilitate release of cargoes. Therefore, the magnetoelastic membrane can be used as a container to carry and protect volatile or toxic molecules and release them in targeted region labelled by the external magnetic fields. This work was supported as part of the Center for Bio-Inspired Energy Science, an Energy Frontier Research Center funded by the US Department of Energy, Office of Science, Basic Energy Sciences under Award DE-SC0000989. [^1]: For details of derivation, please refer to SI.
{ "pile_set_name": "ArXiv" }
--- author: - | Arata Yamamoto\ Department of Physics, The University of Tokyo, Tokyo 113-0033, Japan\ E-mail: title: Lattice QCD simulation of the Berry curvature --- Berry curvature =============== Let us consider the Hamiltonian $H(p)$ described by a parameter $p$. From the eigenvalue equation $$H(p) \Phi(p) = E(p) \Phi(p) ,$$ the eigenfunction $\Phi(p)$ is obtained. The Berry connection is defined as $$\tilde{A}_\mu(p) = -i \Phi^\dagger(p) \frac{\partial}{\partial p^\mu} \Phi(p)$$ and the Berry curvature is defined as $$\tilde{F}_{\mu\nu}(p) = \frac{\partial}{\partial p^\mu} \tilde{A}_\nu(p) - \frac{\partial}{\partial p^\nu} \tilde{A}_\mu(p) .$$ These definitions indicate that the Berry connection and curvature correspond to the gauge connection and curvature in parameter space, respectively. The Berry curvature is quite general in theoretical physics [@Berry:1984]. It can be defined in any parameter space. In this study, we focus on the spatial momenta of the ground-state fermions. The Berry curvature of fermions is essential for describing several physical phenomena. For example, the Berry curvature of chiral fermions describes the chiral magnetic and vortical effects [@Son:2012wh], and the Berry curvature of electrons describes the quantum Hall effect [@Thouless:1982zz] and topological insulators [@Kane:2005]. Although the Berry curvature has been calculated in many theoretical works, most of the calculations were done in non-interacting approximation. For the Berry curvature including interaction effects, we need numerical simulations. For this purpose, we formulate the computational scheme to calculate the Berry curvature in lattice QCD. This presentation is based on the recent paper [@Yamamoto:2016rfr]. We would like to skip technical details and overview only the outline. Formalism ========= To calculate the Berry curvature, we need the fermion ground state $\Phi(p)$ as a function of the spatial momentum $p$. The ground state is obtained by the standard ground-state projection in lattice QCD, which is used for the ground-state hadron mass calculation. We construct a single-fermion state with a fixed spatial momentum by the spatial Fourier transformation $$\phi (p,\tau) = \sum_{x,x'} e^{ip\cdot(x-x')} D^{-1}(x,\tau|x',0) \phi_{\rm init} ,$$ where $D^{-1}(x,\tau|x',\tau')$ is a fermion propagator and $\phi_{\rm init}$ is an initial state. When the imaginary-time separation $\tau$ is large enough, this state is independent of the choice for $\phi_{\rm init}$ and goes to the ground state $$\Phi(p) = \lim_{\tau \to \infty} \phi (p,\tau) .$$ We here consider a non-degenerate ground state for simplicity. When $\Phi(p)$ is not degenerate, the Berry curvature is Abelian. In general, the Berry curvature can be defined for any state. For example, degenerate states have the non-Abelian Berry curvature [@Wilczek:1984dh]. The formulation will be easily extended to the non-Abelian case. Since lattice QCD simulations are done in a finite volume, the corresponding momentum space is also a finite-volume lattice. Thus we should formulate the Berry curvature as lattice gauge theory in momentum space [@Fukui:2005wr]. The schematic figure is shown in Fig. \[figlattice\]. The coordinate-space lattice with the spacing $a$ and the length $L$ is mapped onto the momentum-space lattice with the spacing $\tilde{a}=2\pi/L$ and the length $\tilde{L}=2\pi/a$. Since gauge connection is introduced as link variable, the Berry connection is introduced as the Berry link variable $$\tilde{U}_\mu(p) = e^{i\tilde{a} \tilde{A}_\mu(p)} = \frac{\Phi^\dagger(p) \Phi(p+\tilde{\mu})}{|\Phi^\dagger(p) \Phi(p+\tilde{\mu})|} ,$$ where $\tilde{\mu}$ is the unit vector in the $\mu$ direction on the momentum-space lattice. The Berry curvature is given by the Berry plaquette $$\tilde{P}_{\mu\nu}(p) = e^{i\tilde{a}^2 \tilde{F}_{\mu\nu}(p)} = \tilde{U}_\mu(p) \tilde{U}_\nu(p+\tilde{\mu}) \tilde{U}^*_\mu(p+\tilde{\nu}) \tilde{U}^*_\nu(p) .$$ In the Monte Carlo simulation, we calculate this Berry curvature for each configuration, and take the ensemble average over configurations. ![\[figlattice\] The schematic figure of the standard lattice gauge theory (left) and the lattice Berry field theory (right). ](figlattice.pdf) The Berry connection has the local U(1) gauge degree of freedom, which exists even in a non-interacting case. The Berry curvature is independent of the U(1) gauge choice because the plaquette is gauge invariant in lattice gauge theory. In an interacting case, there is additional gauge, i.e., the local SU(3) gauge of gluons. Since a single fermion is gauge dependent, the Berry curvature depends on the SU(3) gauge choice. Thus we did not fix the U(1) gauge but fixed the SU(3) gauge in the simulation below. Example ======= We performed the first numerical test in a simple example. We considered the (2+1)-dimensional Wilson fermion $$D(x,x') = (ma+3) \delta_{x,x'}- \frac{1}{2} \sum_{\mu=1}^{3} \bigg[ \left(1-\sigma_\mu \right) U_\mu(x) \delta_{x+\hat{\mu},x'} + \left(1+\sigma_\mu \right) U_\mu^\dagger(x') \delta_{x-\hat{\mu},x'} \bigg] ,$$ where $\hat{\mu}$ is the unit vector in the $\mu$ direction on the coordinate-space lattice. In 2+1 dimensions ,the Berry link variable is described by two-dimensional U(1) lattice gauge theory. The analysis is the same as the two-dimensional U(1) lattice gauge theory [@Panagiotakopoulos]. We calculated the Berry curvature $$\tilde{a}^2 \tilde{F}_{xy}(p) = {\rm Im \ ln} \ \tilde{P}_{xy}(p) .$$ This corresponds to the topological charge density in the two-dimensional U(1) lattice gauge theory. The integral of the topological charge density gives topological charge $$N = \frac{1}{2\pi} \sum_p \tilde{a}^2 \tilde{F}_{xy}(p) ,$$ which is called the first Chern number. The first Chern number is topological charge and thus an integer. The first Chern number of the non-interacting Wilson fermion is known. It is depicted in Fig. \[figfree\]. For positive mass $m>0$, topology is trivial, i.e., $N=0$. On the other hand, for negative mass $m<0$, topology can be nontrivial, i.e., $N \ne 0$. This behavior can be easily understood by counting the numbers of massless modes. The Chern number changes when massless modes appear. The (2+1)-dimensional Wilson fermion has one physical mode and seven doublers. The physical mode is massless at $m=0$. This gives the change from $N=1$ to $N=0$ at $m=0$. The doublers are massless only in $m<0$. These give the changes in $m<0$. In particle physics, these doubler poles are unphysical because particle mass must be $m>0$. In condensed matter physics, however, materials with $m<0$ can be generated. Actually, the non-relativistic version of the Wilson fermion is used for a model of the quantum Hall effect [@Qi:2006]. The first Chern number explains the quantization of the Hall resistivity $R_{xy} = 2\pi/e^2 N$ [@Thouless:1982zz]. ![\[figfree\] The first Chern number $N$ of the free (2+1)-dimensional Wilson fermion as a function of the bare mass parameter $m$. ](figfree.pdf) In Fig. \[figBC\], the momentum-space distribution of the Berry curvature is shown. We see the peaks at $p = 0$, which are related to the physical pole at $m=0$. The peak is negative at $ma=0.5$ and positive at $ma=-0.5$. This change indicates the topological transition at $m=0$. The Chern number $N$ is given by the integral of the Berry curvature. The Chern number of the ground state is obtained by taking imaginary time $\tau$ large enough. The imaginary-time dependence is shown in Fig. \[figN\]. For $m>0$, the result is independent of $\tau$, and thus trivially $N=0$. For $m<0$, the result depends on $\tau$. We look at the region of $\tau \ge 12$ and conclude $N=1$. These results are consistent with our expectation in Fig. \[figfree\]. The data of non-interacting simulation and quenched Monte Carlo simulation are shown in Fig. \[figN\]. The results are the same in the present parameters, while they can be different by interaction effects in general. ![\[figBC\] The Berry curvature $\tilde{a}^2 \tilde{F}_{xy}(p)$ at $ma=0.5$ (left) and $ma=-0.5$ (right). ](figBC.pdf) ![\[figN\] The first Chern number $N$ as a function of imaginary time $\tau$ at $ma=0.5$ (left) and $ma=-0.5$ (right). ](figN.pdf) Summary ======= The lattice QCD simulation of the Berry curvature was formulated. The validity of the formulation was successfully confirmed in a simple example, the (2+1)-dimensional Wilson fermion. The formulation will be applicable to physical phenomena in realistic systems: the chiral magnetic and vortical effects in QCD, the quantum Hall effect and topological insulators in condensed matter physics, etc. Acknowledgments {#acknowledgments .unnumbered} =============== The author is supported by JSPS KAKENHI Grant Number JP15K17624. The numerical simulations were carried out on SX-ACE in Osaka University. [99]{} M. V. Berry, Proc. Roy. Soc. A [**392**]{}, 45 (1984). D. T. Son and N. Yamamoto, Phys. Rev. Lett.  [**109**]{}, 181602 (2012) \[arXiv:1203.2697 \[cond-mat.mes-hall\]\]; I. Zahed, Phys. Rev. Lett.  [**109**]{}, 091603 (2012) \[arXiv:1204.1955 \[hep-th\]\]; J. W. Chen, S. Pu, Q. Wang and X. N. Wang, Phys. Rev. Lett.  [**110**]{}, 262301 (2013) \[arXiv:1210.8312 \[hep-th\]\]; G. Basar, D. E. Kharzeev and I. Zahed, Phys. Rev. Lett.  [**111**]{}, 161601 (2013) \[arXiv:1307.2234 \[hep-th\]\]. D. J. Thouless, M. Kohmoto, M. P. Nightingale, and M. den Nijs, Phys. Rev. Lett.  [**49**]{}, 405 (1982). C. L. Kane and E. J. Mele, Phys. Rev. Lett. [**95**]{}, 146802 (2005) \[arXiv:cond-mat/0506581 \[cond-mat.mes-hall\]\]; Phys. Rev. Lett. [**95**]{}, 226801 (2005) \[arXiv:cond-mat/0411737 \[cond-mat.mes-hall\]\]. A. Yamamoto, Phys. Rev. Lett. [**117**]{}, 052001 (2016) \[arXiv:1604.08424 \[hep-lat\]\]. F. Wilczek and A. Zee, Phys. Rev. Lett.  [**52**]{}, 2111 (1984). T. Fukui, Y. Hatsugai, and H. Suzuki, J. Phys. Soc. Jap.  [**74**]{}, 1674 (2005) \[arXiv:cond-mat/0503172 \[cond-mat.mes-hall\]\]. C. Panagiotakopoulos, Nucl. Phys. [**B251**]{}, 61 (1985). X. L. Qi, Y. S. Wu, and S. C. Zhang, Phys. Rev. B [**74**]{}, 045125 (2006) \[arXiv:cond-mat/0604071 \[cond-mat.mes-hall\]\].
{ "pile_set_name": "ArXiv" }
--- abstract: 'In this paper we prove that given two sets $E_1,E_2 \subset {\mathbb{Z}}$ of positive density, there exists $k \geq 1$ which is bounded by a number depending only on the densities of $E_1$ and $E_2$ such that $k{\mathbb{Z}}\subset (E_1-E_1)\cdot(E_2-E_2)$. As a corollary of the main theorem we deduce that if $\alpha,\beta > 0$ then there exist $N_0$ and $d_0$ which depend only on $\alpha$ and $\beta$ such that for every $N \geq N_0$ and $E_1,E_2 \subset {\mathbb{Z}}_N$ with $|E_1| \geq \alpha N, |E_2| \geq \beta N$ there exists $d \leq d_0$ a divisor of $N$ satisfying $d \, {\mathbb{Z}}_N \subset (E_1-E_1)\cdot(E_2-E_2)$.' address: 'School of Mathematics and Statistics, University of Sydney, Australia' author: - Alexander Fish date: 8 February 2017 title: On product of difference sets for sets of positive density --- **introduction** ================ One of the main themes of additive combinatorics is sum-product estimates. It goes back to Erdös and Szemerédi [@ES] who conjectured that for any finite set $A \in {\mathbb{Z}}$ (or in ${\mathbb{R}}$), for every ${\varepsilon}> 0$ we have $$|A+A| + |A \cdot A| \gg |A|^{2-{\varepsilon}},$$ where the $A+A = \{a+b \, | \, a,b \in A\}$, and $A\cdot A = \{ a b \, | \, a, b \in A\}$. Currently the best known estimate is due to Konyagin-Shkredov [@KS] and it is based on the beautiful previous breakthrough work by Solymosi [@So]: $$|A+A| + |A \cdot A| \gg |A|^{4/3 + c},$$ for any $c < 5/9813$. In this paper we study a slightly twisted, but nevertheless related, sum-product phenomenon. Namely, we address the following For a given **infinite** set $E \subset {\mathbb{Z}}$, how much structure does possess the set $(E-E) \cdot (E-E)$? We will restrict our attention to sets having positive density, see the definition below. Furstenberg [@Fu] noticed a intimate connection between difference sets for sets of positive density, and the sets of return times of a set of positive measure in measure-preserving systems. In this paper we will establish an arithmetic richness of a set of return times of a set of a positive measure to itself within a measure-preserving system. Recall that a triple $(X,\mu,T)$ is a measure-preserving system if $X$ is a compact metric space, $\mu$ is a probability measure on the Borel $\sigma$-algebra of $X$, and $T:X \to X$ is a bi-measurable map which preserves $\mu$. For a measurable set $A \subset X$ with $\mu(A) > 0$ the set of return times from $A$ to itself is: $$R(A) = \{ n \in {\mathbb{Z}}\, | \, \mu(A \cap T^n A) > 0\}.$$ We will denote by $E^2= \{ e^2 \, | e \in E \}$ the set of squares of $E \subset {\mathbb{Z}}$. It has been proved by Björklund and the author [@BF] that for any three sets of positive measure $A,B,$ and $C$ in measure-preserving systems there exists $k \ge 1$ (depending on the sets $A,B,$ and $C$) such that $ k \, {\mathbb{Z}}\subset R(A)\cdot R(B) - R(C)^2$. One of the motivations for this work was to show that $k$ in the latter statement depends only on the measures of the sets $A,B,$ and $C$. We prove the latter, and even more surprisingly, we show that $R(C)$ can be omitted. We have \[main\_thm\] Let $(X,\mu,T)$ and $(Y,\nu,S)$ be measure-preserving systems, and let $A \subset X, B \subset Y$ be measurable sets with $\mu(A) > 0,$ and $\nu(B) > 0$. Then there exist $k_0$ depending only on $\mu(A)$ and $\nu(B))$, and $k \le k_0$ such that $k\, {\mathbb{Z}}\subset R(A) \cdot R(B)$. This result has a few combinatorial consequences. To state the first application, we recall that the upper Banach density of a set $E \subset {\mathbb{Z}}$ is defined by $$d^*(E) =\limsup_{N \to \infty} \sup_{a \in {\mathbb{Z}}} \frac{|E \cap \{a,a+1,\ldots,a+(N-1)\}|}{N}.$$ Through Furstenberg’s correspondence principle [@Fu], we obtain \[cor1\] Let $E_1,E_2 \subset {\mathbb{Z}}$ be sets of positive upper Banach density. Then there exist $k_0$ which depends only on the densities of $E_1$ and $E_2$ and $k \le k_0$ such that $$k \, {\mathbb{Z}}\subset (E_1 - E_1)\cdot (E_2 - E_2).$$ Another application of Theorem \[main\_thm\] is the following result. \[cor2\] For any $\alpha,\beta > 0$ there exist $N_0$ and $d_0$, depending only on $\alpha$ and $\beta$, such that for every $N \geq N_0$ and $E_1,E_2 \subset {\mathbb{Z}}_N$ with $|E_1| \geq \alpha N, |E_2| \geq \beta N$ there exists $d \leq d_0$ which is a divisor of $N$ and $d \, {\mathbb{Z}}_N \subset (E_1-E_1)\cdot(E_2-E_2)$. Corollary \[cor2\] implies also that if $p$ is a large enough prime and $E_1,E_2 \subset {\mathbb{Z}}_p$ satisfy $|E_1| \ge \alpha p, |E_2| \ge \beta p$, then $(E_1 - E_1) \cdot (E_2 - E_2) = {\mathbb{Z}}_p$. This also follows from a result by Hart-Iosevich-Solymosi [@HIS] who proved that if $E \subset {\mathbf{F}}_{q}$ (where ${\mathbf{F}}_q$ is a field with $q$ elements) with $|E| \ge q^{3/4 + {\varepsilon}}$ then for $q$ large enough $(E-E) \cdot (E-E) = {\mathbf{F}}_q$. *Acknowledgment:* The work has been carried out during a research visit to Weizmann Institute, Israel. The author would like to thank Feinberg visiting program and Mathematics Department at Weizmann Institute for their support. The author is indebted to Omri Sarig for his constant encouragement and support, Eliran Subag and Igor Shparlinski for enlightening discussions, and Ilya Shkredov for his useful comments on the first version of the paper and for allowing us to reproduce his simplified proof of Lemma \[lem1\]. **Proof of Theorem \[main\_thm\]** ================================== Let us assume that $(X,\mu,T)$ is a measure-preserving system, and let $A \subset X$ be a measurable set with $\mu(A) > 0$. Recall that the set of return times of $A$ is defined by $$R(A) = \{ n \in {\mathbb{Z}}\, | \, \mu(A \cap T^n A) > 0 \}.$$ The theorem will follow from the following statement. \[lem1\] For every $L \ge 1$ and every $b \in {\mathbb{Z}}\setminus \{ 0 \}$ there exists $m \le \lfloor \frac{1}{\mu(A)^{L }}\rfloor + 1$ such that $$\{mb, 2mb, \ldots, L m b\} \subset R(A).$$ Indeed, let $R(A)$ and $R(B)$ be sets of return times for measurable sets $A$ and $B$ of positive measures. Then choose $N = \lfloor \frac{1}{\nu(B)}\rfloor+1$. Then for every $b \in {\mathbb{Z}}\setminus \{ 0 \}$ there exist $1 \leq i < j \le N$ such that $\nu((S^b)^i B \cap (S^b)^j B) > 0$. Then by $S$-invariance of $\nu$ it follows that there exists $1 \leq m \leq N$ ($m = j - i$) such that $mb \in R(B)$. Let us define $L = N!$. By Lemma \[lem1\] there exists $n = n(L,\mu(A))$ such that for every $b \in {\mathbb{Z}}\setminus \{ 0 \}$ there exists $m \leq n$ with $\{mb,2mb,\ldots,Lmb\} \in R(A)$. Let us define $k = L \cdot n!$. Take any $b \in {\mathbb{Z}}\setminus \{ 0 \}$. By the choice of $n$, there exists $m \le n$ such that $\{mb,2mb,\ldots, Lmb\} \in R(A)$. By the choice of $N$ it follows that there exists $1 \le j \le N$ such that $j \cdot \frac{k}{Lm} \in R(B)$. Also, $\frac{Lm}{j}$ is an integer less or equal than $Lm$, therefore $\frac{Lm}{j} b \in R(A)$. Thus $kb = \frac{Lm}{j} b \cdot j \frac{k}{Lm} \in R(A) \cdot R(B)$. This finishes the proof of Theorem \[main\_thm\]. *Proof[^1] of Lemma \[lem1\].* Let $(X,\mu,T)$ be a measure-preserving system, and let $A \subset X$ be a measurable set, and let $b \in {\mathbb{Z}}\setminus \{ 0\}$. We introduce a new product system $Z = \prod_{i=1}^{L } X$ with the transformation $S = \prod_{i=1}^{L } T^{ib}$ and the product measure $\nu = \prod_{i=1}^{L } \mu$. Then $(Z,\nu,S)$ is a measure-preserving system, and the set $\tilde{A} = \prod_{i=1}^{L } A$ has measure $$\nu\left( \tilde{A}\right) = \mu(A)^{L } > 0.$$ Then by Poincaré lemma there exists $m \leq \lfloor \frac{1}{\mu(A)^{L }}\rfloor + 1$ such that $$\nu(\tilde{A} \cap S^m \tilde{A}) > 0.$$ The latter means that for every $1 \le i \le L $ we have $$\mu(A \cap T^{ibm} A) > 0.$$ Therefore, we have $\{bm, 2bm,\ldots,L b m \} \in R(A)$ for $m \le \lfloor \frac{1}{\mu(A)^{L }}\rfloor + 1$. **Proofs of Corollaries \[cor1\] and \[cor2\]** =============================================== Furstenberg [@Fu] in his seminal work on Szemerédi’s theorem showed: **Correspondence Principle.** *Given a set $E \subset {\mathbb{Z}}$ there exists a measure-preserving system $(X,\mu,T)$ and a measurable set $A \subset X$ such that for all $n \in {\mathbb{Z}}$ we have $$d^*\left( E \cap (E + n) \right) \ge \mu(A \cap T^n A),$$ and $$d^*(E) = \mu(A).$$* *Proof of Corollary \[cor1\].* Let $E_1, E_2 \subset {\mathbb{Z}}$ be sets of positive densities. Then by Furstenberg’s correspondence principle there exist measure-preserving systems $(X,\mu,T)$ and $(Y,\nu,S)$ and measurable sets $A \subset X$, $B \subset Y$ that satisfy $$\mu(A) = d^*(E_1), \mbox{ } \nu(B) = d^*(E_2)$$ and $$R(A) \subset E_1 - E_1, \mbox{ } R(B) \subset E_2 - E_2.$$ By Theorem \[main\_thm\] there exist $k(\mu(A),\nu(B))$ and $k \le k(\mu(A),\nu(B))$ such that $k \, {\mathbb{Z}}\subset R(A) \cdot R(B)$. The latter statement implies the conclusion of the corollary. *Proof of Corollary \[cor2\].* Let $\alpha > 0,$ and $\beta > 0$ and let $E_1,E_2 \subset {\mathbb{Z}}_N$ with $|E_1| \geq \alpha N$, and $|E_2| \ge \beta N$. It is clear that $X = {\mathbb{Z}}_N$ with the shift map $Tx = x + 1(\!\!\!\mod N)$ and the uniform measure $\mu$ on $X$ defined by $\mu(E) = \frac{|E|}{N}$ for any $E \subset X$ is a measure-preserving system. It is also clear that for $(X,\mu,T)$ and the sets $E_1,E_2 \subset X$ we have[^2] $R(E_1) = (E_1 - E_1) + N\, {\mathbb{Z}}$ and $R(E_2) = (E_2 - E_2) + N \, {\mathbb{Z}}$. Then by Theorem \[main\_thm\] it follows that if $N \ge N_0$, where $N_0$ depends only on $\alpha$ and $\beta$, then there exist $k(\alpha,\beta)$ and $k \le k(\alpha,\beta)$ such that $k \, {\mathbb{Z}}\subset R(E_1) \cdot R(E_2)$. Then by the Chinese Remainder theorem for $d = \gcd{(k,N)} \le k$ we have $d \, {\mathbb{Z}}\subset (E_1-E_1) \cdot (E_2 - E_2) + N {\mathbb{Z}}$, which implies the statement of the corollary. **Further problems** ==================== To formulate the first problem, we mention a recent result by Björklund-Bulinski [@BB], who proved, in particular, that for any $E \subset {\mathbb{Z}}^3$ of positive density there exists $k \ge 1$, depending on the set $E$ and not only on its density, such that $$k \, {\mathbb{Z}}\subset \{ x^2 - y^2 - z^2 \, | \, (x,y,z) \in E-E\}.$$ Recall, the definition of the upper Banach density of a set $E \subset {\mathbb{Z}}^2$: $$d^*(E) = \limsup_{b - a \to \infty, d - c \to \infty} \frac{|E \cap [a,b)\times[c,d)|}{(b - a)(d- c)}.$$ Is it true that given $E_1,E_2 \subset {\mathbb{Z}}$ of positive density there exist $k_0$, which depends only on $d^*(E_1)$ and $d^*(E_2)$, and $k \le k_0$ such that $k \, {\mathbb{Z}}\subset (E_1-E_1)^2 - (E_2 - E_2)^2$? If yes, can we show that for any set $E \subset {\mathbb{Z}}^2$ of positive density there exist $k_0$, which depends only on $d^*(E)$, and $k \le k_0$ such that $k \, {\mathbb{Z}}\subset \{ x^2 - y^2 \,|\, (x,y) \in E-E\}$? The next two problems arise naturally by Theorem \[main\_thm\] and the following result proved by Björklund and the author in [@BF]: \[thm3\] Let $E \subset Mat_d^0({\mathbb{Z}}) = \{ (a_{ij}) \in {\mathbb{Z}}^{d \times d} \, | \, tr{(a_{ij})} = 0 \}$ be a set of positive density. Then there exists $k \ge 1$ (which a priori depends on the set $E$ and not only on its density) such that for any matrix $A \in k \cdot Mat_d^0({\mathbb{Z}})$ there exists $B \in E-E$ such that the characteristic polynomial of $B$ coincides with the characteristic polynomial of $A$. Is it true that given $E \subset {\mathbb{Z}}^2$ of positive upper Banach density, there exist $k_0$ that depends only on $d^*(E)$ and $k \le k_0$ such that $$k \, {\mathbb{Z}}\subset \{ x y \,\,| \,\, (x,y) \in E-E\}?$$ We also would like to establish the quantitative version of Theorem \[thm3\]: Is it true that the parameter $k$ in Theorem \[thm3\] depends only on the density of the set $E \subset Mat_d^0({\mathbb{Z}})$? In view of Corollary \[cor2\] we believe that a similar statement holds true for any finite commutative ring. Let $\alpha > 0$. Then there exist $N$ and $k$ depending only on $\alpha$ such that for any finite commutative ring $R$ with $|R| \ge N$ and any set $E \subset R$ satisfying $|E| \ge \alpha |R|$ the set $(E-E)\cdot (E-E)$ contains a subring $R_0$ such that $|R| / |R_0| \le k$. [99]{} M. Björklund, K. Bulinski, *Twisted patterns in large subsets of $\mathbb{Z}^N$.* Preprint. M. Björklund, A. Fish, *Characteristic polynomial patterns in difference sets of matrices*, Bull. London Math. Soc. (2016) 48 (2): 300-308. P. Erdös, E. Szemerédi, *On sums and products of integers.* Studies in pure mathematics, 213-218, Birkhäuser, Basel, 1983. D. Hart, A. Iosevich, J. Solymosi, *Sum-product estimates in finite fields via Kloosterman sums*, Int. Math. Res. Not. IMRN 2007, no. 5 H. Furstenberg, *Ergodic behavior of diagonal measures and a theorem of Szemerédi on arithmetic progressions.* J. Analyse Math. 31 (1977), 204–256. S.V. Konyagin, I.D. Shkredov, *New results on sum-products in $\mathbb{R}$*, Preprint, arXiv:1602.03473. J. Solymosi, *Bounding multiplicative energy by the sumset*, Advances in Mathematics Volume 222, Issue 2, (2009), 402-408. E. Szemerédi, [*On sets of integers containing no $k$ elements in arithmetic progression*]{}, Collection of articles in memory of Juriĭ Vladimirovič Linnik, Acta Arith. [^1]: This proof of the lemma has been proposed to the author by I. Shkredov. [^2]: We identify here the ring ${\mathbb{Z}}_N$ with the set $\{0,1,\ldots,N-1\}$.
{ "pile_set_name": "ArXiv" }
--- abstract: | We present a new stochastic approach to describe and remodel the conversion process of a wind farm at a sampling frequency of 1Hz. The method is trained on data measured on one onshore wind farm for an equivalent time period of 55 days. Three global variables are defined for the wind farm: the 1-Hz wind speed $u(t)$ and ten-minute average direction $\bar{\phi}$ both averaged over all wind turbines, as well as the cumulative 1-Hz power output $P(t)$. When conditioning on various wind direction sectors, the dynamics of the conversion process $u(t) \to P(t)$ appear as a fluctuating trajectory around an average IEC-like power curve, see section \[sec:wf\]. Our approach is to consider the wind farm as a dynamical system that can be described as a stochastic drift/diffusion model, where a drift coefficient describes the attraction towards the power curve and a diffusion coefficient quantifies additional turbulent fluctuations. These stochastic coefficients are inserted into a Langevin equation that, once properly adapted to our particular system, models a synthetic signal of power output for any given wind speed/direction signals, see section \[sec:model\]. When combined with a pre-model for turbulent wind fluctuations, the stochastic approach models the power output of the wind farm at a sampling frequency of 1Hz using only ten-minute average values of wind speed and directions. The stochastic signals generated are compared to the measured signal, and show a good statistical agreement, including a proper reproduction of the intermittent, gusty features measured. In parallel, a second application for performance monitoring is introduced in section \[sec:monitor\]. The drift coefficient can be used as a sensitive measure of the global wind farm performance. When monitoring the wind farm as a whole, the drift coefficient registers some significant deviation from normal operation if one of twelve wind turbines is shut down during less than $4\%$ of the time. Also, intermittent anomalies can be detected more rapidly than when using ten-minute averaging methods. Finally, a probabilistic description of the conversion process is proposed and modeled in appendix \[sec:proba\], that can in turn be used to further improve the estimation of the stochastic coefficients. Published as:\ P. Milan, M. W[ä]{}chter, J. Peinke, Stochastic modeling and performance monitoring of wind farm power production, J. Renewable Sustainable Energy **6**, 033119 (2014) author: - Patrick Milan - 'Matthias W[ä]{}chter' - Joachim Peinke title: Stochastic modeling and performance monitoring of wind farm power production --- Introduction {#sec:intro} ============ Wind energy is currently the fastest growing energy sector in Europe [@WWEA2010]. The fast integration of the complex wind resource in electric networks raises new challenges in terms of grid stability [@Liu2011; @Ayodele2012; @Milan2013a]. Smart grid concepts are being developed to cope with the turbulent nature of wind power, that involve e.g. energy storage or smart control strategies. The challenge is on the one hand technical, as the current grid structure must evolve from a strongly polarized source/load configuration with few large power sources towards a delocalized distribution of small renewable sources [@Rohden2012]. New grid concepts must be designed to handle the increasing amount of fluctuating wind power, see e.g. Ref. [@Liu2013] for an example on active wind power regulation. On the other hand, a fundamental understanding of the wind resource still lacks. Wind energy planing still focuses mostly on large-scale meteorological wind dynamics. Dynamics at faster time scales in minutes are typically simplified to a Gaussian wind field [@IEC], bypassing the complex dynamics of turbulence [@Lovejoy2001; @Boettcher2007a; @Morales2010a]. Yet it is observed on measured data that fast wind fluctuations deviate from Gaussianity [@Morales2010a], and are instead intermittent multifractals [@Lovejoy2001; @Lovejoy2009]. This observation must be coupled with the fact that modern wind turbine designs optimize power performance [@Bianchi2006] by following these fast, intermittent wind fluctuations. Besides the obvious impact on the mechanical loads acting on the turbine machinery [@Tavner2011; @Muecke2011], such performance-oriented control strategy implies feeding intermittent gusts into the power grid [@Milan2013a]. Our increasing dependence on wind energy stresses the necessity to reliably predict first the wind dynamics, and second the resulting wind power production. The first aspect is a fundamental challenge that traditionally involves both fields of meteorology for large-scale dynamics and turbulence for fast fluctuations. The central problem is that wind dynamics involve a wide range of spatio-temporal scales that are entangled in an energy cascade. The standard picture involves a three-dimensional downward cascade at small scales and a two-dimensional inverse cascade at mesoscales. In between, a hypothetical spectral gap [@vdHoven1957] would separate the two regimes. Some recent studies contradict this hypothesis and propose a universal cascade model instead [@Schertzer2012; @Fitton2011]. Besides solving directly Navier-Stokes equation including all external influences such as thermal transfers or topography (that is far from achievable), no simpler alternative is recognized as a valid method to describe the entire wind dynamics. Instead various meteorological models parametrize all the influencing effects, but they are limited in their resolution due to their high computational cost, nowadays resolving spatial and temporal scales of typically $1$ km$^2$ and $1$ hour. The smaller structures (that remain quite large) are usually not modeled using CFD, but with a statistical model of turbulence, e.g. a Gaussian correlated field [@IEC]. It should be noted that high-frequency wind models have been developed for decades [@Kaminsky1991], and the turbulence community has proposed several approaches in recent years [@Muzy2010b; @Calif2012a; @Calif2012b]. The second aspect consists in modeling the conversion process operated by the wind installation, seen here as the conversion of a wind speed $u$ conditioned on the wind direction $\phi$ into an electrical power output $P$. The complete conversion process is commonly simplified to an average power curve [@IEC-12] in the case of a single wind turbine. Being an average curve, this approach is a good description of the long-term behavior, but naturally fails to describe the dynamics in the faster time scales of minutes [@Boettcher2007a] [^1]. Ref. [@Wu2014] recently proposed a prediction model for the low-frequency (30-minute) trend of a wind farm power output from weather data using a mixed statistical/CFD approach. It has been recently shown that the cumulative power output of a wind farm [@Milan2013a] or a 300-km large wind installation [@Kamps2012] contain intermittent fluctuations at time scales of minutes and even seconds that are typical of a multifractal process [^2]. The standard power curve method overlooks such effects, so it is impossible to predict precisely how stable the power grid will be at the short time scales at which it is operated. For the optimal exploitation of wind energy, it is also important to minimize downtimes. Servicing and spare parts were found to account for one fourth of operation and maintenance costs for onshore installations, i.e. about $1\%$ of the total investment [@Blanco2009]. Ref. [@Hahn2007] made a 15-year-long study and showed that the 1500 onshore wind turbines studied had an average availability of $98\%$ (downtimes occurred during $2\%$ of the time). About half of the recorded failures were attributed to electrical and control systems, the other half coming from mechanical systems. It is interesting to note that failure rates were especially high for wind turbines of 1MW and more. Recent results [@IWES2012] show that the availability of offshore installations is much lower, in the order of $70-90\%$. This makes the development of procedures for early warning, identification and reparation of damages essential, see Ref. [@Ciang2008] for an overview of various detection methods. In particular, methods for the early prediction of emerging failures are very valuable in order to react optimally. There are strong indications that the turbulent nature of wind flows affects the mechanical fatigue of wind turbines [@Muecke2011]. This underlines the necessity to include the fast dynamics of turbulence in wind energy methods in order to further improve the reliability of wind energy systems. In this paper, we propose an alternative approach that describes and remodels the conversion process of a wind farm at a frequency of $1$ Hz. Our analysis is performed on a wind farm that is described in section \[sec:wf\]. We present a stochastic model for the conversion process in section \[sec:model\], where the dynamics are intuitively characterized through a set of stochastic estimates. We show in section \[sec:monitor\] that these estimates are highly reactive to dynamical changes, and we promote them as well for condition monitoring of the power performance for a wind farm. Also, a probabilistic description of the power production is presented in appendix \[sec:proba\]. Wind farm dynamics {#sec:wf} ================== Data description {#sec:data} ---------------- All results presented in this paper were extracted from measurement data for one wind farm. The wind farm is installed onshore over an area covering roughly $4km^2$, and is surrounded by flat rural terrain. It consists of 12 identical variable-speed, pitch-regulated wind turbines. The rated power of each turbine is in the order of $2$MW. Its exact value cannot be published following an agreement with the farm manager. For this reason, all values of power output will be normalized to the rated power of the entire wind farm $P_r = 100\% \simeq 12 \times 2$MW. Measurements were conducted synchronously at each wind turbine. The measured signals are the net electrical power output generated by each wind turbine, and the wind speed and direction measured on each nacelle by a cup anemometer and a wind vane. The operational status of each wind turbine was also provided, so that the data was rejected when one or more turbines were not operating in normal conditions. All measurements were performed at a sampling frequency $f_s=1$Hz. Unless stated otherwise, all the time series presented in this paper have a sampling frequency of $1$Hz. The measurement campaign was conducted over a period of eight months, from June 2009 till February 2010. The measurements were regularly interrupted, eventually leaving $7,775$ ten-minute periods ($4,665,000$ samples, or $53$ days $23$ hours and $50$ minutes). Wind farm observables {#sec:observables} --------------------- This paper aims towards wind farm dynamics, and not dynamics of single wind turbines. In order to describe the entire wind farm, some new observables were defined from the data measured on all single wind turbines. Ref. [@Cutler2011] observes that the average of all single wind speed and direction measurements is a more precise measure for power curve modeling than an eventual met mast measurement. Following this observation, we define three observables for the wind farm. The average wind speed over the entire wind farm is defined as $$\begin{aligned} u(t) =\frac{1}{N} \sum_{i=1}^N u_i(t) \, , \label{eq:u}\end{aligned}$$ where $u_i(t)$ is the wind speed measured at $1$Hz by the cup anemometer on the nacelle of turbine $i$. Similarly, the average wind direction over the entire wind farm is defined following $$\begin{aligned} \phi(t) =\frac{1}{N} \sum_{i=1}^N \phi_i(t) \, , \label{eq:phi}\end{aligned}$$ where $\phi_i(t)$ is the wind direction measured at $1$Hz by the wind vane on the nacelle of turbine $i$. Finally, the total power output fed by the wind farm into the electric grid is $$\begin{aligned} P(t) = \sum_{i=1}^N P_i(t) \, , \label{eq:P}\end{aligned}$$ where $P_i(t)$ is the net electrical power output of turbine $i$ at $1$Hz. $N$ represents the total number of turbine measurements considered, in our case $N=12$ corresponding to all the turbines in the wind farm. The three signals $\{u(t),\phi(t),P(t)\}$ are sampled at $1$Hz. From now on, these three effective observables for the wind farm are named the [*wind speed $u(t)$*]{}, the [*wind direction $\phi(t)$*]{} and the [*power output $P(t)$*]{}. Ten-minute dynamics {#sec:dyn10min} ------------------- While our analysis focuses on 1-Hz dynamics, we first analyze the wind farm dynamics on the basis of ten-minute averaged results. The ten-minute analysis gives a coarse picture of the wind farm dynamics. Within each ten-minute time period, the average wind speed $\bar{u}$ is calculated from the 600 samples $u(t)$ following $$\begin{aligned} \bar{u}=\frac{1}{600} \sum_{t=1}^{600} u(t) \, . \label{eq:u10min}\end{aligned}$$ Similarly, the ten-minute average wind direction $\bar{\phi}$ and power output $\bar{P}$ are calculated from $\phi(t)$ and $P(t)$. Histograms are presented in figures \[fig:um\_hist\] and \[fig:pm\_hist\] for $\bar{u}$ and $\bar{P}$. From the $7,775$ ten-minute periods measured, we can observe the Weibull-like shape of the wind speed histogram, as usually observed for atmospheric wind measurements. This is an indication that averaging over all turbine anemometers following equation (\[eq:u\]) conserves the wind speed histogram. The histogram of power output shows that power values below $20\%$ are most probable, and that the probability of measuring a higher power value decays rapidly. However, it is interesting to note that power values up to $105\%$ of the farm rated power are measured, in which case all turbines deliver slightly above their rated power. ![Histogram of ten-minute average power output $\bar{P}$ with a resolution $\Delta P=5$ %.[]{data-label="fig:pm_hist"}](figures/um_hist.pdf){width="0.8\linewidth"} ![Histogram of ten-minute average power output $\bar{P}$ with a resolution $\Delta P=5$ %.[]{data-label="fig:pm_hist"}](figures/pm_hist.pdf){width="0.8\linewidth"} A histogram of wind directions $\bar{\phi}$ is presented in figure \[fig:am\_hist\] for 12 wind sectors of size $\Delta\bar{\phi}=30^{\circ}$. The most frequent wind directions lie within the South-West sector, as expected from the wind farm location. From now on, the data will be systematically separated for the 12 wind sectors $k$ such that $(k-1) \cdot 30^{\circ} \leq \bar{\phi} < k \cdot 30^{\circ}$. We define the IEC-like power curve for the wind farm by adapting the norm IEC 61400-12-1 [@IEC-12] originally designed for a single wind turbine [^3]. For each wind sector $k$ of size $\Delta \bar{\phi}=30^{\circ}$, we sort the ten-minute averages of wind speed $\bar{u}$ and power output $\bar{P}$. We then average the ten-minute averages in each wind speed bin $j$ of size $\Delta \bar{u}=0.5$ m/s. This yields an IEC power curve $P_{IEC}(\bar{u}_j,\bar{\phi}_k)$ for each wind sector, as presented in figure \[fig:iec\_sectors\]. The power curves look similar to the typical IEC power curve of a wind turbine [@Burton2001]. A cut-in wind speed is found at around $4$ m/s, and the rated farm power is reached at roughly $13$ m/s, as specified for the single wind turbines in the farm. We observe some deviations in power performance of up to $10\%$ depending on the wind direction. This is due to the various wake effects (reduced wind speed and power downstream of wind turbines) that change with the inflow direction. This implies that the global performance of the farm depends on the inflow direction, and modeling or monitoring applications must be direction-dependent. ![IEC-like power curve $P_{IEC}(\bar{u}_j,\bar{\phi}_k)$ with corresponding error bars for the wind sectors $k\in \{1,12\}$.[]{data-label="fig:iec_sectors"}](figures/am_hist.pdf){width="0.8\linewidth"} ![IEC-like power curve $P_{IEC}(\bar{u}_j,\bar{\phi}_k)$ with corresponding error bars for the wind sectors $k\in \{1,12\}$.[]{data-label="fig:iec_sectors"}](figures/iec_sectors.pdf){width="0.8\linewidth"} 1Hz dynamics {#sec:dyn1Hz} ------------ We focus in this paper on wind farm dynamics at $1$Hz. This focus is motivated by observations of measurement data, as presented in figure \[fig:series8b\]. The effective wind speed $u(t)$ and power output $P(t)$ display fast fluctuations at $1$Hz. While the ten-minute averages $\bar{u}$ and $\bar{P}$ give a rough estimate, the actual dynamics of the wind farm happen on a faster time scale. The wind farm filters the fastest wind fluctuations in time scales of few seconds due to its inertia. However, wind speed changes happening within seconds and minutes are converted into power output changes. The high frequency data also shows wind gusts that are underestimated by ten-minute averages. It is the case in figure \[fig:series8b\] e.g. at time $100$ min $<$ $t$ $<$ $110$ min, where the wind speed $u(t)$ drops by $5$m/s (while $\bar{u}$ only drops by $2$m/s). In the same time, the power output $P(t)$ drops by $60\%$ (while $\bar{P}$ only drops by $30\%$). This indicates first that rapid and large changes are measured for the wind speed and power output, and second that ten-minutes averages obviously underestimate the amplitude of these rapid changes. The $\{u,P\}$ space is common for power curve methods because it is the phase space representing the input/output variables $u/P$ for a wind turbine. We use this representation for the wind farm in figure \[fig:uP\_1Hz\]. The fast 1Hz data fluctuates around the IEC power curve with deviations in power output of up to $\pm 20\%$. This shows that large errors up to $20\%$ can be generated by using an IEC power curve for power estimation / modeling. The 2h sample shown in figure \[fig:series8b\] is overlaid in figure \[fig:uP\_1Hz\], illustrating the volatility of the conversion process. We also observe that the conversion process $u(t)\to P(t)$ is a highly dynamical process that portrays well how the wind farm functions. We promote the two variables $u(t)$ / $P(t)$ as input / output for the wind farm, and present a stochastic model for the conversion process $(u(t),\bar{\phi})\to P(t)$ in the next section. It should be noted that from now on, the analysis is systematically conditioned on the wind sector $k$ defined by the ten-minute average direction $\bar{\phi}$ [^4]. ![Measured data $\{u(t);P(t)\}$ (black trajectory) at a sampling frequency of $1$Hz for sector $k=8$. The 2-hour excerpt from figure \[fig:series8b\] is overlaid (gray trajectory). The IEC power curve of figure \[fig:iec\_sectors\] is displayed for this sector (red curve with error bars).[]{data-label="fig:uP_1Hz"}](figures/series8d.pdf){width="0.8\linewidth"} ![Measured data $\{u(t);P(t)\}$ (black trajectory) at a sampling frequency of $1$Hz for sector $k=8$. The 2-hour excerpt from figure \[fig:series8b\] is overlaid (gray trajectory). The IEC power curve of figure \[fig:iec\_sectors\] is displayed for this sector (red curve with error bars).[]{data-label="fig:uP_1Hz"}](figures/uP_1Hz.pdf){width="0.8\linewidth"} Modeling the conversion process of a wind farm {#sec:model} ============================================== We propose in this section a stochastic approach to model the conversion process of the wind farm $(u(t),\bar{\phi})\to P(t)$. We extend a stochastic model originally developed for wind turbines [@Milan2011a] to an entire wind farm, based on observations in section \[sec:dyn1Hz\]. This method converts the wind speed / direction signals into a stochastic signal of power output, after a proper parametrization during a training period. Three variations of the stochastic method are presented, see table \[table:models\]. Model 1 (estimated) generates a power output signal $P_1(t)$ at 1Hz using a wind speed signal $u(t)$ at 1Hz. The parameters of model 1 are estimated directly during the training period, see subsection \[sec:model1\]. Model 2 (parametric) functions similarly to model 1 except that the parameters are no longer those estimated from the training period but a parametric version of them, see subsection \[sec:model2\]. Model 3 (parametric 10-min) extends model 2 by including a pre-model that first models a 1Hz wind speed signal from ten-minute wind data, then follows instructions from model 2, see subsection \[sec:model3\]. Model 3 can be used when only ten-minute wind data is available. Additionally, a fourth model based on the IEC power curve is used to compare to the stochastic models 1-3. All power output signals (one measured and four modeled) are compared in subsection \[sec:model\_results\]. -------------------------------------- --------------------------- ------------------------------ ---------------------- **OUTPUT** **wind speed** **wind direction** **power output** **reference (measurement)** measured $u(t)$ measured 10-min $\bar{\phi}$ measured $P(t)$ **model 1 (estimated)** measured $u(t)$ measured 10-min $\bar{\phi}$ modeled $P_1(t)$ **model 2 (parametric)** measured $u(t)$ measured 10-min $\bar{\phi}$ modeled $P_2(t)$ **model 3 (parametric+10-min wind)** measured 10-min $\bar{u}$ measured 10-min $\bar{\phi}$ modeled $P_3(t)$ **IEC power curve** measured $u(t)$ measured 10-min $\bar{\phi}$ modeled $P_{IEC}(t)$ -------------------------------------- --------------------------- ------------------------------ ---------------------- : Overview of the different models presented. Each model converts the wind speed and wind direction input signals into an output power signal.[]{data-label="table:models"} Stochastic estimation of the conversion process {#sec:model1} ----------------------------------------------- In this subsection, the model parameters are estimated directly from data measured during the training period. This approach is the direct extension of the wind turbine model presented in [@Milan2011a]. We observed in figure \[fig:uP\_1Hz\] that the 1Hz trajectory in $\{u,P\}$ space fluctuates around the IEC power curve $P_{IEC}(\bar{u},\bar{\phi})$. Simply put, turbulent wind fluctuations drive the farm dynamics away from the power curve, that represents the average behavior aimed by the wind farm. Ref. [@Anahua2007; @Gottschall2008] recognized this property for single wind turbines and proposed to describe these dynamics with a relaxation model towards the power curve. The power curve is simply seen as an attractor, around which the system fluctuates. Such dynamics can be reproduced by a stochastic drift / diffusion model. In this paradigm, the attractor towards some hypothetical power curve $P(u)$ is represented by a drift coefficient $D^{(1)}$, while a diffusion coefficient $D^{(2)}$ models additional fluctuations. These two coefficients (also called Kramers-Moyal coefficients in stochastic theory) can be estimated directly from measured data [@Risken1996; @Friedrich2011] collected during the training period. They describe the dynamics of the power output $P(t)$ conditioned on the 1Hz wind speed $u(t)$ and the ten-minute average wind direction $\bar{\phi}$. The system evolves in a three-dimensional space with variables $P(t)$, $u(t)$ and $\bar{\phi}$. The drift and diffusion coefficients must then be defined as the three-dimensional arrays $$\begin{aligned} D^{(n)}(P(t),u(t),\bar{\phi}) &=& \frac{1}{n!} \lim_{\tau \to 0} \frac{1}{\tau}M^{(n)}(P(t),u(t),\bar{\phi},\tau) \\ &=& \frac{1}{n!} \frac{\partial M^{(n)}(P(t),u(t),\bar{\phi},\tau)}{\partial \tau}\Bigg|_{\tau=0} \, , \label{eq:kmc}\end{aligned}$$ with $n=1$ for the drift coefficient and $n=2$ for the diffusion coefficient. As shown in Ref. [@Boettcher2006], they are the partial derivatives with respect to the time increment $\tau$ of the conditional moments[^5] $$\begin{aligned} M^{(n)}(P(t),u(t),\bar{\phi},\tau) &=& \int\limits_{-\infty}^{\infty} \big[\,P(t+\tau)-P(t) \,\big]^n \,\,\, f(P(t+\tau)|P(t),u(t),\bar{\phi}) \,\,\, dP(t+\tau) \\ &=& \Big \langle \,\big[\,P(t+\tau)-P(t) \,\big]^n\,\, \Big|\,\,P(t) \, ,\, u(t),\, \bar{\phi}\, \Big \rangle \, , \label{eq:cond_mom}\end{aligned}$$ where the operator $\langle A|B \rangle$ represents the conditional mean of $A$ for condition $B$. The conditional moment $M^{(n)}(P(t),u(t),\bar{\phi},\tau)$ is the $n$-th moment of the conditional probability $f(P(t+\tau)|P(t),u(t),\bar{\phi})$ of the power increment $P(t+\tau)-P(t)$ conditioned on the values of $P(t)$, $u(t)$ and $\bar{\phi}$. It is not possible to calculate the derivative in equation (\[eq:kmc\]) exactly from measured data. The time increment $\tau$ is limited to the smallest value min$(\tau)=1/f_s$, where $f_s$ is the sampling frequency. In our case, min$(\tau)=1$s, and the limit $\tau \to 0$ cannot be performed. However, a Taylor expansion of the conditional moments gives in first-order approximation $$\begin{aligned} M^{(n)}(*,\tau) &=& M^{(n)}(*,\tau=0) + \tau \cdot \frac{\partial M^{(n)}(*,\tau)}{\partial \tau}\Bigg|_{\tau=0} + o(\tau^2)\\ &=& \tau \cdot \frac{\partial M^{(n)}(*,\tau)}{\partial \tau}\Bigg|_{\tau=0} + o(\tau^2)\\ &=& n! \tau \cdot D^{(n)}(*) + o(\tau^2) \, , \label{eq:cond_mom_taylor}\end{aligned}$$ where $*$ stands for the variables $P(t)$, $u(t)$ and $\bar{\phi}$. The relation $M^{(n)}(*,\tau=0)=0$ is obvious from equation (\[eq:cond\_mom\]). For small non-zero $\tau$ values, the derivative in equation (\[eq:kmc\]) becomes to a first-order approximation $$\begin{aligned} D^{(n)}(P(t),u(t),\bar{\phi}) \simeq \frac{M^{(n)}(P(t),u(t),\bar{\phi},\tau)}{n! \tau} \, . \label{eq:kmc2}\end{aligned}$$ Additionally, the discrete nature of measured data has further implications on the applicability of theories developed for continuous stochastic processes. It was observed by [@Sura2002; @Gottschall2008b] that when estimating the coefficients from a dataset that is not sampled with a high enough sampling frequency, an artificial parabolic term is added to the diffusion coefficient. A simple ansatz exists in order to remove this artifact from $D^{(2)}$, where equation (\[eq:kmc2\]) is used with a second conditional moment $M^{(2)}$ modified as [@Friedrich2011] $$\begin{aligned} M^{(2)}(P(t),u(t),\bar{\phi},\tau) = \Big \langle \,\big[\,P(t+\tau)-P(t) - \tau \, D^{(1)}(P(t),u(t),\bar{\phi}) \,\big]^2 \,\, \Big|\,\,P(t) \, ,\, u(t),\, \bar{\phi}\, \Big \rangle \, . \label{eq:cond_mom_finite_sampling}\end{aligned}$$ This approach is used systematically in this paper in order to minimize finite sampling artifacts. Concretely speaking, the three measurement signals $P(t)$, $u(t)$ and $\bar{\phi}$ are collected during the training period and sorted in the three-dimensional phase space $\{P,u,\bar{\phi}\}$. Each axis, e.g. the P-axis is split into $N_P$ equidistant intervals of size $\Delta P$. The phase space is then cut into $N_P \times N_u \times N_{\bar{\phi}}$ bins (small cubes) of size $\{\Delta P,\Delta u,\Delta \bar{\phi}\}$. Let us consider one bin $(i,j,k)$ which contains all the data samples that satisfy $P(t) \in (P_i \pm \Delta P/2$), $u(t) \in (u_j \pm \Delta u/2$) and $\bar{\phi} \in (\bar{\phi}_k \pm \Delta \bar{\phi}/2$). From all these data samples, the conditional moments $M^{(n)}(P_i,u_j,\bar{\phi}_k,\tau)$ are calculated following equation (\[eq:cond\_mom\]). Then the coefficients $D^{(n)}(P_i,u_j,\bar{\phi}_k)$ can be estimated from the conditional moments following equation (\[eq:kmc2\]). This operation is then repeated for each of the $N_P \times N_u \times N_{\bar{\phi}}$ bins, such that the drift and diffusion coefficients can be estimated in the entire three-dimensional phase space [^6]. Here we choose $\Delta u=0.5$ m/s, $\Delta \bar{\phi}=30^{o}$ and $\Delta P=P_r/50=2\%$. The binning procedure allows to characterize the dynamics locally in the phase space. The information is averaged following equation (\[eq:cond\_mom\]) in each local bin (rather than in time like the IEC procedure). This means that in each bin of the power-speed-direction space, the power dynamics are described by the drift and diffusion coefficients. In bin $(i,j,k)$, the drift coefficient $D^{(1)}(P_i,u_j,\bar{\phi}_k)$ represents the first moment, i.e. the mean value of the power change $\frac{dP(t)}{dt}=\lim_{\tau \to 0} \frac{P(t+\tau)-P(t)}{\tau}$ calculated from all data samples in the bin. Similarly, the diffusion coefficient $D^{(2)}(P_i,u_j,\bar{\phi}_k)$ represents the variance of the power change. In summary, $D^{(1)}$ and $D^{(2)}$ represent the mean and variance of the power change $\frac{dP(t)}{dt}$. They quantify how much the power changes on average (drift), and how much deviation is expected from that average change (diffusion). Although not mathematically exact, one could conceptualize the coefficients as $$\begin{aligned} D^{(1)}(P_i,u_j,\bar{\phi}_k) &\sim& \Bigg \langle \frac{dP(t)}{dt} \Bigg|\,\,P_i,u_j,\bar{\phi}_k\, \Bigg \rangle \, , \\ D^{(2)}(P_i,u_j,\bar{\phi}_k) &\sim& \Bigg \langle \Bigg(\frac{dP(t)}{dt}\Bigg)^2 dt \Bigg|\,\,P_i,u_j,\bar{\phi}_k\, \Bigg \rangle \, . \label{eq:kmc3}\end{aligned}$$ The drift coefficient is presented in figure \[fig:drift\]. In bin $(i,j,k)$, if $D^{(1)}>0$ the drift is represented by a blue arrow pointing up because a positive drift means a power increase. Similarly, if $D^{(1)}<0$ the drift is represented by a red arrow pointing down describing a power decrease. The dynamics of the wind farm appear clearly. In all bins located below the power curve, the drift is positive and the power increases (on average). Analogously, in all bins above the power curve, the power decreases as the drift coefficient is negative. A clear drift towards the power curve appears. The drift coefficient is a map of the conversion process, that extends the information of the ten-minute averaged IEC power curve to 1Hz dynamics. A trajectory is overlaid to the drift coefficient in figure \[fig:drift\] to illustrate the concept of a drift towards the attractive power curve. The drift coefficient is shown in figure \[fig:drift\_potential\](a) for one wind speed / direction bin and confirms what was observed in figure \[fig:drift\]. ![(a) drift coefficient $D^{(1)}(P,u_j;\bar{\phi}_k)$; (b) drift potential $\Psi(P,u_j;\bar{\phi}_k)$ for wind speed bin $j=20$ ($u\simeq 9.75$m/s) and sector $k=8$ ($\bar{\phi}\simeq 225^\circ$) (gray line with error bars in case of the drift coefficient). Symbols are superimposed for positive drift (blue arrows), negative drift (red arrows) and zero drift values (black dot). The zero drift and minimum potential values are indicated (horizontal dashed line), corresponding to the value of the Langevin power curve $P_L(u_j;\bar{\phi}_k)$.[]{data-label="fig:drift_potential"}](figures/drift_iec_path.pdf){width="\columnwidth"} ![(a) drift coefficient $D^{(1)}(P,u_j;\bar{\phi}_k)$; (b) drift potential $\Psi(P,u_j;\bar{\phi}_k)$ for wind speed bin $j=20$ ($u\simeq 9.75$m/s) and sector $k=8$ ($\bar{\phi}\simeq 225^\circ$) (gray line with error bars in case of the drift coefficient). Symbols are superimposed for positive drift (blue arrows), negative drift (red arrows) and zero drift values (black dot). The zero drift and minimum potential values are indicated (horizontal dashed line), corresponding to the value of the Langevin power curve $P_L(u_j;\bar{\phi}_k)$.[]{data-label="fig:drift_potential"}](figures/drift_potential_example.pdf){width="\columnwidth"} One can also define the potential of the drift coefficient as $$\begin{aligned} \Psi(P,u_j,\bar{\phi}_k) = - \int_{-\infty}^P D^{(1)}(P',u_j,\bar{\phi}_k) \, dP' \, . \label{eq:psi}\end{aligned}$$ The potential is a more intuitive representation of the dynamics, see figure \[fig:drift\_potential\](b). The system tends to minimize its potential by drifting towards an attractive fixed point (that corresponds to a local minimum of the potential). Around the attractive fixed point, the potential climbs to indicate the repulsion of the system to move away from the power curve. However, the system keeps following the non-stationary wind condition (changing $j$, $k$ bins) and always drifts towards a new fixed point. Beyond the drift coefficient, the power values towards which the system drifts are of interest. We define this set of attractive fixed points as a [*Langevin power curve*]{} $P_L(u_j,\bar{\phi}_k)$ following $$\begin{aligned} P_L(u_j,\bar{\phi}_k)\Leftrightarrow \begin{cases} D^{(1)}(P_L(*),*)=0 \\ \Big( \frac{\partial D^{(1)}(P,*)}{\partial P} \Big)_{P=P_L(*)} <0 \end{cases} \label{eq:lpc}\end{aligned}$$ where $*$ stands for $(u_j,\bar{\phi}_k)$. At a fixed point, the drift coefficient is zero. For the fixed point to be attractive, the drift must be decreasing (negative derivative with respect to P). On the contrary, a fixed point with increasing drift would be repulsive. The system drifts towards the Langevin power curve $P_L(u_j,\bar{\phi}_k)$, that represents the basin of attraction for the wind speed $u_j$ and wind sector $\bar{\phi}_k$. The Langevin and IEC power curves have similar values for the data set measured here, see figure \[fig:drift\]. The dynamics can be seen as a random-like trajectory circulating around the power curve. This is the signature of a process that contains both deterministic and stochastic dynamics. The true reason for the random nature of the process is not obvious, as wind turbines have deterministic control strategies. At this point, it is important to note that the power dynamics are driven by the wind speed dynamics. In an idealized case with a spatially homogeneous, laminar wind inflow, the system should react in a deterministic manner. If this spatially homogeneous flow would see its speed change everywhere, the drift field should describe the corresponding change in power output satisfyingly. However, this ideal case is not realistic because the wind inflow is turbulent, and contains spatial fluctuations. The wind farm does not react to the effective wind speed $u(t)$ averaged over the whole area, but to the local wind fluctuations acting on the local blade elements of all the rotors, so to say the wind field $\vec{U}(\vec{x},t)$. The effective wind speed $u(t)$ is a macroscopic estimate of the many microscopic influences driving the wind farm (analogously to temperature being a macroscopic estimate of thermodynamic fluctuations). When projecting the system dynamics over this macroscopic variable, the microscopic degrees of freedom become stochastic fluctuations. A macroscopic variable driven by many microscopic interactions can often be approximated by a low-dimensional Langevin process [^7]. We can write a Langevin equation that describes the time evolution of the power output $P(t)$ following $$\begin{aligned} \frac{dP(t)}{dt}=D^{(1)}(P(t),u(t),\bar{\phi})+\sqrt{D^{(2)}(P(t),u(t),\bar{\phi})}\cdot \Gamma(t) \, , \label{eq:langevin}\end{aligned}$$ where $\Gamma(t)$ is a Gaussian-distributed noise that is uncorrelated, i.e. $\langle \Gamma(t) \Gamma(t') \rangle=2\delta(t-t')$ and with a mean value $\langle \Gamma(t) \rangle=0$. The Langevin equation can be integrated numerically in the It[ô]{} sense [@Risken1996] $$\begin{aligned} P(t+\Delta t)=P(t)+\Delta t \cdot D^{(1)}(P(t),u(t),\bar{\phi})+\sqrt{\Delta t \cdot D^{(2)}(P(t),u(t),\bar{\phi})}\cdot \eta(t) \, \label{eq:langevin_euler}\end{aligned}$$ where $\eta(t)$ is a Gaussian-distributed random variable with mean value $\langle \eta(t) \rangle=0$ and variance $\langle \eta(t)^2 \rangle=2$. [^8] The assumption of Gaussianity can be quantified by higher-order Kramers-Moyal coefficients being zero, i.e. $D^{(n)}=0$ for $n>2$ following Pawula theorem [@Risken1996]. The structure of the Langevin equation indicates that the power output $P(t)$ drifts towards the power curve following $D^{(1)}$, but stochastic fluctuations are superimposed using the stochastic noise $\Gamma(t)$ amplified by the diffusion coefficient $D^{(2)}$. This way the wind farm dynamics captured by the drift / diffusion coefficients can be remodeled by solving the Langevin equation. Given any new wind speed signal $u(t)$ and wind direction $\bar{\phi}$ as input, equation (\[eq:langevin\_euler\]) generates a synthetic power output signal $P_1(t)$. This serves as a model for the conversion process ($u(t),\bar{\phi}) \to P_1(t)$. Some results are presented in subsection \[sec:model\_results\] in comparison with other models. Parametric form of the stochastic model {#sec:model2} --------------------------------------- The drift and diffusion coefficients $D^{(n)}(P_i,u_j;\bar{\phi}_k)$ estimated from equation (\[eq:kmc2\]) describe the dynamics of the conversion process. A parametric simplification of the coefficients is presented here in an effort to reduce the number of model parameters, see figure \[fig:KMC\_parametric\]. We observe that for a given wind speed bin $u_j$ and direction bin $\bar{\phi}_k$, the drift coefficient $D^{(1)}(P,u_j,\bar{\phi}_k)$ is a tolerably linear function of $P$. The stable fixed points are given by the Langevin power curve $P_{L}(u_j,\bar{\phi}_k)$,[^9] so that the drift coefficient can be simplified as $$\begin{aligned} D^{(1)}(P,u_j,\bar{\phi}_k) \simeq \alpha_{jk} \cdot \big( P - P_{L}(u_j,\bar{\phi}_k) \big) \, , \label{eq:D1_parametric}\end{aligned}$$ where $\alpha_{jk}$ is its slope [^10]. The drift slope $\alpha_{jk}$ is a direct measure of how fast the wind farm can follow wind fluctuations. A more negative slope (steeper drift) corresponds to a faster reaction of the wind farm [^11]. Similarly for the diffusion term $D^{(2)}(P,u_j,\bar{\phi}_k)$, some parametrization can be performed. As illustrated in figure \[fig:KMC\_parametric\], the diffusion coefficient is almost constant, such that $$\begin{aligned} D^{(2)}(P,u_j,\bar{\phi}_k) \simeq \beta_{jk} = \frac{1}{N_P} \sum_{i=1}^{N_P} D^{(2)}(P_i,u_j,\bar{\phi}_k) \, , \label{eq:D2_parametric}\end{aligned}$$ where the constant $\beta_{jk}$ is the diffusion coefficient averaged over the $N_P$ power bins. The mean diffusion term $\beta_{jk}$ quantifies the random nature of the power fluctuations. Higher values indicate stronger power fluctuations [^12]. The estimation of $\beta_{jk}$ is further optimized following appendix \[sec:opt\].\ ![(upper) drift coefficient $D^{(1)}(P,u_j,\bar{\phi}_k)$ estimated (symbols and dashed line) for various wind speed bins $j$ (various colors) and sector $k=8$ ($\bar{\phi}\simeq 225^\circ$). The power curve value $P_{IEC}(u_j,\bar{\phi}_k)$ is indicated (colored symbol $+$). The linear fit proposed in equation (\[eq:D1\_parametric\]) is also drawn (bold colored line); (lower) similarly for the diffusion coefficient $D^{(2)}(P,u_j,\bar{\phi}_k)$. The mean value $\beta_{jk}$ of the diffusion coefficient is represented (colored bold line) following equation (\[eq:D2\_parametric\]). []{data-label="fig:KMC_parametric"}](figures/KMC_parametric.pdf){width="\columnwidth"} All the values of $\alpha_{jk}$ and $\beta_{jk}$ are presented in figure \[fig:KMC\_parameters\]. Both terms depend only weakly on the wind direction, especially at lower wind speeds where wake effects might be less pronounced. $\alpha_{jk}$ grows rather cubically, i.e. $\alpha_{jk}\propto u^3$. This indicates that the reaction time of the wind farm (the inverse of $\alpha$) drops cubically with $u$, stressing that the wind farm changes power values much faster at large wind speeds. the diffusion mean $\beta_{jk}\propto u^6$ to counteract the growing drift. At large wind speeds $u>13$m/s, the diffusion term drops because the pitch mechanism of the turbines regulates the power output that fluctuates very little. It is important to note that a larger diffusion term stresses the greater need for a stochastic model. At large wind speeds, the stochastic model becomes better-suited than a simpler deterministic model because the diffusion term is large. This illustrates the flexibility of a drift/diffusion model, where the drift quantifies the reaction time (inertia) of the system, and the diffusion/drift ratio controls the strength of the fluctuations. ![(left) slope of the drift coefficient $\alpha_{jk}$ and; (right) mean of the diffusion coefficient $\beta_{jk}$ for all direction bins $k$, as a function of the wind speed (bin j).[]{data-label="fig:KMC_parameters"}](figures/KMC_parameters.pdf){width="0.7\columnwidth"} The parametrization greatly reduces the number of model parameters. While the two estimated coefficients contain each $N_P \times N_u \times N_{\bar{\phi}}$ values, their two parametric forms only requires $N_u \times N_{\bar{\phi}}$ values each. This parametrization also gives more insight on the dynamics of the wind farm. The coefficients simplified with $\alpha_{jk}$ and $\beta_{jk}$ are used within the Langevin equation (\[eq:langevin\]) to model the conversion process $(u(t),\bar{\phi}) \to P_2(t)$. Some results are presented in subsection \[sec:model\_results\] in comparison with other models. In addition, a further simplification of the model could be achieved given the fact that $P_L\propto u^3$ for $P_L<P_r$. Then $\alpha_{jk}\simeq \alpha_{0} P_L$ and $\beta_{jk}\simeq \beta_{0} P_L^2$, where the parameters $\alpha_0$ and $\beta_0$ depend weakly on the wind direction. If this weak direction dependence is dropped, the Langevin model can be further simplified to the two-parameter form $$\begin{aligned} \frac{dP(t)}{dt}=\alpha_0 \, P_L\big(u(t)\big) \cdot \Big(P(t)-P_L(u(t))\Big)+\sqrt{\beta_0} \, P_L\big(u(t)\big) \cdot \Gamma(t) \, . \label{eq:langevin_norm}\end{aligned}$$ The two parameters are $\alpha_0\simeq -(6.48\pm 0.25)\times 10^{-4}\, \%^{-1}s^{-1}$ and $\beta_0\simeq (7.42\pm 0.21) \times 10^{-5}\, s^{-1}$ for the wind farm. The question of the dependence of these two parameters on the nature and size of the wind farm arises, and would be of interest for further studies. One should note that for better accuracy, $\beta_0$ should not stay constant but be reduced for $u>13$m/s to reproduce the effect of pitching. Results of the simplified model are not presented here. Pre-modeling of turbulent wind fluctuations {#sec:model3} ------------------------------------------- Both models 1 and 2 replicate the conversion process $(u(t),\bar{\phi}) \to P(t)$. A practical application consists in predicting the power production when the wind condition is known. However, models 1 and 2 require a wind speed signal $u(t)$ sampled at around 1Hz and a ten-minute mean wind direction $\bar{\phi}$. Yet practical applications summarize the wind condition to ten-minute (or even one-hour) averages, and the 1Hz information is not known. In order to overcome such limitation, one must be able to model the wind dynamics at 1Hz realistically from the ten-minute (or even slower) wind data. Such a model is presented in this subsection. We developed a stochastic model that generates high-frequency fluctuations of wind speed [^13]. The model superimposes high-frequency fluctuations to the ten-minute average, following the description of atmospheric wind speed data from [@Boettcher2007; @Morales2010a]. Simply put, the wind speed signal is decomposed as $$\begin{aligned} u(t) &=& \bar{u} + \sigma \cdot u'(t) \, \\ &=& \bar{u} \, \big( \,1 + TI \cdot u'(t) \,\big) \, , \label{eq:udn}\end{aligned}$$ where $\bar{u}$ and $\sigma$ are respectively the ten-minute mean value and standard deviation of the wind speed $u(t)$, and $TI$ is the turbulence intensity $TI=\sigma / \bar{u}$. In this description, the slow wind dynamics are described by $\bar{u}$ and $TI$ (that change every ten minutes), and the fast turbulent fluctuations are given by $u'(t)$ [^14]. We observed that the fluctuations $u'(t)$ extracted from the measured wind speed $u(t)$ have a distribution reasonably close to the normal distribution [^15]. The autocorrelation function of $u'(t)$ is a rapidly decaying, exponential-like function $$\begin{aligned} R_{u'u'}(\tau)=\big\langle u'(t+\tau)u'(t)\big\rangle \simeq exp(-\gamma \tau). \label{eq:udn_acf}\end{aligned}$$ Such exponentially-correlated, Gaussian process is similar to the so-called [*Ornstein-Uhlenbeck process*]{} [@Risken1996] that is a special class of Langevin processes. The wind fluctuation signal $u'(t)$ can be modeled as an Ornstein-Uhlenbeck process following $$\begin{aligned} \frac{du'(t)}{dt}=-\gamma u'(t) + \sqrt{\gamma}\cdot \Gamma(t) \, , \label{eq:udn_OUP}\end{aligned}$$ where $\Gamma(t)$ is a Langevin noise, that was already introduced in equation (\[eq:langevin\]). Equation (\[eq:udn\_OUP\]) can be integrated numerically using the Euler scheme, see equation (\[eq:langevin\_euler\]). In summary, we propose a simple model for the fluctuations $u'(t)$ of the wind speed signal $u(t)$ (representing the effective wind speed over the wind farm). One should be aware that turbulence is in general much more complex, as was shown e.g. by n-point statistics [@Stresing2010]. But here we show that this ansatz already works quite well for our particular application. While finer approaches can be developed, this simple method suffices here. $u'(t)$ is modeled by the stochastic equation (\[eq:udn\_OUP\]), where the unique parameter $\gamma$ is extracted from the measured wind speed following equation (\[eq:udn\_acf\]) [^16]. The signal $u'(t)$ is then used in combination with the measured ten-minute mean wind speed $\bar{u}$ and turbulence intensity $TI$ following equation (\[eq:udn\]) to construct a synthetic signal of wind speed $u(t)$ at 1Hz. When this wind speed model is combined with the wind farm model presented in subsection \[sec:model2\], we can model the farm power output $P_3(t)$ at 1Hz knowing only the ten-minute mean wind speed, turbulence intensity and mean wind direction. Some results are presented in subsection \[sec:model\_results\] in comparison with other models. Statistical validation of the stochastic approach {#sec:model_results} ------------------------------------------------- We compare the power output signals modeled to the reference power output $P(t)$ measured. Three variations of our stochastic model are presented in subsections \[sec:model1\], \[sec:model2\] and \[sec:model3\], yielding three power output signals $P_1(t)$, $P_2(t)$ and $P_3(t)$, see table \[table:models\]. A fourth model is tested using the IEC power curve directly to convert at 1Hz the measured wind speed $u(t)$ into a power output $P_{IEC}(t)\equiv P_{IEC}(u(t))$. The five signals are compared in identical conditions over the total period of the measurement campaign (54 days spread over a total period of eight months, during which all turbines operate). Excerpts of the five signals are presented in figure \[fig:model\_series\]. The time period shown was chosen because it covers a wide range of power values (the time series are displayed as a first illustration, whereas the statistical analysis covers the entire measurement period). The power signal measured $P(t)$ shows fluctuations at many scales and complex statistical features such as volatility clustering in the first five hours. Large power changes of about $50\%$ are observed within minutes. These dynamics measured are well reproduced by the first two stochastic models for $P_1(t)$ and $P_2(t)$. The third stochastic model for $P_3(t)$ also follows the slow trend with similar fluctuations, but shows some minor variations due to the fact that only ten-minute information of the wind was used (the wind fluctuations at 1 Hz were modeled, see subsection \[sec:model3\]). The IEC method for $P_{IEC}(t)$ yields mixed results, as the slow trend is reproduced, but fast fluctuations are strongly over-represented. These results can be explained easily: the first two stochastic models reproduce the measured features because they exploit 1Hz wind data and they model the wind farm dynamics, including its inertia; the third stochastic model also models the farm dynamics, but only has access to ten-minute wind data, thus the fastest fluctuations are synthetic; finally, the IEC method does not model the fast farm dynamics but only the long-term behavior, as expected from using an average curve.\ ![ Excerpts of power output signals sampled at 1 Hz over a continuous time period of 10h. The measured signal $P(t)$ is compared to the four modeled signals $P_1(t)$, $P_2(t)$, $P_3(t)$ and $P_{IEC}(t)$ (top to bottom). For each signal the vertical axis spans the range of power values from $0\%$ to $100\%$.[]{data-label="fig:model_series"}](figures/series1.pdf){width="0.6\columnwidth"} The probability density functions (PDFs) $f(P)$ of the power output signals are presented in figure \[fig:model\_hist\]. Except for some minor deviations from the parametric results $P_2(t)$ and $P_3(t)$, all the models span the same values as the measurement with equal probabilities. ![Relative difference of energy production $(E_i(t)-E(t))/E(t)$ for the various models $i=1,2,3,IEC$ in \[$\%$\].[]{data-label="fig:model_energy"}](figures/model_hist.pdf){width="0.8\linewidth"} ![Relative difference of energy production $(E_i(t)-E(t))/E(t)$ for the various models $i=1,2,3,IEC$ in \[$\%$\].[]{data-label="fig:model_energy"}](figures/model_energy_diff.pdf){width="0.8\linewidth"} We present the cumulative energy produced as a function of time $E_i(t)=\int_0^t P_i(t')dt'$ in figure \[fig:model\_energy\] ($E(t)$ is calculated for the measured power $P(t)$). The four models produce some deviation of up to $ 9\%$ in the first 4 days of the simulation, then these deviations drop rapidly over longer integration times. The IEC model produces the best results all along for long-term energy production, and matches mostly perfectly the measurement over the total simulation time of 54 days, with a deviation of only $0.02\%$. It is to be expected that the IEC power curve, being an average curve, can reproduce the total energy (i.e. mean power) of the data set from which it is estimated. It is followed closely by model 1 (estimated), that only deviates by $0.3\%$ of the total energy production over the total simulation time. Models 2 and 3 show a slight deviation of $0.8\%$ and $0.7\%$ from the measurement over the simulation time. While the IEC power curve remains optimal for long-term energy production, minor deviations are found using the estimated model 1. Larger deviations are introduced by the parametrization in models 2 and 3. The stochastic models show satisfactory results, considering that an integration over about 5 million stochastic samples deviates from the measurement at most by $1\%$. This is an important validation that the stochastic models follow the measurement over time without diverging (as tested tens of times for the three stochastic variations). This is due to the nature of the model, and particularly of the drift field that constantly drives the dynamics towards the power curve. Properly reproducing slow dynamics is a necessary (but not sufficient) condition for a model of the high-frequency dynamics. The integral power is a first-oder quantity of the power conversion process. Next we aim to present a deeper statistical analysis of the data. In particular we focus on two-time quantities, namely the power spectrum, the autocorrelation function as well as the probability distribution of power increments and its corresponding structure functions, for further details on the analysis scheme see Ref. [@Morales2010a].\ The power spectral density commonly referred to as [*spectrum*]{} $|\hat{P}(f)|^2$ [^17] is presented in figure \[fig:model\_spec\] for the power signals. Each spectral value $|\hat{P}(f)|^2$ indicates the energy of the corresponding wave of frequency $f$. The spectra can be described reasonably well by a power law $f^{-\beta}$ (linear behavior in log-log scale). The three stochastic models have similar spectra, very close to a power law. They are close to that of the measurement, although the measured spectrum is more complex, with different scaling behaviors in various frequency bands. The IEC power curve model deviates from the measurement, having spectral values up to two orders of magnitude too high; it over-represents the intensity of the fluctuations for $f>2.10^{-4}$Hz, corresponding to fast fluctuations up to roughly 80 minutes. Although not shown here, all spectra match at lower frequencies $f<2.10^{-4}$Hz, indicating that they all follow the trend of the wind dynamics at 80 minutes and longer time scales. ![autocorrelation function $R_{PP}(\tau)$ for the five power signals, displayed in linear-log scale as a function of the time lag $\tau$.[]{data-label="fig:model_acf"}](figures/model_spectrum.pdf){width="0.8\linewidth"} - ![autocorrelation function $R_{PP}(\tau)$ for the five power signals, displayed in linear-log scale as a function of the time lag $\tau$.[]{data-label="fig:model_acf"}](figures/model_acf.pdf){width="0.8\linewidth"} The spectrum is related to the autocorrelation function $R_{PP}(\tau)$ [^18]. The autocorrelation functions of the power signals are presented in figure \[fig:model\_acf\]. The power output measured has long time power-law-like correlations. As an example, a power output signal remains correlated to $97\%$ with itself when shifted by one hour. The three stochastic models match the measurement very well, indicating that they reproduce the time correlations of the measured power. They display power-law correlations close to that of the measurement, as already observed for the power spectral densities. In contrary to the good results of the stochastic models, the IEC power curve model deviates from the measurement and generates a signal that is less correlated. This deviation can be related to results of the power spectral density, where the IEC model was shown to deviate significantly from all the other signals. Particularly, the large drop of correlations at $\tau<10$min corresponds mostly to spectral deviations at $f<2.10^{-3}$Hz, as the low-frequency Fourier modes control the coarse-grained dynamics. Our final statistical test consists in comparing the PDFs of the power increment $P_\tau(t)=P(t+\tau)-P(t)$. The power increment $P_\tau(t)$ is the change (increment) of power output $P(t)$ when waiting a time $\tau$. The PDF $f(P_\tau(t))$ indicates with what probability the power output can change over a time scale $\tau$ [^19]. The PDFs of the power increments are presented in figure \[fig:model\_incpdf\] for four time scales. The power output measured has exponentially-distributed increments (linearly decaying tails in linear-log scale). This is typical for non-Gaussian, intermittent processes that have some quiet time periods alternating with some periods of strong fluctuations, as observed for the time series in figure \[fig:model\_series\]. This notion relates to the intermittent nature of the wind [@Morales2010a] that is converted into power intermittency by the wind farm [@Milan2013a]. At the shortest time scale available $\tau=1$ s, the stochastic models reproduce the increment PDF of the measurement mostly well, with a slight over-representation of large power increments for model 1. Model 2 and 3 match the measurement. The IEC procedure generates much too large power fluctuations at this time scale, with increments up to $\pm 20\%$ within one second, when the measurement can change by at most $\pm 5\%$. At the time scale $\tau=10$ s, the stochastic models under-represent large power increments and the IEC procedure still largely over-represents them. The measured power output can change by $\pm 20\%$ within ten seconds. At the next time scale $\tau=60$ s, the stochastic models match the measurement very well, except for slightly too many extreme events at $P_\tau>30\%$, and the IEC procedure comes closer to the measurement. Note that for these strong fluctuations the probabilities are obtained from 10 and less events per bin, thus error bars larger than $30\%$ are expected. The measured power output can change by $\pm 30\%$ within one minute. At the time scale $\tau =600$ s, the four models match the measurement very well, except for the IEC procedure that still slightly over-represents extreme events. The measured power output can change by $\pm60\%$ within ten minutes. We want to stress that the probabilities also show that more extreme events are expected when longer data sets are considered. Note, a power fluctuation of $5\%$ corresponds to a change of more than $1$MW here. Fluctuations of several MW may effect the local grid stability essentially. The non-Gaussianity of the increment PDFs tells us that even more extreme values become likely over longer time periods. This is important for grid stability and puts additional constraints on the design of back up methods such as energy storage or overload security. Power gustiness is an important feature of wind power production. Rapid power changes affect grid stability, especially on networks that rely strongly on wind energy. When considering the entire dataset, we observe power changes of up to $60\%$ within 3 minutes, and up to $85\%$ within 10 minutes. This stresses the importance of power intermittency on grid stability, and how necessary it is to properly model it. We promote the three stochastic models that generate some power gusts that have a similar amplitude to that of the power gusts measured. These gusts also occur with similar probabilities. Thus our model time series are clearly better than the time series obtained from the IEC procedure, which strongly exaggerates fast power gusts in time scales of seconds. For time scales of ten minutes and more both methods become similar. This confirms that the IEC procedure is valid for slow dynamics, and fluctuates too strongly for fast dynamics within one hour. ![increment PDFs $f(P_\tau(t))$ at various scales $\tau$ for the five power signals, displayed in linear-log scale.[]{data-label="fig:model_incpdf"}](figures/model_incpdf.pdf){width="0.7\columnwidth"} Gustiness and intermittency are related to the scaling of the increments $P_\tau$. We define the structure functions following [@Arneodo1996] $$\begin{aligned} S_q(\tau)=\langle |P_\tau|^q \rangle \sim \tau^{\zeta_q}. \label{eq:strucfunc}\end{aligned}$$ The power law scaling $S_q(\tau) \sim \tau^{\zeta_q}$ is observed for all the power output signals, indicating their fractal nature. Additionally, the scaling exponent $\zeta_q$ indicates the nature of the fractality [^20]. A linear dependence $\zeta_q \propto q$ indicates a monofractal signal with self-similar features along scales, implying that the form of the PDFs $f(P_\tau)$ would not change. In this case only the $\tau$-dependency of the variance would be needed to characterize the statistics. If $\zeta_q$ is instead a nonlinear, concave function of $q$, the process is multifractal with an intermittent, walk-like behavior at small scales and a noise-like behavior at large scales. Consequently we need to characterize $f(P_\tau)$ for all $\tau$. The scaling exponents are presented in figure \[fig:model\_zeta\], compared to Kolmogorov’s 1941 monofractal and 1962 multifractal models for ideal local isotropic turbulence. The measured power lies in between the two models, being intermittent but less than K62. Models 2 and 3 reproduce very well the multifractality of the measurement. Model 1 is too multifractal, probably owing to the (still too) quadratic form of the diffusion coefficient estimated. We observe that model 3 (that uses Gaussian wind fluctuations) is multifractal, showing that power intermittency for the wind farm comes from the conversion process, rather than from the intermittency in the wind. The IEC model is less multifractal, as the small-scale intermittency introduced by the inertia of the wind turbine rotors is not reproduced. ![flatness $F(\tau)$ for the five power signals, displayed in log-linear scale as a function of the time lag $\tau$. $F=3$ indicates a Gaussian process (horizontal dotted line).[]{data-label="fig:model_flatness"}](figures/zeta_err_power.pdf){width="0.8\linewidth"} ![flatness $F(\tau)$ for the five power signals, displayed in log-linear scale as a function of the time lag $\tau$. $F=3$ indicates a Gaussian process (horizontal dotted line).[]{data-label="fig:model_flatness"}](figures/flatness_power.pdf){width="0.8\linewidth"} The flatness (or kurtosis) $F(\tau)=\frac{S_4(\tau)}{(S_2(\tau))^2}$ [@Press2007] quantifies directly the non-Gaussian character of the distribution $f(P_\tau(t))$, see figure \[fig:model\_flatness\]. We measure a high flatness up to a value of $12$ at short time scales in seconds confirming the strong intermittency at short time scales. This value is reduced by a factor two within two hours, and reaches the Gaussian limit $F=3$ for $\tau>3$ days [^21] [^22]. Model 1 matches the measurement moderately well for $\tau<100$ s, and models 2 and 3 deviate more at those scales. At larger scales $\tau>200$ s, the three stochastic models match the measurement very well. The IEC model is not intermittent enough for $\tau<3$ hours, then matches the measurement at larger scales. Performance monitoring {#sec:monitor} ====================== After characterizing and modeling the power output based on the wind speed, the stochastic approach has a second important application for the overall monitoring of the wind park performance based on 1Hz data. Performance monitoring consists in tracking the power performance of the wind installation. A monitoring procedure verifies in real time whether the installation is running in its optimal configuration, or if maintenance is needed to fix a potential anomaly. Many components of e.g. aerodynamic, mechanical or electrical nature interact in complex ways in a wind turbine. A detailed monitoring of each single component yields a huge amount of data to analyze in real time. This [*microscopic*]{} monitoring can be advantageously simplified to a [*macroscopic*]{} procedure, where a set of global estimates describes the overall system. This reduction of the many variables into a set of global estimates is even more meaningful for a large array of wind turbines such as a wind farm. The reduced amount of information must be balanced by insightful estimates that properly extract the essential information. These estimates must be sensitive enough to detect anomalies, yet robust enough to cope with the expected fluctuations. The wind speed/power output signals are commonly displayed together as a power curve. The IEC power curve (see section \[sec:dyn10min\]) gives a rapid overview of the overall performance of the conversion process $u \to P$ for a single wind turbine. The extensive averaging applied greatly reduces the amount of data, but seriously lessens the sensitivity to fine changes. Based on our analysis above, the IEC power curve is expected not to be an optimal estimate of the dynamical performance. Instead, we promote the drift coefficient $D^{(1)}(P,u;\phi)$ (see section \[sec:model1\]) as a flexible and accurate alternative. Our goal here is to test the sensitivity of both the IEC and the stochastic estimates, see resp. subsections \[sec:monitor\_iec\] and \[sec:monitor\_drift\]. First we want to find out how long an anomaly must last, so that it is surely detected. In particular, the progressive modification of the estimates is studied in relation to the anomaly duration for both methods. As a second aspect, the impact of continuous (uninterrupted) anomalies as well as intermittent anomalies (in the sense that they only last during short time periods, here of 1 min) is analyzed. A dataset is selected within the wind sector $k=8$ ($\bar{\phi}\simeq 225^\circ$) spanning a discontinuous time period of 14 days, when all 12 turbines function properly. For each anomaly, we artificially shut down one or two wind turbines (set their power output $P_i(t)=0$) during a given duration and recalculate the cumulative farm power. We developed around 50 anomaly scenarios and tested their impact on the IEC and stochastic estimates. Four of these scenarios are selected here to illustrate the two central aspects: how long must the anomaly last; and how robust are the IEC and stochastic methods to anomalies occurring intermittently during short time periods? A proper definition of the four scenarios is presented in table \[table:scenarios\]. anomaly description shut down anomaly duration normal operation anomaly type --------- -------------------- ------------ ------------------ ------------------- -------------- (a) 1shutdown\_12h 1 turbine 12h ($3.6\%$) 13d12h ($96.4\%$) continuous (b) 1shutdown\_4d 1 turbine 4d ($28.6\%$) 10d ($71.4\%$) continuous (c) 1shutdown\_4d\_int 1 turbine 4d ($28.6\%$) 10d ($71.4\%$) intermittent (d) 2shutdown\_7d 2 turbines 7d ($50\%$) 7d ($50\%$) continuous : overview of the four anomaly scenarios implemented over a 14-day time period. For each scenario (a)-(d), a description is given, as well as how many of the twelve turbines are shut down in the wind farm. The duration of the anomaly and of the normal operation are given in time (and in percent). Finally the anomaly type is given, whether continuous in time (at the end of the 14-day period) or intermittent (5760 one-minute-long anomalies are spread over the total time period, representing a total anomaly time of 4 days).[]{data-label="table:scenarios"} We monitor the cumulative farm power output for each scenario, see figure \[fig:compare\_monitoring\_signal\] and compare it to the measured (normal) farm output. It is important to note that the wind/power conditions constantly change over the 14-day period. The four anomalies cover different time periods, such that each scenario affects the power performance at different locations in the $\{u;P\}$ space. It makes it difficult to compare different scenarios from one another. Instead, we separately compare each scenario from the measured normal case and a [*full anomaly*]{} case where the shut down (of one or two turbines) occurs during the entire 14-day period. ![Wind farm power output $P(t)$ for the wind sector $k=8$ ($\bar{\phi}\simeq 225^\circ$) for scenarios (a-d) (left to right). In each anomaly scenario, the power output (black line) is compared to the power output measured (gray line in background). The period of the anomaly is marked by the gray area (except for (c) where the anomaly happens intermittently all over the 14-day period).[]{data-label="fig:compare_monitoring_signal"}](figures/compare_monitoring_signal.pdf){width="\columnwidth"} Monitoring procedure based on the IEC power curve {#sec:monitor_iec} ------------------------------------------------- We compare in figure \[fig:compare\_monitoring\_iec\] the impact of each anomaly on the ten-minute means of power output $\bar{P}$ and on the corresponding IEC power curve $P_{IEC}(u_j,\bar{\phi}_k)$. We observe that for scenario (a), most ten-minute means lie around the power curve of the normal case, and the resulting IEC power curve in figure \[fig:compare\_monitoring\_iec\](a2) does not deviate noticeably from the normal case. The anomaly only lasts $3.6\%$ of the total time, which is not enough for the IEC procedure to detect it. For scenario (b), enough ten-minute means deviate from the normal power curve, such that the IEC power curve is changed significantly by a step-like structure at wind speeds higher than $10$ m/s. We found that this is the minimal anomaly duration, i.e. $28.6\%$ of the time for the IEC method to detect an anomaly. In scenario (c), the anomaly also lasts 4 days but intermittently. In this case the ten-minute means deviate only slightly from the normal case, and no outliers are observed. The resulting IEC power curve does deviate slightly from the normal case, but less so than for the equally long but uninterrupted anomaly in scenario (b). This is expected from a ten-minute averaging that mixes 9 minutes of normal operation with 1 minute of anomaly. This is a limitation of the IEC procedure, that cannot separate scenarios alternating faster than ten minutes. Finally for scenario (d) two turbines are shut down during $50\%$ of the time. Half of the ten-minute samples are outliers that deviate clearly from the normal case. The IEC power curve lies in between the two states, as each state takes place half of the time. The anomaly is clearly detected. ![(upper) ten-minute means of farm power $\bar{P}$ versus wind speed $\bar{u}$ (dots) and; (lower) IEC power curve $P_{IEC}(u_j,\bar{\phi}_k)$ (bold black line) for the wind sector $k=8$ ($\bar{\phi}\simeq 225^\circ$) for scenarios (a-d) (left to right). The IEC power curves are also given for the measured normal case (full gray line) and for the full anomaly case (dashed gray line).[]{data-label="fig:compare_monitoring_iec"}](figures/compare_monitoring_iec_samples.pdf "fig:"){width="\columnwidth"} ![(upper) ten-minute means of farm power $\bar{P}$ versus wind speed $\bar{u}$ (dots) and; (lower) IEC power curve $P_{IEC}(u_j,\bar{\phi}_k)$ (bold black line) for the wind sector $k=8$ ($\bar{\phi}\simeq 225^\circ$) for scenarios (a-d) (left to right). The IEC power curves are also given for the measured normal case (full gray line) and for the full anomaly case (dashed gray line).[]{data-label="fig:compare_monitoring_iec"}](figures/compare_monitoring_iec.pdf "fig:"){width="\columnwidth"} Monitoring procedure based on the drift coefficient {#sec:monitor_drift} --------------------------------------------------- The drift coefficient $D^{(1)}(P,u_j,\bar{\phi}_k)$ presented in section \[sec:model1\] quantifies the conversion dynamics of the system to a first order approximation. It is a good candidate to detect fine changes in the farm dynamics. Additionally, its potential $\Psi(P,u_j,\bar{\phi}_k)$ gives a more intuitive representation of the attractors of the system, that define the Langevin power curve $P_L(u_j,\bar{\phi}_k)$. We compare the impact of each scenario on the drift coefficient, potential and Langevin power curve in figure \[fig:compare\_monitoring\_drift\]. As seen in figure \[fig:drift\_potential\], the measured normal case (full gray line) corresponds to a linear drift towards the Langevin power value. The extreme case of a full anomaly (dashed gray line) also displays a fairly linear drift towards a new, reduced power value. Shutting down some turbines reduces the power production, and the farm naturally drifts towards a lower power value. In the four scenarios, the normal and anomaly situations coexist, and the drift coefficient systematically records these two states. The anomaly duration acts as a weighting factor that determines how much the drift is distorted close to the reduced, anomalous power value. For scenario (a), $3.6\%$ anomaly time is sufficient to modify the drift coefficient and potential noticeably at power values $P<95\%$. We observe for scenario (b) that a longer anomaly time of $28.6\%$ brings new fixed points at $P\simeq 95\%$ in the rated power region. The drift coefficient displays multiple stable fixed points, and the potential has several local minima. The intermittent anomaly in scenario (c) is also detected in the drift coefficient and potential. Less fixed points are created in the Langevin power curve, due to the fact that the anomaly scatters over the entire state space $\{u;P\}$. It is not as localized in the state space as the continuous anomaly in scenario (b), so that it affects a larger region of the state space, but to a lesser extent. Finally for scenario (d), both dynamical states are recorded by the drift coefficient that clearly detects the anomaly superposed to the normal operation. ![(upper) drift coefficient $D^{(1)}(P,u_j,\bar{\phi}_k)$ and; (middle) drift potential $\Psi(P,u_j,\bar{\phi}_k)$ for the wind speed bin $j=28$ ($u\sim 13.75$m/s); (bottom) Langevin power curve $P_L(u_j,\bar{\phi}_k)$ for the wind sector $k=8$ ($\bar{\phi}\simeq 225^\circ$) for scenarios (a-d) (left to right). In each plot, the anomaly case (bold black line / black crosses) is compared to the normal case (gray line / full gray dots) and to the full anomaly case (dashed gray line / open gray dots). The drift potential is arbitrarily shifted upwards for comparison (minimum set to zero). In the (lower) figures, the dotted vertical line indicates the wind speed bin $j=28$.[]{data-label="fig:compare_monitoring_drift"}](figures/compare_lpc.pdf){width="\columnwidth"} At last, the intermittent anomaly is considered again, yet this time the duration of the anomaly is varied, see figure \[fig:compare\_potential\_intermittent\]. One sees that when the intermittent anomaly only takes place 6 hours ($1.8\%$) or less, the potential does not change. Yet if the intermittent anomaly occurs over a duration of 12 hours ($3.6\%$) or longer, the potential clearly deviates from the normal case. The deviation increases with increasing duration, as the potential approaches the full anomaly case. This type of dynamical behavior is reminiscent of phases transitions. It is similar to the cusp catastrophe described in catastrophe theory, see e.g. Ref. [@Zeeman1979]. Thus, the potential follows the generic law of the form $\Psi(P)=P^4+aP^2+bP$, where $a$ causes the bistability of the system. For a bistable system, the slope of the drift coefficient changes from negative to positive to negative again (instead of being only negative in the presence of only one attractive fixed point). In our case, we see how the parameter $b$ causes the phase transition by shifting $D^{(1)}$ along the $y-$axis and also causes the transition from one to two stable fixed points. This kind of bifurcation caused by the weighted mixing of two phases was also observed, e.g. for the differential resistance of some nonequilibrium semi- and super-conductor systems [@Peinke1987]. ![Drift potential $\Psi(P,u_j,\bar{\phi}_k)$ for the wind speed bin $j=28$ ($u\sim 13.75$m/s) and wind sector $k=8$ ($\bar{\phi}\simeq 225^\circ$) for an intermittent anomaly of varying duration (see legend). The drift potentials are given for the measured normal case (full red line) and the case when the anomaly lasts during the total 14-day period (dashed red line). The curves are arbitrarily shifted upwards (minimum set to zero) for comparison.[]{data-label="fig:compare_potential_intermittent"}](figures/compare_potential_uj=28_intermittent.pdf){width="0.5\columnwidth"} Summary {#sec:monitoring_summary} ------- In summary, we conclude from the sensitivity analysis that the drift coefficient is a better estimate than the IEC power curve to monitor the power performance of the wind farm. It reacts faster to a change in the farm dynamics than a procedure based on ten-minute averaging. This comes from the fact that the IEC method gives an average result that deviates slowly. In contrast to this the drift estimate can detect the bistable dynamics, which provides much more insight onto the cause of the anomaly, as an emerging bistability clearly indicates that something does not function properly. We observe in scenario (a) that the drift coefficient of the entire farm changes significantly when one of the twelve turbines is shut down during $3.6\%$ of the time. Scenario (b) illustrates that the same turbine must be shut down during at least $28.6\%$ of the time for the IEC method to detect it. Additionally, scenario (c) shows that if the anomaly switches the turbine on and off intermittently, the IEC power curve can no longer detect it affirmatively. The drift coefficient detects the intermittent anomalies as well as the continuous ones, in both cases the anomaly needs only last $3.6\%$ of the time to be detected. For this reason, we promote the drift coefficient and the corresponding Langevin power curve as improved macroscopic estimates for wind farm performance monitoring. Conclusion {#sec:conclu} ========== The conversion process of an onshore wind farm is analyzed at a sampling frequency of 1Hz. Clear dynamics appear when considering the effective wind speed $u$ and ten-minute average direction $\bar{\phi}$ averaged over all wind turbines, as well as their cumulative power production $P$. Fast power fluctuations of up to $\pm 20\%$ around the power curve are measured continuously, stressing the need for a more evolved description. Feature-rich dynamics are observed at the time scales of few seconds where the wind farm reacts to wind fluctuations. We propose a new approach where this conversion process $(u(t),\bar{\phi}) \to P(t)$ is modeled as a stochastic Langevin process. Two model coefficients are estimated from the data set measured, and describe intuitively the wind farm dynamics: the drift coefficient describes the attraction towards an attractive power curve, and the diffusion coefficient represents additional turbulent fluctuations. In addition, we introduce an additional pre-model of wind speed fluctuations, such that only ten-minute wind data is needed to model the power output at 1Hz. A statistical analysis confirms that the Langevin model reproduces the complex properties of the measurement well, including its intermittent features. Such high-frequency dynamics cannot be modeled correctly using a power curve method, that only describes the long-term dynamics. Beyond predicting the trend of the wind power production, our model predicts realistic power fluctuations at 1Hz, so that the impact of wind fluctuations on the grid stability can be forecast. Also, the added insight helps understand the dynamical response of large wind installations to wind fluctuations, such that smart grid concepts can be further refined. Thanks to their general flexible structure, such stochastic methods can be upscaled easily in order to understand and model the large wind installations that will populate future electric networks. As a second application, we show that the drift coefficient is a compact measure of the global performance of a wind installation, making it a powerful tool to monitor the power performance of a wind farm. It captures more information about the dynamical behavior of the wind farm than the IEC power curve. For example, the drift coefficient detects a significant change in the global wind farm dynamics if one of twelve turbines is shut down during $4\%$ of the time, whereas an anomaly time of almost $30\%$ is needed for the IEC power curve to detect it. The IEC power curve is also at a disadvantage when in presence of anomalies that intermittently switch off a turbine during a short time, because the time averaging it applies strongly reduces the sensitivity to dynamical changes faster than ten minutes. On the contrary, the drift coefficient is equally reactive to intermittent and continuous anomalies. Such stochastic analysis can be useful for the real-time monitoring of wind farms. It can be applied in addition to existing tools to give some first information about the entire farm. If a potential anomaly is detected, the stochastic analysis can then be performed in greater detail for single wind turbines. If an anomaly still appears then, a full (microscopic) analysis of all the turbine sensors is justified. Also here, a stochastic analysis may be helpful too for damage detection, see Ref. [@Rinn2012]. Our approach is fruitful as a first tool to alert of an occurring anomaly. Acknowledgements {#sec:acknowledgements} ================ We thank the land of Lower Saxony, the German Ministry for Environment and the German Ministry for Education and Research for funding this research project. We also thank Deutsche Windtechnik AG Bremen for providing us with wind turbine data. We would also like to thank Philip Rinn, Mehrnaz Anvari, David Bastine, Benjamin Wahl, Mohammad Reza Rahimi Tabar, Allan Morales, Tanja M[ü]{}cke and Michael H[ö]{}lling for the many fruitful discussions. Probabilistic description of the wind farm power production {#sec:proba} =========================================================== The Fokker-Planck equation {#sec:fp} -------------------------- A stochastic approach was presented in section \[sec:model\] to generate time series of power output for the wind farm. The stochastic signals are generated by solving a Langevin equation (\[eq:langevin\]) in time. The Langevin process generated is a stochastic drift / diffusion process. The drift coefficient $D^{(1)}(P,u,\bar{\phi})$ and the diffusion coefficient $D^{(2)}(P,u,\bar{\phi})$ approximate to a first and second order how the wind farm converts a wind field $(u(t),\bar{\phi})$ into a power output $P(t)$. The Langevin equation being by nature stochastic, the process simulated $P(t)$ has a random character. This means that if the same Langevin equation is solved $N$ times with identical wind conditions, the new signal simulated each time will always be different to some degree, owing to the random nature of the Langevin noise $\Gamma(t)$ in equation (\[eq:langevin\]) (but its statistics won’t change). While the value of $P(t)$ at a given time $t$ is random, the probability $f(P,t)$ of having a power value $P$ at time $t$ is fixed. The probability density function of the Langevin process $P(t)$ at any given time $t$ is uniquely defined as the solution of the Fokker-Planck equation [@Risken1996; @Friedrich2011] $$\begin{aligned} \frac{\partial f(P(t),t)}{\partial t}=-\frac{\partial}{\partial P(t)} &\Big( D^{(1)}\big(P(t),u(t),\bar{\phi}\big) \cdot f\big(P(t),t\big) \Big) \nonumber \\ &+ \frac{1}{2}\frac{\partial^2}{\partial P(t)^2} \Big(D^{(2)}\big(P(t),u(t),\bar{\phi}\big) \cdot f\big(P(t),t\big) \Big) \, , \label{eq:fp}\end{aligned}$$ where the Kramers-Moyal coefficients are estimated following equation (\[eq:kmc\]). The Fokker-Planck equation is the equivalent in the probability domain of the Langevin equation in the time domain. If the probability $f(P,t)$ is desired instead of simply one sample value $P(t)$, the Fokker-Planck equation should be solved over time [^23]. An example is presented in figure \[fig:L\_FP\] where the Fokker-Planck equation is solved at 1Hz using as initial condition a Dirac distribution $f(P,t=0)=\delta(P(t=0))$. We see at six exemplary times $t$ what the measured power value $P(t)$ is, and the probability $f(P,t)$ estimated by the Fokker-Planck equation. ![Histogram of the difference $\delta P = max \, f(P,t) - P(t)$ estimated over 36000 samples (10h).[]{data-label="fig:FP_deviation"}](figures/L_FP_offset=0.pdf){width="0.8\columnwidth"} ![Histogram of the difference $\delta P = max \, f(P,t) - P(t)$ estimated over 36000 samples (10h).[]{data-label="fig:FP_deviation"}](figures/FP_deviation_measurement.pdf){width="0.8\columnwidth"} The difference $\delta P$ between the most probable power value given by the Fokker-Planck equation $max\,f(P,t)$ and the measurement $P(t)$ is calculated for 36000 samples, and its histogram is presented in figure \[fig:FP\_deviation\]. $\delta P$ is rather centered around zero, has a mean value $-0.11\%$ and a standard deviation $2.54\%$. The standard deviation of $P(t)$ is equal to $\sigma_P=22.63\%$, that is 9 times as large as the standard deviation of the difference. This means that the most probable power value predicted by the Fokker-Planck equation $max\,f(P,t)$ deviates from the measurement by only $\sigma_P/9$ on average. This concludes that the PDF simulated by the Fokker-Planck equation follows the measurement faithfully. Probability-based optimization of the Kramers-Moyal coefficients {#sec:opt} ---------------------------------------------------------------- ![PDF of measured power $f(P|u_j,\bar{\phi}_k)$ (dots) and PDF $f_{st}(P)$ generated by model 2 (dashed line) and model 2 optimized (bold full line) for various wind speed bins $j \in [9,27]$ (various colors) for the wind sector $k=8$ ($\bar{\phi}\simeq 225^\circ$).[]{data-label="fig:power_histo_model2"}](figures/power_histo_model2_optimization_D2.pdf){width="0.5\columnwidth"} The drift coefficient $D^{(1)}(P,u,\bar{\phi})$ and the diffusion coefficient $D^{(2)}(P,u,\bar{\phi})$ are defined in each bin of the three-dimensional phase space $\{P_i,u_j,\bar{\phi}_k\}$. A parametric form is presented in subsection \[sec:model2\] where in each sub-domain $\{u_j,\bar{\phi}_k\}$, the drift is simplified to a linear function $D^{(1)}(P,u_j,\bar{\phi}_k) \simeq \alpha_{jk} \cdot ( P - P_{L}(u_j,\bar{\phi}_k) )$ and the diffusion to a constant $D^{(2)}(P,u_j,\bar{\phi}_k) \simeq \beta_{jk}$. The two parameters $\alpha_{jk}$ and $\beta_{jk}$ are fitted from the drift and diffusion coefficients estimated. However, some artifacts such as low data sampling or the presence of measurement noise within the data affect the estimated Kramers-Moyal coefficients, which then deviate from the true coefficients of the system. We propose an optimization procedure that corrects this discrepancy by refining the estimation of $D^{(1)}$ and $D^{(2)}$. As long as the system stays inside the given sub-domain $\{u_j,\bar{\phi}_k\}$, the local dynamics are dictated by the local values of drift and diffusion. During that period of time, the dynamics do not change and the system fluctuates around the power curve value $P_{L}(u_j,\bar{\phi}_k)$. This corresponds to the PDF $f(P,t)$ relaxing to a stationary form $f_{st}(P)$ that is given by $$\begin{aligned} 0 &=& \frac{\partial f_{st}(P)}{\partial t} \\ 0 &=& -\Bigg( \frac{\partial}{\partial P(t)} D^{(1)}\big(P(t),u(t),\bar{\phi}\big) \cdot f_{st}\big(P\big) \Bigg) + \frac{1}{2}\Bigg(\frac{\partial^2}{\partial P(t)^2} D^{(2)}\big(P(t),u(t),\bar{\phi}\big) \cdot f_{st}\big(P\big) \Bigg) \, , \label{eq:fp_zero}\end{aligned}$$ giving $$\begin{aligned} f_{st}(P)=\frac{N}{D^{(2)}\big(P,u,\bar{\phi}\big)} \,\, exp \Bigg( \int^P \frac{D^{(1)}\big(P',u,\bar{\phi}\big)}{D^{(2)}\big(P',u,\bar{\phi}\big)} dP' \Bigg) \, , \label{eq:fp_stat}\end{aligned}$$ where $N$ is a normalization constant such that $\int_{-\infty}^{\infty} f_{st}(P) \, dP=1$. $f_{st}(P)$ is the stationary solution of the Fokker-Planck equation, that corresponds to the stationary case where $D^{(1)}$ and $D^{(2)}$ do not change in time. In the limit of a system that reacts infinitely fast to the changing wind condition, one can consider that between two changes of the wind condition, the system has time to relax to its stationary state and wait for the next change. We assume that our system is reasonably close to this limit, owing to the low dispersion of the data around the power curve (see figure \[fig:uP\_1Hz\]). This means that all the data $(P|u_j,\bar{\phi}_k)$ contained in the sub-domain $\{u_j,\bar{\phi}_k\}$ is assumed to have relaxed towards its local power curve value $P_{L}(u_j,\bar{\phi}_k)$, and its PDF is $$\begin{aligned} f(P|u_j,\bar{\phi}_k) \sim f_{st}(P)=\frac{N}{\beta_{jk}} \,\, exp \Bigg( - \frac{\alpha_{jk} \big(P-P_{L}(u_j,\bar{\phi}_k)\big)^2}{2 \, \beta_{jk}} \Bigg) \, . \label{eq:fp_stat_opt}\end{aligned}$$ The two PDFs $f(P|u_j,\bar{\phi}_k)$ and $f_{st}(P)$ are presented in figure \[fig:power\_histo\_model2\] for various wind speed bins. We observe a reasonably good agreement between the two PDFs, which can be improved. The optimization is based on the fact that the PDFs are reasonably close to a Gaussian distribution $$\begin{aligned} f(P|u_j,\bar{\phi}_k) \simeq \frac{1}{\sqrt{2\pi \sigma_P \, ^2}} \,\, exp \Bigg( - \frac{ \big(P-\bar{P})\big)^2}{2 \, \sigma_P \, ^2} \Bigg) \, , \label{eq:fp_stat_opt_gauss}\end{aligned}$$ with $\bar{P}$ and $\sigma_P$ the mean and standard deviation of $(P|u_j,\bar{\phi}_k)$. Observing the analogy between equations (\[eq:fp\_stat\_opt\]) and (\[eq:fp\_stat\_opt\_gauss\]), we optimize the drift and diffusion coefficients following $$\begin{aligned} D^{(1)}_{opt}(P,u_j,\bar{\phi}_k) &=& \alpha_{jk} \cdot ( P - \bar{P} )\\ D^{(2)}_{opt}(P,u_j,\bar{\phi}_k) &=& \alpha_{jk} \cdot \sigma_P \,^2 \, . \label{eq:kmc_opt}\end{aligned}$$ The stationary PDF calculated from the optimized coefficients is presented in figure \[fig:power\_histo\_model2\]. We observe a better agreement to the measured PDFs, validating the optimization procedure [^24]. To summarize, we relate the stationary solution of the Fokker-Planck equation to the PDF of the measured power data, in each wind speed/direction sub-domain. This way we can fine-tune the parametric form of the drift and diffusion coefficients to reproduce exactly the distribution of the measured data. This optimization is used for models 2 and 3 presented resp. in subsections \[sec:model2\] and \[sec:model3\]. [10]{} World Wind Energy Association. World wind energy report 2010. Technical report, World Wind Energy Association, 2011. J. Liu, B. H. Krogh, and B. E. Ydstie. Passivity-based robust control for power systems subject to wind power variability. In [*American Control Conference*]{}, 2011. TR Ayodele, AA Jimoh, JL Munda, and JT Agee. The impact of wind power on power system transient stability based on probabilistic weighting method. , 4, 2012. Patrick Milan, Matthias Wächter, and Joachim Peinke. Turbulent character of wind energy. , 110:138701, Mar 2013. Martin Rohden, Andreas Sorge, Marc Timme, and Dirk Witthaut. Self-organized synchronization in decentralized power grids. , 109:064101, Aug 2012. Dewei Liu, Jianbo Guo, Yuehui Huang, and Weisheng Wang. An active power control strategy for wind farm based on predictions of wind turbine’s maximum generation capacity. , 5:013121, 2013. . . International Electrotechnical Commission, 2005. /61400-1, 3rd Edition. S. Lovejoy, D. Schertzer, and J. D. Stanway. Direct evidence of multifractal atmospheric cascades from planetary scales down to 1 km. , 86(22):5200–5203, May 2001. F. B[ö]{}ttcher, J. Peinke, D. Kleinhans, and R. Friedrich. Handling systems driven by different noise sources – implications for power estimations. In [*Wind Energy*]{}, pages 179–182. Springer, Berlin, 2007. A Morales, M W[ä]{}chter, and J Peinke. Characterization of wind turbulence by higher-order statistics. , 15(3):391–406, 2012. S. Lovejoy, D. Schertzer, V. Allaire, T. Bourgeois, S. King, J. Pinel, and J. Stolle. Atmospheric complexity or scale by scale simplicity. , 36:L01801, January 2009. F.D. Bianchi, H. [De Battista]{}, and R.J. Mantz. . Springer, Berlin, 2nd edition, 2006. Peter Tavner, Yingning Qiu, Athanasios Korogiannos, and Yanhui Feng. The correlation between wind turbine turbulence and pitch failure. In [*Proceedings of EWEA 2011*]{}, 2011. Tanja M[ü]{}cke, David Kleinhans, and Joachim Peinke. Atmospheric turbulence and its influence on the alternating loads on wind turbines. , 14(2):301–316, 2011. Isaac Van der Hoven. Power spectrum of horizontal wind speed in the frequency range from 0.0007 to 900 cylcles per hour. , 14(2):160–164, 1957. D. Schertzer, I. Tchiguirinskaia, S. Lovejoy, and A. F. Tuck. Quasi-geostrophic turbulence and generalized scale invariance, a theoretical reply. , 12(1), 2012. G. Fitton, I. Tchiguirinskaia, D. Schertzer, and S. Lovejoy. The anisotropic multifractal model and wind turbine wakes. In [*7th PhD Seminar on Wind Energy in Europe*]{}, 2011. F. C. Kaminsky, R. H. Kirchhoff, C. Y. Syu, and J. F. Manwell. A comparison of alternative approaches for the synthetic generation of a wind speed time series. , 113(4):280–289, 1991. Rachel Ba[ï]{}le, Philippe Poggi, and Jean-Francois Muzy. Intermittency model for surface layer wind speed fluctuations. applications to short term forecasting and calibration of the wind resource. In [*Proceedings of EWEC*]{}, 2010. Rudy Calif. Pdf models and synthetic model for the wind speed fluctuations based on the resolution of langevin equation. , 99(0):173 – 182, 2012. R. Calif and F.G. Schmitt. Modeling of atmospheric wind speed sequence using a lognormal continuous stochastic equation. , 109(0):1 – 8, 2012. . . International Electrotechnical Commission, 2005. T. Burton, D. Sharpe, N. Jenkins, and E. Bossanyi. . Wiley, New York, 2001. Bingheng Wu, Mengxuan Song, Kai Chen, Zhongyang He, and Xing Zhang. Wind power prediction system for wind farm based on auto regressive statistical model and physical model. , 6(1):013101, 2014. Oliver Kamps. Characterizing the fluctuations of wind power production by multi-time statistics. In [*Proceedings of the Euromech Colloquium 528*]{}, 2012. Jean-Francois Muzy, Rachel Ba[ï]{}le, and Philippe Poggi. Intermittency of surface-layer wind velocity series in the mesoscale range. , 81(5):056308, May 2010. Mar[í]{}a Isabel Blanco. The economics of wind energy. , 13(6):1372–1382, 2009. Berthold Hahn, Michael Durstewitz, and Kurt Rohrig. Reliability of wind turbines. In [*Wind energy*]{}, pages 329–332. Springer, 2007. Sebastian Pfaffel, Volker Berkhout, Stefan Faulstich, Paul K[ü]{}hn, Katrin Linke, Philipp Lyding, and Renate Rothkegel. Wind energy report germany 2011. Technical report, Fraunhofer Institute for Wind Energy and Energy System Technology (IWES), 2012. Chia Chen Ciang, Jung-Ryul Lee, and Hyung-Joon Bang. Structural health monitoring for a wind turbine system: a review of damage detection methods. , 19(12):122001, 2008. Nicholas J. Cutler, Hugh R. Outhred, and Iain F. MacGill. Using nacelle-based wind speed observations to improve power curve modeling for wind power forecasting. , pages 245–258, 2011. Patrick Milan, Matthias W[ä]{}chter, and Joachim Peinke. Stochastic modeling of wind power production. In [*Proceedings of EWEA 2011*]{}, Brussels, 2011. Edgar Anahua, Stephan Barth, and Joachim Peinke. Markovian power curves for wind turbines. , 11(3), 2008. Julia Gottschall and Joachim Peinke. How to improve the estimation of power curves for wind turbines. , 3(1):015005 (7pp), 2008. Hannes Risken. . Springer, 1996. Rudolf Friedrich, Joachim Peinke, Muhammad Sahimi, and M. Reza Rahimi Tabar. Approaching complexity by stochastic methods: From biological systems to turbulence. , 2011. Frank B[ö]{}ttcher, Joachim Peinke, Rudolf Friedrich, David Kleinhans, Pedro G. Lind, and Maria Haase. Reconstruction of complex dynamical systems affected by strong measurement noise. , 97, 2006. Philip Sura and Joseph Barsugli. A note on estimating drift and diffusion parameters from timeseries. , 305:304–311, 2002. J. Gottschall and J. Peinke. On the definition and handling of different drift and diffusion estimates. , 10:083034 (20pp), 2008. H Haken. Advanced synergetic, instability hierarchies of self-organizing systems and devices, 1983. Frank B[ö]{}ttcher, Stephan Barth, and Joachim Peinke. Small and large fluctuations in atmospheric wind speeds. , 21:299–308, 2007. Robert Stresing and J Peinke. Towards a stochastic multi-point description of turbulence. , 12(10):103046, 2010. William H. Press, Saul A. Teukolsky, William T. Vetterling, and Brian P. Flannery. . Cambridge University Press, third edition, 2007. A. Arneodo, C. Baudet, F. Belin, R. Benzi, B. Castaing, B. Chabaud, R. Chavarria, S. Ciliberto, R. Camussi, F. Chilla, B. Dubrulle, Y. Gagne, B. Hebral, J. Herweijer, M. Marchand, J. Maurer, J. F. Muzy, A. Naert, A. Noullez, J. Peinke, F. Roux, P. Tabeling, W. van de Water, and H. Willaime. Structure functions in turbulence, in various flow configurations, at reynolds number between 30 and 5000, using extended self-similarity. , 34(6), 1996. R. Benzi, S. Ciliberto, R. Tripiccione, C. Baudet, F. Massaioli, and S. Succi. Extended self-similarity in turbulent flows. , 48(1), 1993. R.G. Kavasseri and R. Nagarajan. Evidence of crossover phenomena in wind speed data. , 51 (11):2255–2262, November 2004. E Christopher Zeeman. . Springer, 1979. J Peinke, DB Schmid, B R[ö]{}hricht, and J Parisi. Positive and negative differential resistance in electrical conductors. , 66(1):65–73, 1987. P. Rinn, H. Hei[ß]{}elmann, M. W[ä]{}chter, and J. Peinke. Stochastic method for in-situ damage analysis. , 86(1):1–5, 2012. [^1]: More accurate approaches involve an aero-mechanical description of each wind turbine [@Burton2001], which becomes excessively demanding for everyday prediction. [^2]: This observation contradicts the intuitive yet false argument claiming that the fluctuations of neighboring wind turbines cancel out. Considering many wind installations separated by a distance shorter than the correlation length of atmospheric winds (in hundreds of kilometers according to Ref. [@Muzy2010a]), these [*neighboring*]{} installations are driven by correlated wind fields and produce correlated outputs. They are not independent power sources, and their cumulative output does not sum up to a Gaussian process, that is the central limit theorem cannot be applied. [^3]: From now on, we refer to the IEC-like power curve as [*IEC power curve*]{}. We stress that it is not the original procedure, but our own adaptation of the IEC standard to a wind farm. [^4]: While we have access to the effective wind direction $\phi(t)$ at $1$Hz, our results show that the ten-minute average $\bar{\phi}$ is a more representative measure. Wake effects are expected to be slow over the farm size of several km, explaining why we favor $\bar{\phi}$ over the fluctuating direction $\phi(t)$. [^5]: This definition of the coefficients based on a derivation allows for a more robust estimation in the presence of measurement noise that spoils the data. [^6]: The estimation is only possible in regions of the phase space that were visited during the training period, i.e. sufficient events are in this bin so that the corresponding mean values can be estimated. The training period should then be long enough for the measurement to span all the regions of interest. [^7]: It is a direct analogy to the motion of a macroscopic particle within a flow, that was described by Robert Brown in 1827 as [*Brownian motion*]{}, or a consequence of synergetic [@Haken1983]. [^8]: The noise term $\Gamma(t)$ fluctuates much faster than $\sqrt{D^{(2)}(P(t),u(t),\bar{\phi})}$, that is considered constant during the integration step $\Delta t$. When integrating the stochastic Langevin equation following It[ô]{} definition, $\int_t^{t+\Delta t} \sqrt{D^{(2)}(P(t'),u(t'),\bar{\phi})}\cdot \Gamma(t') dt'=\sqrt{D^{(2)}(P(t),u(t),\bar{\phi})} \int_t^{t+\Delta t} \Gamma(t') dt'$. Ref. [@Friedrich2011] shows that $\int_t^{t+\Delta t} \Gamma(t') dt'=\sqrt{\Delta t} \cdot \eta(t)$. [^9]: As observed in figure \[fig:drift\], the IEC power curve $P_{IEC}(u_j,\bar{\phi}_k)$ is very close to the Langevin power curve $P_{L}(u_j,\bar{\phi}_k)$ for the data set measured. For this reason, the IEC power curve could be used instead of the Langevin power curve in equation (\[eq:D1\_parametric\]) if necessary. [^10]: The condition $\alpha_{jk}<0$ must be fulfilled for the model to be stable. This condition corresponds to a converging drift that drives the dynamics towards a stable attractor, the Langevin power curve. [^11]: For example, when the wind speed is around $12$ m/s, $\alpha_{jk}\simeq-0.05$/s. If the power output $P$ is $20\%$ lower than the ideal value $P_{L}$, equation (\[eq:D1\_parametric\]) indicates that $D^{(1)} \simeq \alpha_{jk} \cdot ( P - P_{L} ) \simeq 1\%/$s. When in such conditions, the power output changes on average by $\sim 1\%$ each second. [^12]: Although $\beta_{jk}$ quantifies the intensity of the diffusion coefficient $D^{(2)}(P,u_j,\bar{\phi}_k)$, a measure for the variance of the fluctuations is the ratio $\beta_{jk}/\alpha_{jk}$. That is, the competition between diffusion and drift determines the range of expected fluctuations, see appendix \[sec:opt\]. [^13]: The IEC norm [@IEC] defines wind speed fluctuations as a Gaussian field with spectral properties described by either a Kaimal or a von Karman spectrum. [^14]: The question whether wind dynamics can be easily separated into small-scale (turbulence) and large-scale (meteorology) effects is still an open question. The historical yet controversial hypothesis of a spectral gap [@vdHoven1957] implied such a separation. Some recent studies go beyond such simplification and present more complex descriptions such as e.g. multifractal cascade models. [^15]: The normal distribution is commonly observed in natural phenomena, including the wind speed of some turbulent flows. Yet this oversimplified description of turbulence is deceived by a more complex multifractal scaling and small-scale intermittency. Here we consider the effective wind speed $u(t)$ averaged over the entire wind farm, whose fluctuations $u'(t)$ are close to a Gaussian monofractal process, see also Ref. [@Morales2010a]. We believe that a random Gaussian model is in this particular case acceptable. [^16]: The decay parameter $\gamma$ should be estimated once for the given location of the wind farm as a measure of the local wind correlations. [^17]: The Fourier transform $\hat{P}(f)$ was calculated using the [*Fastest Fourier Transform in the West*]{} algorithm. The spectrum was then estimated using Bartlett’s method to obtain a denoised spectrum [@Press2007]. [^18]: The autocorrelation function of a stationary signal $x(t)$ is given by $R_{xx}(\tau)=\frac{\langle (x(t+\tau)-\bar{x})(x(t)-\bar{x}) \rangle}{\sigma_x \,^2}$, where $\sigma_x \,^2$ is the variance of $x(t)$. $R_{xx}(\tau)$ is the inverse Fourier transform of the spectrum $|\hat{x}(f)|^2$ of $x(t)$, following Wiener-Khinchin theorem. Because the power signals we study are non-stationary, we calculated $R_{PP}(\tau)$ over time windows of $8192$s. [^19]: The second-order 2-point moment $<P_\tau(t)^2>$ corresponds to the autocorrelation and the power spectrum. The PDFs $f(P_\tau(t))$ incorporate information about all the higher-order 2-point statistics $< P_\tau(t)^n>=\int P_\tau(t)^n f(P_\tau) dP_\tau$. This allows for a thorougher validation of our models. [^20]: More precisely, we use the [*extended self-similarity*]{} method from Ref. [@Benzi1993] to remove a slightly anomalous scaling observed in the structure functions. This method extracts the scaling exponents from the relation $S_q(\tau) \sim S_3(\tau)^{\zeta_q}$, forcing $\zeta_3=1$. [^21]: It is interesting to note that the correlation length of some atmospheric wind datasets was found to be around $4-5$ days in [@Kavasseri2004; @Muzy2010a], matching our observation for the wind farm output. The autocorrelation function shows $R_{PP}(\tau=4$days$) \simeq 0.25$ and $R_{PP}(\tau>15$days$) \simeq 0$. [^22]: A local peak is observed for the flatness of the measured data for $\tau \sim 10$ minutes. This time scale corresponds to the characteristic size of the farm of about $3$ km. It could be due to the fact that the farm converts atmospheric structures at this scale without filtering them, yielding a local increase in intermittency. This feature is not reproduced by any of the models. [^23]: We solve the Fokker-Planck equation using the [*path integral method*]{} presented in Ref. [@Risken1996]. Given an initial condition $f(P,t=0)$, we calculate $f(P,t>0)$ by repeatedly solving the equation $f(P,t+dt)=\int f(P,t+dt|P',t) f(P',t) dP'$ over an infinitesimal time step $dt$, where the conditional probability $f(P,t+dt|P',t)$ is completely described by $D^{(1)}$ and $D^{(2)}$. One can visualize how the drift and diffusion coefficients concurrently contract and stretch the PDF over time. [^24]: A chi-square test was performed to quantify the deviation between measured and reconstructed PDFs. When averaging the chi-square value $\chi^2_j$ over all wind speed bins $j$, we find for the average $< \chi^2_j >_j$ that the optimization reduces the value by a factor $6.8$.
{ "pile_set_name": "ArXiv" }
--- author: - Sam Kriegman - Nick Cheney - Josh Bongard bibliography: - 'main.bib' title: How morphological development can guide evolution --- Supplementary References {#supp-refs .unnumbered} ======================== 1\. Ancel, L. W. Undermining the baldwin expediting effect: does phenotypic plasticity accelerate evolution? *Theor. population biology* **58**, 307–319 (2000).
{ "pile_set_name": "ArXiv" }
--- abstract: | In this paper, we adapt the well-known *local* uniqueness results of Borg-Marchenko type in the inverse problems for one dimensional Schrödinger equation to prove *local* uniqueness results in the setting of inverse *metric* problems. More specifically, we consider a class of spherically symmetric manifolds having two asymptotically hyperbolic ends and study the scattering properties of massless Dirac waves evolving on such manifolds. Using the spherical symmetry of the model, the stationary scattering is encoded by a countable family of one-dimensional Dirac equations. This allows us to define the corresponding transmission coefficients $T(\lambda,n)$ and reflection coefficients $L(\lambda,n)$ and $R(\lambda,n)$ of a Dirac wave having a fixed energy $\lambda$ and angular momentum $n$. For instance, the reflection coefficients $L(\lambda,n)$ correspond to the scattering experiment in which a wave is sent from the *left* end in the remote past and measured in the same left end in the future. The main result of this paper is an inverse uniqueness result local in nature. Namely, we prove that for a fixed $\lambda \not=0$, the knowledge of the reflection coefficients $L(\lambda,n)$ (resp. $R(\lambda,n)$) - up to a precise error term of the form $O(e^{-2nB})$ with $B>0$ - determines the manifold in a neighbourhood of the left (resp. right) end, the size of this neighbourhood depending on the magnitude $B$ of the error term. The crucial ingredients in the proof of this result are the Complex Angular Momentum method as well as some useful uniqueness results for Laplace transforms. *Keywords*. Inverse Scattering, Black Holes, Dirac Equation.\ *2010 Mathematics Subject Classification*. Primaries 81U40, 35P25; Secondary 58J50. author: - 'Thierry Daudé [^1], Damien Gobin and François Nicoleau [^2]' title: Local inverse scattering at fixed energy in spherically symmetric asymptotically hyperbolic manifolds --- Introduction and statement of the results ========================================= The aim of this short paper is to extend the *local* inverse uniqueness results of Borg-Marchenko type for one dimensional Schrödinger equation obtained first in [@Si], and improved in [@Be; @GS; @Te], to the setting of inverse *metric* problems, that is inverse problems on three or four dimensional curved manifolds whose unknown - the object we wish to determine by observing waves at infinity - is the (Riemanniann or Lorentzian) metric itself. We shall consider for the moment a very specific and simple class of $3$D-Riemanniann manifolds that we name Spherically Symmetric Asymptotically Hyperbolic Manifolds, in short SSAHM. Precisely, these are described by the set $$\Sigma = {\mathbb{R}}_x \times {\mathbb{S}}^2_{\theta,\varphi},$$ equipped with the Riemanniann metric $$\sigma = dx^2 + a^{-2}(x) \,d\omega^2,$$ where $d\omega^2 = \left( d\theta^2 + \sin^2\theta \,d\varphi^2 \right)$ is the euclidean metric on the $2$D-sphere ${\mathbb{S}}^2$. The assumptions on the function $a(x)$ - that determines completely the metric - are: $$\label{RegulA} a \in C^2({\mathbb{R}}), \quad a > 0,$$ and $$\label{AsympA} \exists \, a_\pm > 0, \ \kappa_+ < 0, \ \kappa_- > 0, \quad \quad \begin{array}{ccc} a(x) & = & a_\pm e^{\kappa_\pm x} + O(e^{3\kappa_\pm x}), \quad x \to \pm \infty, \\ a'(x) & = & a_\pm \kappa_\pm e^{\kappa_\pm x} + O(e^{3\kappa_\pm x}), \quad x \to \pm \infty. \end{array}$$ Under these assumptions, $(\Sigma,\sigma)$ is clearly a spherically symmetric Riemanniann manifold with two asymptotically hyperbolic ends $\{x = \pm \infty\}$. Note indeed that the metric $\sigma$ is asymptotically a *small* perturbation of the “hyperbolic like” metrics $$\sigma_\pm = dx^2 + e^{-2\kappa_\pm x} d\omega_\pm^2, \quad x \to \pm \infty,$$ where $d\omega_\pm^2 = 1/(a_\pm^2) d\omega^2$ are fixed metrics on ${\mathbb{S}}^2$. From this, we see easily that the sectional curvature of $\sigma$ tends to the *constant negative values* $- (\kappa_\pm)^2$ on the corresponding ends $\{x = \pm \infty\}$. Hence the name “asymptotically hyperbolic” for this kind of geometry. Note in passing that we allow $\kappa_\pm$ to take different values leading to different sectional curvatures in the two ends. We emphasize at last that such SSAHM are very particular cases (because of our assumption of spherical symmetry) of the much broader class of asymptotically hyperbolic manifolds for instance described in [@IK; @JSB; @SB] (to cite only a few papers that deal with inverse problems). On the manifold $(\Sigma,\sigma)$, we are interested in studying how (scalar, electromagnetic, Dirac, …) waves evolve, scatter at late times and ultimately, in trying to answer the question: can we determine the metric by observing these waves at the infinities of the manifold (in our model, the two ends $\{x=\pm\infty\}$). For definiteness, we shall consider in this paper how massless Dirac waves propagate and scatter towards the two asymptotically hyperbolic ends. Note that the same results should hold with the Dirac equation replaced by the wave equation. Precisely, let us consider the massless Dirac equation $$\label{DE} i \partial_t \psi = \mathbb{D}_{\sigma} \psi,$$ where $\mathbb{D}_{\sigma}$ denotes a representation of the Dirac operator on $(\Sigma,\sigma)$ and the $2$-spinor solution $\psi$ belongs to $L^2(\Sigma; {\mathbb{C}}^2)$. It will be shown in Section \[TM\] that we have a very simple connection between $\mathbb{D}_{\sigma}$ and the function $a(x)$ appearing in the metric $\sigma$, precisely $$\label{DO} \mathbb{D}_{\sigma} = {\Gamma^1}D_x + a(x) \mathbb{D}_{{\mathbb{S}}^2},$$ where $\mathbb{D}_{{\mathbb{S}}^2}$ denotes the intrinsic Dirac operator on ${\mathbb{S}}^2$, represented here by the expression $$\label{DiracS2} \mathbb{D}_{{\mathbb{S}}^2} = {\Gamma^2}\left( D_{\theta} + \frac{i \cot{\theta}}{2} \right) + {\Gamma^3}\frac{1}{\sin{\theta}} D_{\varphi},$$ with $D_x = -i\partial_x, \ D_\theta = -i\partial_\theta, \ D_\varphi = -i\partial_\varphi$ and where the $2 \times 2$- Dirac matrices ${\Gamma^1}, {\Gamma^2}, {\Gamma^3}$ satisfy the usual anti-commutation relations $$\label{Anticom} \Gamma^i \Gamma^j + \Gamma^j \Gamma^i = 2 \delta_{ij}.$$ Due to the spherical symmetry of the problem and the existence of generalized spherical harmonics $\{Y_{kl}\}$ that “diagonalize” $\mathbb{D}_{{\mathbb{S}}^2}$, we can decompose the energy Hilbert space ${\mathcal{H}}= L^2(\Sigma; {\mathbb{C}}^2)$ onto a Hilbert sum of partial Hilbert spaces ${\mathcal{H}}_{kl}$ with the property that the ${\mathcal{H}}_{kl}$’s are let invariant through the action of the Dirac operator (\[DO\]). More precisely, if we introduce the set of indices $I = \{ k \in 1/2 + {\mathbb{Z}}, \ l \in 1/2 + {\mathbb{N}}, \ |k| \leq l\}$, we have $${\mathcal{H}}= \oplus_{kl} {\mathcal{H}}_{kl}, \quad {\mathcal{H}}_{kl} = L^2({\mathbb{R}};{\mathbb{C}}^2) \otimes Y_{kl},$$ and $$\mathbb{D}_{\sigma}^{kl} := \mathbb{D}_{\sigma \ |{\mathcal{H}}_{kl}} = {\Gamma^1}D_x - (l+1/2) a(x) {\Gamma^2}.$$ Note that the partial Dirac operator $\mathbb{D}_{\sigma}^{kl}$’s only depend on the *angular momentum* $l + 1/2 \in {\mathbb{N}}^*$. For simplicity, we shall denote $l+{\frac{1}{2}}$ by $n$ (hence the new parameter $n$ runs over the integers ${\mathbb{N}}^*$) and also $$\label{DO-1D} \mathbb{D}_{\sigma}^n = {\Gamma^1}D_x - n a(x) {\Gamma^2},$$ for the partial Dirac operators on each generalized spherical harmonic $Y_{kl}$. We are thus led to consider the restriction of the Dirac equation (\[DE\]) to each partial Hilbert space ${\mathcal{H}}_{kl}$ separatly and study the properties of the family of $1$D Dirac Hamiltonians $\mathbb{D}_{\sigma}^n, \ n \in {\mathbb{N}}^*$ in order to obtain spectral, direct and inverse scattering results for the complete Dirac Hamiltonian $\mathbb{D}_{\sigma}$. This has been done in [@DN3] in a very similar context[^3] (see also [@AKM; @Da1; @DN2]). Let us summarize here these results. We refer to Section \[SS\] for more explanations. First, the Dirac Hamiltonian $\mathbb{D}_{\sigma}$ is selfadjoint on the Hilbert space ${\mathcal{H}}= L^2(\Sigma; {\mathbb{C}}^2)$ and has absolutely continuous spectrum. In particular, the pure point spectrum of $\mathbb{D}_{\sigma}$ is empty. As a consequence, the energy of massless Dirac fields cannot remain trapped on any compact subsets of $\Sigma$, *i.e.* for all compact subset $K \subset {\mathbb{R}}$, $$\lim_{t \to \pm \infty} \| \mathbf{1}_{K}(x) e^{-it\mathbb{D}_{\sigma}} \psi \| = 0.$$ In other words, the massless Dirac fields *scatter* towards the asymptotic ends $\{ x = \pm \infty \}$ of the manifold $\Sigma$ at late times. Second, a complete direct scattering theory can be established for $\mathbb{D}_{\sigma}$ on $(\Sigma,\sigma)$. For all energy $\lambda \in {\mathbb{R}}$, we denote the scattering matrix at energy $\lambda$ by $S(\lambda)$. It is a unitary operator on $L^2({\mathbb{S}}^2; {\mathbb{C}}^2)$ and thus has the structure of an operator valued $2 \times 2$ matrix, *i.e.* $$\label{SM} S(\lambda) = \left[ \begin{array}{cc} T_L(\lambda) & R(\lambda)\\ L(\lambda)&T_R(\lambda) \end{array} \right],$$ where $T_L, T_R$ are the transmission operators and $R, L$ the reflection operators. The formers measure the part of a signal having energy $\lambda$ transmitted from an end to the other end in a scattering process whereas the latters measure the part of a signal of energy $\lambda$ reflected from an end to itself ($\{x = -\infty\}$ for $L$ and $\{ x = +\infty\}$ for $R$). Due to the spherical symmetry of the model, the scattering matrix lets invariant all the partial Hilbert spaces ${\mathcal{H}}_{kl}$ and can be thus decomposed into a Hilbert sum of unitary operators acting ${\mathbb{C}}^2$. We write as a shorthand $$S(\lambda) = \sum_{k,l \in I} S_{kl}(\lambda), \quad S_{kl}(\lambda) := S(\lambda)_{|{\mathcal{H}}_{kl}}.$$ Since the $1$D Dirac operator (\[DO-1D\]) only depends on $n = l+1/2 \in {\mathbb{N}}^*$, the partial scattering matrices $S_{kl}(\lambda)$ also only depend on $n$. We shall thus use the notation $$\label{SM-n} S(\lambda,n) = \left[ \begin{array}{cc} T(\lambda,n)&R(\lambda,n)\\ L(\lambda,n)&T(\lambda,n) \end{array} \right].$$ For all $n \in {\mathbb{N}}^*$, we emphasize that the partial scattering matrices $S(\lambda,n)$ are unitary matrices that encode the stationary scattering at a fixed energy $\lambda$ on a given generalized spherical harmonics ${\mathcal{H}}_{kl}$ with $n = l + 1/2$. As above, the transmission coefficients $T(\lambda,n)$ correspond to the transmitted part of a signal (from one end to the other) whereas the left $L(\lambda,n)$ and right $R(\lambda,n)$ reflection coefficients correspond to the reflected part of a signal in a given end. In [@DN3], we addressed the question whether it was possible to determine uniquely the metric from the knowledge of the reflection coefficients $L(\lambda,n)$ or $R(\lambda,n)$. Using essentially the Complex Angular Momentum method (see [@Re] for the first appearance of this method and [@Ra] for an application to Schrödinger inverse scattering), we were able to answer positively to the question with some interesting improvements in the hypotheses. Precisely, we state here the inverse scattering uniqueness result proved in [@DN3]. \[globaluniqueness\] Let $\Sigma = {\mathbb{R}}\times {\mathbb{S}}^2$ be a SSAHM equipped with the Riemanniann metric $$\sigma = dx^2 + a^{-2}(x) d\omega^2,$$ where the function $a(x)$ satisfies the assumptions (\[RegulA\]) - (\[AsympA\]). Let $\mathbb{D}_{\sigma} = {\Gamma^1}D_x + a(x) \mathbb{D}_{{\mathbb{S}}^2}$ be an expression of the massless Dirac operator associated to $(\Sigma,\sigma)$. To the evolution equation $i\partial_t \psi = \mathbb{D}_{\sigma} \psi$ with $\psi \in {\mathcal{H}}= L^2(\Sigma; {\mathbb{C}}^2)$, we associate the countable family of partial waves scattering matrices $S(\lambda,n)$ for $\lambda \in {\mathbb{R}}$ and $n \in {\mathbb{N}}^*$ as above. Consider also a subset $\mathcal{L}$ of ${\mathbb{N}}^*$ that satisfies a Müntz condition $$\sum_{n \in \mathcal{L}} \frac{1}{n} = \infty.$$ Then the knowledge of either $R(\lambda,n)$ or $L(\lambda,n)$ for a fixed $\lambda \ne 0$ and for all $n \in \mathcal{L}$ determines uniquely the function $a(x)$ (and thus the metric $\sigma$) up to a discrete set of translations. First, we emphasize that the above result is not true if $\lambda=0$, (see Remark 3.7, [@DN3]). Secondly, in [@DN3], Theorem 1.1, it is claimed that the knowledge of the transmission coefficients $T(\lambda,n)$ for a fixed $\lambda \ne 0$ and for all $n \in \mathcal{L}$ also determines uniquely the function $a(x)$ up to a translation. The crucial ingredient of the proof can be found in the Proposition 3.13 of [@DN3] which states that if $T(\lambda,n) = \tilde{T}(\lambda,n)$ for all $n \in \mathcal{L}$, then the corresponding reflection coefficients $L(\lambda,n)$ and $\tilde{L}(\lambda,n)$ (resp. $R(\lambda,n)$ and $\tilde{R}(\lambda,n)$) coincide up to a multiplicative constant. The proof of this result given in [@DN3] is unfortunately incomplete and therefore, this last point is not so clear and could even be false. However, in this paper, in Proposition \[unicitetransmission\], Addendum B, we prove that the knowledge of the transmission coefficients $T(\lambda,n)$ for all $n \in \mathcal{L}$ together with the knowledge of the reflection coefficients $L(\lambda,k)$ for a finite number of integer $k$, (and a technical assumption on the sectional curvatures), uniquely determines the function $a(x)$ up to a translation. The question whether these last hypotheseses are necessary remains open. For more general Asymptotically Hyperbolic Manifolds (AHM in short) with no particular symmetry, difficult direct and inverse scattering results for *scalar waves* have been proved by Joshi, Sá Barreto in [@JSB], by Sá Barreto in [@SB] and by Isozaki, Kurylev in [@IK]. In [@JSB] for instance, it is shown that the asymptotics of the metric of an AHM are uniquely determined (up to certain diffeomorphisms) by the scattering matrix $S(\lambda)$ at a fixed energy $\lambda$ off a discrete subset of ${\mathbb{R}}$. In [@SB], it is proved that the metric of an AHM is uniquely determined (up to certain diffeomorphisms) by the scattering matrix $S(\lambda)$ for every $\lambda \in {\mathbb{R}}$ off an exceptional subset. Similar results are obtained recently in [@IK] for even more general classes of AHM. At last, we also mention [@BP] where related inverse problems - inverse resonance problems - are studied in certain subclasses of AHM. The new inverse scattering results of this paper are local in nature, in the same spirit as [@Be; @GS; @Si; @Te]. Instead of assuming the full knowledge of one of the reflection operators, we instead assume the knowledge of one of these operators up to some precise error remainder (see below). Using the particular analytic properties of the scattering coefficients $ L(\lambda,z)$ and $R(\lambda,z)$ with respect to the complex angular momentum $z$ and some well known uniqueness properties of the Laplace transform (see [@Ho1; @Si]), we are able to prove the following improvement of our previous result. \[Reflection\] Let $(\Sigma, \sigma)$ and $(\Sigma,\tilde{\sigma})$ two a priori different SSAHM. We denote by $a(x)$ and $\tilde{a}(x)$ the two radial functions defining the metrics $\sigma$ and $\tilde{\sigma}$. We define $$A = \int_{\mathbb{R}}a(x) dx, \quad \tilde{A} = \int_{\mathbb{R}}\tilde{a}(x) dx,$$ as well as the diffeomorphisms $$\begin{array}{ccl} g: \ {\mathbb{R}}& \longrightarrow & (0, A), \\ x & \longrightarrow & g(x) = \int_{-\infty}^x a(s) ds, \end{array}, \quad \quad \begin{array}{ccl} \tilde{g}: \ {\mathbb{R}}& \longrightarrow & (0, \tilde{A}), \\ x & \longrightarrow & \tilde{g}(x) = \int_{-\infty}^x \tilde{a}(s) ds, \end{array}.$$ We also denote by $h=g^{-1} : (0,A) \longrightarrow {\mathbb{R}}, \ \tilde{h}=\tilde{g}^{-1} : (0,\tilde{A}) \longrightarrow {\mathbb{R}}$ their inverse diffeomorphisms. As above, we define $S(\lambda,n)$ and $\tilde{S}(\lambda,n)$ the corresponding partial scattering matrices. Let $\lambda\not= 0$ be a fixed energy and $0< B < \min\ (A, \tilde{A})$. Then the following assertions are equivalent: $$\label{Ln} (i) \quad L(\lambda,n) = \tilde{L}(\lambda,n) + \ O\left(e^{-2nB} \right),\ n \rightarrow +\infty.$$ $$(ii) \quad \exists k \in {\mathbb{Z}}, \quad a(x) = \tilde{a}(x + \frac{k\pi}{\lambda}), \quad \forall \ x \leq h(B)= \tilde{h}(B) - \frac{k\pi}{\lambda}.$$ Symmetrically, the following assertions are also equivalent: $$\label{Rn} (iii) \quad R(\lambda,n) = \tilde{R}(\lambda,n) + \ O\left(e^{-2nB} \right), \ n \rightarrow +\infty.$$ $$(iv) \quad \exists k \in {\mathbb{Z}}, \quad a(x) = \tilde{a}(x + \frac{k\pi}{\lambda}), \quad \forall \ x \geq h(A-B)= \tilde{h}(\tilde{A}-B) - \frac{k\pi}{\lambda}.$$ The above result asserts that the *partial* knowledge of the reflection coefficients in the sense of (\[Ln\]) or (\[Rn\]) allows to determine uniquely the metric $\sigma$ in the neighbourhoods of the two ends $\{x = \pm \infty\}$. The size of these neighbourhoods depend on the magnitude of the error terms in (\[Ln\]) - (\[Rn\]). Of course, $h(B)$ (resp. $h(A-B)$) depends on the metric $a(x)$ which is a priori unknown. But, it is not difficult to prove using (\[AsympA\]) that $h(B) \sim \frac{1}{\kappa_-} \ \log B$ when $B \rightarrow 0$. In the same way, $h(A-B) \sim - \frac{1}{\kappa_+} \ \log (A-B)$ when $B \rightarrow A$. We also emphasize that the “surface gravities” $\kappa_{\pm}$ can be explicitly recover from the asymptotics of $L(\lambda, n)$ or $R(\lambda,n)$, $n \rightarrow + \infty$, (see [@DN3], Theorem 4.22). As a direct consequence, we obtain immediately the following *global uniqueness* result for the metric of a SSAHM. This result slightly improves our earlier version obtained in [@DN3] and stated in Theorem \[globaluniqueness\]. Assume that for a given $C \geq \min (A, \tilde{A})$ and $\lambda\not=0$ a fixed energy, one of the following assertions holds : $$(i) \quad L(\lambda,n) = \tilde{L}(\lambda,n) + \ O\left(e^{-2nC} \right),\ n \rightarrow +\infty.$$ $$(ii) \quad R(\lambda,n) = \tilde{R}(\lambda,n) + \ O\left(e^{-2nC} \right) ,\ n \rightarrow +\infty.$$ Then, there exists $k \in {\mathbb{Z}}, \ a(x) = \tilde{a}(x+ \frac{k\pi}{\lambda}), \ \forall \ x \in {\mathbb{R}}.$ Let us treat for instance the case $(i)$ and assume that $A \leq \tilde{A}$. From our hypothesis, for all $B<A$, we have $$L(\lambda,n) = \tilde{L}(\lambda,n) + \ O\left(e^{-2nB} \right).$$ Hence Theorem \[Reflection\] implies that $a(x) = \tilde{a}(x + \frac{k\pi}{\lambda}), \ \forall \ x \leq h(B)$. The result follows letting $B$ tend to $A$ and using that ${\displaystyle{\lim_{X \to A} h(X) = +\infty}}$. Note that we also obtain $A= \tilde{A}$. Let us give here a possible interpretation of the above *local uniqueness* result Theorem \[Reflection\]. Consider for instance the reflection coefficients $L(\lambda,n)$ and recall that it encodes the following scattering experiment: a wave having energy $\lambda$ is sent from the end $\{x = - \infty\}$ in the past and evolves on the SSAHM. Then $L(\lambda,n)$ measures the part of this wave that is reflected to the same end $\{ x = -\infty \}$ in the far future. Now our result asserts that if we know $L(\lambda,n)$ up to a precise error term of the form $O \left(e^{-2nB}\right)$, then the metric is uniquely determined in a neighbourhood of $\{x = -\infty\}$, the size of the neighbourhood depending only on the constant $B$ defining the error term. We infer thus that, under our assumption, the wave sent from $\{ x = -\infty \}$ hasn’t the time to travel through the whole manifold before being measured back in the end $\{ x = - \infty\}$. This explains heuristically why the partial knowledge of $L(\lambda,n)$, in the precise sense given by our assumption, is not enough to determine the full metric.\ At last, when using the transmission coefficients as the starting point of our inverse problem, we get a result different in nature than the one obtained with the reflection coefficients. Precisely, we obtain a *global uniqueness* result. Moreover, as we have said before, we have to assume that the reflection coefficients $L(\lambda,n)$ are equal for a finite number of integer $n$ and we make a technical assumption on the sectional curvatures $\kappa_\pm$. The question whether these last hypotheses are necessary remains open. \[Transmission\] Assume that $$\frac{1}{\kappa_+} + \frac{1}{\kappa_-} < 0, \quad \frac{1}{\tilde{\kappa}_+} + \frac{1}{\tilde{\kappa}_-} < 0,$$ and that for a fixed energy $\lambda\not=0$ and for some $B > \max(A,\tilde{A})$, $$\label{Tn} \quad T(\lambda,n) = \tilde{T}(\lambda,n) + \ O\left(e^{-2nB} \right),\ n \rightarrow +\infty.$$ Assume also that $$\label{rr11} L(\lambda,n) = \tilde{L}(\lambda,n),$$ for a finite but large enough number of indices $n \in {\mathbb{N}}$. Then there exists a constant $\sigma \in {\mathbb{R}}$ such that $$\tilde{a}(x) = a(x + \sigma).$$ In consequence, the two SSAHM $(\Sigma,g)$ and $(\tilde{\Sigma},\tilde{g})$ coincide up to isometries. Let us make a few comments on this result. First, we provide an heuristic reason why we don’t have a *local uniqueness* result when we assume the knowledge of the transmission coefficients up to a precise error. The transmission coefficients - by definition - measure the part of a wave transmitted from one end, say $\{x = -\infty\}$, to the other end $\{x = +\infty\}$. In our case where the SSAHM has only two ends, the transmitted wave has thus the time to propagate into the whole manifold. It is then natural that the transmission coefficients encode *all* the information of the SSAHM. Second, the asymptotics of the transmission coefficients when $n$ tends to infinity are computed in [@DN3]. Precisely, we have $$\mid T(\lambda,n) \mid \sim C \ e^{-nA}, \quad \mid \tilde{T}(\lambda,n) \mid \sim \tilde{C} \ e^{-n\tilde{A}}.$$ Hence, the condition on $B$ in Theorem \[Transmission\] cannot be weaker that $B > \frac{1}{2} \max (A, \tilde{A})$. Note then that the assertion (i) implies immediately $A = \tilde{A}$ from the above asymptotics. We mention that a global uniqueness inverse result in the case where $\frac{1}{2} \max (A, \tilde{A}) < B \leq \max (A, \tilde{A})$ is still an open question. This paper is organised as follows. In Section \[TM\], we recall how to compute the Dirac equation on a curved manifold and apply this formalism to obtain a representation of a massless Dirac operator on a SSAHM that is suitable for us. In Section \[SS\], we recall the main results from [@DN3] where a complete description of the stationary scattering corresponding to massless Dirac fields evolving in a SSAHM was obtained. In Section \[LIS\], we prove our main results, Theorem \[Reflection\], Theorem \[Transmission\]. Eventually, we include the last Section \[BH\] in which an application of our local inverse uniqueness results on SSAHM is given in the context of black hole spacetimes, precisely on Reissner-Nordström-de-Sitter black holes. The model ========= \[TM\] Since the Dirac equation is by essence a relativistic equation, we prefer to work directly on the four dimensional Lorentzian manifold $(M,\tau)$ defined by $$M = {\mathbb{R}}_t \times \Sigma,$$ and equipped with the metric $$\tau = dt^2 - \sigma,$$ where $(\Sigma,\sigma)$ is the SSAHM we aim to study. Below we recall how to compute the massless Dirac equation on such a $4$D curved background and obtain a representation of it that we put under the generic Hamiltonian form $$i \partial_t \psi = \mathbb{D}_{\tau} \psi.$$ We shall see in Remark \[Link\] that this procedure leads to an equivalent form of the massless Dirac equation than the one $$i \partial_t \psi = \mathbb{D}_{\sigma} \psi,$$ we would have obtained working on the $3$D-Riemanniann manifold $(\Sigma,\sigma)$. In other words, the Dirac operators $\mathbb{D}_{\tau}$ and $\mathbb{D}_{\sigma}$ that we obtain by these two formalisms are shown to be unitarily equivalent. We prefer to work with the relativistic point of view nevertheless since we are also interested in applications of our inverse results to spacetimes coming from General Relativity, namely black hole spacetimes. We postpone this parenthesis till Section \[BH\]. Orthonormal frame formalism for the massless Dirac equation in $4$D curved space-time ------------------------------------------------------------------------------------- To calculate the massless Dirac equation in a 4D curved spacetime $M$ equipped with a Lorentzian metric $\tau$ of signature $(1,-1,-1,-1)$, we use Cartan’s orthonormal frame formalism as explained for instance in [@CP] or in a more relativistic setting in [@Ni]. Let us denote by $\{e_A\}_{A=0,1,2,3}$ a given local Lorentz frame, *i.e.* a set of vector fields satisfying $\tau(e_A, e_B) = \eta_{AB}$ where $\eta_{AB} =$ diag$(1,-1,-1,-1)$ is the flat (Lorentz) metric. We also denote by $\{e^A\}_{A=0,1,2,3}$ the set of dual $1$-forms of the frame $\{e_A\}$. Latin letters A,B will denote in what follows local Lorentz frame indices, while Greek letters $\mu, \nu$ run over four-dimensional space-time coordinate indices. The *massless* Dirac equation takes then the generic form $$\label{AbstractDiracEq} \mathbb{D} \phi = \gamma^A (\partial_A + \Gamma_A) \phi = 0.$$ Here, the $\gamma^A$’s are the gamma Dirac matrices satisfying the anticommutation relations $$\label{Clifford} \{\gamma^A, \gamma^B\} = \gamma^A \gamma^B + \gamma^B \gamma^A = 2 \eta^{AB}.$$ The differential operators $\partial_A$’s are given by $\partial_A = e_A^\mu \partial_\mu$ in terms of the local differential operators and the $\Gamma_A$’s are the components of the spinor connection $\Gamma = \Gamma_A e^A = \Gamma_\mu dx^\mu$ in the local Lorentz frame. In order to derive the latter, we first compute the spin-connection $1$-form $\omega_{AB} = \omega_{AB\mu} dx^\mu = f_{ABC} e^C$ thanks to Cartan’s first structural equation and the skew-symmetric condition $$\label{Cartan} d e^A + \omega^A_{\ B} \wedge e^B = 0, \quad \omega_{AB} = \eta_{AC} \omega^C_{\ B} = - \omega_{BA}.$$ Note here that we use the flat metric $\eta_{AB}$ or its inverse $\eta^{AB}$ to raise or lower Latin indices. We also use Einstein summation convention. With this definition, the spinor connection $\Gamma$ is then defined as $$\label{Gamma} \Gamma = \frac{1}{8} [\gamma^A, \gamma^B] \omega_{AB} = \frac{1}{4} \gamma^A \gamma^B \omega_{AB} = \frac{1}{4} \gamma^A \gamma^B f_{ABC} e^C.$$ The Dirac equation on a SSAHM ----------------------------- We now apply this formalism to calculate the massless Dirac equation on the $4$D- Lorentzian manifold $(M,\tau)$ given by $$M = {\mathbb{R}}_t \times {\mathbb{R}}_x \times {\mathbb{S}}^2_{\theta,\varphi},$$ and $$\tau = dt^2 - dx^2 - a^{-2}(x) \left( d\theta^2 + \sin^2\theta \,d\varphi^2 \right),$$ where $$a \in C^2({\mathbb{R}}), \quad a > 0.$$ The spherical symmetry of the metric leads to the natural choice of local Lorentz frame $$\label{LocalFrame} e_0 = \partial_t, \quad e_1 = \partial_x, \quad e_2 = a(x) \partial_\theta, \quad e_3 = \frac{a(x)}{\sin\theta} \partial_\varphi.$$ The dual $1$-forms are then given by $$\label{LocalDualFrame} e^0 = dt, \quad e^1 = dx, \quad e^2 = a^{-1}(x) d\theta, \quad e^3 = \frac{\sin\theta}{a(x)} d\varphi.$$ The exterior derivatives of the $e^A$’s are readily computed $$d e^0 = 0, \quad de^1 = 0, \quad de^2 = \frac{a'(x)}{a(x)} \, e^2 \wedge e^1, \quad de^3 = \frac{a'(x)}{a(x)} \, e^3 \wedge e^1 - a(x) \cot\theta \, e^3 \wedge e^2.$$ Using (\[Cartan\]), we then easily get $$\omega^0_{\ 1} = \omega^0_{\ 2} = \omega^0_{\ 3} = 0, \quad \omega^1_{\ 2} = \frac{a'(x)}{a(x)} \, e^2, \quad \omega^1_{\ 3} = \frac{a'(x)}{a(x)} \, e^3, \quad \omega^2_{\ 3} = - a(x) \cot \theta \, e^3,$$ or equivalently $$\omega_{01} = \omega_{02} = \omega_{03} =0, \quad \omega_{12} = - \frac{a'(x)}{a(x)} \, e^2, \quad \omega_{13} = - \frac{a'(x)}{a(x)} \, e^3, \quad \omega_{23} = a(x) \cot \theta \, e^3.$$ Hence we deduce $$\label{SpinConnection} \Gamma = \frac{1}{4} \gamma^A \gamma^B \omega_{AB} = \left( - \frac{a'(x)}{2 a(x)} \gamma^1 \gamma^2 \right) e^2 + \left( - \frac{a'(x)}{2 a(x)} \gamma^1 \gamma^3 + \frac{a(x) \cot \theta}{2} \, \gamma^2 \gamma^3 \right) e^3= \Gamma_A e^A.$$ The massless Dirac equation $\gamma^A (\partial_A + \Gamma_A) \phi = 0$ on $(M,\tau)$ thus takes the form $$\left[ \gamma^0 \partial_t + \gamma^1 \partial_x + \gamma^2 \left( a(x) \partial_\theta - \frac{a'(x)}{2 a(x)} \gamma^1 \gamma^2 \right) + \gamma^3 \left( \frac{a(x)}{\sin \theta} \partial_\varphi - \frac{a'(x)}{2 a(x)} \gamma^1 \gamma^3 + \frac{a(x) \cot \theta}{2} \, \gamma^2 \gamma^3 \right) \right] \phi = 0,$$ or using (\[Clifford\]) $$\left[ \gamma^0 \partial_t + \gamma^1 \partial_x + a(x) \left( \left( \partial_\theta - \frac{\cot \theta}{2} \right) \gamma^2 + \frac{1}{\sin \theta} \partial_\varphi \gamma^3 \right) - \frac{a'(x)}{a(x)} \gamma^1 \right] \phi = 0.$$ We can get rid of some potentials by considering the weighted spinor $$\label{WeightedSpinor} \psi = a^{-1}(x) \phi.$$ Then $\psi$ satifies the equation $$\left[ \gamma^0 \partial_t + \gamma^1 \partial_x + a(x) \left( \left( \partial_\theta - \frac{\cot \theta}{2} \right) \gamma^2 + \frac{1}{\sin \theta} \partial_\varphi \gamma^3 \right) \right] \psi = 0.$$ We finally put this equation under Hamiltonian form. The spinor $\psi$ thus satisfies $$i \partial_t \psi = \mathbb{D} \psi,$$ where the Dirac operator $\mathbb{D}$ is given by $$\mathbb{D} = \gamma^0 \gamma^1 D_x + a(x) \left[ \left( D_\theta + \frac{i \cot \theta}{2} \right) \gamma^0 \gamma^2 + \frac{1}{\sin \theta} D_\varphi \gamma^0 \gamma^3 \right],$$ and $$D_x = -i \partial_x, \quad D_\theta = -i \partial_\theta, \quad D_\varphi = -i \partial_\varphi.$$ Let us introduce some notations. We denote $$\Gamma^1 = \gamma^0 \gamma^1, \quad \Gamma^2 = \gamma^0 \gamma^2, \quad \Gamma^3 = \gamma^0 \gamma^3.$$ From (\[Clifford\]), it is clear that the Dirac matrices $\Gamma^1, \Gamma^2, \Gamma^3$ satisfy the usual anticommutation relations $$\label{AntiCom} \{ \Gamma^i, \, \Gamma^j \} = 2 \delta_{ij}, \quad \forall i,j = 1, 2, 3.$$ We choose the following representation for these Dirac matrices $$\label{DiracMatrices} {\Gamma^1}= \left( \begin{array}{cc} 1&0 \\0&-1 \end{array} \right), \quad {\Gamma^2}= \left( \begin{array}{cc} 0&1 \\1&0 \end{array} \right), \quad {\Gamma^3}= \left( \begin{array}{cc} 0&i \\-i&0 \end{array} \right).$$ We also denote $$\label{DiracSphere} \mathbb{D}_{{\mathbb{S}}^2} = {\Gamma^2}\left( D_{\theta} + \frac{i \cot{\theta}}{2} \right) + {\Gamma^3}\frac{1}{\sin{\theta}} D_{\varphi},$$ which turns out to be an expression of the intrinsic Dirac operator on ${\mathbb{S}}^2$. With all these notations, the massless Dirac equation on $(M,\tau)$ takes its final Hamiltonian form $$\label{DiracEq} i \partial_t \psi = \mathbb{D} \psi, \quad \mathbb{D} = {\Gamma^1}D_x + a(x) \mathbb{D}_{{\mathbb{S}}^2}.$$ \[Link\] Consider the Riemanniann manifold $\Sigma = {\mathbb{R}}_x \times {\mathbb{S}}^2_{\theta,\varphi}$ equipped with the metric $\sigma = dx^2 + a^{-2}(x) d\omega^2$ where $d\omega^2$ denotes the euclidean metric on ${\mathbb{S}}^2$. Using the same Cartan’s orthonormal frame formalism as described above, we could associate to $(\Sigma,\sigma)$ a Dirac operator $\mathbb{D}_{\sigma}$ and consider the associated Dirac equation $$\label{DiracEq-SV} i \partial_t \psi = \mathbb{D}_{\sigma} \psi,$$ with $\psi$ given by (\[WeightedSpinor\]). The Dirac equation (\[DiracEq-SV\]) is the one we obtain if we adopt the Schrödinger viewpoint, namely if we consider the evolution of Dirac fields on the fixed $3$D-Riemanniann manifold $(\Sigma,\sigma)$. On the other hand, the Dirac equation (\[DiracEq\]) is the one we obtain if we adopt the relativistic viewpoint, that is the natural Dirac equation associated to the $4$D-Lorentzian manifold $(M = {\mathbb{R}}_t \times \Sigma, \tau = dt^2 - \sigma)$. It turns out that the two points of view are equivalent in the sense that the corresponding Dirac operators $\mathbb{D}$ and $\mathbb{D}_{\sigma}$ are unitarily equivalent. This can be seen by a direct calculation. To each point of $(\Sigma,\sigma)$, we associate the orthonormal local frame $$e_1 = \partial_x, \quad e_2 = a(x) \partial_\theta, \quad e_3 = \frac{a(x)}{\sin\theta} \partial_\varphi.$$ Note that the vector fields $\{e_A\}_{A = 1,2,3}$ satisfy $\sigma(e_A, e_B) = \delta_{AB}$ where $\delta_{AB} =$ diag$(1,1,1)$ is the flat (Riemanniann) $3$D metric. Now following the same procedure as above (still introducing the spinor weight (\[WeightedSpinor\])), we obtain the following Dirac equation $$i \partial_t \psi = \mathbb{D} \psi, \quad \mathbb{D} = \gamma^1 D_x + a(x) \left[ \gamma^2 \left( D_{\theta} + \frac{i \cot{\theta}}{2} \right) + \gamma^3 \frac{1}{\sin{\theta}} D_{\varphi} \right],$$ where the gamma Dirac matrices $\gamma^1, \gamma^2, \gamma^3$ satisfy the anticommutation formulae (\[AntiCom\]). Hence we conclude that the Dirac equations (\[DiracEq\]) and (\[DiracEq-SV\]) only differ by a choice of equivalent representation of the gamma Dirac matrices satisfying (\[AntiCom\]). But it is well known that such two different choices of representation lead to unitarily equivalent Dirac operators (see [@Th]). The stationary scattering ========================= \[SS\] In this section, we recall the construction of the stationary representation of the scattering matrix $S(\lambda,n)$ for a fixed energy $\lambda \in {\mathbb{R}}$ and all angular momentum $n \in {\mathbb{N}}$, (we refer to [@AKM] and [@DN3] for details). Let us consider first the stationary solutions of equation (\[DiracEq\]) restricted to each spin weighted spherical harmonic, *i.e.* the solutions of $$\label{PartialSE} [ {\Gamma^1}D_x - n a(x) {\Gamma^2}] \psi = \lambda \psi, \quad \forall n \in {\mathbb{N}}^*.$$ For $\lambda \in {\mathbb{R}}$, we define the Jost solution from the left $F_L(x,\lambda,n)$ and the Jost solution from the right $F_R(x,\lambda,n)$ as the $2\times2$-matrix solutions of (\[PartialSE\]) satisfying the following asymptotics $$\begin{aligned} F_L(x,\lambda,n) & = & e^{i{\Gamma^1}\lambda x} (I_2 + o(1)), \ x \to +\infty, \label{FL}\\ F_R(x,\lambda,n) & = & e^{i{\Gamma^1}\lambda x} (I_2 + o(1)), \ x \to -\infty. \label{FR}\end{aligned}$$ From (\[PartialSE\]), (\[FL\]) and (\[FR\]), it is easy to see that such solutions (if there exist) must satisfy the integral equations $$\label{IE-FL} F_L(x,\lambda,n) = e^{i{\Gamma^1}\lambda x} - i n {\Gamma^1}\int_x^{+\infty} e^{-i{\Gamma^1}\lambda (y-x)} a(y) {\Gamma^2}F_L(y,\lambda,n) dy,$$ $$\label{IE-FR} F_R(x,\lambda,n) = e^{i{\Gamma^1}\lambda x} + i n {\Gamma^1}\int_{-\infty}^x e^{-i{\Gamma^1}\lambda (y-x)} a(y) {\Gamma^2}F_R(y,\lambda,n) dy.$$ Since the potential $a$ belongs to $L^1({\mathbb{R}})$, it follows that the integral equations (\[IE-FL\]) and (\[IE-FR\]) are uniquely solvable by iteration and that $$\|F_L(x,\lambda,n)\| \leq e^{n \int_x^{+\infty} a(s) ds}, \quad \|F_R(x,\lambda,n)\| \leq e^{n \int_{-\infty}^x a(s) ds}.$$ Since the Jost solutions are fundamental matrices of (\[PartialSE\]), there exists a $2\times2$-matrix $A_L(\lambda,n)$ such that $F_L(x,\lambda,n) = F_R(x,\lambda,n) \, A_L(\lambda,n)$. From (\[FR\]) and (\[IE-FL\]), we get the following expression for $A_L(\lambda,n)$ $$\label{ALRepresentation} A_L(\lambda,n) = I_2 - i n {\Gamma^1}\int_{\mathbb{R}}e^{-i{\Gamma^1}\lambda y} a(y) {\Gamma^2}F_L(y,\lambda,n) dy.$$ Moreover, the matrix $A_L(\lambda,n)$ satisfies the following equality (see [@AKM], Proposition 2.2) $$\label{AL-Relation} A_L^*(\lambda,n) {\Gamma^1}A_L(\lambda,n) = {\Gamma^1}, \quad \forall \lambda \in {\mathbb{R}}, \ n \in {\mathbb{N}}.$$ Using the notation $$\label{ScatCoef} A_L(\lambda,n) = \left[\begin{array}{cc} a_{L1}(\lambda,n)&a_{L2}(\lambda,n)\\a_{L3}(\lambda,n)&a_{L4}(\lambda,n) \end{array} \right],$$ the equality (\[AL-Relation\]) can be written in components as $$\label{ALUnitarity} \left. \begin{array}{ccc} |a_{L1}(\lambda,n)|^2 - |a_{L3}(\lambda,n)|^2 & = & 1, \\ |a_{L4}(\lambda,n)|^2 - |a_{L2}(\lambda,n)|^2 & = & 1, \\ a_{L1}(\lambda,n) \overline{a_{L2}(\lambda,n)} - a_{L3}(\lambda,n) \overline{a_{L4}(\lambda,n)} & = & 0. \end{array} \right.$$ The matrices $A_L(\lambda,n)$ encode all the scattering information of equation (\[PartialSE\]). In particular, it is shown in [@AKM] that the scattering matrix $S(\lambda,n)$ has the representation $$\label{SR-SM1} S(\lambda,n) = \left[ \begin{array}{cc} T(\lambda,n)&R(\lambda,n)\\ L(\lambda,n)&T(\lambda,n) \end{array} \right],$$ where $$\label{SR-SM2} T(\lambda,n) = \frac{1}{a_{L1}(\lambda,n)}, \quad R(\lambda,n) = - \frac{a_{L2}(\lambda,n)}{a_{L1}(\lambda,n)}, \quad L(\lambda,n) = \frac{a_{L3}(\lambda,n)}{a_{L1}(\lambda,n)}.$$ It follows from (\[IE-FL\]) that if we define the new potential $\tilde{a}(x)=a(x+c)$, the associated Jost solutions satisfy $$\label{FLdecale} \tilde{F_L}(x,\lambda,n) = F_L(x + c,\lambda,n) e^{-i {\Gamma^1}\lambda c}.$$ Hence, it follows from (\[ALRepresentation\]) that (with obvious notations) $$\label{ALdecale} \tilde{A_L}(\lambda,n) = e^{i{\Gamma^1}\lambda c} A_L(\lambda,n) e^{-i{\Gamma^1}\lambda c},$$ and so, using (\[SR-SM1\]) and (\[SR-SM2\]), we conclude that $$\tilde{S}(\lambda,n) = e^{i{\Gamma^1}\lambda c} S(\lambda,n) e^{-i{\Gamma^1}\lambda c},$$ or in components $$\label{Sdecale} \left[ \begin{array}{cc} \tilde{T}(\lambda,n)& \tilde{R}(\lambda,n)\\ \tilde{L}(\lambda,n) & \tilde{T}(\lambda,n) \end{array} \right] = \left[ \begin{array}{cc} T(\lambda,n)& e^{2i \lambda c} R(\lambda,n)\\ e^{-2i \lambda c} L(\lambda,n)&T(\lambda,n) \end{array} \right].$$ Hence the transmission coefficients $T(\lambda,n)$ are invariant under any radial translations of the potential $a$, whereas the reflection coefficients $L(\lambda,n)$ and $R(\lambda,n)$ are invariant under the discrete set of radial translations $\tilde{a}(x) = a(x + \frac{k \pi}{\lambda})$ for $k \in {\mathbb{Z}}$ and $\lambda \ne 0$. Following an original idea due to Regge [@Re], we shall allow the angular momentum $n \in {\mathbb{N}}$ to take *complex values* $z$ and study the analytic properties of the above scattering data with respect to $z \in {\mathbb{C}}$. Precisely, it was shown in [@DN3] that we can define for $z \in {\mathbb{C}}$, the Jost solutions $F_L(x,\lambda,z)$ and $F_R(x,\lambda,z)$ which are the unique solutions of the stationary equation $$\label{PartialSE1} [ {\Gamma^1}D_x - z a(x) {\Gamma^2}] \psi = \lambda \psi, \quad \forall z \in {\mathbb{C}}.$$ with the asymptotics (\[FL\]) and (\[FR\]). Similarly, we can define the matrix $A_L(\lambda,z)$ for all $z \in {\mathbb{C}}$. All these matrix-functions are analytic in the complex variable $z \in {\mathbb{C}}$. Moreover, they satisfy the following properties: \[AL-Analytic\] (i) Set $A = \displaystyle\int_{\mathbb{R}}a(x) dx$. Then $$\begin{aligned} |a_{L1}(\lambda,z)|, \ |a_{L4}(\lambda,z)| \leq \cosh(A|z|), \quad \forall z \in {\mathbb{C}}, \label{AL-ExpType1}\\ |a_{L2}(\lambda,z)|, \ |a_{L3}(\lambda,z)| \leq \sinh(A|z|), \quad \forall z \in {\mathbb{C}}. \label{AL-ExpType2} \end{aligned}$$ (ii) The functions $a_{L1}(\lambda,z)$ and $a_{L4}(\lambda,z)$ are entire and even in $z$ whereas the functions $a_{L2}(\lambda,z)$ and $a_{L3}(\lambda,z)$ are entire and odd in $z$. Moreover they satisfy the symmetries $$\begin{aligned} a_{L1}(\lambda,z) & = & \overline{a_{L4}(\lambda,\bar{z})}, \quad \forall z \in {\mathbb{C}}, \label{ALSym1}\\ a_{L2}(\lambda,z) & = & \overline{a_{L3}(\lambda,\bar{z})}, \quad \forall z \in {\mathbb{C}}. \label{ALSym2} \end{aligned}$$ (iii) The following relations hold for all $z \in {\mathbb{C}}$ $$\begin{aligned} a_{L1}(\lambda,z) \overline{a_{L1}(\lambda,\bar{z})} - a_{L3}(\lambda,z) \overline{a_{L3}(\lambda,\bar{z})} & = & 1, \label{SymAL1-AL3}\\ a_{L4}(\lambda,z) \overline{a_{L4}(\lambda,\bar{z})} - a_{L2}(\lambda,z) \overline{a_{L2}(\lambda,\bar{z})} & = & 1. \label{SymAL2-AL4} \end{aligned}$$ At this stage, we have proved that the components of the matrix $A_L(\lambda,z)$ are entire functions of exponential type in the variable $z$. Precisely, from (\[AL-ExpType1\]) and (\[AL-ExpType2\]), we have $$\label{AL-ExpType} |a_{Lj}(\lambda,z)| \leq e^{A |z|}, \quad \forall z \in {\mathbb{C}}, \ j=1,..,4.$$ Using the relations (\[SymAL1-AL3\]), (\[SymAL2-AL4\]) and the parity properties of the $a_{Lj}(\lambda,z)$, we can improve these estimates (see Lemma 3.4. in [@DN3]). \[MainEsti\] Let $\lambda \in {\mathbb{R}}$ be fixed. Then for all $z \in {\mathbb{C}}$ $$\label{MainEst} |a_{Lj}(\lambda,z)| \leq e^{A |Re(z)|}, \quad j=1,..,4.$$ It follows from Lemma \[MainEsti\] that the functions $z \rightarrow a_{Lj}(\lambda, z)$ belong to the Nevanlinna class in the right half-plane (see for instance [@Ru] for a definition). We emphasize that this property is the key point to prove Theorem \[globaluniqueness\] (see [@DN3]). Similarly, if we use the notation $$F_L(x,\lambda,z) = \left[\begin{array}{cc} f_{L1}(x,\lambda,z)&f_{L2}(x,\lambda,z)\\f_{L3}(x,\lambda,z)&f_{L4}(x,\lambda,z) \end{array} \right], \quad F_R(x,\lambda,z) = \left[\begin{array}{cc} f_{R1}(x,\lambda,z)&f_{R2}(x,\lambda,z)\\f_{R3}(x,\lambda,z)&f_{R4}(x,\lambda,z) \end{array} \right],$$ we have the corresponding estimates for the Jost functions $f_{Lj}(x,\lambda,z)$ and $f_{Rj}(x,\lambda,z)$ for $j= 1,\dots,4$. Precisely \[MainEstiF\] For all $j=1,..,4$ and for all $x \in {\mathbb{R}}$, $$\begin{aligned} |f_{Lj}(x,\lambda,z)| \leq C \, e^{|Re(z)| \int_x^{\infty} a(s) ds}, \\ |f_{Rj}(x,\lambda,z)| \leq C \, e^{|Re(z)| \int_{-\infty}^x a(s) ds}. \end{aligned}$$ Finally, we shall need later the asymptotic expansion of the scattering data when the angular momentum $z \rightarrow +\infty$, $z$ real. The main tool to obtain these asymptotics easily is a simple change of variable $X=g(x)$, called the Liouville transformation which we precise here. Let us define $$\label{Liouville} X= g(x)=\int_{-\infty}^x a(t) \ dt.$$ Clearly, since $a > 0$ and continuous, $g:{\mathbb{R}}\rightarrow ]0,A[$ is a $C^1$-diffeomorphism where $$\label{DefA} A=\int_{\mathbb{R}}a(t) \ dt.$$ In what follows, we denote by $h=g^{-1}$ the inverse diffeomorphism of $g$ and we use the notation ${\displaystyle{f'(X) = \frac{\partial f}{\partial X} (X)}}$. We also define for $j=1,...,4$, and for $X \in ]0,A[$, $$\label{Fj} {f_j (X, \lambda,z)}= f_{Lj} (h(X), \lambda,z),$$ $$\label{Gj} {g_j (X, \lambda,z)}= f_{Rj} (h(X), \lambda,z).$$ Observe at last that in the variable $X$, Lemma \[MainEstiF\] can be written as $$\label{estimatefg} \forall z > 0, \quad |f_j(X,\lambda,z)| \leq C \, e^{z(A-X)} \ \ ,\ \ |g_j(X,\lambda,z)| \leq C \, e^{z X}.$$ The interest in introducing the variable $X$ is that the components ${f_j (X, \lambda,z)}$ and ${g_j (X, \lambda,z)}$ of the Jost solutions satisfy now singular Sturm-Liouville differential equations in the variable $X$, in which the complex angular momentum $z$ plays the role of the spectral parameter. More precisely, we have the following lemma. \[SturmLiouville\] 1. For $j=1,2$, ${f_j (X, \lambda,z)}$ and ${g_j (X, \lambda,z)}$ satisfy on $]0,A[$ the Sturm-Liouville equation $$\label{SL1} y'' +q(X)y = z^2 y.$$ 2. For $j=3,4$, ${f_j (X, \lambda,z)}$ and ${g_j (X, \lambda,z)}$ satisfy on $]0,A[$ the Sturm-Liouville equation $$\label{SL2} y'' +\overline{q(X)}y = z^2 y,$$ where the potential $${\displaystyle{q(X) = \lambda^2 h'(X)^2 -i \lambda h''(X)= \frac{\lambda^2}{a^2(x)} +i\lambda \frac{a'(x)}{a^3(x)}}},$$ has the asymptotics $$\begin{aligned} \label{omega} & &q(X) - \frac{\omega_-}{X^2} = O(1)\ , \ X \rightarrow 0 \ ,\quad \rm{with} \ \ \omega_- = \frac{\lambda^2}{\kappa_-^2} + i \frac{\lambda}{\kappa_-}, \\ & &q(X) - \frac{\omega_+}{(A-X)^2} = O(1) \ , \ X \rightarrow A \ ,\quad \rm{with} \ \ \omega_+ = \frac{\lambda^2}{\kappa_+^2} + i \frac{\lambda}{\kappa_+}. \label{omega1} \end{aligned}$$ A short glance at Lemma \[SturmLiouville\] suggests that the Jost functions $f_j$ and $g_j$ can be constructed as small perturbations of usual modified Bessel functions $I_{\nu}(z(A-X))$ and $I_{\mu}(zX)$ for suitable $\mu, \ \nu$. This was done in details in [@DN3]. As a consequence of this construction and using the well known asymptotic expansion of the modified Bessel functions, the large $z$ asymptotics of the scattering data $a_{Lj}(\lambda,z)$ were calculated in [@DN3]. More precisely, if we set $$\nu_+ = \frac{1}{2} - i\frac{\lambda}{\kappa_+} \ ,\ \mu_- = \frac{1}{2} + i\frac{\lambda}{\kappa_-},$$ the following asymptotics hold. \[asymtoticsutiles\] 1. For $X \in ]0,A[$ fixed and $z \in S_\theta$ where $S_\theta = \{ z \in {\mathbb{C}}, \ |\arg(z)| \leq \theta\}$ for a given $0 < \theta < \frac{\pi}{2}$, we have for the Jost solutions $f_1(X)$ and $g_2(X)$ $$\begin{aligned} f_1(X,\lambda, z) &=& \frac{2^{-\nu_+}}{\sqrt{2\pi}}\ (-\frac{\kappa_+}{a_+})^{\frac{i\lambda}{\kappa_+}} \ \Gamma(1-\nu_+) \ z^{-\frac{i\lambda}{\kappa_+}} \ e^{z(A-X)} \ \Big(1+O(\frac{1}{z})\Big). \label{asymflun} \\ g_2(X,\lambda, z) &=& i \ \frac{2^{-\mu_-}}{\sqrt{2\pi}}\ (\frac{\kappa_-}{a_-})^{-\frac{i\lambda}{\kappa_-}} \ \Gamma(1-\mu_-) \ z^{\frac{i\lambda}{\kappa_-}} \ e^{zX} \ \Big(1+O(\frac{1}{z})\Big). \label{asymgdeux}\end{aligned}$$ 2. For the scattering data $a_{L1}(\lambda,z)$ and $a_{L3}(\lambda,z)$, we have $$\begin{aligned} {a_{L1}(\lambda,z)}& = & \frac{1}{2\pi}\ \left(-\frac{\kappa_+}{a_+}\right)^{\frac{i\lambda}{\kappa_+}} \left(\frac{\kappa_-}{a_-}\right)^{-\frac{i\lambda}{\kappa_-}} \Gamma\left(\frac{1}{2}-\frac{i\lambda}{\kappa_-}\right) \Gamma\left(\frac{1}{2}+\frac{i\lambda}{\kappa_+}\right) \nonumber \\ & & \hspace{1cm} \times \left(\frac{z}{2}\right)^{i\lambda (\frac{1}{\kappa_-} - \frac{1}{\kappa_+})} \ e^{zA} \ \left(1 + O(\frac{1}{z})\right), \label{asymptoticalun} \\ {a_{L3}(\lambda,z)}& = & \frac{i}{2\pi}\ \left(-\frac{\kappa_+}{a_+}\right)^{\frac{i\lambda}{\kappa_+}} \left(\frac{\kappa_-}{a_-}\right)^{\frac{i\lambda}{\kappa_-}} \Gamma\left(\frac{1}{2}+\frac{i\lambda}{\kappa_-}\right) \Gamma\left(\frac{1}{2}+\frac{i\lambda}{\kappa_+}\right) \nonumber \\ & & \hspace{1cm} \times \left(\frac{z}{2}\right)^{-i\lambda (\frac{1}{\kappa_+} + \frac{1}{\kappa_-})} e^{zA} \ \left(1 + O(\frac{1}{z})\right). \nonumber\end{aligned}$$ 3. For the scattering coefficients $T(\lambda,z)$ and $L(\lambda,z)$, we have $$\begin{aligned} \label{asymptoticTL} T(\lambda,z) & = & 2\pi \frac{\left(-\frac{a_+}{\kappa_+}\right)^{\frac{i\lambda}{\kappa_+}} \left(\frac{a_-}{\kappa_-}\right)^{-\frac{i\lambda}{\kappa_-}} }{\Gamma\left(\frac{1}{2}-\frac{i\lambda}{\kappa_-}\right) \Gamma\left(\frac{1}{2}+\frac{i\lambda}{\kappa_+}\right)} \ \left(\frac{z}{2}\right)^{i\lambda (\frac{1}{\kappa_+} - \frac{1}{\kappa_-})} e^{-zA} \left(1 + O(\frac{1}{z})\right), \nonumber \\ L(\lambda,z) & = & i\ \left(\frac{\kappa_-}{a_-}\right)^{\frac{2i\lambda}{\kappa_-}}\ \frac{\Gamma\left(\frac{1}{2}+\frac{i\lambda}{\kappa_-}\right)}{\Gamma\left(\frac{1}{2}-\frac{i\lambda}{\kappa_-}\right)}\ \left(\frac{z}{2}\right)^{-\frac{2i\lambda}{\kappa_-}} \left(1 + O(\frac{1}{z})\right).\end{aligned}$$ Proofs of the local inverse scattering results ============================================== \[LIS\] Proof of Theorem \[Reflection\], $(i) \Rightarrow (ii)$. -------------------------------------------------------- Assume that $L(\lambda,n) = \tilde{L}(\lambda,n) + \ O\left(e^{-2nB} \right)$, $n \rightarrow + \infty.$ Our first step is to extend these asymptotics (which are true for $n$ integer $\rightarrow + \infty$) to the case of $z \rightarrow +\infty$ ($z$ real and positive). To do this, we shall use some well-known uniqueness results for Laplace transforms obtained in [@Ho1; @Si]. We begin with an elementary result for functions of the complex variable belonging to the Hardy class. We recall (see for instance [@Lev], Lecture 19) that the Hardy class $H_+^2$ is the set of analytic functions $F$ in the right half-plane $\Omega = \{ z \in {\mathbb{C}}\ ,\ Re \ z >0 \}$, satisfying the condition $$\label{Hardy} \sup_{x >0} \ \int_{{\mathbb{R}}} \ \mid F(x+i y) \mid ^2 \ dy \ < \infty,$$ and equipped with the norm $$\label{norme} \mid \mid F \mid \mid = \left( \sup_{x >0} \ \int_{{\mathbb{R}}} \ \mid F(x+i y) \mid ^2 \ dy \right)^{\frac{1}{2}}.$$ The Paley-Wiener Theorem asserts that a function $F(z)$ belongs to the Hardy space $H_+^2$ if and only if there exists a function $f \in L^2(0,+\infty)$ such that $$\label{PaleyWiener} F(z) = \frac{1}{\sqrt{2\pi}}\ \int_0^{+\infty} e^{-tz} \ f(t) \ dt \ ,\ \forall z \in \Omega.$$ Moreover, we have $$\label{normeL2} \mid \mid F \mid \mid \ = \ \mid \mid f \mid \mid_{L^2(0,\infty)}.$$ Let us also recall a uniqueness result for Laplace transforms given in [@Ho1], Prop. 2.4., (see also [@Si] for a continuous version): \[uniciteLaplace\] Let $f \in L^1(0,a)$. If for all $\epsilon >0$, $$\int_0^a \ e^{-nt} f(t) \ dt \ =\ O(e^{-an(1-\epsilon)})\ , \ \ n \rightarrow +\infty,$$ then $f=0$ a.e. We now put together all the previous results and prove the following Proposition. \[PropHardy\] Let $F$ be a function in the Hardy class $H_+^2 $. Assume that for some $B>0$, we have $F(n) = O\left( e^{-Bn} \right), \ n \rightarrow +\infty$, ($n$ integer). Then, $$\label{majoration} \mid F(z) \mid \ \leq \ \frac{\mid \mid F \mid \mid}{\sqrt{4\pi Re z}} \ e^{-B Re z} \ , \ \forall z \in \Omega.$$ For $n\in {\mathbb{N}}$, the Paley-Wiener theorem and the Cauchy-Schwarz inequality imply $$\begin{aligned} \int_0^B e^{-nt} f(t) \ dt &=& \sqrt{2\pi} \ F(n) - \int_B^{+\infty} e^{-nt} f(t) \ dt \\ &=& \sqrt{2\pi} \ F(n) +O\left(e^{-nB} \right) \ = \ O\left(e^{-nB} \right).\end{aligned}$$ So, Proposition \[uniciteLaplace\] entails that $f = 0$ a.e in $(0,B)$. Using (\[PaleyWiener\]) again and (\[normeL2\]), we obtain at once (\[majoration\]). Let us give a direct consequence (which could be certainly improved) of the previous result to our inverse problem. \[estimL\] Assume that the reflection coefficients $L(\lambda, n)$ and $\tilde{L}(\lambda, n)$ satisfy for some $0 < B < \min(A,\tilde{A})$, $$L(\lambda,n) = \tilde{L}(\lambda,n) + O(e^{-2nB}) \ , \ n \rightarrow + \infty, \ \ n \ \textrm{integer}.$$ Then $$L(\lambda,z) = \tilde{L}(\lambda,z) + O(\sqrt{z} \ e^{-2zB}) \ , \ z \rightarrow + \infty, \ \ z \ \textrm{real}.$$ First, let us recall that $$L(\lambda, n) = \frac{a_{L3}(\lambda, n)}{a_{L1}(\lambda,n)} \ \ ,\ \ \tilde{L}(\lambda, n) = \frac{\tilde{a}_{L3}(\lambda, n)}{\tilde{a}_{L1}(\lambda,n)}.$$ Using Lemma \[MainEsti\], we obtain immediately $$\label{estim} a_{L3}(\lambda, n) \tilde{a}_{L1}(\lambda,n) - \tilde{a}_{L3}(\lambda, n) a_{L1}(\lambda, n) = O(e^{n(A+\tilde{A}-2B)}).$$ For $z \in \Omega$, we set $$\label{fonctionF} F(z) = \frac{ a_{L3}(\lambda, z) \tilde{a}_{L1}(\lambda,z) - \tilde{a}_{L3}(\lambda, z) a_{L1}(\lambda, z)} {z+1} \ e^{-z(A+\tilde{A})}.$$ Clearly, $F$ is holomorphic in $\Omega$, and by Lemma \[MainEsti\], we have $$\mid F(z) \mid \ \leq \ \frac{2}{\mid z+1 \mid}.$$ It follows that $F \in H_+^2$ and by (\[estim\]), we have $F(n) = O(e^{-2nB})$. Using Proposition \[PropHardy\], we see that $F(z) = O( z^{-\frac{1}{2}} e^{-2zB})$, $z \rightarrow +\infty$. For $z>0$, we write $$L(\lambda, z) - \tilde{L}(\lambda,z) = \frac{(z+1)\ e^{z(A+\tilde{A})}} {a_{L1}(\lambda, z) \ \tilde{a}_{L1}(\lambda,z)} \ F(z).$$ We conclude the proof using (\[asymptoticalun\]). This concludes the first step of the proof of Theorem \[Reflection\], $(i) \Rightarrow (ii)$. The second step of the proof consists in an adaption of the strategy used to prove the local Borg-Marchenko Theorem in [@Si; @Be] for one-dimensional Schrödinger operators to our setting of Dirac operators on a SSAHM. This strategy is relatively close to the proof of Theorem 1.1 given in [@DN3], itself inspired by [@FY]. Let us introduce for $X \in ]0,B[$ the matrix $$P(X,\lambda,z) = \left( \begin{array}{cc} P_1(X,\lambda,z) & P_2(X,\lambda,z) \\ P_3(X,\lambda,z) & P_4(X,\lambda,z) \end{array} \right),$$ defined by $$\label{matricepassage} P(X,\lambda,z) \ \tilde{F}_R (\tilde{h} (X), \lambda, z) \ = \ F_R (h(X), \lambda, z),$$ where $F_R =(f_{Rk})$ and $\tilde{F}_R =(\tilde{f}_{Rk})$ are the Jost solutions from the right associated with $a(x)$ and $\tilde{a}(x)$. To simplify the notations, for $k=1, ...,4$, we set as previously: $$\begin{aligned} f_{k} (X,\lambda, z) = f_{Lk} (h (X), \lambda, z), & \tilde{f}_{k} (X,\lambda, z) = \tilde{f}_{Lk} (\tilde{h} (X), \lambda, z),\\ g_{k} (X,\lambda, z) = f_{Rk} (h (X), \lambda, z), & \tilde{g}_{k} (X,\lambda, z) = \tilde{f}_{Rk} (\tilde{h} (X), \lambda, z).\end{aligned}$$ Using that det $F_R =1$ and det $\tilde{F}_R = 1$, we obtain the following equalities : $$\label{premiereformule} \left\{ \begin{array}{ccc} P_1(X,\lambda, z) &=& \ g_{1}\ \tilde{g}_{4} - \ g_{2} \ \tilde{g}_{3}, \\ P_2(X,\lambda, z) &=& - \ g_{1} \ \tilde{g}_{2} + \ g_{2} \ \tilde{g}_{1} . \end{array} \right.$$ It follows from (\[premiereformule\]) and the analytical properties of the Jost functions that, for $j=1,2$, the applications $z \rightarrow P_j(X,\lambda,z)$ are analytic on ${\mathbb{C}}$ and of exponential type. Moreover, by Lemma \[MainEstiF\], these applications are bounded on the imaginary axis $i {\mathbb{R}}$. We shall now prove that the applications $z \rightarrow P_j(X,\lambda,z)$ are also bounded on the real axis. To do this, we first perform some elementary algebraic transformations on $P_j(X,\lambda,z)$. Since $F_L(x,\lambda,z)=F_R (x, \lambda, z)\ A_L (\lambda, z)$, we easily get for $z>0$, $$\begin{aligned} g_1 &=& \frac{f_1}{a_{L1}} - L(\lambda, z) g_2, \\ \tilde{g}_3 &=& \frac{\tilde{f}_3}{\tilde{a}_{L1}} - \tilde{L}(\lambda, z) \tilde{g}_4 .\end{aligned}$$ Thus, $$P_1(X,\lambda, z) = (\tilde{L}(\lambda,z) -L(\lambda,z)) \ g_2 \tilde{g}_4 + \left( \frac{f_1 \tilde{g}_4}{a_{L1}} - \frac{ \tilde{f}_3 g_2}{\tilde{a}_{L1}} \right).$$ Using (\[estimatefg\]) and (\[asymptoticalun\]), it is easy to see that the function [$\displaystyle{ z \rightarrow \left( \frac{f_1 \tilde{g}_4}{a_{L1}} - \frac{ \tilde{f}_3 g_2}{\tilde{a}_{L1}} \right)}$]{} is bounded on ${\mathbb{R}}^+$ for all fixed $X\in ]0,A[$. Moreover, (\[estimatefg\]) and Proposition \[estimL\] imply $$\mid (\tilde{L}(\lambda,z) -L(\lambda,z))\ g_2 \tilde{g}_4 \mid \ \leq \ C \sqrt{z} e^{-2z (B-X)},$$ and thus, this term remains bounded when $z \to +\infty$ for all $X \in ]0,B[$. Summarizing, for all fixed $X \in ]0,B[$, the function $ z \rightarrow P_1 (X,\lambda, z)$ is bounded on ${\mathbb{R}}^+$. Similarly, we have $$P_2(X,\lambda, z) = (\tilde{L}(\lambda,z) -L(\lambda,z)) \ g_2 \tilde{g}_2 + \left(\frac{ \tilde{f}_1 g_2}{\tilde{a}_{L1}} - \frac{f_1 \tilde{g}_2}{a_{L1}}\right),$$ and using the same arguments as above, we obtain that, for all fixed $X \in ]0,B[$, $ z \rightarrow P_2 (X,\lambda, z)$ is bounded on ${\mathbb{R}}^+$. Clearly, these last results remain true on ${\mathbb{R}}$ by an elementary parity argument. Finally, applying the Phragmen-Lindelöf’s Theorem ([@Bo], Thm 1.4.2.) on each quadrant of the complex plane, we deduce that $z \rightarrow P_j (X,\lambda,z)$ is bounded on ${\mathbb{C}}$. By Liouville’s Theorem, and a standard continuity argument in the variable $X$, we have thus obtained $$\label{thliouville} P_j (X,\lambda,z)=P_j (X,\lambda,0) \ \ ,\ \ \forall z \in {\mathbb{C}}\ ,\ \forall X \in ]0,B].$$ Now, we return to the definition of $P_j(X,\lambda,z)$ for $z=0$. We observe first that ${\displaystyle{F_R (x,\lambda, 0) = e^{i \lambda \Gamma^1 x} }} $ and similarly ${\displaystyle{\tilde{F}_R (x,\lambda, 0) = e^{i \lambda \Gamma^1 x} }} $. This is immediate from the definition of the Jost functions. Thus we deduce from (\[matricepassage\]) that $$\label{egalitemat} P(X,\lambda, 0) = e^{i \lambda \ ( h(X)-\tilde{h}(X))\ \Gamma^1 }.$$ Then, putting (\[egalitemat\]) and (\[thliouville\]) into (\[matricepassage\]) we get $$\label{egalitejost} \left\{ \begin{array}{ccc} \tilde{g}_{1}(X,\lambda, z) &=& e^{i \lambda \ (\tilde{h}(X) - h (X))} \ g_{1}(X,\lambda, z), \\ \tilde{g}_{2}(X,\lambda, z) &=& e^{i \lambda \ (\tilde{h}(X) - h (X))} \ g_{2}(X,\lambda, z). \end{array} \right.$$ By Lemma 4.2 in [@DN3], the Wronskians $W(g_{1} , g_{2}) = W(\tilde{g}_{1} , \tilde{g}_{2}) = iz$. Then, a straightforward calculation gives $$\label{egalitephase} e^{2i \lambda \ (\tilde{h}(X) - h (X))} \ =\ 1.$$ Thus, by a standard continuity argument, there exists $k \in {\mathbb{Z}}$ such that $$\label{diffeos} \tilde{h}(X)= h(X) + \frac{k \pi}{\lambda} \ \ ,\ \ \forall X \in ]0,B].$$ Note that, for the particular choice $X=B$, we obtain ${\displaystyle{\tilde{h}(B)= h(B) + \frac{k \pi}{\lambda}}}$. Let us differentiate (\[diffeos\]) with respect to $X$. We obtain easily $$\frac{1}{a(\tilde{h}(X))} = \frac{1}{a(h(X))},$$ and using again (\[diffeos\]), we have $$\label{UnicitePot} a(x) = \tilde{a}(x + \frac{k \pi}{\lambda}) \ \ , \ \ \forall x \in ]-\infty,h(B)].$$ Thus, we have proved the first part of Theorem \[Reflection\]. $\Box$ Proof of Theorem \[Reflection\], $(ii) \Rightarrow (i)$. -------------------------------------------------------- Let us assume there exists $k \in {\mathbb{Z}}$ such that $a(x)= \tilde{a}(x + \frac{k \pi}{\lambda}), \ \forall x \leq h(B)$. It follows immediately from the definition of the diffeomorphisms $h$ and $\tilde{h}$, that ${\displaystyle{\tilde{h}(B)= h(B) + \frac{k \pi}{\lambda}}}$. Moreover, if we set $\breve{a}(x) = \tilde{a}(x + \frac{k \pi}{\lambda}), \ \forall x \in {\mathbb{R}}$, and using (\[Sdecale\]), we see that (with obvious notation), $$\breve{L}(\lambda,n) = e^{-2i\lambda \ \frac{k \pi}{\lambda}} \ \tilde{L}(\lambda, n) = \tilde{L}(\lambda, n).$$ Thus, it remains to prove the implication $(ii) \Rightarrow (i)$ in the case $k=0$. Now, let us begin with an obvious lemma (whose proof is omitted) : Assume that $a(x) = \tilde{a}(x), \ \forall x \leq h(B)=\tilde{h}(B)$. Then, $$\label{uniciteparam} a_- = \tilde{a}_- \ ,\ \kappa_- = \tilde{\kappa}_-.$$ and $$\label{uniciteg} g_j(X,\lambda,z) = \tilde{g}_j(X,\lambda, z), \ \forall X \leq B, \ \forall j=1, \ldots 4.$$ Using again the relation $F_L(x,\lambda,z) = F_R(x,\lambda,z) A_L(\lambda, z)$ and $(\ref{uniciteg})$, we have for $z>0$ and $X \leq B$, $$\frac{f_1}{a_{L1}} \ =\ g_1 +L(\lambda,z)\ g_2 \ ,\ \frac{\tilde{f}_1}{\tilde{a}_{L1}} \ =\ g_1 +\tilde{L}(\lambda,z)\ g_2.$$ For $z>0$ large enough, (\[asymgdeux\]) implies that $g_2 \not=0$. So, for such $z$, we can write $$L(\lambda,z) - \tilde{L}(\lambda,z) \ =\ \frac{1}{g_2} \left( \frac{f_1}{a_{L1}} - \frac{\tilde{f}_1}{\tilde{a}_{L1}} \right).$$ Now, using Theorem \[asymtoticsutiles\], we obtain easily: $$L(\lambda,z) - \tilde{L}(\lambda,z) \ =\ O \left( e^{-2zX} \right) \ ,\ \forall X \in ]0, B],$$ and taking $X=B$, the proof is complete. $\Box$ Proof of Theorem \[Reflection\], $(iii) \Leftrightarrow (iv)$. -------------------------------------------------------------- The local uniqueness result for the reflection coefficient $R(\lambda,n)$ is actually a by-product of the previous one using the following trick. If we set $a^{\star}(x) = a(-x)$, a straightforward calculation using (\[IE-FL\]) - (\[IE-FR\]) shows that the associated Jost solutions satisfy $$\label{egalitestar} \left\{ \begin{array}{ccc} F_{R}^{\star}(x,\lambda,n) &=& F_{L}(-x,-\lambda,-n), \\ F_{L}^{\star}(x,\lambda,n) &=& F_{R}(-x,-\lambda,-n). \\ \end{array} \right.$$ It follows immediately that $A_{L}^{\star}(\lambda,n) = A_{L}^{-1}(-\lambda,-n)$ which implies the equality $R^{\star}(\lambda,n)= - \overline{L(-\lambda,n)}$. Thus, it suffices to use the previous result for the reflection coefficients $L$, with $\lambda$ replaced by $-\lambda$, to prove the equivalence $(iii) \Leftrightarrow (iv)$ of Theorem \[Reflection\]. $\Box$ Proof of Theorem \[Transmission\]. ---------------------------------- Assume that $$T(\lambda,n) = \tilde{T}(\lambda,n) + \ O\left(e^{-2nB} \right),$$ with $B > \max(A,A')$. Using the asymptotics in Theorem \[asymtoticsutiles\], we obtain $A=\tilde{A}$ and $$a_{L1}(\lambda,n) - \tilde{a}_{L1}(\lambda,n) = O\left(e^{-2n(B-A)} \right).$$ Now, we set for $z \in \Omega$, $$F(z) \ =\ \frac{ a_{L1}(\lambda,z) - \tilde{a}_{L1}(\lambda,z)}{z+1} \ e^{-zA}.$$ As previously, we see that $F$ belongs to the Hardy space $H_+^2$ and $F(n) = O\left(e^{-(2B-A)n} \right)$. By Proposition \[PropHardy\], we have $$\label{majorationF} \mid F(z) \mid \ \leq \ \frac{\mid \mid F \mid \mid}{\sqrt{4 \pi Re z}} \ e^{-(2B-A) Re z} \ , \ \forall z \in \Omega.$$ It follows that there exists $C>0$ such that for all $z>0$, $$\mid a_{L1}(\lambda,z) - \tilde{a}_{L1}(\lambda,z) \mid \ \leq \ C \sqrt{z} \ e^{-2z(B-A)}.$$ Thus, $f(z) := a_{L1}(\lambda,z) - \tilde{a}_{L1}(\lambda,z)$ is bounded on ${\mathbb{R}}^+$. Moreover, this function is of exponential type, and bounded on $i{\mathbb{R}}$. The Phragmen - Lindelöf Theorem implies that $f$ is bounded on $\Omega$ and consequently, is also bounded on ${\mathbb{C}}$ using parity arguments. Hence Liouville’s Theorem entails that $f(z) = f(0) = 0$. We conclude the proof using Proposition \[unicitetransmission\]. $\Box$. Inverse uniqueness results in Reissner-Nordström-de-Sitter black holes ====================================================================== \[BH\] In this Section, we adapt the previous local inverse uniqueness results to the setting of general relativity and more precisely to Reissner-Nordström-de-Sitter black holes. We emphasize that the link between such black holes and SSAHM was already given in [@DN3]. Considering the scattering of massless Dirac fields evolving in the outer region of a RN-dS black holes, we shall prove that the partial knowledge of the corresponding reflection coefficients in the sense of (\[Ln\]) - (\[Rn\]) not only determines the metric of such black holes in the neighbourhood of the event and cosmological horizons (see below for the definition), but in fact determines the whole metric. This is due to the fact that the metric of RN-dS black holes only depend on $3$ parameters - their mass, electric charge and positive cosmogical constant - parameters that can be deduced from the explicit form of the metric in the neighbourhoods of the horizons. Reissner-Nordstöm-de-Sitter black holes --------------------------------------- Refering to Wald [@W] for more general details on black hole spacetimes, we summarize here the essential features of Reissner-Nordström-de-Sitter (RN-dS) black holes given in [@DN2; @DN3]. First, RN-dS are spherically symmetric electrically charged exact solutions of the Einstein-Maxwell equations. In Schwarzschild coordinates, the exterior region of a RN-dS black hole is described by the four-dimensional manifold ${\mathcal{M}}= \mathbb{R}_{t} \times ]r_-,r_+[_r \times {\mathbb{S}}_{\theta,\varphi}^{2}$ equipped with the Lorentzian metric $$\label{Metric} \tau = F(r)\,dt^{2} - F(r)^{-1} dr^{2} - r^{2} \big(d\theta^{2}+\sin^{2}\theta \, d\varphi^{2}\big),$$ where $$\label{F} F(r) = 1-\frac{2M}{r}+\frac{Q^{2}}{r^{2}} - \frac{\Lambda}{3} r^2.$$ The constants $M>0$, $Q \in {\mathbb{R}}$ appearing in (\[F\]) are interpreted as the mass and the electric charge of the black hole and $\Lambda > 0$ is the cosmological constant of the universe. We assume here that the function $F(r)$ has three simple positive roots $0<r_c<r_-<r_+$ and a negative one $r_n <0$. This is always achieved if we suppose for instance that $Q^2<\frac{9}{8} M^2$ and that $\Lambda M^2$ be small enough (see [@L]). The sphere $\{r=r_c\}$ is called the Cauchy horizon whereas the spheres $\{r=r_-\}$ and $\{r=r_+\}$ are the event and cosmological horizons respectively. These horizons which appear as singularities of the metric (\[Metric\]) are in fact mere coordinates singularities. This means that using appropriate coordinates system, these horizons can be understood as regular null hypersurfaces that can be crossed one way but would require speeds greater than that of light to be crossed the other way: hence their names horizons. In what follows, we shall only consider the exterior region of the black hole, that is the region $\{r_- < r <r_+\}$ lying between the event and cosmological horizons. Note that the function $F$ is positive there. The point of view implicitly adopted here is indeed that of static observers located far from the event and cosmological horizons of the black hole. We think typically of a telescope on earth aiming at the black hole or at the cosmological horizon. We understand these observers as living on worldlines $\{r = r_0\}$ with $r_- << r_0 << r_+$. The variable $t$ corresponds to their true perception of time. From the point of view of our static observers, the event and cosmological horizons turn out to be the boundaries of the *observable* world. This can be more easily understood if we remark that the event and cosmological horizons are never reached in a finite time $t$ by incoming and outgoing radial null geodesics, the trajectories followed by classical light-rays aimed radially at the black hole or at the cosmological horizon. Both horizons are thus perceived as *asymptotic regions* by our static observers. Instead of working with the radial variable $r$, we make the choice to describe the exterior region of the black hole by using the Regge-Wheeler (RW) radial variable which is more natural when studying the scattering properties of any fields. The RW variable $x$ is defined implicitly by $\frac{dx}{dr} = F^{-1}(r)$, or explicitly by $$\label{RW} x = \frac{1}{2\kappa_n} \ln(r-r_n) + \frac{1}{2\kappa_c} \ln(r-r_c) +\frac{1}{2\kappa_-} \ln(r-r_-) + \frac{1}{2\kappa_+} \ln(r_+-r) \ + \ c,$$ where $c$ is any constant of integration and the quantities $\kappa_j, \ j=n,c,-,+$ are defined by $$\label{SurfaceGravity} \kappa_n = \frac{1}{2} F'(r_n), \ \kappa_c = \frac{1}{2} F'(r_c), \ \kappa_- = \frac{1}{2} F'(r_-), \ \kappa_+ = \frac{1}{2} F'(r_+).$$ The constants $\kappa_->0$ and $\kappa_+<0$ are called the surface gravities of the event and cosmological horizons respectively. Note from (\[RW\]) that the event and cosmological horizons $\{r=r_\pm\}$ are pushed away to the infinities $\{x = \pm \infty\}$ using the RW variable. Let us also emphasize that the incoming and outgoing null radial geodesics become straight lines $\{x=\pm t\}$ in this new coordinates system, a fact that provides a natural manner to define the scattering data simply by mimicking the usual definitions in Minkowski-spacetime. The Dirac equation in RN-dS --------------------------- As waves, we consider massless Dirac fields propagating in the exterior region of a RN-dS black hole. We refer to [@Me1; @Ni] for a detailed study of this equation in this background including a complete time-dependent scattering theory. We shall use the expression of the equation obtained in these papers as the starting point of our study. Thus the considered massless Dirac fields are represented by 2 components spinors $\psi$ belonging to the Hilbert space $L^2({\mathbb{R}}\times {\mathbb{S}}^2; \, {\mathbb{C}}^2)$ which satisfy the evolution equation $$\label{FullEq} i \partial_{t} \psi = \Big({\Gamma^1}D_x + a(x) D_{{\mathbb{S}}^2} \Big) \psi,$$ The symbol $D_x$ stands for $-i{\partial}_x$ whereas $D_{{\mathbb{S}}^2}$ denotes the Dirac operator on ${\mathbb{S}}^2$ which, in spherical coordinates, takes the form (\[DiracSphere\]). The potential $a$ is the scalar smooth function given in term of the metric (\[Metric\])-(\[F\]) by $$\label{Potential} a(x) = \frac{\sqrt{F(r(x))}}{r(x)},$$ where $r(x)$ is the inverse diffeomorphism of (\[RW\]). It was shown in [@DN3] that the potential $a$ verifies the hypotheses (\[AsympA\]). Finally, the matrices ${\Gamma^1}, {\Gamma^2}, {\Gamma^3}$ appearing in (\[FullEq\]) and (\[DiracSphere\]) are usual $2 \times 2$ Dirac matrices that satisfy the anticommutation relations (\[AntiCom\]). As before, we shall work with the following representations of the Dirac matrices $$ {\Gamma^1}= \left( \begin{array}{cc} 1&0 \\0&-1 \end{array} \right), \quad {\Gamma^2}= \left( \begin{array}{cc} 0&1 \\1&0 \end{array} \right), \quad {\Gamma^3}= \left( \begin{array}{cc} 0&i \\-i&0 \end{array} \right).$$ Hence, the massless Dirac equation on the exterior region of a RN-dS black hole can be put under the same form as the massless Dirac equation on a SSHAM studied in the previous Sections. We can thus define the transmission coefficients $T(\lambda,n)$ and reflection coefficients $L(\lambda,n)$ and $R(\lambda,n)$ for a fixed energy $\lambda \in {\mathbb{R}}$ and all angular momenta $n \in {\mathbb{N}}^*$ as before. Moreover, Theorem \[Reflection\] remains true in this new setting. Taking advantage of the particular form of the potential $a$ given in (\[Potential\]), we can slightly improve these results. Uniqueness of the parameters ---------------------------- Using the notations of the Theorem \[Reflection\], the following assertions are equivalents : $$(i) \quad L(\lambda,n) = \tilde{L}(\lambda,n) + \ O\left(e^{-2nB} \right).$$ $$(ii) \quad R(\lambda,n) = \tilde{R}(\lambda,n) + \ O\left(e^{-2nB} \right).$$ $$(iii) \quad M=\tilde{M}, \quad Q^{2} = \tilde{Q}^{2} \quad \mathrm{and} \quad \Lambda = \tilde{\Lambda}.$$ $$(iv) \quad \exists k \in {\mathbb{Z}}, \quad a(x) = \tilde{a}(x + \frac{k\pi}{\lambda}), \quad \forall \ x \in \mathbb{R}.$$ We first use Theorem \[Reflection\] to obtain the equality of the potential $a$ and $\tilde{a}$ on a half-line $]- \infty,b_1]$ (respectively $[b_2,+\infty[$), $b_1 < b_2 \in \mathbb{R}$. Then, a line by line inspection of the proof given in [@DN3] p. 43-44 shows that this information is enough to prove the uniqueness of the mass $M$, the square of the charge $Q^2$ and the cosmological constant $\Lambda$ of the black hole. Finally, since the parameters of the black hole determine uniquely the metric, we obtain the equality of the potentials $a$ and $\tilde{a}$ on $\mathbb{R}$ (up to a discrete set of translations as stated in (iv)). Other formulation of the main inverse uniqueness results ======================================================== In this Section, we formulate our main Theorems \[Reflection\] in a more global way, avoiding the use of a decomposition onto generalized spherical harmonics. More precisely, we replace the main assumptions (\[Ln\]) and (\[Rn\]) by $L^2({\mathbb{S}}^2)$-operator norms conditions on the global reflection coefficients. We recall first the definition and essential properties of these operators. \[TkRkLk\] For all $(n,k) \in I$ where $I = \{ n \in {\mathbb{N}}^*, \ k \in 1/2 + {\mathbb{Z}}, \ |k| \leq n - \frac{1}{2}\}$, we use the notation $Y_{kn} = (Y_{kn}^1, Y_{kn}^2)$ for the corresponding generalized spherical harmonics. Then,\ 1) The families $\{Y_{kn}^1\}_{(n,k) \in I}$ and $\{Y_{kn}^2\}_{(n,k) \in I}$ form Hilbert bases of $\mathfrak{l} = L^2({\mathbb{S}}^2; {\mathbb{C}})$; precisely for all $\psi \in \mathfrak{l}$, we can decompose $\psi$ as $$\psi = \sum_{n,k \in I} \psi_{kn}^{j} Y_{kn}^{j}, \quad j=1,2,$$ with $$\| \psi \|^2 = \frac{1}{2} \sum_{n,k \in I} |\psi_{kn}^{j} |^2.$$ 2\) Let $\lambda \in {\mathbb{R}}$ be a fixed energy. Then, the transmission operators $T_L(\lambda)$ and $T_R(\lambda)$ are defined as operators from $\mathfrak{l}$ to $\mathfrak{l}$ as follows. For all $\psi = \sum_{n,k \in I} \psi_{kn}^{j} Y_{kn}^{j} \in \mathfrak{l}$ $$\label{TL} T_L(\lambda) \psi = T_L(\lambda) \left( \sum_{n,k \in I} \psi_{kn}^1 Y_{kn}^1(\lambda) \right) = \sum_{n,k \in I} \left( T(\lambda,n) \psi_{kn}^1 \right) Y_{kn}^1(\lambda),$$ and $$\label{TR} T_R(\lambda) \psi = T_R(\lambda) \left( \sum_{n,k \in I} \psi_{kn}^2 Y_{kn}^2(\lambda) \right) = \sum_{n,k \in I} \left( T(\lambda,n) \psi_{kn}^2 \right) Y_{kn}^2(\lambda),$$ where $T(\lambda,n)$ are the transmission coefficients defined in (\[SR-SM1\]) - (\[SR-SM2\]). In short, we write $$\label{TLR} T_L(\lambda) Y_{kn}^1 = T(\lambda,n) Y_{kn}^1, \quad \quad T_R(\lambda) Y_{kn}^2 = T(\lambda,n) Y_{kn}^2, \quad \forall n,k \in I,$$ and thus the operators $T_L(\lambda)$ (resp. $T_R(\lambda)$) are diagonalizable on the Hilbert basis of eigenfunctions $(Y_{kn}^1)_{k,n \in I}$ (resp. $(Y_{kn}^2)_{k,n \in I}$) associated to the eigenvalues $T(\lambda,n)$ (in both cases).\ 3) Let $\lambda \in {\mathbb{R}}$ be a fixed energy. Then, the reflection operators $L(\lambda)$ and $R(\lambda)$ are defined as operators from $\mathfrak{l}$ to $\mathfrak{l}$ as follows. For all $\psi = \sum_{n,k \in I} \psi_{kn}^{j} Y_{kn}^{j} \in \mathfrak{l}$ $$\label{R} R(\lambda) \psi = R(\lambda) \left( \sum_{n,k \in I} \psi_{kn}^2 Y_{kn}^2(\lambda) \right) = \sum_{n,k \in I} \left( R(\lambda,n) \psi_{kn}^2 \right) Y_{kn}^1(\lambda),$$ and $$\label{L} L(\lambda) \psi = L(\lambda) \left( \sum_{n,k \in I} \psi_{kn}^1 Y_{kn}^1(\lambda) \right) = \sum_{n,k \in I} \left( L(\lambda,n) \psi_{kn}^1 \right) Y_{kn}^2(\lambda),$$ where $R(\lambda,n)$ and $L(\lambda,n)$ are defined in (\[SR-SM1\]) - (\[SR-SM2\]). In short, we write $$\label{Rek} R(\lambda) Y_{kn}^2 = R(\lambda,n) Y_{kn}^1, \quad \quad L(\lambda) Y_{kn}^1 = L(\lambda,n) Y_{kn}^2, \quad \forall n,k \in I.$$ It is immediate from the above definitions to express the $L^2({\mathbb{S}}^2)$-operator norms of the transmission operators $T_L(\lambda), T_R(\lambda)$ and of the reflection operators $L(\lambda), R(\lambda)$ in terms of the coefficients $T(\lambda,n), L(\lambda,n)$ and $R(\lambda,n)$. For a fixed $\lambda \in {\mathbb{R}}$, we have $$\label{LinkTLR} \| T_L(\lambda)\| = \|T_R(\lambda)\| = \| T(\lambda,n) \|_\infty, \quad \|L(\lambda)\| = \| L(\lambda,n) \|_\infty, \quad \|R(\lambda)\| = \| R(\lambda,n) \|_\infty.$$ To reformulate the assumptions (\[Ln\]) and (\[Rn\]) by $L^2({\mathbb{S}}^2)$-operator norms conditions, we observe that the selfadjoint operator $|{\mathbb{D}_{{\mathbb{S}}^2}}|$ acts as multiplication by $n$ on each generalized spherical harmonics $Y_{kn}$ in the Hilbert decomposition $L^2({\mathbb{S}}^2, {\mathbb{C}}^2) = \oplus_{n,k \in I} {\mathbb{C}}^2 \otimes Y_{kn}$. We still denote by $|{\mathbb{D}_{{\mathbb{S}}^2}}|$ the restriction of this operator to $\mathfrak{l} = L^2({\mathbb{S}}^2, {\mathbb{C}})$ and thus, $|{\mathbb{D}_{{\mathbb{S}}^2}}|$ acts as multiplication by $n$ on each generalized spherical harmonics $Y^1_{kn}$ or $Y^2_{kn}$ in the two Hilbert decompositions $\mathfrak{l} = \oplus_{n,k \in I} {\mathbb{C}}\otimes Y^j_{kn}, \ j=1,2$. Now, let $\lambda \in {\mathbb{R}}$ and $0 < B < \min(A,\tilde{A})$. Then, using (\[LinkTLR\]), the assumptions (\[Ln\]) and (\[Rn\]) for the reflection coefficients can be written as $$\label{L-OpNorm} (\ref{Ln}) \quad \Longleftrightarrow \quad \left\| e^{2B|{\mathbb{D}_{{\mathbb{S}}^2}}|} \left(L(\lambda) - \tilde{L}(\lambda) \right) \right\|_{{\mathcal{B}}(\mathfrak{l})} = O(1),$$ and $$\label{R-OpNorm} (\ref{Rn}) \quad \Longleftrightarrow \quad \left\| e^{2B|{\mathbb{D}_{{\mathbb{S}}^2}}|} \left(R(\lambda) - \tilde{R}(\lambda) \right) \right\|_{{\mathcal{B}}(\mathfrak{l})} = O(1).$$ Similarly, let $B > \max(A,\tilde{A})$. Then we get for the assumption (\[Tn\]) on the transmission coefficients the equivalences $$\begin{aligned} \label{T-OpNorm} (\ref{Tn}) \ & \Longleftrightarrow & \ \left\| e^{2B|{\mathbb{D}_{{\mathbb{S}}^2}}|} \left(T_L(\lambda) - \tilde{T_L}(\lambda) \right) \right\|_{{\mathcal{B}}(\mathfrak{l})} = O(1), \\ & \Longleftrightarrow & \ \left\| e^{2B|{\mathbb{D}_{{\mathbb{S}}^2}}|} \left(T_R(\lambda) - \tilde{T_R}(\lambda) \right) \right\|_{{\mathcal{B}}(\mathfrak{l})} = O(1). \nonumber \end{aligned}$$ Addendum on the inverse scattering problem from the transmission coefficients $T(\lambda,n)$ ============================================================================================ In [@DN3], Theorem 1.1, it is claimed that the knowledge of the transmission coefficients $T(\lambda,n)$ for a fixed $\lambda \ne 0$ and for all $n \in \mathcal{L}$ where $\mathcal{L}$ is a subset of ${\mathbb{N}}$ satisfying the Müntz condition $$\sum_{n \in \mathcal{L}} \frac{1}{n} = +\infty,$$ also determines uniquely the function $a(x)$ up to a translation. The crucial ingredient of the proof can be found in the Proposition 3.13 of [@DN3] which states that\ “If $T(\lambda,n) = \tilde{T}(\lambda,n)$ for all $n \in \mathcal{L}$, then the corresponding reflection coefficients $L(\lambda,n)$ and $\tilde{L}(\lambda,n)$ (resp. $R(\lambda,n)$ and $\tilde{R}(\lambda,n)$) coincide up to a multiplicative constant”.\ The proof of this result given in [@DN3] is unfortunately incomplete. In fact, this last point is not so clear and could even be false. We shall try in this Appendix to give some insights of what happens when we try to determine the metric from the transmission coefficient $T(\lambda,n)$. We first give a correct version of the above result that is weaker than the Proposition 3.13. given in [@DN3]. \[Transmission1\] Let $(\Sigma, g)$ ans $(\tilde{\Sigma}, \tilde{g})$ be two SSAHM whose metrics depend on the functions $a(x)$ and $\tilde{a}(x)$ satisfying the assumptions (\[RegulA\]) - (\[AsympA\]). For a fixed energy $\lambda \ne 0$, consider the corresponding countable family of transmission coefficients $T(\lambda,n)$ and $\tilde{T}(\lambda,n)$ for all $n \in {\mathbb{N}}^*$. Consider also a subset $\mathcal{L}$ of ${\mathbb{N}}^*$ that satisfies a Müntz condition $\displaystyle\sum_{n \in \mathcal{L}} \frac{1}{n} = \infty$. Assume that $$T(\lambda,n) = \tilde{T}(\lambda,n), \quad \forall n \in \mathcal{L}.$$ Then $$T(\lambda,z) = \tilde{T}(\lambda,z), \quad \forall z \in {\mathbb{C}}.$$ Assume moreover that $\frac{1}{\kappa_+} + \frac{1}{\kappa_-} < 0$. - If $\frac{1}{\tilde{\kappa_+}} + \frac{1}{\tilde{\kappa_-}} < 0$, there exists a rational function $g(z)$ such that $$L(\lambda,z) = g(z) \tilde{L}(\lambda,z), \quad \forall z \in {\mathbb{C}}.$$ - If $\frac{1}{\tilde{\kappa_+}} + \frac{1}{\tilde{\kappa_-}} > 0$ and $\left( \frac{\tilde{a_-}}{a_+} \right)^{\frac{i \lambda}{\kappa_+}} = \left( \frac{\tilde{a_+}}{a_-} \right)^{\frac{i \lambda}{\kappa_-}}$, there exists a rational function $h(z)$ such that $$L(\lambda,z) = h(z) \tilde{R}(\lambda,z), \quad \forall z \in {\mathbb{C}}.$$ By definition of the transmission coefficients and using Corollary 3.9. and Theorem 3.10. in [@DN3], our assumption implies that $$T(\lambda,z) = \tilde{T}(\lambda,z), \quad \forall z \in {\mathbb{C}},$$ or equivalently $$\label{rr0} a_{L1}(\lambda,z) = \tilde{a}_{L1}(\lambda,z), \quad \forall z \in {\mathbb{C}}.$$ Now, we set ${\displaystyle{f(z)= \frac{a_{L3}(\lambda,z)}{z}}}$. Using that $a_{L3}(\lambda,0)=0$, we see that $f(z)$ is an even entire function of order $1$ thanks to Lemma \[AL-Analytic\]. Thus, we can write $f(z)=g(z^2)$ where $g$ is an entire function of order $\frac{1}{2}$. Using Hadamard’s factorization Theorem, we obtain the following expression for $f$ $$\label{fz} f(z) = G \, z^{2m}\ \prod_{n=1}^\infty \Big( 1 - \frac{z^2}{z_n^2} \Big),$$ where $2m$ is the multiplicity of $0$, $G$ is a constant and the $z_n$ are the zeros of $f$ belonging to $C^+ = \{z \in {\mathbb{C}}, \ \Im(z) > 0, \ \textrm{or} \ \Im(z) = 0, \ \Re(z) > 0\}$ counted according to their multiplicity. From (\[rr0\]) and Lemma \[AL-Analytic\], (iii), we have $$f(z)\overline{f(\bar{z})} = \tilde{f}(z)\overline{\tilde{f}(\bar{z})},$$ where ${\displaystyle{\tilde{f}(z)= \frac{\tilde{a_{L3}}(\lambda,z)}{z}}}$. Thus we get $$\mid G\mid^2 \ z^{4m} \ \prod_{n=1}^\infty \Big( 1 -\frac{z^2}{z_n^2} \Big) \Big( 1 -\frac{z^2}{\bar{z_n}^2} \Big) = \ \mid \tilde{G}\mid^2 \ z^{4\tilde{m}} \ \prod_{n=1}^\infty \Big( 1 -\frac{z^2}{\tilde{z_n}^2} \Big) \Big( 1 -\frac{z^2}{\bar{\tilde{z_n}}^2} \Big).$$ It follows that $\mid G \mid =\mid \tilde{G}\mid , \ m=\tilde{m}$ and $$\label{rr1} z_n = \pm \tilde{z_n} \ \textrm{or} \ z_n = \pm \overline{\tilde{z_n}}, \quad \forall n \in {\mathbb{N}}^*.$$ 1\) The equation (\[rr1\]) is where we made an error in [@DN3], Proposition 3.13. since we asserted that $$z_n = \tilde{z_n}, \quad \forall n \in {\mathbb{N}}^*.$$ 2) If $\Im(z_n) > 0$, then we must have $$\label{rr2} z_n = \tilde{z_n} \ \textrm{or} \ z_n = - \overline{\tilde{z_n}}, \quad \forall n \in {\mathbb{N}}^*.$$ Hence, the zeros $z_n$ and $\tilde{z_n}$ with positive imaginary parts coincide *up to “-” complex conjugation*. On the other hand, if $\Im(z_n) = 0$ and $\Re(z_n) > 0$, then $$\label{rr3} z_n = \tilde{z_n}, \textrm{or} \ z_n = \overline{\tilde{z_n}},$$ holds. In some cases, we can prove that the large zeros $z_n$ and $\tilde{z_n}$ coincide using the asymptotics of $a_{L3}(\lambda,z)$ for large $z$ in the complex plane. We shall use \[LargeAsymp-aL3\] For $|z|$ large in the complex plane, we have $$\begin{array}{rl} a_{L1}(\lambda,z) = & \frac{1}{2\pi} \left( - \frac{\kappa_+}{a_+} \right)^{\frac{i \lambda}{\kappa_+}} \left( \frac{\kappa_-}{a_-} \right)^{-\frac{i\lambda}{\kappa_-}} \Gamma(1-\nu_+) \Gamma(1-\mu_-) \left( \frac{z}{2} \right)^{i \lambda \left(\frac{1}{\kappa_-}-\frac{1}{\kappa_+} \right)} \\ & \times \left( e^{zA} + e^{-zA} e^{-\mathrm{sg}(\mathrm{Im}(z)) \pi \lambda \left(\frac{1}{\kappa_+} - \frac{1}{\kappa_-} \right)} \right) \ \left( 1 + O(\frac{1}{z}) \right),\\ a_{L3}(\lambda,z) = & \frac{i}{2\pi} \left( - \frac{\kappa_+}{a_+} \right)^{\frac{i \lambda}{\kappa_+}} \left( \frac{\kappa_-}{a_-} \right)^{\frac{i \lambda}{\kappa_-}} \Gamma(1-\nu_+) \Gamma(1-\nu_-) \left( \frac{z}{2} \right)^{-i\lambda \left(\frac{1}{\kappa_-}+\frac{1}{\kappa_+} \right)} \\ & \times \left( e^{zA} + e^{-zA} e^{\mathrm{sg}(\mathrm{Im}(z)) i \pi \left(1+ i\lambda \left( \frac{1}{\kappa_+} + \frac{1}{\kappa_-} \right) \right)} \right) \ \left( 1 + O(\frac{1}{z}) \right). \end{array}$$ We refer to [@DaKaNi; @FY] where similar asymptotics have been obtained. Using Rouche’s Theorem and a standard argument (see [@FY]), we obtain from Lemma \[LargeAsymp-aL3\] the following asymptotics for the large zeros $z_n$ with positive imaginary part. \[Asymp-Large-zn\] There exists $p \in {\mathbb{Z}}$ such that for large $n$, we have $$z_n = i \frac{\pi}{A} ( n + p) - \frac{\lambda \pi}{2A} \left( \frac{1}{\kappa_+} + \frac{1}{\kappa_-} \right) + O(\frac{1}{n}).$$ We conclude from Corollary \[Asymp-Large-zn\] and the previous Remark that there exists $N \in {\mathbb{N}}$ large enough such that for all $n > N$, we have $$z_n = \tilde{z_n} \ \textrm{or} \ z_n = - \overline{\tilde{z_n}}.$$ Assume from now on that $\frac{1}{\kappa_+} + \frac{1}{\kappa_-} < 0$. We conclude that the large zeros of $a_{L3}(\lambda, z)$ with positive imaginary part are located in the quadrant $I = \{z \in {\mathbb{C}}, \ \Re(z) > 0, \ \Im(z) > 0 \}$. By parity, the zeros with negative imaginary part are located in the quadrant $III = \{z \in {\mathbb{C}}, \ \Re(z) < 0, \ \Im(z) < 0 \}$. Since the $\tilde{z_n}$’s with positive imaginary part also satisfy the asymptotics in Corollary \[Asymp-Large-zn\], we get the following dichotomy. - If $\frac{1}{\tilde{\kappa}_+} + \frac{1}{\tilde{\kappa}_-} < 0$, then the zeros $\tilde{z_n}$’s with positive imaginary part are located in the quadrant $I$. Hence, using (\[rr2\]), we have the following. There exists a $N \in {\mathbb{N}}$ such that $$\label{rr4} z_n = \tilde{z_n}, \quad \forall n > N.$$ Using (\[rr0\]), Lemma \[LargeAsymp-aL3\] and Corollary \[Asymp-Large-zn\], we get in this case $$\frac{1}{\tilde{\kappa}_-} - \frac{1}{\tilde{\kappa}_+} = \frac{1}{\kappa_-} + \frac{1}{\kappa_+}, \quad \frac{1}{\tilde{\kappa}_-} + \frac{1}{\tilde{\kappa}_+} = \frac{1}{\kappa_-} + \frac{1}{\kappa_+},$$ which gives $$\tilde{\kappa}_- = \kappa_-, \quad \tilde{\kappa}_+ = \kappa_+.$$ Also, we use (\[fz\]) and (\[rr4\]) to obtain $$\label{rr5} f(z) = \frac{G}{\tilde{G}} \prod_{n = 1}^N \frac{\Big( 1 -\frac{z^2}{z_n^2} \Big)}{\Big( 1 -\frac{z^2}{\tilde{z_n}^2} \Big)} \tilde{f}(z).$$ Finally, denote by $E_N = \left\{n \in \{1,\dots,N\}, \ z_n \ne \tilde{z_n}, \ \textrm{and} \ z_n \ne -\tilde{z_n} \right\}$. Then we obtain from (\[rr5\]) $$\label{rr6} f(z) = \frac{G}{\tilde{G}} \prod_{n \in E_N} \frac{\Big( 1 -\frac{z^2}{\overline{\tilde{z_n}}^2} \Big)}{\Big( 1 -\frac{z^2}{\tilde{z_n}^2} \Big)} \tilde{f}(z).$$ Denoting by $g(z)$ the rational function $g(z) = \frac{G}{\tilde{G}} \prod_{n \in E_N} \frac{\Big( 1 -\frac{z^2}{\overline{\tilde{z_n}}^2} \Big)}{\Big( 1 -\frac{z^2}{\tilde{z_n}^2} \Big)}$, we finally get $$a_{L3}(\lambda,z) = g(z) \tilde{a_{L3}}(\lambda,z),$$ and thus $$L(\lambda,z) = g(z) \tilde{L}(\lambda,z).$$ - If $\frac{1}{\tilde{\kappa}_+} + \frac{1}{\tilde{\kappa}_-} > 0$, then the zeros $\tilde{z_n}$’s with positive imaginary part are located in the quadrant $II = \{ z \in {\mathbb{C}}, \ \Re(z) < 0, \ \Im(z) > 0 \}$. Hence, using (\[rr2\]), we have the following. There exists a $N \in {\mathbb{N}}$ such that $$\label{rr7} z_n = - \overline{\tilde{z_n}}, \quad \forall n > N.$$ Using (\[rr0\]), Lemma \[LargeAsymp-aL3\] and Corollary \[Asymp-Large-zn\], we get in this case $$\frac{1}{\tilde{\kappa}_-} - \frac{1}{\tilde{\kappa}_+} = \frac{1}{\kappa_-} + \frac{1}{\kappa_+}, \quad \frac{1}{\tilde{\kappa}_-} + \frac{1}{\tilde{\kappa}_+} = - \frac{1}{\kappa_-} - \frac{1}{\kappa_+},$$ which gives $$\tilde{\kappa}_- = -\kappa_+, \quad \tilde{\kappa}_+ = -\kappa_-.$$ Using again (\[rr0\]) and the asymptotics of $a_{L1}(\lambda,z)$ from Lemma \[LargeAsymp-aL3\], we get the necessary condition $$\left( \frac{\tilde{a_-}}{a_+} \right)^{\frac{i \lambda}{\kappa_+}} = \left( \frac{\tilde{a_+}}{a_-} \right)^{\frac{i \lambda}{\kappa_-}}.$$ Also, we use (\[fz\]) and (\[rr7\]) to obtain $$\label{rr8} f(z) = \frac{G}{\overline{\tilde{G}}} \prod_{n = 1}^N \frac{\Big( 1 -\frac{z^2}{z_n^2} \Big)}{\Big( 1 -\frac{z^2}{\overline{\tilde{z_n}}^2} \Big)} \overline{\tilde{f}(\bar{z})}.$$ Finally, denote by $F_N = \{n \in \{1,\dots,N\}, \ z_n \ne \overline{\tilde{z_n}}, \ \textrm{and} \ z_n \ne -\overline{\tilde{z_n}} \}$. Then we obtain from (\[rr5\]) $$\label{rr9} f(z) = \frac{G}{\overline{\tilde{G}}} \prod_{n \in F_N} \frac{\Big( 1 -\frac{z^2}{\tilde{z_n}^2} \Big)}{\Big( 1 -\frac{z^2}{\overline{\tilde{z_n}}^2} \Big)} \overline{\tilde{f}(\bar{z})}.$$ Denoting by $h(z)$ the rational function $h(z) = \frac{G}{\overline{\tilde{G}}} \displaystyle\prod_{n \in F_N} \frac{\Big( 1 -\frac{z^2}{\tilde{z_n}^2} \Big)}{\Big( 1 -\frac{z^2}{\overline{\tilde{z_n}}^2} \Big)}$, we finally get $$a_{L3}(\lambda,z) = h(z) \overline{\tilde{a_{L3}}(\lambda,\bar{z})} = h(z) \tilde{a_{L2}}(\lambda,z),$$ and thus $$L(\lambda,z) = - h(z) \tilde{R}(\lambda,z).$$ Both above cases prove the results stated in the Proposition. Even in the case when the reflection coefficients $L(\lambda,z)$ and $\tilde{L}(\lambda,z)$ (resp. $R(\lambda,z)$ and $\tilde{R}(\lambda,z)$) coincide up to a *rational function* in the $z$ variable, we cannot conclude from this fact the result stated in [@DN3], that is the uniqueness of the function $a(x)$ and $\tilde{a}(x)$ up to a translation. This question remains thus open and we conjecture that this is false. We refer to the last Section of [@DaKaNi] for more details about this point in a similar and more general model. What we can prove however is the following weaker statement. \[unicitetransmission\] Assume that $$\frac{1}{\kappa_+} + \frac{1}{\kappa_-} < 0, \quad \frac{1}{\tilde{\kappa}_+} + \frac{1}{\tilde{\kappa}_-} < 0.$$ Let $\mathcal{L}$ be a subset of ${\mathbb{N}}$ such that $\displaystyle\sum_{n \in \mathcal{L}} \frac{1}{n} = \infty$. Assume that $$T(\lambda,n) = \tilde{T}(\lambda,n), \quad \forall n \in \mathcal{L}.$$ Assume also that $$\label{rr10} L(\lambda,k) = \tilde{L}(\lambda,k),$$ for a finite but large enough number of indices $k \in {\mathbb{N}}$. Then there exists a constant $\sigma \in {\mathbb{R}}$ such that $$\tilde{a}(x) = a(x + \sigma).$$ In consequence, the two SSAHM $(\Sigma,g)$ ans $(\tilde{\Sigma},\tilde{g})$ coincide up to isometries. From Proposition \[Transmission1\], we know that there exists a rational function $g(z)$ such that $$L(\lambda,z) = g(z) \tilde{L}(\lambda,z).$$ From (\[rr10\]), we infer that $g(z) = 1$ for all $z\in {\mathbb{C}}$ and thus $$L(\lambda,z) = \tilde{L}(\lambda,z).$$ Hence the Proposition is proved using Theorem 1.1. in [@DN3]. [99]{} Aktosun T., Klaus M., van der Mee C., *Direct and inverse scattering for selfadjoint hamiltonian systems on the line*, Integr. Equa. Oper. Theory $\mathbf{38}$ (2000), 129-171. Bennewitz C., *A proof of the local Borg-Marchenko Theorem*, Comm. Math. Phys. $\mathbf{211}$, (2001), 131-132. Boas R.P., *Entire Functions*, Academic Press, (1954). Borthwick D., Perry P.A., *Inverse scattering results for manifolds hyperbolic near infinity*, J. of Geom. Anal. $\mathbf{21}$, No. 2, (2011), 305-333. Cohen J.M., Powers R.T., *The general relativistic hydrogen atom*, Comm. Math. Phys. $\mathbf{86}$, (1982), 69-86. Daudé T. Kamran N., Nicoleau F., *Inverse scattering at fixed energy on asymptotically hyperbolic Liouville surfaces*, preprint (2014), arXiv:1409.6229. Daudé T., *Time-dependent scattering theory for massive charged Dirac fields by a Reissner-Nordström black hole*, J. Math. Phys. $\mathbf{51}$, (2010), 102504. Daudé T., Nicoleau F., *Inverse scattering in (de Sitter)-Reissner-Nordström black hole spacetimes*, Rev. Math. Phys. $\mathbf{22}$ (4), (2010), 431-484. Daudé T., Nicoleau F., *Inverse scattering at fixed energy in de Sitter-Reissner-Nordström black holes*, Annales Henri Poincaré $\mathbf{12}$, (2011), 1-47. Freiling G., Yurko V., *Inverse problems for differential operators with singular boundary conditions*, Math. Nachr. $\mathbf{278}$ no. 12-13, (2005), 1561-1578. Gesztesy F., Simon B., *On local Borg-Marchenko uniqueness results*, Comm. Math. Phys. $\mathbf{211}$, (2000), 273-287. Horváth M., *Partial identification of the potential from phase shifts*, J. Math. Anal. Appl., (2010), doi:10.1016/j.jmaa.2010.10.071. Isozaki H., Kurylev J., *Introduction to spectral theory and inverse problems on asymptotically hyperbolic manifolds*, preprint (2011), arXiv:1102.5382. Joshi M.S., Sá Barreto A., *Inverse scattering on asymptotically hyperbolic manifolds*, Acta Mathematica, $\mathbf{184}$, (2000), 41-86. Lake K., *Reissner-Nordström-de Sitter metric, the third law, and cosmic censorship*, Phys. Rev. D $\mathbf{19}$, (1979), no.2, 421-429. Levin B. Y., *Lectures on entire functions*, Translations of Mathematical Monograph, $\mathbf{150}$, American Mathematical Society (1996). Melnyk F., *Scattering on Reissner-Nordström metric for massive charged spin $\frac{1}{2}$ fields*, Ann. Henri Poincaré $\mathbf{4}$, (2003), no. 5, 813-846. Nicolas J.-P., *Scattering of linear Dirac fields by a spherically symmetric black hole*, Ann. Inst. Henri Poinc. $\mathbf{62}$, (1995), no.2, 145-179. Ramm A.G., *An Inverse Scattering Problem with part of the Fixed-Energy Phase shifts*, Comm. Math. Phys. $\mathbf{207}$, (1999), no.1. 231-247. Regge T., *Introduction to complex orbital momenta*, Nuevo Cimento $\mathbf{XIV}$, (1959), no.5, 951-976. Sá Barreto A., *Radiation fields, scattering and inverse scattering on asymptotically hyperbolic manifolds*, Duke Math. Journal $\mathbf{129}$, no. 3, (2005), 407-480. Rudin W., *Real and Complex Analysis, Third edition*, McGraw-Hill Book Company, (1986). Simon B., *A new approach to inverse spectral theory, I. fundamental formalism*, Annals of Math. $\mathbf{150}$, (1999), 1-29. Teschl G., *Mathematical Methods in Quantum Mechanics*, Graduate Studies in Mathematics Vol. 99, AMS Providence, Rhode Island, (2009). Thaller B., *The Dirac equation*, Texts and Monographs in Physics, Springer-Verlag, (1992). Wald R.M., *General Relativity*, The University of Chicago Press, (1984). [^1]: Département de mathématiques, Université de Cergy-Pontoise, UMR CNRS 8088, 2 Av. Adolphe Chauvin, 95302 Cergy-Pontoise cedex. Email: [email protected]. Research supported by the French National Research Project AARG, No. ANR-12-BS01-012-01 [^2]: Département de Mathématiques, Université de Nantes, 2, rue de la Houssinière, BP 92208, 44322 Nantes Cedex 03. Emails: [email protected], [email protected]. Research supported by the French National Research Project NOSEVOL, No. ANR- 2011 BS0101901 [^3]: Note that the models studied in [@DN3] (see also [@Da1; @DN2]) come from General Relativity. More precisely, the direct and inverse scattering of Dirac waves propagating in the exterior region of a Reissner-Nordström-de Sitter black holes were studied therein. It turns out that these “relativistic” models and the one presented in this paper are equivalent. This was briefly mentioned in [@DN3] and made rigorous in the next Section \[TM\].
{ "pile_set_name": "ArXiv" }
--- abstract: 'We show that if $F(s)$ is a nondegenerate ordinary Dirichlet series with nonnegative coefficients and $F(k)$ is a rational number for all large enough positive integers $k$, then the denominators of those rational numbers are unbounded. In particular, our result holds for the Riemann zeta function over any arithmetic progression. These results are derived via upper bounds on associated Hankel determinants.' address: - | School of Math. and Phys. Sciences\ University of Newcastle\ Callaghan\ Australia - | School of Math. and Phys. Sciences\ University of Newcastle\ Callaghan\ Australia author: - Michael Coons - Daniel Sutherland title: 'Towards the (ir)rationality of values of Dirichlet series' --- Introduction ============ The values of the Riemann zeta function at positive even integers were determined by Euler nearly 300 years ago. Together with Lindemann’s proof of the transcendence of $\pi$, for over 130 years we have known that $\zeta(2n)$ is transcendental for $n\geqslant 1$. The complementary irrationality results for zeta values at odd integers has been of great importance in the mathematical community for some time. Comparitively recently (only 35 years ago), Apéry [@Apery] showed that $\zeta(3)$ was irrational, though unfortunately, his proof does not extend to other odd zeta values; see also Beukers [@Beukers]. The story ends here for irrationality of specific zeta values, though one can say more with less specified outcomes. Rivoal [@Rivoal] has shown that infinitely many odd zeta values are irrational, and Zudilin [@Zudilin] has shown that one of $\zeta(5)$, $\zeta(7)$, $\zeta(9)$, or $\zeta(11)$ is irrational. Similar to Rivoal’s result, this paper concerns (ir)rationality of ordinary Dirichlet series, with particular attention to zeta values over arithmetic progressions. Concerning the specific case of Riemann’s zeta function, in this paper we contribute the following result. \[mainzeta\] Let $a$ and $b$ be positive integers and let $\zeta_{a,b}(k)=\sum_{m\geqslant 1}{m^{-(ak+b)}}$. Suppose for some $R>0$ that $\zeta_{a,b}(k)\in\Q$ for all integers $k\geqslant R$. For each $k\geqslant R$, define the positive pair of coprime integers $p_k$ and $q_k$ by ${p_k}/{q_k}=\zeta_{a,b}(k)$. Then the sequence $\{q_k\}_{k\geqslant R}$ is unbounded. An analogous result to Theorem \[mainzeta\] is true replacing $\zeta_{a,b}(s)$ with any nondegenerate ordinary Dirichlet series with nonnegative coefficients. We prove the following generalisation in this paper. \[thm:main\] Let $F(s)=\sum_{n\geqslant 1} f(n)n^{-s}$ be a nondegenerate ordinary Dirichlet series with $f(n)\geqslant 0$ for all $n$, which is convergent for $\Re(s)\geqslant s_0$. Suppose that $F(k)\in\Q$ for all integers $k\geqslant R\geqslant s_0-2$, and for each $k\geqslant R$, define the positive pair of coprime integers $p_k$ and $q_k$ by ${p_k}/{q_k}=F(k)$ and set $$\D_m[F] = {\rm lcm}(q_{R+2},q_{R+3},\ldots,q_{R+m}).$$ Then the common denominator $\D_m[F]$ grows at least exponentially in $m$. In the case of $\zeta_{a,b}(s)$, if one keeps track of constants, Theorem \[thm:main\] implies that for each $\varepsilon\in(0,1)$ that $\D_m[\zeta_{a,b}]>(2-\varepsilon)^m$ for large enough $m$ depending on $\varepsilon$. This in turn implies that the $q_k$, as defined in Theorem \[mainzeta\], satisfy $$\max_{k\leqslant m}\{q_k\}>m\log(2-\varepsilon)$$ for any $\varepsilon\in(0,1)$ and $m$ large enough. Theorem \[mainzeta\] follows immediately from Theorem \[thm:main\]; we prove Theorem \[thm:main\] via a result on Hankel determinants. For a given sequence of real numbers $\{h(k)\}_{k\geqslant 2}$ and integers $n$ and $r$, we define the Hankel determinant of size $n$ starting at offset $r$ by $$H_n^{(r)}[h] = \det_{1\leqslant i,j\leqslant n}\Big(h(i+j+r)\Big).$$ The growth of these determinants plays an important role in investigating the (ir)rationality of zeta values. Theorems \[mainzeta\] and \[thm:main\] are easily deduced from analogous statements on the decay of Hankel determinants. Indeed, Theorem \[thm:main\] is a consequence of the following result on the growth of Hankel determinants of nondegenerate ordinary Dirichlet series with nonnegative coefficients. \[main\] Let $F(s)=\sum_{n\geqslant 1} f(n)n^{-s}$ be a nondegenerate ordinary Dirichlet series with $f(n)\geqslant 0$ for all $n$, which is convergent for $\Re(s)\geqslant s_0$. Then $H_n^{(r)}[F]>0$ for all $n\geqslant 1$ and $r\geqslant s_0-2$, and there is a $c>0$ such that $$\log H_n^{(r)}[F] < -c n^2,$$ for all sufficiently large integers $n$ and $r$. We note that Theorem \[main\] is valid for $r$ large enough, where “$r$ large enough” can in reality be “$r$ very small.” In the case of the Riemann zeta function, the theorem holds for all $r\geqslant 0$. In fact, our result implies that $$\log H_n^{(0)}[\zeta]<-n^2\log(2-\varepsilon),$$ for any positive $\varepsilon$ close to zero and large enough $n$. Experimental work suggests that the asympotitic decay of the Hankel determinants of the Riemann zeta function for $r=0,1$ is a bit better than what Theorem \[main\] concludes. In a preprint, Monien [@Monien] has produced a heuristic, which suggests that $$\log H_n^{(0)}[\zeta]\sim\log H_n^{(1)}[\zeta]\sim -n^2\log(2n)=-n^2\log 2-n^2\log n.$$ Monien also found experimentally, that $$-\frac{H_{n-1}^{(0)}[\zeta]H_{n}^{(1)}[\zeta]}{H_{n}^{(0)}[\zeta]H_{n+1}^{(1)}[\zeta]} = -\frac{1}{2n+1}+\frac{2}{(2n+1)^{2}}-\frac{7}{3}\frac{1}{(2n+1)^{3}}+\cdots,$$ and $$-\frac{H_{n+1}^{(0)}[\zeta]H_{n-1}^{(1)}[\zeta]}{H_{n}^{(0)}[\zeta]H_{n}^{(1)}[\zeta]} = -\frac{1}{2n}-\frac{1}{(2n)^{2}}+\frac{2}{3}\frac{1}{(2n)^{3}}-\frac{6}{5}\frac{1}{(2n)^{4}}+\frac{56}{45}\frac{1}{(2n)^{5}}+\cdots.$$ Additionally, according to Monien, detailed numerical experiments by Zagier suggest that $$H_{n}^{(0)}[\zeta]=A^{(0)}\left(\frac{2n+1}{e\sqrt{e}}\right)^{-(n+\frac{1}{2})^{2}}\left(1+\frac{1}{24}\frac{1}{(2n+1)^{2}}-\frac{12319}{259200}\frac{1}{(2n+1)^{4}}+\ldots\right),$$ and $$H_{n-1}^{(1)}[\zeta]=A^{(1)}\left(\frac{2n}{e\sqrt{e}}\right)^{-n^{2}+\frac{3}{4}}\left(1-\frac{17}{240}\frac{1}{(2n)^{2}}-\frac{199873}{7257600}\frac{1}{(2n)^{4}}-\ldots\right),$$ where $A^{(0)}\approx0.351466738331\ldots$ and $A^{(1)} =\frac{e^{9/8}}{\sqrt{6}}A^{(0)}.$ Hankel Determinants of some specialised sequences {#Special} ================================================= This section contains two results, which are of paramount importance to our investigation. The first is Dodgson condensation—sometimes known as “Lewis Carroll’s identity.” \[prop:rr\] Let $n\geqslant 2$ and $r\geqslant 0$ be integers and $\{h(k)\}_{k\geqslant 0}$ be a sequence of real numbers. Then $$H_{n+1}^{({r})}[h]\cdot H_{n-1}^{(r+2)}[h]=H_n^{(r)}[h]\cdot H_n^{(r+2)}[h]-\left(H_n^{(r+1)}[h]\right)^2.$$ While the setting in Lemma \[prop:rr\] is quite specific to Hankel determinants, Dodgson condensation is fully generalizable for evaluation of any determinant. Indeed, this is the way that Dodgson used it. For a good reference on Hankel determinants see Pólya and Szegő [@Polya Part 7]. Our next result provides a way to bound the Hankel determinants of certain sequences by using bounds on the growth of those sequences and their consecutive ratios. \[prop:decreaseN\] Let $K\geqslant 2$ be a fixed integer and $h(k)>0$ for $k\geqslant K$. Suppose that there exist positive functions $A(k)$, $B(k)$ and $\lambda(k)$, with $A(k)<B(k+1)$ for all $k\geqslant K$, such that 1. $A(k) \leqslant \frac{h(k+1)}{h(k)}<B(k)$ for all $k\geqslant K$, and 2. $h(k+2) < \lambda(k+2)\cdot\frac{B(k+1)}{B(k+1)-A(k)},$ for all $k\geqslant K$. If $H_n^{(r)}[h]>0$ for all $n\geqslant 1$ and for $r\geqslant K-2$, then $$H_n^{(r)}[h] < h(2+r)\prod_{k=2}^n\lambda(2k+r),$$ for all $n\geqslant 2$ and for $r\geqslant K-2$. Fix $r\geqslant K-2$. By definition, $H_2^{(r)}[h]=h(2+r)h(4+r)-h(3+r)^2$ and $H_1^{(r)}[h]=h(2+r)$. Thus using the positivity of $h(k)$ along with assumptions $(i)$ and $(ii)$, we have $$\begin{aligned} H_2^{(r)}[h] &= h(2+r)h(4+r)\left(1-\frac{h(3+r)}{h(4+r)}\frac{h(3+r)}{h(2+r)}\right) \\ &< h(2+r)h(4+r)\left(1-\frac{A(2+r)}{B(3+r)}\right)\\ &< H_1^{(r)}[h]\cdot\lambda(4+r). \end{aligned}$$ Suppose that $n\geqslant 2$. By Lemma \[prop:rr\] and the positivity of $H_n^{(r)}[h]$, we have $$H_{n+1}^{({r})}[h]=\frac{H_n^{(r)}[h]\cdot H_n^{(r+2)}[h]-\left(H_n^{(r+1)}[h]\right)^2}{H_{n-1}^{(r+2)}[h]}<H_n^{(r)}[h]\cdot \frac{H_n^{(r+2)}[h]}{H_{n-1}^{(r+2)}[h]}.$$ Combining this with the analogous result for $H_n^{(r+2)}[h]$ gives $$H_{n+1}^{({r})}[h]<H_n^{(r)}[h]\cdot \frac{H_{n-1}^{(r+2)}[h]\cdot \frac{H_{n-1}^{(r+4)}[h]}{H_{n-2}^{(r+4)}[h]}}{H_{n-1}^{(r+2)}[h]}=H_n^{(r)}[h]\cdot\frac{H_{n-1}^{(r+4)}[h]}{H_{n-2}^{(r+4)}[h]}.$$ We continue this process to keep reducing $n$ (while increasing $r$) until we can apply the $n=1$ case to give $$H_{n+1}^{(r)}[h] < H_{n}^{(r)}[h]\cdot \frac{H_2^{(r+2(n-1))}[h]}{H_1^{(r+2(n-1))}[h]} < H_n^{(r)}[h]\cdot\lambda(2(n+1)+r).$$ Now let $n\geqslant 3$. The previous inequality gives $H_{n}^{(r)}[h]<H_{n-1}^{(r)}[h]\cdot \lambda(2n+r)$. Repeated application of this inequality gives $$H_{n}^{(r)}[h] < H_{1}^{(r)}[h]\cdot \prod_{k=2}^n\lambda(2k+r)=h(2+r)\prod_{k=2}^n\lambda(2k+r),$$ which is the desired result. In what follows we will only be interested in decreasing sequences $\{h(k)\}_{k\geqslant2}$. We note that, if additionally $h(k)$ is bounded for all $k$, say by $M$, then condition [*(ii)*]{} of Proposition \[prop:decreaseN\] will be satisfied by defining $B(k)=B$ for some constant $B>0$ and $\lambda(k)=(B-A(k-2))M/B.$ This will be the setting in the following section, though Proposition \[prop:decreaseN\] holds for increasing functions as well. By way of example, consider the sequence $\{h(k)\}_{k\geqslant2}$ where $h(k)=(k-2)!$. Denote by $n\$$ the superfactorial of $n$; that is, $$n\$=\prod_{k=1}^n k!.$$ Strehl has shown that $H_n^{(r)}[h]>0$ for all $n\geqslant1$ and $r\geqslant0$; in fact, exact values for these determinants are known—see Radoux [@Radoux] for details. To gain an upper bound, we can apply Proposition \[prop:decreaseN\] with $K=2$, $A(k)=k-1$, $B(k)=k$ and $\lambda(k)=2(k-2)!$ to obtain $$H_n^{(r)}[h]<2^{n-1}\prod_{k=1}^n(2k+r-2)!=2^{n-1}\cdot r!(r+2)!(r+4)!\cdots(r+2n-2)!,$$ for all $n\geqslant1$ and $r\geqslant0$. Hankel determinants of ordinary Dirichlet series {#Ordinary} ================================================ In this section, we consider general ordinary Dirichlet series with nonnegative coefficients. There is a natural split between what can be considered a degenerate case, and a nondegenerate case. As the degenerate case, we consider a Dirichlet series having only finitely many nonzero coefficients, and as the nondegenerate case, we consider Dirichlet series having infinitely many nonzero coefficients. We use the following result for the nonvanishing of values of Hankel Determinants. \[thm:M\] Let $F(s)=\sum_{n\geqslant 1} f(n)n^{-s}$ be an ordinary Dirichlet series, which is convergent for $\Re(s)\geqslant s_0$. Then for any $n\geqslant 1$ and $r\geqslant s_0-2$, we have $$H_{n}^{(r)}[F]=\frac{1}{n!}\sum_{m_{1},m_{2},\ldots m_{n}=1}^{\infty}\prod_{i=1}^{n}\frac{f(m_{i})}{m_{i}^{2n+r}}\prod_{i<j}(m_{i}-m_{j})^{2}.$$ While Theorem \[thm:M\] holds in great generality, a classical result due to Kronecker (see also Pólya and Szegő [@Polya Part 7]) provides a nice dichotomy. Let $\{a(n)\}_{n\geqslant 0}$ be a sequence of real numbers. Then the function $\sum_{n\geqslant 0}a(n)x^n$ is rational if and only if finitely many of the determinants $H_n^{(r)}[a]$ are nonzero. This leads to the following result. \[nonzero\] Let $F(s)=\sum_{n\geqslant 1} f(n)n^{-s}$ be an ordinary Dirichlet series with $f(n)\geqslant 0$ for all $n$, which is convergent for $\Re(s)\geqslant s_0$. Then 1. if there are only finitely many $f(n)\neq 0$, then $H_n^{(r)}[F]=0$ for sufficiently large $n$ and any $r\geqslant s_0-2$; 2. if there are infinitely many $f(n)\neq 0$, then $H_n^{(r)}[F]>0$ for any $n\geqslant 1$ and $r\geqslant s_0-2$. Theorem \[thm:M\] implies both [*(i)*]{} and [*(ii)*]{} immediately. Part [*(i)*]{} also follows from Kronecker’s theorem. To further examine the Hankel determinants in the nondegenerate case, we require a lower bound on the ratio of successive values of $F(s)$. \[MandN\] Let $F(s)=\sum_{n\geqslant 1} f(n)n^{-s}$ be a nondegenerate ordinary Dirichlet series with $f(n)\geqslant 0$ for all $n$, which is convergent for $\Re(s)\geqslant s_0$. If $N<M$ are the two minimal indices of the nonzero coefficients of $F(s)$, then the following hold: 1. if $N=1$, then $$\lim_{s\to\infty}\frac{M^{s+1}}{(M-1)f(M)}\left(1-\frac{F(s+1)}{F(s)}\right)=1;$$ 2. if $N>1$, then $$\lim_{s\to\infty}\frac{f(N)N^{(\alpha-1)s}}{f(M)(1-N^{1-\alpha})}\left(1-N\cdot \frac{F(s+1)}{F(s)}\right)=1,$$ where $\alpha=\log M/\log N.$ We treat the two cases, $N=1$ and $N>1$, separately. [Case $N=1$.]{} Note that by definition, we have $$F(s)=1+f(M)\cdot M^{-s}+O\left(M^{-\delta s}\right),$$ where $\delta=\log(M+1)/\log M>1.$ Thus $$\frac{F(s+1)}{F(s)}=\frac{1+\frac{f(M)}{M}\cdot M^{-s}+O\left(M^{-\delta s}\right)}{1+f(M)\cdot M^{-s}+O\left(M^{-\delta s}\right)}=1-(M-1)f(M) M^{-(s+1)}+O\left(M^{-\delta s}\right),$$ and so $$\frac{M^{s+1}}{(M-1)f(M)}\left(1-\frac{F(s+1)}{F(s)}\right)=1+O\left(\frac{M^{s+1}}{M^{\delta s}}\right).$$ Since $\delta>1$, taking $s$ to infinity gives the required result. [Case $N>1$.]{} By definition, we have $$\begin{aligned} F(s)&=f(N)\cdot N^{-s}+f(M)\cdot N^{-\alpha s}+O\left(N^{-\beta s}\right)\\ &=f(N)\cdot N^{-s}\left(1+\frac{f(M)}{f(N)}\cdot N^{(1-\alpha) s}+O\left(N^{(1-\beta) s}\right)\right),\end{aligned}$$ where $\alpha=\log M/\log N$ and $\beta=\log(M+1)/\log N$. Note that $\beta>\alpha>1$. Thus $$\begin{aligned} \frac{F(s+1)}{F(s)}&=\frac{f(N)\cdot N^{-s-1}\left(1+\frac{f(M)}{f(N)}\cdot N^{(1-\alpha)(s+1)}+O\left(N^{(1-\beta)s}\right)\right)}{f(N)\cdot N^{-s}\left(1+\frac{f(M)}{f(N)}\cdot N^{(1-\alpha) s}+O\left(N^{(1-\beta) s}\right)\right)}\\ &=\frac{1}{N}\left(1-\frac{f(M)}{f(N)}\left(1-N^{1-\alpha}\right)N^{(1-\alpha)s}+o\left(N^{(1-\alpha)s}\right)\right),\end{aligned}$$ and so $$\frac{f(N)N^{(\alpha-1)s}}{f(M)(1-N^{1-\alpha})}\left(1-N\cdot \frac{F(s+1)}{F(s)}\right)=1+o\left(1\right).$$ This completes the proof of the lemma. This lemma provides the following immediate corollary. \[govergasym\] Let $F(s)=\sum_{n\geqslant 1} f(n)n^{-s}$ be a nondegenerate ordinary Dirichlet series with $f(n)\geqslant 0$ for all $n$, which is convergent for $\Re(s)\geqslant s_0$. If $N<M$ are the two minimal indices of the nonzero coefficients of $F(s)$, then the following hold: 1. if $N=1$, then there are positive constants $c_1>0$ and $K_1\geqslant s_0$ such that for all $s\geqslant K_1\geqslant s_0$, we have $$0<1-\frac{c_1}{M^{s}} \leqslant \frac{F(s+1)}{F(s)}< 1;$$ 2. if $N>1$, then there are positive constants $c_2>0$ and $K_2\geqslant s_0$ such that for all $s\geqslant K_2\geqslant s_0$, we have $$0<\frac{1}{N}-\frac{c_2}{N^{(\alpha -1)s}} \leqslant \frac{F(s+1)}{F(s)}< \frac{1}{N},$$ where $\alpha=\log M/\log N.$ We are now in a position to prove the Theorem \[main\]. As in the proof of Lemma \[MandN\], we split this proof into the two cases, $N=1$ and $N>1$, where $N$ is the minimal index of nonzero coefficients of $F(s)$. [Case $N=1$.]{} Let $c_1,K_1>0$ be such that the conclusion of Corollary \[govergasym\][*(i)*]{} holds, and for all $k$, set $$A(k)=1-\frac{c_1}{M^{k}}.$$ Then we have $A(k)\leqslant F(k+1)/F(k)<1$ and $A(k)$ is a positive function for $k\geqslant K_1$. Also, $$\frac{1}{1-A(k-2)}=\frac{M^{k-2}}{c_1}.$$ Set $$\lambda(k)=2\cdot F(s_0)\cdot\big(1-A(k-2)\big)=2\cdot F(s_0)\cdot\frac{c_1}{M^{k-2}}.$$ Then $$\lambda(k)\cdot \frac{1}{1-A(k-2)} =2\cdot F(s_0)> F(k),$$ since $F(s)$ is a monotonically decreasing function of $s\geqslant s_0$. We may now apply Proposition \[prop:decreaseN\], with $K=K_1$, so that $$H_n^{(r)}[F] < F(2+r)\prod_{k=2}^n\lambda(2k+r)\leqslant F(s_0)\prod_{k=2}^n\lambda(2k),$$ for all integers $n\geqslant 2$ and $r\geqslant K$. Thus $$\begin{aligned} \log H_n^{(r)}[F] &< \log\left(F(s_0)\prod_{k=2}^n\lambda(2k)\right) \\ &=\log F(s_0)+\sum_{k=2}^n \log \lambda(2k)\\ &=\log F(s_0)+\sum_{k=2}^n \log \left(2\cdot F(s_0)\cdot \frac{c_1}{M^{2k-2}}\right)\\ &=\log F(s_0)+(n-1)\log(2 F(s_0) c_1 M^2) -2\log M\cdot\frac{n(n+1)-2}{2}\\ &\sim -n^2 \log M.\end{aligned}$$ This completes the $N=1$ case. [Case $N>1$.]{} Let $c_2,K_2>0$ be such that conclusion of Corollary \[govergasym\][*(i)*]{} holds, and for all $k$, set $$A(k)=\frac{1}{N}-\frac{c_2}{N^{(\alpha -1)k}}.$$ Then we have $A(k)\leqslant F(k+1)/F(k)<1/N$ and $A(k)$ is a positive function for $k\geqslant K_2$. Set $$\lambda(k)=2N\cdot F(s_0)\cdot\big(1/N-A(k-2)\big)=2N\cdot F(s_0)\cdot\frac{c_2}{N^{(\alpha-1)(k-2)}}.$$ Then $$\lambda(k)\cdot \frac{1/N}{1/N-A(k-2)} =2\cdot F(s_0)> F(k),$$ since $F(s)$ is a monotonically decreasing function of $s\geqslant s_0$. We may now apply Proposition \[prop:decreaseN\], with $K=K_2$, so that $$H_n^{(r)}[F] < F(2+r)\prod_{k=2}^n\lambda(2k+r)\leqslant F(s_0)\prod_{k=2}^n\lambda(2k),$$ for all integers $n\geqslant 2$ and $r\geqslant K$. Thus $$\begin{aligned} \log H_n^{(r)}[F] &< \log\left(F(s_0)\prod_{k=2}^n\lambda(2k)\right) \\ &=\log F(s_0)+\sum_{k=2}^n \log \lambda(2k)\\ &=\log F(s_0)+\sum_{k=2}^n \log \left(2N\cdot F(s_0)\cdot\frac{c_2}{N^{(\alpha-1)(k-2)}}\right)\\ &=\log F(s_0)+(n-1)\log(2N F(s_0) c_2 N^{2(\alpha-1)})\\ &\qquad\qquad -2(\alpha-1)\log N\cdot\frac{n(n+1)-2}{2}\\ &\sim -n^2\cdot 2(\alpha-1)\log N.\end{aligned}$$ Recalling that $\alpha>1$ completes the $N>1$ case, and with that, the proof of the theorem. To see how Theorem \[thm:main\] is deduced from Theorem \[main\], suppose that the conclusion of Theorem \[main\] holds; that is, let $N_0,R>0$ be such that for $n\geqslant N_0$ and $r\geqslant R$, we have $\log H_n^{(r)}[F] < -c n^2$. Now suppose that $F(k)$ is rational for all integers $k\geqslant R+2$. As in the statement of Theorem \[thm:main\], write ${p_k}/{q_k}=F(k)$ and set $$\D_m[F] = {\rm lcm}(q_{R+2},q_{R+3},\ldots,q_{R+m}).$$ Then, using the positivity of $H_n^{(r)}[F]$ given by Theorem \[main\], we have $$\big(\D_{2n}[F]\big)^n\cdot H_n^{(r)}[F]\in\mathbb{N},$$ so that $$\label{g0} \log H_n^{(r)}[F]+n\log \D_{2n}[F] \geqslant 0,$$ for all integers $n\geqslant 1$. If the sequence $\D_m[F]$ satisfied $\D_m\leqslant C^m$ for some $C\in(1,e^c)$ where $c>0$ is as given by Theorem \[main\] for large enough $m$, then using Theorem \[main\], for large enough $n$ there is a $\delta>0$ such that $$0\leqslant \log H_n^{(r)}[F]+n\log \D_{2n}[F] =\log H_n^{(r)}[F]+(c-\delta)n^2<-\delta n^2,$$ which is a contradiction for $n$ large enough. Concluding remarks ================== Note that repeated use of Dodgson condensation gives $$\begin{aligned} H_{n+1}^{(r)}[h]&=H_{n}^{(r)}[h]\cdot \frac{H_{n}^{(r+2)}[h]}{H_{n-1}^{(r+2)}[h]}\cdot\left(1-\frac{\left(H_{n}^{(r+1)}[h]\right)^2}{H_{n}^{(r)}[h]\cdot H_{n}^{(r+2)}[h]}\right)\\ &=H_{n}^{(r)}[h]\cdot \frac{H_{2}^{(r+2(n-1))}[h]}{H_{1}^{(r+2(n-1))}[h]}\cdot\prod_{j=0}^{n-2}\left(1-\frac{\left(H_{n-j}^{(r+(j+1))}[h]\right)^2}{H_{n-j}^{(r)}[h]\cdot H_{n-j}^{(r+2j)}[h]}\right),\end{aligned}$$ so, replacing $n+1$ with $n$, we have $$\begin{aligned} H_{n}^{(r)}[h] &=H_{n-1}^{(r)}[h]\cdot \frac{H_{2}^{(r+2(n-2))}[h]}{H_{1}^{(r+2(n-2))}[h]}\cdot\prod_{j=0}^{n-3}\left(1-\frac{\left(H_{n-j-1}^{(r+(j+1))}[h]\right)^2}{H_{n-j-1}^{(r)}[h]\cdot H_{n-j-1}^{(r+2j)}[h]}\right)\\ &=h(2+r)\cdot \left(\prod_{i=2}^n\frac{H_{2}^{(r+2(i-2))}[h]}{H_{1}^{(r+2(i-2))}[h]}\right)\cdot\prod_{i=2}^n\prod_{j=0}^{i-3}\left(1-\frac{\left(H_{i-j-1}^{(r+(j+1))}[h]\right)^2}{H_{i-j-1}^{(r)}[h]\cdot H_{i-j-1}^{(r+2j)}[h]}\right)\\ &=h(2+r)\cdot \left(\prod_{i=2}^n\frac{H_{2}^{(r+2(i-2))}[h]}{h(r+2(i-1))}\right)\cdot\prod_{i=2}^n\prod_{j=0}^{i-3}\left(1-\frac{\left(H_{i-j-1}^{(r+(j+1))}[h]\right)^2}{H_{i-j-1}^{(r)}[h]\cdot H_{i-j-1}^{(r+2j)}[h]}\right).\end{aligned}$$ In the case of a nondegenerate ordinary Dirichlet series $F(s)$ with nonnegative coefficients with $r\geqslant s_0-2$, using asymptotics for $F(s)$, gives some $c>0$ such that $$\log\left(\prod_{i=2}^n\frac{H_{2}^{(r+2(i-2))}[F]}{F(r+2(i-1))}\right)\sim -cn^2.$$ As our method is based off using this term with a constant upper bound for the double product above, Theorem \[main\] cannot be improved using our method. Getting upper bounds on the double product involves obtaining lower bounds on $\log H_n^{(r)}[F]$; we hope to address this in future work.\ **Acknowledgements.** We thank Wadim Zudilin for valuable remarks on an earlier version of this paper. [1]{} Roger Apéry, *Irrationalité de $\zeta(2)$ et $\zeta(3)$*, Astérisque **61** (1979), 11–13. F. Beukers, *A note on the irrationality of [$\zeta (2)$]{} and [$\zeta (3)$]{}*, Bull. London Math. Soc. **11** (1979), no. 3, 268–272. [MR ]{}[554391 (81j:10045)]{} C. L. Dodgson, *Condensation of determinants, being a new and brief method for computing their arithmetical values*, Proceedings of the Royal Society of London **15** (1866), pp. 150–155 (English). H. Monien, *[Hankel determinants of Dirichlet series]{}*, ArXiv e-prints (2009). G. P[ó]{}lya and G. Szeg[ő]{}, *Problems and theorems in analysis ii*, Classics in Mathematics, Springer-Verlag, Berlin, 1998, Translated from the German by C. E. Billigheimer, Reprint of the 1976 English translation. [MR ]{}[1492448]{} C. Radoux, *Déterminants de hankel et théorème de sylvester*, Séminaire Lotharingien de Combinatoire **28** (1992), 115–122. Tanguy Rivoal, *La fonction zêta de [R]{}iemann prend une infinité de valeurs irrationnelles aux entiers impairs*, C. R. Acad. Sci. Paris Sér. I Math. **331** (2000), no. 4, 267–270. [MR ]{}[1787183 (2001k:11138)]{} Wadim Zudilin, *Arithmetic of linear forms involving odd zeta values*, J. Théor. Nombres Bordeaux **16** (2004), no. 1, 251–291. [MR ]{}[2145585 (2006j:11102)]{}
{ "pile_set_name": "ArXiv" }
--- abstract: | We present preliminary results regarding the Fundamental Plane (FP) for galaxies in the two rich clusters Abell 665 and Abell 2218. Both clusters have a redshift of 0.18. We have compared the FP for A665 and A2218, and for the cluster CL0024+16 at z=0.39, with the FP for the Coma cluster. The scatter around the FP is similar for all four clusters. There may be indications that the slope of the FP is more shallow for the intermediate redshift clusters than for the Coma cluster. More complete samples of galaxies in intermediate redshift clusters are needed to map in detail the possible change of the slope as function of redshift. The mass-to-light (M/L) ratio as measured by the FP changes with redshift. At z=0.18 the M/L ratio (in Gunn r) is $16\pm 9$% smaller than for the Coma cluster. Together with earlier results reported for CL0024+16 this implies that the M/L ratio changes with redshift as $\Delta \log M/L_r \sim -0.4 \Delta z$. The results presented here are in agreement with passive evolution of a stellar population, which formed at a redshift larger than one. However, the possible presence of more recent bursts of star formation complicates the interpretation of the data. author: - 'IngerJørgensen[^1] and JensHjorth' title: 'The Fundamental Plane at z=0.18' --- epsf.tex [To appear in the proceedings of the 3rd ESO-VLT Workshop “Galaxy Scaling Relations”, eds. da Costa et al., Springer]{} Introduction ============ Observational studies show that the formation and the evolution of galaxies are complex processes, which may involve interactions, starbursts and infall (e.g., Dressler et al. 1994ab; Lilly et al. 1995; Moore et al. 1996). Some nearby E and S0 galaxies may also have experienced recent star formation. Caldwell et al. (1993) found a fraction of the E and S0 galaxies in the Coma cluster to have post-starburst spectra, and Faber et al. (1995) suggest that nearby field E and S0 galaxies have a substantial variation in the mean age of their stellar populations. In order to investigate the evolution of galaxies, studies are needed of both the morphological evolution and the evolution of the luminosities and the mass-to-light (M/L) ratios of the galaxies. High-resolution imaging with the [*Hubble Space Telescope*]{} (HST) and from the ground, combined with spectroscopy from the ground make it possible to carry out this kind of studies. Studies of the morphological evolution show that the fraction of spiral galaxies and irregular galaxies in clusters was higher at larger redshifts (e.g., Dressler et al. 1994ab; Oemler et al. 1997). The luminosity evolution of disk galaxies has recently been studied by Vogt et al. (1996) who established the Tully-Fisher (1977) relation for a sample of field galaxies with redshifts between 0.1 and 1. The Fundamental Plane (FP) (Dressler et al. 1987; Djorgovski & Davis 1987) for elliptical galaxies makes it possible to study how the M/L ratios change with redshift. Also, S0 galaxies in nearby clusters follow the FP (e.g., J[ø]{}rgensen et al. 1996, hereafter JFK96). The FP relates the effective radius, $r_{\rm e}$, the mean surface brightness within this radius, $< \hspace{-3pt} I \hspace{-3pt}>_{\rm e}$, and the (central) velocity dispersion, $\sigma$, in a tight relation, which is linear in log-space. For 226 E and S0 galaxies in nearby clusters JFK96 found $\log r_{\rm e} = 1.24 \log \sigma - 0.82 \log < \hspace{-3pt} I \hspace{-3pt}>_{\rm e} + {\rm cst}$. This relation can be interpreted as $M/L \propto M^{0.24}$, see also Faber et al. (1987). The scatter of the FP is very low, equivalent to a scatter of 23% in the M/L ratio (e.g., JFK96). Thus, the FP offers the possibility of detecting even small differences in the M/L ratios by observing five to ten galaxies in a distant cluster. The FP for intermediate redshift clusters has been established by van Dokkum & Franx (1996) and Kelson et al. (1997). Both studies find the M/L ratios of the galaxies to increase between redshifts of $z=0.3$ to 0.6 and the present. In this paper we establish the FP for the two rich clusters Abell 665 and Abell 2218. We use the data for these two clusters together with earlier published data for CL0024+16 (van Dokkum & Franx 1996) and data for the Coma cluster to study how the M/L ratios of the E and S0 galaxies change with redshift. Observational data for A665 and A2218 ===================================== The central parts of the two rich clusters Abell 665 and Abell 2218 were both observed with the Nordic Optical Telescope (NOT), La Palma, in March 1994. Observations were done in the I-band and the V-band. The total size of the observed fields are $3\farcm 4 \times 2\farcm 7$ for A665 and $2\farcm 8 \times 3\farcm 5$ for A2218. The basic reductions and the standard calibration are described in detail in J[ø]{}rgensen et al. (1997). Determination of the effective radius, $r_{\rm e}$, and the mean surface brightness within this radius, ${< \hspace{-3pt} \mu \hspace{-3pt}>}_{\rm e}$, was done following the technique described by van Dokkum & Franx (1996). This technique uses the full 2-dimensional information in the image, and takes the point-spread-function into account. The magnitudes were calibrated to Gunn r in the rest frame of the clusters. We are in the process of analysing HST images of the two clusters. The final discussion of the FP for these clusters will be based on photometry from both the NOT and the HST, cf. J[ø]{}rgensen et al. (1997). The spectra of 7 galaxies in A665 were obtained with the Multiple Mirror Telescope in January 1991. The total integration time was 9 hours. The observations include an E+A galaxy. The spectra of 10 galaxies in A2218 were obtained with the KPNO 4m telescope in June 1994; total integration time 12 hours. We have photometry of 8 of these galaxies. Detailed description of the reductions and the determination of the velocity dispersions can be found in J[ø]{}rgensen et al. (1997). Figure 1 shows two of the spectra obtained of galaxies in A2218. Determination of velocity dispersions of galaxies is usually done with spectra of template stars obtained with the same instrumentation as the galaxy spectra. Due to the large redshifts of A665 and A2218 this will not work for these clusters. A further complication is that the instrumental resolution of the spectra of A665 and A2218 varies with wavelength and from slit-let to slit-let. Therefore, the instrumental resolution as a function of wavelength and slit-let was mapped based on calibration lamp exposures. Spectra of template stars were then convolved to the exact same resolution. The velocity dispersions were determined by fitting the galaxy spectra with the template spectra using the Fourier Fitting Method (Franx et al. 1989). The velocity dispersions were aperture corrected following the technique described by J[ø]{}rgensen et al. (1995b). We corrected the velocity dispersions to a circular aperture with diameter 1.19h$^{-1}$kpc, equivalent to 3.4 arcsec at the distance of the Coma cluster. =11.5cm Data for the Coma cluster ========================= The Coma cluster is in the following used as a low redshift reference for the purpose of determination of changes in the FP as function of redshift. We use the sample of E and S0 galaxies, which has photometry in Gunn r from J[ø]{}rgensen et al. (1995a), see also J[ø]{}rgensen & Franx (1994). The sample covers the central $64'\times 70'$ of the cluster. The velocity dispersions from the literature (Davies et al. 1987; Dressler 1987; Lucey et al. 1991) were calibrated to a consistent system and aperture corrected. We use the values as listed by J[ø]{}rgensen et al. (1995b). Further, we use new velocity dispersions measurements from J[ø]{}rgensen (1997). A total of 116 galaxies have available photometry and spectroscopy. The sample is 93% complete to a magnitude limit of $m_{\rm T} = 15\fm 05$ in , equivalent to $M_{\rm r_T} = -20\fm 75$ ($H_0=50~{\rm km\, s^{-1}\, Mpc^{-1}}$ and $q_0=0.5$). =8.8cm The Fundamental Plane ====================== Figure 2 shows the FP for A665 and A2218. The figure also shows the FP for the Coma cluster and for the cluster CL0024+16 with a redshift of 0.39. The data for CL0024+16 are from van Dokkum & Franx (1996), and the photometry has been calibrated to rest frame Gunn r, see J[ø]{}rgensen et al. (1997). We adopt the coefficients for the FP from JFK96, ($\alpha$,$\beta$)=($1.24\pm 0.07$,$-0.82 \pm 0.02$). The coefficients were derived for a sample of 226 galaxies in 10 nearby clusters. Photometry in Gunn r was used. The relations shown on Figure 2 are FPs with these coefficients. Figure 3 shows the FP face-on and in two edge-on views. This figure includes the Coma cluster galaxies as well as the galaxies in A665, A2218 and CL0024+16. The mean surface brightnesses have been corrected for the dimming due to the expansion of the Universe. The effective radii are in kpc ($H_0=50~{\rm km\, s^{-1}\, Mpc^{-1}}$ and $q_0 =0.5$ were used). The scatter around the FP is the same for A665 and A2218 (0.091 and 0.115 in $\log r_{\rm e}$) as for the Coma cluster (0.095 in $\log r_{\rm e}$). The scatter for CL0024+16 is slightly lower, 0.064, though the difference is not statistically significant. The scatter for the Coma cluster reported here is somewhat larger than previous estimates based on smaller samples (e.g., J[ø]{}rgensen et al. 1993, 1996; Lucey et al. 1991). The difference is due to inclusion of more galaxies, which have post-starburst spectra (cf., Caldwell et al. 1993), see also J[ø]{}rgensen (1997). =8.8cm The slope of the FP ------------------- The samples in the intermediate redshift clusters are heavily biased towards intrinsically bright galaxies. This makes it difficult to test for possible differences in the slope of the FP for Coma and for the intermediate redshift clusters. The selection bias is clearly visible in the face-on view of the FP, Figure 3a. In addition to the difference in selection effects the intermediate redshift clusters seem to contain a larger number of galaxies with low surface brightness and large effective radius than the Coma cluster. The samples are still too small and incomplete to perform any statistical test of this. Figure 4 shows the FP as the relation between the M/L ratio and the mass. This figure and Figure 3c indicate that the slope of the FP may be slightly different for the intermediate redshift clusters than for the Coma cluster. In order to test if the coefficients of the FP depend on the redshift we fit the FP to the three clusters A665, A2218, and CL0024+16 as parallel planes, under the assumption that the FPs for these clusters have the same coefficients. This gives ($\alpha$,$\beta$)=($0.89\pm 0.14$,$-0.78 \pm 0.04$). Omitting the E+A galaxy in A665 does not change the fit. Formally the value of $\alpha$ is different from the coefficient for nearby cluster galaxies (JFK96). A fit to the Coma cluster galaxies alone gives ($\alpha$,$\beta$)=($1.28\pm 0.08$,$-0.83 \pm 0.03$). In order to limit the effect of the different selection criteria we repeated the fit to the Coma cluster galaxies enforcing a magnitude limit of $M_{\rm r_T} = -21\fm 65$. This does not give a result significantly different from the fit for the whole sample. The difference in $\alpha$ between the fit to the Coma cluster sample and the fit to the three intermediate redshift clusters is formally significant on the $2.5\sigma$ level. Still, the weight of the low luminosity galaxies is much larger in the Coma cluster sample than for the intermediate redshift clusters. The coefficients we find for the FP for the intermediate redshift clusters imply (assuming structural homology of the galaxies) that $$M/L \propto r_{\rm e} ^{0.28} \sigma ^{0.86} \propto M^{0.43} r_{\rm e}^{0.15}$$ This should be compared with $M/L \propto M^{0.24} r_{\rm e}^{-0.02}$ found for nearby cluster (JFK96). The difference may indicate that the low mass galaxies show a stronger luminosity evolution that the more massive galaxies. We emphasize that these results regarding the slope of the FP are preliminary. We discuss the issue in greater length in J[ø]{}rgensen et al. (1997), where also data from Kelson et al. (1997) are included in the analysis. The evolution of the M/L ratio ------------------------------- The zero point of the FP depends on the cosmological effects (surface brightness dimming and the value of $q_0$), and the evolution of the galaxies. The FP can in principle be used to test for the expansion of the Universe (Kj[æ]{}rgaard et al. 1993). In the following we correct the data for the expansion of the Universe, and we assume $q_0=0.5$. The FP shown in Figure 3 is corrected for the expansion of the Universe. The zero point of the FP can then be used to study the mean evolution of the galaxies in the clusters. Figure 5 shows the zero point differences between the intermediate redshift clusters and the Coma cluster as the offset in the M/L ratio. The uncertainties of the zero point offsets have contributions from the photometric calibration, the calibration of the velocity dispersion, and uncertainties in the zero points for the FP for the Coma cluster and for the intermediate redshift clusters. The error bars shown on Figure 5 include these contributions, see also J[ø]{}rgensen et al. (1997). The evolution of the M/L ratio with redshift is consistent with $\Delta \log M/L \approx (-0.4\pm 0.1) \Delta z$. =8.8cm The change of the M/L ratio with redshift is expected based on stellar populations models. Models of single burst populations show that for passive evolution $$M/L \propto t_{\rm age} ^{\kappa}$$ From models by Vazdekis et al. (1996) we find that $\kappa \approx 0.9-0.2x$ for photometry in Gunn r. Here $x$ is the slope of the initial mass function (IMF) of the stars. A Salpeter (1956) IMF has $x=1.35$. Eq. 2 can be used to relate the change in the M/L ratio with redshift to the formation redshift, $z_{\rm form}$, and the value of $q_0$ $$\Delta {\rm ln} M/L = - \kappa (1+q_0+z_{\rm form}^{-1})\,\Delta z$$ (Franx 1995; van Dokkum & Franx 1996). The slope we find for the change of the M/L ratio with redshift implies $$\kappa (1+q_0+z_{\rm form}^{-1}) = 0.9\pm 0.2$$ The result is in agreement with the result found by van Dokkum & Franx (1996) based on CL0024+16, only. Eq. 4 is consistent with passive evolution of a single stellar population model with a Salpeter IMF, $q_0=0.5$ and $z_{\rm form} = \infty$. A model with $z_{\rm form}=1.2$ deviates on the $3\sigma$ level. The model constraints given here should only be taken as rough guidelines. The correct interpretation of the data is most likely rather more complicated than indicated here. The correct value of $\kappa$ is not known. Further, the evolution of E and S0 galaxies cannot be viewed as a single burst event. The presence of younger stellar populations in the galaxies would imply a larger formation redshift for the old stellar populations in the galaxies. The most fundamental assumption for the interpretation of the data is that the observed E and S0 galaxies in the intermediate redshift clusters in fact evolve into galaxies similar to the present day E and S0 galaxies. It is possible that our selection of E and S0 galaxies in the intermediate redshift clusters is biased to select already aged galaxies, see the discussion by Franx (1995, and this volume). Larger and more complete samples of galaxies in intermediate redshift clusters are needed in order to address this problem in detail. Conclusions =========== We have established the Fundamental Plane for the two rich clusters Abell 665 and Abell 2218, both at a redshift of 0.18. The photometric parameters were derived from ground based observations obtained with the Nordic Optical Telescope, La Palma. The photometry was calibrated to Gunn r in the rest frame of the clusters. Central velocity dispersions were measured for seven galaxies in A665 and ten galaxies in A2218. The results presented here are preliminary. The final analysis of the FP for the two clusters will include photometry based on HST data, see J[ø]{}rgensen et al. (1997). The FP for the two clusters were compared to the FP for nearby clusters derived by JFK96, and to the FP for the Coma cluster (J[ø]{}rgensen 1997). The scatter around the FP for A665 and A2218 is similar to the scatter found for nearby clusters. We have used the data for A665 and A2218 together with data for CL0024+16 (van Dokkum & Franx 1996) to test if the slope of the FP changes with redshift. All photometry was calibrated to Gunn r in the rest frame of the clusters. There may be indications that the coefficient for $\log \sigma$ is significantly smaller for the intermediate redshift clusters than for the Coma cluster. However, severe selection effects and possibly also real differences in the distributions within the FP make it difficult to draw definite conclusions based on the current data. If the smaller value of the coefficient is confirmed in later studies it implies that $M/L \propto M^{0.43}$ for the intermediate redshift clusters, while $M/L \propto M^{0.24}$ for nearby clusters. This may indicate that the low mass galaxies in the clusters evolve faster than the more massive galaxies. The zero point offsets in the FP for the intermediate redshift clusters were used to investigate the average evolution of the M/L ratio of the galaxies. The M/L ratios of the galaxies in A665 and A2218 are $16\pm 9$% smaller than the M/L ratios of galaxies in the Coma cluster. From all four clusters we find that the evolution of the M/L ratio with redshift is consistent with $\Delta \log M/L \approx (-0.4\pm 0.1) \Delta z$. This can be used to constrain the formation redshift of the galaxies. The interpretation depends on the crucial assumption that the E and S0 galaxies observed in the intermediate redshift clusters evolve into galaxies similar to present day E and S0 galaxies, and that the samples in the intermediate redshift clusters form a representative sample of all the galaxies that will end as present day E and S0 galaxies. For a single burst population, a Salpeter IMF and $q_0=0.5$ the evolution in the M/L ratio implies that the formation redshift is larger than one. A more complete analysis of our data for A665 and A2218, which also involves photometry based on HST images, is presented in J[ø]{}rgensen et al. (1997). Acknowledgements: We thank M. Franx for permission to use spectroscopic data for A665 before publication. The staff at NOT, KPNO and MMT are thanked for assistance during the observations. Financial support from the Danish Board for Astronomical Research is acknowledged. This research was supported through Hubble Fellowship grant number HF-01073.01.94A from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. Bender, R., Burstein, D., Faber, S. M. 1992, ApJ, 399, 462 Caldwell, N., Rose, J. A., Sharples, R. M., Ellis, R. S., Bower, R. G. 1993, AJ, 106, 473 Davies, R. L., Burstein, D., Dressler, A., Faber, S. M., Lynden-Bell, D., Terlevich, R. J., Wegner, G. 1987, ApJS, 64, 581 Djorgovski, S., Davis, M. 1987, ApJ, 313, 59 Dressler, A. 1987, ApJ, 317, 1 Dressler, A., Lynden-Bell, D., Burstein, D., Davies, R. L., Faber, S. M., Terlevich, R. J., Wegner, G. 1987, ApJ, 313, 42 Dressler, A., Oemler, A. Jr., Butcher, H., Gunn, J. E. 1994a, ApJ, 430, 107 Dressler, A., Oemler, A. Jr., Sparks, W. B., Lucas, R. A. 1994b, ApJ, 435, L23 Faber, S. M., Dressler, A., Davies, R. L., Burstein, D., Lynden-Bell, D., Terlevich, R. J., Wegner, G. 1987, in Nearly Normal Galaxies, ed. S. M. Faber, Springer-Verlag, New York, p. 175 Faber, S. M., Trager, S. C., Gonzales, J. J., Worthey, G. 1995, in Stellar Populations, eds. van der Kruit, P. C., Gilmore, G., IAU coll. 164, Kluwer, Dordrecht, p. 249 Franx, M. 1995, in Stellar Populations, eds. van der Kruit, P. C., Gilmore, G., IAU coll. 164, Kluwer, Dordrecht, p. 269 Franx, M., Illingworth, G., Heckman, T. 1989, ApJ, 344, 613 J[ø]{}rgensen, I. 1997, in preparation J[ø]{}rgensen, I., Franx, M. 1994, ApJ, 433, 553 J[ø]{}rgensen, I., Franx, M., Hjorth, J., van Dokkum, P. G. 1997, in preparation J[ø]{}rgensen, I., Franx, M., Kj[æ]{}rgaard, P. 1993, ApJ, 411, 34 J[ø]{}rgensen, I., Franx, M., Kj[æ]{}rgaard, P. 1995a, MNRAS, 273, 1097 J[ø]{}rgensen, I., Franx, M., Kj[æ]{}rgaard, P. 1995b, MNRAS, 276, 1341 J[ø]{}rgensen, I., Franx, M., Kj[æ]{}rgaard, P. 1996, MNRAS, 280, 167 (JFK96) Kelson, D., van Dokkum, P. G., Franx, M., Illingworth, G. D., Fabricant, D. 1997, ApJ, in press Kj[æ]{}rgaard, P., J[ø]{}rgensen, I., Moles, M. 1993, ApJ, 418, 617 Le Borgne, J. F., Pelló, R., Sanahuja, B. 1992, A&AS, 95, 87 Lilly, S. J., Tresse, L., Hammer, F., Crampton, D., Le Fèvre, O. 1995, ApJ, 455, 108 Lucey, J. R., Guzmán, R., Carter, D., Terlevich, R. J. 1991, MNRAS, 253, 584 Moore, B., Matz, N., Lake, G., Dressler, A., Oemler, A. Jr. 1996, Nature, 379, 613 Oemler, A. Jr., Dressler, A., Butcher, H. 1997, ApJ, 474, 561 Salpeter, E. E. 1956, ApJ, 121, 161 Tully, R. B., Fisher, J. R. 1977, A&A, 54, 661 van Dokkum, P. G., Franx, M. 1996, MNRAS, 281, 985 Vazdekis, A., Casuso, E., Peletier, R. F., Beckman, J. E. 1996, ApJS, 106, 307 Vogt, N. P., Forbes, D. A., Phillips, A. C., Gronwall, C., Faber, S. M., Illingworth, G. D., Koo, D. C. 1996, ApJ, 465, L15 [^1]: Hubble Fellow
{ "pile_set_name": "ArXiv" }
--- abstract: | For an even dimensional, compact, conformal manifold without boundary we construct a conformally invariant differential operator of order the dimension of the manifold. In the conformally flat case, this operator coincides with the critical [GJMS]{} operator of Graham-Jenne-Mason-Sparling. We use the Wodzicki residue of a pseudo-differential operator of order $-2,$ originally defined by A. Connes, acting on middle dimension forms. Math.Subj.Clas.: 53A30. Keywords: Wodzicki’s residue, GJMS operators. Research supported in part by NSF grant DMS-9983601 author: - 'William J. Ugalde' title: 'A construction of critical [GJMS]{} operators using Wodzicki’s residue' --- Introduction ============ In [@Con1] Connes uses his quantized calculus to find a conformal invariant. A central part of the explicit computation of this conformal invariant in the 4-dimensional case is the study of a trilinear functional on smooth functions over the manifold $M$ given by the relation $$\tau(f_0,f,h) = \operatorname{Wres}(f_0[F,f][F,h]),$$ where $\operatorname{Wres}$ represents the Wodzicki residue, and $F$ is a pseudo-differential operator of order 0 acting on 2-forms over $M.$ $F$ is given as $2{\mathcal{D}}-1$ with ${\mathcal{D}}$ the orthogonal projection on the image of $d$ in ${\mathcal{H}}={\Lambda}^2(T^*M)\ominus H^2$ with $H^2$ the space of harmonic forms. Using the general formula for the total symbol of the product of two pseudo-differential operators, Connes computed a natural bilinear differential functional of order $4,$ $B_4$ acting on $C^\infty(M).$ This bilinear differential functional $$B_4(f,h) = 4\Delta(\<df,dh>) -2\Delta f \Delta h + 4\<{\nabla}df, {\nabla}dh> +8\<df,dh>J,$$ is symmetric: $B_4(f,h) = B_4(h,f),$ conformally invariant: $\widehat{B_4}(f,h) = e^{-4\eta}B_4(f,h),$ for $\widehat{g} = e^{2\eta}g,$ and uniquely determined, for every $f_i$ in $C^\infty(M),$ by the relation: $$\tau(f_0,f_1,f_2) = \int_M f_0 B_4(f_1,f_2)\,d x.$$ On an $n$ dimensional manifold, we use $J$ to represent the normalized scalar curvature $2(n-1)J=\operatorname{Sc}.$ The [GJMS]{} operators [@GJMS] are invariant operators on conformal densities $$P_{2k}:\mathcal{E}[-n/2+k] \to \mathcal{E}[-n/2-k]$$ with principal parts $\Delta^k$, unless the dimension is even and $2k > n.$ If $n$ is even, the $n$-th order operator $$P_n:\mathcal{E}[0] \to \mathcal{E}[-n]$$ is called *critical* [GJMS]{} operator. As noted in [@GJMS], $\mathcal{E}[0]=C^\infty(M)$ and $\mathcal{E}[-n]$ is the bundle of volume densities on $M.$ In the 4-dimensional case, Connes has shown that $$\begin{aligned} P_4(f) &= \Delta^2 f +2 \Delta(f)\,J + 4\<{{\mbox{\sf P}}},{\nabla}df> + 2\<dJ,df> \\ &= {\delta}(d{\delta}+(n-2) J - 4{{\mbox{\sf P}}})d\, f,\end{aligned}$$ the Paneitz operator (*critical* [GJMS]{} for $n=4$), can be derived from $B_4$ by the relation $$\int_M B_4(f,h)\,d x = \frac{1}{2} \int_M f P_4(h)\,d x.$$ The GJMS operators $P_{2k},$ of Graham-Jenne-Mason-Sparling [@GJMS] by construction have the following properties. - $P_{2k}$ exists for all $k$ if $n$ is odd, and if $n$ is even exists for $1\leq k \leq n/2.$ - $P_{2k}$ is formally self-adjoint. - $P_{2k}$ is conformally invariant in the sense that $$\widehat P_{2k}=e^{-(n/2+k)\eta} P_{2k} e^{(n/2-k)\eta}$$ for conformally related metrics $\widehat g = e^{2\eta}g.$ - $P_{2k}$ has a polynomial expression in ${\nabla}$ and $R$ in which all coefficients are rational in the dimension $n$. - $P_{2k}=\Delta^k + \operatorname{lot}$. (Here and below, $\operatorname{lot}= $“lower order terms”.) - $P_{2k}$ has the form $${\delta}S_{2k} d + \bigl(\dfrac{n}2-k\bigr) Q_{2k},$$ where $Q_{2k}$ is a local scalar invariant, and $S_{2k}$ is an operator on 1-forms of the form $$(d\delta)^{k-1}+\operatorname{lot}\mbox{~or~}\Delta^{k-1}+\operatorname{lot}.$$ In this last expression, $d$ and ${\delta}$ are the usual de Rham operators and $\Delta$ is the form Laplacian $d \delta + \delta d$. (o), (ii), and (iv) are proved in [@GJMS]. The fact that (o) is exhaustive is proved in [@GoHi]. (i), (iii), and (v) are proved in [@Bra5]. The original work [@GJMS] uses the “ambient metric construction” of [@FeGra-amc] to prove their existence. Since their appearance, there has been a considerable interest in constructive ways to obtain formally selfadjoint, conformally invariant powers of the Laplacian on functions with properties as stated. Examples can be found in [@GraZw] with scattering theory, in [@FeGra] with Poincaré metrics, and in [@GoPe] with tractor calculus. In [@mio1] (see also [@mio2]) we have proved the following result: \[theoremonOmeganinmio1\] Let $M$ be a compact conformal manifold without boundary. Let $S$ be a pseudo-differential operator of order 0 acting on sections of a vector bundle over $M$ such that $S^2 f_1 = f_1 S^2 $ and the pseudo-differential operator $P=[S,f_1][S,f_2]$ is conformal invariant for any $f_i \in C^\infty(M).$ There exists a unique, symmetric, and conformally invariant, bilinear, differential functional $B_{n,S}$ of order $n$ such that $$\operatorname{Wres}(f_0[S,f_1][S,f_2]) = \int_M f_0 B_{n,S}(f_1,f_2)\, d x$$ for all $f_i \in C^\infty(M).$ This result, aiming to extend part of the work of Connes [@Con1], is based on the study of the formula for the total symbol $\sigma(P_1 P_2)$ of the product of two pseudo-differential operators $P_1$ and $P_2$ given by $$\sum \frac{1}{{\alpha}!} \del^{\alpha}_\xi(\sigma(P_1)) \,D^{\alpha}_x(\sigma(P_2))$$ where ${\alpha}=({\alpha}_1,\dots,{\alpha}_n)$ is a multi-index, ${\alpha}! = {\alpha}_1!\cdots {\alpha}_n!$ and $D^{\alpha}_x = (-i)^{|{\alpha}|} \del^{\alpha}_x.$ This expression for ${\sigma}(P_1P_2)$ is highly asymmetric on $\del_\xi$ and $\del_x.$ Even though, some control can be achieved when $P_1$ say, is given as a “multiplication” operator. If $F$ is the operator $2{\mathcal{D}}-1$ where ${\mathcal{D}}$ is the orthogonal projection on the image of $d$ inside the Hilbert space of square integrable forms of middle dimension without the harmonic ones, ${\mathcal{H}}= L^2(M,\Lambda^{n/2}_{{\mathbb{C}}} T^*M)\ominus H^{n/2}$ as defined in [@Con1] then, this differential functional $B_n=B_{n,F}$ provides a way of constructing operators like the critical [GJMS]{} operators on compact conformal manifolds of even dimension, out of the Wodzicki residue of a commutator operator acting on middle dimension forms. We say here ‘like’ because up to date, we do not know the relation in between the operators we construct and the “ambient metric construction” of [@FeGra-amc]. The following is the main result of this paper: \[theoremonP\_n\] Let $M$ be an even dimensional, compact, conformal manifold without boundary and let $({\mathcal{H}},F)$ be the Fredholm module associated to $M$ by Connes [@Con1]. Let $P_n$ be the differential operator given by the relation $$\label{omegatoP} \int_M B_n(f,h) \,d x = \int_M f P_n(h) \,d x$$ for all $f,h \in C^\infty(M).$ Then, - $P_n$ is formally selfadjoint. - $P_n$ is conformally invariant in the sense $\widehat{P_n}(h) = e^{-n\eta} P_n(h),$ if $\widehat g = e^{2 \eta}g.$ - $P_n$ is expressible universally as polynomial in the ingredients ${\nabla}$ and $R$ with coefficients rational in $n.$ - $P_n(h) = c_n \Delta^{n/2}(h) + \operatorname{lot}$ with $c_n$ a universal constant and $\operatorname{lot}$ meaning “lower order terms”. - $P_n$ has the form $\delta S_n d$ where $S_n$ is an operator on 1-forms given as a constant multiple of $\Delta^{n/2-1} + \operatorname{lot},$ or $(d\delta)^{n/2-1} + \operatorname{lot}.$ - $P_n$ and $B_n$ are related by: $$P_n(fh) - fP_n(h) - hP_n(f) = -2B_n(f,h).$$ The structure of this paper is as follows. In Section \[notations\] we provide the setting, definitions, and conventions we will use throughout the rest of the paper plus a small explanation of the origin of $\operatorname{Wres}(f_0[F,f][F,h]).$ In Section \[bilineardfiffunc\] we review Theorem \[theoremonOmeganinmio1\] to make this paper as self contained as possible. In Section \[thesymbolofF\], by a recursive computation of the symbol expansion of $F,$ we show that $B_n(f,h)$ has a universal expression as a polynomial on ${\nabla},$ $R,$ $df,$ and $dh.$ In Section \[PnandWres\] we prove Theorem \[theoremonP\_n\]. In [@Dani] we will present the computations conducting to the explicit expression in the 6-dimensional case together with other calculations related to subleading symbols of pseudo-differential operators. Special thanks to Thomas Branson for support and guidance, to the Erwin Schrödinger International Institute for Mathematical Physics for a pleasant environment where some of the last stages of this work took place, and also to Joseph Várilly for suggestion on how to improve an earlier version. Notations and conventions {#notations} ========================= In this paper, $R$ represents the Riemann curvature tensor, $\operatorname{Rc}_{ij} = R^{k}{}_{ikj}$ represents the Ricci tensor, and $\operatorname{Sc}=\operatorname{Rc}^{i}{}_{i}$ the scalar curvature. ${{\mbox{\sf P}}}= (\operatorname{Rc}- Jg)/(n - 2)$ is the Schouten tensor. The relation between the Weyl tensor and the Riemann tensor is given by $$W^i{}_{jkl} = R^i{}_{jkl} + {{\mbox{\sf P}}}_{jk}{{\delta}_K}^i{}_l - {{\mbox{\sf P}}}_{jl}{{\delta}_K}^i{}_k + {{\mbox{\sf P}}}^i{}_l g_{jk} - {{\mbox{\sf P}}}^i{}_k g_{jl}$$ where ${\delta}_K$ represents the Kronecker delta tensor. If needed, we will “raise” and “lower” indices without explicit mention, for example, $g_{mi}R^i{}_{jkl}=R_{mjkl}$. A differential operator acting on $C^\infty(M)$ is said to be natural if it can be written as a universal polynomial expression in the metric $g,$ its inverse $g^{-1},$ the connection ${\nabla},$ and the curvature $R;$ using tensor products and contractions. The coefficients of such natural operators are called natural tensors. If $M$ has a metric $g$ and if $D$ is a natural differential operator of order $d$, then $D$ is said to be conformally invariant with weight $w$ if after a conformal change of the metric $\widehat{g} = e^{2\eta} g$, $D$ transforms as $\widehat{D} = e^{-(w+d) \eta} D e^{w\eta}.$ In the particular case in which $w=0$ we say that $D$ is conformally invariant and the previous reduces to $\widehat{D} = e^{-d\eta} D.$ For a natural bilinear differential operator $B$ of order $n,$ conformally invariance with bi-weight $(w_1,w_2)$ means $$\label{confinv1} \widehat{B}(f,h) = e^{-(w_1+w_2+n)\eta} B(e^{w_1\eta}f,e^{w_2\eta}h)$$ for every $f,h \in C^\infty(M).$ In this work will consider only the case $w_1=w_2=0.$ In this particular situation, because $\widehat{d x} = e^{n\eta} d x$ we have $$\label{confinv2} \widehat B(f,h)\,\widehat{d x} =B(f,h)\,d x.$$ Some times, when the context allows us to do so, we will abuse of the language and say that $B$ is conformally invariant making reference to instead of . When working with a pseudo-differential operator $P$ of order $k,$ its total symbol (in some given local coordinates) will be represented as a sum of $r \times r$ matrices of the form ${\sigma}(P) \sim {\sigma}_k^P + {\sigma}_{k-1}^P + {\sigma}_{k-2}^P+\cdots,$ where $r$ is the rank of the vector bundle $E$ on which $P$ is acting. It is important to note that the different ${\sigma}_j^P$ are defined only in local charts and are not diffeomorphism invariant [@Ku]. However, Wodzicki [@Wo] has shown that the term ${\sigma}_{-n}^P$ enjoys a very special significance. For a pseudo-differential operator $P,$ acting on sections of a bundle $E$ over a manifold $M,$ there is a $1$-density on $M$ expressed in local coordinates by $$\label{wresP} \operatorname{wres}(P) = \int_{||\xi||=1}\operatorname{trace}\bigl({\sigma}_{-n}^P(x,\xi)\bigr)\,d \xi \,d x,$$ where ${\sigma}^P_{-n}(x,\xi)$ is the component of order $-n$ in the total symbol of $P,$ $||\xi||=1$ means the Euclidean norm of the coordinate vector $(\xi_1,\cdots,\xi_n)$ in ${\mathbb{R}}^n,$ and $d \xi$ is the normalized volume on $\set{||\xi||=1}.$ This *Wodzicki residue density* is independent of the local representation. An elementary proof of this matter can be found in [@Fedosov]. The Wodzicki residue, $\operatorname{Wres}(P)=\int \operatorname{wres}(P)$ is then independent of the choice of the local coordinates on $M$, the local basis of $E$, and defines a trace (see [@Wo]). For a compact, oriented manifold $M$ of even dimension $n=2l,$ endowed with a conformal structure, there is a canonically associated Fredholm module $({\mathcal{H}},F)$ [@Con1]. ${\mathcal{H}}$ is the Hilbert space of square integrable forms of middle dimension with an extra copy of the harmonic forms ${\mathcal{H}}=L^2(M,\Lambda^{n/2}_{{\mathbb{C}}} T^*M)\oplus H^{n/2}$. Functions on $M$ act as multiplication operators on ${\mathcal{H}}.$ $F$ is a pseudo-differential operator of order 0 acting in ${\mathcal{H}}.$ On ${\mathcal{H}}\ominus H^{n/2}\ominus H^{n/2} $ it is obtained from the orthogonal projection ${\mathcal{D}}$ on the image of $d,$ by the relation $F=2{\mathcal{D}}-1.$ On $H^{n/2} \oplus H^{n/2}$ it only interchanges the two components. We are only concerned with the non-harmonic part of ${\mathcal{H}}.$ From the Hodge decomposition theorem it is easy to see that in terms of a Riemannian metric compatible with the conformal structure of $M$, $F$ restricted to ${\mathcal{H}}\ominus H^{n/2}\ominus H^{n/2}$ can be written as: $$F = \frac{d {\delta}-{\delta}d}{d \delta + \delta d}.$$ Why $\operatorname{Wres}(f_0[F,f][F,h])$? ----------------------------------------- As pointed out in [@OPS], the quantum geometry of strings is concerned, among other ideas, with determinants of Laplacians associated to a surface with varying metrics. For a compact surface $M$ without boundary and metric $g,$ it is possible to associate to its Laplacian $\Delta_g$ a determinant $$\det{}' \Delta_g = \prod_{\lambda_j \not= 0} \lambda_j$$ where $0 \leq \lambda_0 < \lambda_1 \leq \lambda_2 \leq \cdots$ are the eigenvalues of $\Delta_g$. To make sense out of this formal infinite product, some regularization procedure is needed. The key to study this expression as a function of $g$ is the *Polyakov action* [@Pol] which gives a formula for the variation of $\log \det' \Delta_g$ under a conformal change of $g.$ For a Riemann surface $M$, a map $f = (f^i)$ from $M$ to ${\mathbb{R}}^2$ and metric $g_{ij}(x)$ on $M$, the 2-dimensional Polyakov action is given by $$I(f) = \frac{1}{2\pi} \int_M g_{ij} \,df^i \wedge \star d f^j.$$ By considering instead of $d f$ its quantized version $[F,f],$ Connes [@Con1] quantized the Polyakov action as a Dixmier trace: $$\begin{aligned} \frac{1}{2 \pi} \int_M g_{ij} d f^i \wedge \star d f^j = -\frac{1}{2} \operatorname{Tr}_\omega\bigl(g_{ij}[F,f^i][F,f^j]\bigr).\end{aligned}$$ Because Wodzicki’s residue extends uniquely the Dixmier trace as a trace on the algebra of PsDOs, this quantized Polyakov action has sense in the general even dimensional case. Connes’ trace theorem [@ConAction] states that the Dixmier trace and the Wodzicki residue of an elliptic PsDO of order $-n$ in an $n$-dimensional manifold $M$ are proportional by a factor of $n(2\pi)^n.$ In the 2-dimensional case the factor is $8\pi^2$ and so the quantized Polyakov action can be written as, $$\begin{aligned} -16\pi^2 I=\operatorname{Wres}\bigl(g_{ij} [F,f^i][F,f^j]\bigr).\end{aligned}$$ A bilinear differential functional {#bilineardfiffunc} ================================== For a vector bundle $E$ of rank $r$ over a compact manifold without boundary $M$ of dimension $n,$ let $S$ be a pseudo-differential operator of order $k$ acting on sections of $E.$ Consider $P$ as the pseudo-differential operator given by the product $P= f_0[S,f][S,h]$ with each $f_0,f,h \in C^\infty(M)$. $P$ is acting on the same vector bundle as $S$, where smooth functions on $M$ act as multiplication operators. Each commutator $[S,f]$ defines a pseudo-differential operator of order $k-1,$ thus $P$ has order $2k-2$. Assuming $2k-2 \geq -n,$ in a given system of local coordinates the total symbol of $P$, up to order $-n$, is represented as a sum of $r \times r$ matrices of the form ${\sigma}_{2k-2}^P+{\sigma}_{2k-3}^P+\cdots+{\sigma}_{-n}^P.$ We aim to study $$\operatorname{Wres}(P) = \int_M \int_{||\xi||=1}\operatorname{trace}\bigl({\sigma}^P_{-n}(x,\xi)\bigr)\,d \xi\,d x.$$ Unless otherwise stated, we assume a given system of local coordinates. For the operator multiplication by $f$ we have ${\sigma}(f) = f I$ with $I$ the identity operator on the sections on which it is acting. In this way ${\sigma}(f P_2) = f{\sigma}(P_2)$ and in particular ${\sigma}_{-n}(f P_2) = f{\sigma}_{-n}(P_2).$ As a consequence we obtain the relation $$\operatorname{wres}(f_0 P_2) = f_0 \operatorname{wres}(P_2).$$ The proof of the following two results is similar to the proof for the case $S$ of order 0 in Lemma 2.2 and Lemma 2.3 in [@mio1] so we omit them in here. $[S,f]$ is a pseudo-differential operator of order $k-1$ with total symbol ${\sigma}([S,f]) \sim \sum_{j \geq 1} {\sigma}_{k-j}([S,f])$ where $$\label{sym-nSf} {\sigma}_{k-j}([S,f]) =\sum_{|{\beta}|=1}^j \frac{D^{\beta}_x(f)}{{\beta}!} \del_\xi^{\beta}\bigl({\sigma}^S_{k-(j-|{\beta}|)}\bigr).$$ \[lemmas\_nSS\] For $2k+n \geq 2,$ with the sum taken over $|{\alpha}'|+|{\alpha}''|+|{\beta}|+|{\delta}|+i+j = n+2k$, $|{\beta}|\geq 1$, and $|{\delta}|\geq 1$, $${\sigma}_{-n}([S,f][S,h]) =\sum \frac{D^{\beta}_x(f) D_x^{{\alpha}'+{\delta}}(h)}{{\alpha}'!{\alpha}''!{\beta}!{\delta}!} \del_\xi^{{\alpha}'+{\alpha}''+{\beta}}({\sigma}^S_{k-i}) \del_\xi^{\delta}(D_x^{{\alpha}''}({\sigma}^S_{k-j})).$$ \[defB\] For every $f,h \in C^\infty(M)$ we define $B_{n,S}(f,h)$ as given by the relation $$B_{n,S}(f,h)\,d x := \operatorname{wres}([S,f][S,h]).$$ Because of Lemma \[lemmas\_nSS\], $B_{n,S}(f,h)$ is explicitly given by: $$\int_{||\xi||=1} \sum \frac{D^{\beta}_x(f) D^{{\alpha}''+{\delta}}_x(h)}{{\alpha}'!{\alpha}''!{\beta}!{\delta}!} \operatorname{trace}\biggl( \del^{{\alpha}'+{\alpha}''+{\beta}}_\xi({\sigma}^S_{k-i}) \del^{{\delta}}_\xi(D^{{\alpha}'}_x({\sigma}^S_{k-j})) \biggr)\,d \xi$$ with the sum taken over $|{\alpha}'|+|{\alpha}''|+|{\beta}|+|{\delta}|+i+j = n+2k$, $|{\beta}|\geq 1$, and $|{\delta}|\geq 1.$ The arbitrariness of $f_0$ in the previous construction, implies that $B_{n,S}$ is uniquely determined by its relation with the Wodzicki residue of the operator $f_0[S,f][S,h]$ as stated in the following theorem. \[theoremonomegan1\] There is a unique bilinear differential functional $B_{n,S}$ of order $n$ such that $$\operatorname{Wres}(f_0[S,f][S,h]) = \int_M f_0 B_{n,S}(f,h) \,d x$$ for all $f_0,$ $f,$ $h$ in $C^\infty(M).$ Furthermore, as the left hand side, the right hand side defines a Hochschild $2$-cocycle over the algebra of smooth functions on $M.$ Linearity is evident and uniqueness follows from the arbitrariness of $f_0.$ Let $b$ denote the Hochschild coboundary operator on $C^\infty(M)$ and consider $$\varphi(f_0,f_1,\cdots,f_r) = \operatorname{Wres}(f_0[S,f_1]\cdots[S,f_r]).$$ From the relation $[S,fh] = [S,f]h + f[S,h]$ we have: $$\begin{aligned} &(b\varphi)(f_0,f_1,\cdots,f_r) = \operatorname{Wres}(f_0 f_1 [S,f_2] \cdots [S,f_r]) \\ &\quad + \sum_{j=1}^{r-1}(-1)^j\operatorname{Wres}(f_0 [S,f_1] \cdots [S,f_j f_{j+1}] \cdots [S,f_r]) \\ &\quad +(-1)^r \operatorname{Wres}(f_rf_0[S,f_1] \cdots [S,f_{r-1}]) \\ &=\operatorname{Wres}(f_0 f_1 [S,f_2] \cdots [S,f_r]) \\ &\quad + \sum_{j=2}^{r}(-1)^{j-1}\operatorname{Wres}(f_0 [S,f_1] \cdots [S,f_{j-1}] f_j [S,f_{j+1}] \cdots [S,f_r]) \\ &\quad + \sum_{j=1}^{r-1}(-1)^{j}\operatorname{Wres}(f_0 [S,f_1] \cdots [S,f_{j-1}] f_j [S,f_{j+1}] \cdots [S,f_r]) \\ &\quad + (-1)^r \operatorname{Wres}(f_rf_0[S,f_1] \cdots [S,f_{r-1}]) \\ =& \operatorname{Wres}(f_0 f_1 [S,f_2] \cdots [S,f_r]) + (-1)^{r-1}\operatorname{Wres}(f_0 [S,f_1] \cdots [S,f_{r-1}] f_r) \\ & - \operatorname{Wres}(f_0 f_1 [S,f_2] \cdots [S,f_r]) +(-1)^r \operatorname{Wres}(f_rf_0[S,f_1] \cdots [S,f_{r-1}]) = 0\end{aligned}$$ because of the trace property of $\operatorname{Wres}$. The result follows as the particular case $r=3.$ So far, by taking $f_0 = 1$, by uniqueness, and by the trace property of $\operatorname{Wres}$, we conclude $$\int_M B_{n,S}(f,h) \, d x = \operatorname{Wres}([S,f][S,h]) = \operatorname{Wres}([S,h][S,f]) = \int_M B_{n,S}(h,f)\, d x.$$ From Definition \[defB\], $B_{n,S}(f,h)\,d x = \operatorname{wres}([S,f][S,h])$ but in general $\operatorname{wres}$ is not a trace, hence asserting that $B_{n,S}(f,h)$ is symmetric is asserting that $$\operatorname{wres}([S,f][S,h])=\operatorname{wres}([S,h][S,f]).$$ To conclude the symmetry of $B_{n,S}(f,h)$ on $f$ and $h$, it is necessary to request more properties on the operator $S$. It is enough to have the property $S^2f=fS^2,$ for every $f\in C^\infty(M).$ The symmetry of $B_{n,S}$ follows from the trace properties of the Wodzicki residue and the commutativity of the algebra $C^\infty(M).$ For a proof of the following result see [@mio1]. \[symmetryofOmega\] If $S^2 f = f S^2$ for every $f \in C^\infty(M)$ then the differential functional $B_{n,S}$ in Theorem \[theoremonomegan1\] is symmetric in $f$ and $h$. In the most of the present work we will use the particular case $S^2= I.$ The functional on conformal manifolds {#omegaandfredmod} ------------------------------------- If we endowed the manifold $M$ of a conformal structure, and ask the operator $P=[S,f_1][S,f_2]$ to be independent of the metric in the conformal class, then we can say a little more about this differential functional $B_{n,S}$. \[theoremonOmeganS\] Let $M$ be a compact conformal manifold without boundary. Let $S$ be a pseudo-differential operator acting on sections of a vector bundle over $M$ such that $S^2 f_1 = f_1 S^2 $ and the pseudo-differential operator $P=[S,f_1][S,f_2]$ is conformal invariant for any $f_i \in C^\infty(M).$ There exists a unique, symmetric, and conformally invariant bilinear differential functional $B_{n,S}$ of order $n$ such that $$\operatorname{Wres}(f_0[S,f_1][S,f_2]) = \int_M f_0 B_{n,S}(f_1,f_2)\, d x$$ for all $f_i \in C^\infty(M).$ Furthermore, $\int_M f_0 B_{n,S}(f_1,f_2)\,d x$ defines a Hochschild 2-cocycle on the algebra of smooth functions on $M.$ Uniqueness follows from . Symmetry follows from Theorem \[symmetryofOmega\] and its conformal invariance, $\widehat{B_{n,S}} = e^{-n\eta} B_{n,S},$ follows from its construction. Indeed, the only possible metric dependence is given by the operator $P.$ Roughly speaking, the cosphere bundle ${\mathbb{S}}^*M,$ as a submanifold of $T^*M,$ depends on a choice of metric in $M.$ But since ${\sigma}_{-n}(P)$ is homogeneous of degree $-n$ in $\xi,$ $\{||\xi||=1\}$ can be replaced by any sphere with respect to a chosen Riemannian metric on $M,$ $\{||\xi||_g = 1\}.$ The formula for the change of variable shows that $$\int_{||\xi||=1} \operatorname{trace}({\sigma}_{-n}(P)) d \xi$$ does not change within a given conformal class except trough a metric dependence of $P$ (this argument is taken from the proof of Theorem 2.3 [@Cordelia]). The Hochschild cocycle property follows from Theorem \[theoremonomegan1\]. We restrict ourselves to the particular case of an even dimensional, compact, oriented, conformal manifold without boundary $M$, and $(E,S)$ given by the canonical Fredholm module $({\mathcal{H}},F)$ associated to $M$ by Connes [@Con1], the pseudo-differential operator $F$ of order 0 is given by $F = (d {\delta}- {\delta}d)(d {\delta}+ {\delta}d)^{-1}$ acting on the space $L^2(M,{\Lambda}^{l}_{\mathbb{C}}T^*M)\ominus H^{n/2},$ with $H^{n/2}$ the finite dimensional space of middle dimension harmonic forms. By definition $F$ is selfadjoint and such that $F^2=1.$ We relax the notation by denoting $B_n = B_{n,F}$ in this particular situation. $P$ is the pseudo-differential operator of order $-2$ given by the product $P= f_0[F,f_1][F,f_2]$ with each $f_i \in C^\infty(M)$. $P$ is acting on middle dimension forms. In this situation, Theorem \[theoremonOmeganS\] is stated as follows: \[theoremonOmegan\] If $M$ is an even dimensional, compact, conformal manifold without boundary and $({\mathcal{H}},F)$ is the Fredholm module associated to $M$ by Connes [@Con1] then, there is a unique, symmetric, and conformally invariant bilinear differential functional $B_n$ of order $n$ such that $$\operatorname{Wres}(f_0[F,f_1][F,f_2]) = \int_M f_0 B_n(f_1,f_2) \,d x$$ for all $f_i \in C^\infty(M).$ Furthermore, $\int_M f_0 B_n(f_1,f_2)\, d x$ defines a Hochschild 2-cocycle on the algebra of smooth functions on $M.$ We must only prove its conformal invariance. In Lemma 2.9 [@mio1] we show that the Hodge star operator restricted to middle dimension forms is conformally invariant in fact, acting on $k$-forms we have $\widehat \star = e^{(2k-n)\eta}\star$ for $\widehat g= e^{2\eta}g.$ The space ${\Lambda}^{n/2} T^* M$ of middle dimension forms has an inner product $$\<\xi_1,\xi_2> = \int_M \overline{\xi_1} \wedge \star \xi_2$$ which is unchanged under a conformal change of the metric, that is, its Hilbert space completion $L^2(M,{\Lambda}^{n/2}T^* M)$ depends only on the conformal class of the metric. Furthermore, the Hodge decomposition: $$\begin{aligned} {\Lambda}^{n/2} T^* M &= \Delta({\Lambda}^{n/2} T^* M)\oplus H^{n/2} \\ &=d\delta ({\Lambda}^{n/2} T^* M)\oplus \delta d({\Lambda}^{n/2} T^* M)\oplus H^{n/2},\end{aligned}$$ is preserved under conformal change of the metric. Indeed, the space of middle dimension harmonic forms given as $H^{n/2}=\operatorname{Ker}(d)\cap\operatorname{Ker}(d\star)$ is conformally invariant. Now if $\omega$ is a middle dimension form orthogonal to the space of harmonic forms we have $\omega=(d\widehat \delta + \widehat \delta d)\omega_1 = (d\delta + \delta d)\omega_2$ with $\omega_1, \omega_2$ middle dimension forms. It is not difficult to check that for a k-form $\xi$ we have: $$\widehat \delta \xi = e^{(2(k-1)-n)\eta}\delta e^{(-2k+n)\eta}\xi,$$ so we must have $$d \delta \omega_2 + \delta d \omega_2 = \omega=(d\widehat \delta + \widehat \delta d)\omega_1 = d(e^{-2\eta}\delta \omega_1) + \delta(e^{-2\eta}d\omega_1)$$ which implies $$\omega_0 := \underbrace{d(e^{-2\eta}\delta \omega_1 - \delta \omega_2)}_{\in\,(\operatorname{Im}\delta d)^\perp} = \underbrace{\delta(d\omega_2 - e^{-2\eta} d\omega_1)}_{\in\,(\operatorname{Im}d\delta)^\perp} \in H^{n/2}.$$ But also $\omega_0$ is orthogonal to any harmonic form so $\omega_0=0$ and hence $$d \widehat \delta \omega_1 = d \delta \omega_2 \qquad \hbox{and} \qquad \widehat \delta d \omega_1 = \delta d \omega_2.$$ We are concerned with $F$ acting on $\Delta({\Lambda}^{n/2})=\widehat{\Delta}({\Lambda}^{n/2}).$ If $$\omega = d\widehat\delta \omega_1 + \widehat\delta d \omega_1 = d\delta \omega_2 + \delta d \omega_2 ,$$ then $$\widehat{F} \omega = d\widehat\delta \omega_1 - \widehat\delta d \omega_1 =d\delta \omega_2 - \delta d \omega_2 = F \omega$$ thus $\widehat{F} = F.$ Last, $\operatorname{Wres}([F,f][F,h])$ does not depend on the choice of the metric in the conformal class and the result follows. The functional in the flat case {#particularocurrence} ------------------------------- In the particular case of a flat metric, we have ${\sigma}_{-k}^F = 0$ for all $k \geq 1$ and so the symbol of $F$ coincides with its principal symbol given by $${\sigma}^F_0(x,\xi) = (\varepsilon_\xi \iota_\xi - \iota_\xi \varepsilon_\xi)||\xi||^{-2}$$ for all $(x,\xi) \in T^*M$, where $\varepsilon_\xi$ denotes exterior multiplication and $\iota_\xi$ denotes interior multiplication. Using this result, it is possible to give a formula for $B_n$ in the flat case using the Taylor expansion of the function $$\psi(\xi,\eta)=\operatorname{trace}\bigl({\sigma}_0^F(\xi) {\sigma}_0^F(\eta)\bigr).$$ Here $\xi,\eta \in T^*_x M \smallsetminus \{0\}$ and ${\sigma}_0^F(\xi) {\sigma}_0^F(\eta)$ is acting on middle-forms. \[theoremtraces0s0\] With ${\sigma}^F_0(\xi) {\sigma}^F_0(\eta)$ acting on $m$-forms we have: $$\psi(\xi,\eta) = \operatorname{trace}({\sigma}^F_0(\xi) {\sigma}^F_0(\eta)) = a_{n,m} \frac{\<{\xi},{\eta}>^2}{||\xi||^2||\eta||^2} + b_{n,m},$$ where $$b_{n,m} = \binom{n-2}{m-2} + \binom{n-2}{m} - 2 \binom{n-2}{m-1} = \binom{n}{m} - a_{n,m}.$$ For the proof see Theorem 4.3 [@mio1]. For the case $n = 4$ and $m = 2$ we obtain $$\psi(\xi,\eta) = 8 \<\xi,\eta>^2(||\xi||||\eta||)^{-2} - 2.$$ Here there is a discrepancy with Connes, he has $4 \<\xi,\eta>^2(||\xi||||\eta||)^{-2} +~\mathrm{constant}$. For $n = 6$, $m = 3$, $$\psi(\xi,\eta) = 24 \<\xi,\eta>^2(||\xi||||\eta||)^{-2} - 4.$$ In the flat case, with $S=F,$ reduces to: $${\sigma}_{-r}([F,f]) = \sum_{|{\beta}|=r} \frac{D_x^{\beta}f}{{\beta}!} \del_\xi^{\beta}({\sigma}_0^F).$$ Using this information we deduce from Lemma \[lemmas\_nSS\] with ${\alpha}' = 0,$ ${\alpha}'' = {\alpha},$ and by the definition of $B_{n,S}:$ $$B_{n\,{\mathrm{flat}}}(f,h) = \int_{||\xi||=1} \sum \frac{D^{\beta}_x f D^{{\alpha}+{\delta}}_x h}{{\alpha}!{\beta}!{\delta}!} \operatorname{trace}\biggl( \del^{{\alpha}+{\beta}}_\xi({\sigma}_0^F)\del^{\delta}_\xi({\sigma}_0^F) \biggr) \,d \xi$$ with the sum taken over $|{\alpha}|+|{\beta}|+|{\delta}| = n$, $1\leq |{\beta}|$, $1\leq |{\delta}|.$ We denote by $T'_n\psi(\xi,\eta,u,v)$ the term of order $n$ in the Taylor expansion of $\psi(\xi,\eta)$ minus the terms with only powers of $u$ or only powers of $v$. That is to say, $$\label{T'} T'_n \psi(\xi,\eta,u,v) := \sum \frac{u^{\beta}}{{\beta}!} \frac{v^{\delta}}{{\delta}!} \operatorname{trace}\bigl(\del^{\beta}_\xi({\sigma}_0^F(\xi)) \del^{\delta}_\eta({\sigma}_0^F(\eta))\bigr),$$ with the sum taken over $|{\beta}|+|{\delta}|=n,$ $|{\beta}|\geq 1,$ and $|{\delta}|\geq 1.$ There is an explicit expression for $B_n$ in the flat case in terms of the Taylor expansion of $\psi(\xi,\eta):$ \[lemmaomeganflat\] $$B_{n\,{\mathrm{flat}}}(f,h) = \sum A_{a,b} (D_x^a f)(D_x^b h),$$ where $$\sum A_{a,b} u^a v^b = \int_{||\xi||=1} \bigl(T'_n\psi(\xi,\xi,u+v,v) - T'_n\psi(\xi,\xi,v,v) \bigr) \,d \xi$$ with $T'_n\psi(\xi,\eta,u,v)$ is given by . For details and a proof of the previous theorem see Section 4 [@mio1]. In the 4 dimensional flat case, we obtain with Maple: $$B_{4,{\mathrm{flat}}}(f,h) = -4\Bigl(\llucovd{f}{i}{j}{j}\ucovd{h}{i} + \llucovd{h}{i}{j}{j}\ucovd{f}{i}\Bigr) -4\llcovd{f}{i}{j} \uucovd{h}{i}{j} -2\lucovd{f}{i}{i} \lucovd{h}{j}{j},$$ and in the 6 dimensional flat case: $$\begin{aligned} &B_{6\,{\mathrm{flat}}}(df,dh) \\ &= 12\,(\lcovd{f}{i}\ululucovd{h}{i}{j}{j}{k}{k} + \lcovd{h}{i}\ululucovd{f}{i}{j}{j}{k}{k}) + 24\,(\llcovd{f}{i}{j}\uulucovd{h}{i}{j}{k}{k} + \llcovd{h}{i}{j}\uulucovd{f}{i}{j}{k}{k}) \\ &\quad + 6\,(\lucovd{f}{i}{i}\lulucovd{h}{j}{j}{k}{k} + \lucovd{h}{i}{i}\lulucovd{f}{j}{j}{k}{k}) + 24\,\llucovd{f}{i}{j}{j}\ulucovd{h}{i}{k}{k} + 16\lllcovd{f}{i}{j}{k}\uuucovd{h}{i}{j}{k},\end{aligned}$$ Here each summand is explicitly symmetric on $f$ and $h.$ \[universalflat\] By looking at the construction of $B_n$ in the flat case and by Theorem \[lemmaomeganflat\], we know that the coefficients of $B_{n\,{\mathrm{flat}}}$ are obtained from the Taylor expansion of the function $\psi(\xi,\xi)$ and integration on the sphere $||\xi||=1.$ Hence they are universal in the flat case. Furthermore, since the volume of the sphere ${\mathbb{S}}^{n-1}$ is $2\pi^{n/2}/{\Gamma}(n/2),$ any integral of a polynomial in the $\xi_i$’s with rational coefficients will be a rational multiple of $\pi^{n/2}.$ To avoid the factor $\pi^{n/2},$ we have assumed in $d \,\xi$ to be normalized. We conclude from Theorem \[lemmaomeganflat\] that the coefficients of $B_{n\,{\mathrm{flat}}}(f,h)$ are all rational in $n.$ In the particular case in which $\widehat g = e^{2\eta}g$ with $g$ the flat metric on $M,$ we have the conformal change equation for the Ricci tensor: $$\label{invariantize} \llcovd{\eta}{i}{j} = - {{\mbox{\sf P}}}_{ij} - \lcovd{\eta}{i}\,\lcovd{\eta}{j} + \frac{1}{2}\lcovd{\eta}{k}\,\ucovd{\eta}{k}\,g_{ij},$$ which allows to replace all higher derivatives on $\eta$ with terms with the Ricci tensor. In this way, using it is possible to obtain the expression for $B_n$ in the conformally flat case out of its expression in the flat metric. We exemplified this process in [@mio2] for the six dimensional case. The universality of the expression for $B_{n\,{\mathrm{flat}}}$ is preserved into the universality of the expression for $B_n$ in the conformally flat case, $B_{n\,{\mathrm{conf.\,flat}}}.$ In the next section we will see in general that there is a universal expression for $B_n.$ Furthermore, using we conclude that also the coefficients of $B_{n\,{\mathrm{conf.\,flat}}}$ are rational in $n.$ A recursive approach to the symbol of $F$ {#thesymbolofF} ========================================= In this section we give a recursive way of computing ${\sigma}_{-j}(F)$ in some given local charts, by considering $F=2{\mathcal{D}}-1$ with ${\mathcal{D}}$ the orthogonal projection on the image of $d$ inside $L^2(M,\Lambda^{n/2}_\mathbb{C} T^*M)$. Both $F$ and ${\mathcal{D}}$ are pseudo-differential operators of order 0 with ${\sigma}_0(F) = 2{\sigma}_0({\mathcal{D}}) - I$ and ${\sigma}_j(F) = 2 {\sigma}_j({\mathcal{D}})$ for each $j < 0$. We shall use the symbols of the differential operators of order 2, $d{\delta}$ and $\Delta$. The purpose of this section is to understand the relation in between $B_n$ and the curvature tensor. Even though the different ${\sigma}_{-j}(F)$ are not diffeomorphic invariant, an explicit understanding of their expressions in a given coordinate chart can be of optimal use, an example is present in [@Kastler] and [@KaWa] where using normal coordinates and recursive computation of symbols they were able to prove that the action functional, that is, the Wodzicki residue of $\Delta^{-n/2+1},$ is proportional to the integral of the scalar curvature by a constant depending on $n:$ $$\operatorname{Wres}\Delta^{-n/2+1} = \frac{(n/2-1)\Omega_n}{6}\int_M \operatorname{Sc}|v_g|,$$ with $\Omega_n$ the volume of the standard $n$-sphere, and $|v_g|$ the 1-density associated to the normalize volume form of $M.$ From Lemma [2.4.2]{} [@Gilkey], it is possible to understand the relation between the ${\sigma}_j^\Delta$ and the metric at a given point $x.$ The same reasoning can be applied to the operator $d{\delta}-{\delta}d$ and by addition the same is true for $d{\delta}.$ We represent ${\sigma}(\Delta),$ ${\sigma}(d{\delta}-{\delta}d),$ and ${\sigma}(d{\delta})$ as ${\sigma}(\Delta) = {\sigma}_2^\Delta + {\sigma}_1^\Delta + {\sigma}_0^\Delta,$ ${\sigma}(d{\delta}-{\delta}d) = {\sigma}_2^{d{\delta}-{\delta}d} + {\sigma}_1^{d{\delta}-{\delta}d} + {\sigma}_0^{d{\delta}-{\delta}d},$ and ${\sigma}(d{\delta}) = {\sigma}_2^{d{\delta}} + {\sigma}_1^{d{\delta}} + {\sigma}_0^{d{\delta}}$ with each ${\sigma}_j$ the component of homogeneity $j$ on the co-variable $\xi$. The conclusion reads: ${\sigma}_2$ only invokes $g_{ij}(x),$ ${\sigma}_1$ is linear in the first partial derivatives of the metric at $x$ with coefficients depending smoothly on the $g_{ij}(x)$, and ${\sigma}_0$ can be written as a term linear in the second partial derivatives of the metric at $x$ with coefficients depending smoothly on the $g_{ij}(x)$ plus a quadratic term linear in the first partial derivatives of the metric at $x.$ Since in terms of a Riemannian metric compatible with the conformal structure of $M$ we have $\Delta {\mathcal{D}}=d{\delta}$ we know ${\sigma}(\Delta {\mathcal{D}}) = {\sigma}(d{\delta}).$ Thus the formula for the total symbol of the product of two pseudo-differential operators implies $$\begin{aligned} {\sigma}_2^{d{\delta}} + {\sigma}_1^{d{\delta}} + {\sigma}_0^{d{\delta}} &= {\sigma}(d{\delta}) = {\sigma}(\Delta {\mathcal{D}}) \sim \sum\frac{1}{{\alpha}!}\del^{\alpha}_\xi {\sigma}(\Delta) D^{\alpha}_x ({\sigma}({\mathcal{D}})) \\ &\sim \sum \frac{1}{{\alpha}!} \del^{\alpha}_\xi({\sigma}_2^\Delta + {\sigma}_1^\Delta + {\sigma}_0^\Delta) D^{\alpha}_x({\sigma}_0^{\mathcal{D}}+ {\sigma}_{-1}^{\mathcal{D}}+ {\sigma}_{-2}^{\mathcal{D}}+ \cdots).\end{aligned}$$ Expanding the right hand side into sum of terms with the same homogeneity we conclude: \[lemmasymbolofcD\] In any given system of local charts, we can express the total symbol of ${\mathcal{D}}$, ${\sigma}({\mathcal{D}}) \sim {\sigma}_0^{\mathcal{D}}+ {\sigma}_{-1}^{\mathcal{D}}+ \cdots$ in a recursive way by the formulae: $$\begin{aligned} {\sigma}_0^{\mathcal{D}}&= ({\sigma}_2^\Delta)^{-1} {\sigma}_2^{d{\delta}}, \\ {\sigma}_{-1}^{\mathcal{D}}&= ({\sigma}_2^\Delta)^{-1}\Bigl({\sigma}_1^{d{\delta}} - {\sigma}_1^\Delta {\sigma}_0^{\mathcal{D}}- \sum_{|{\alpha}|=1} \del^{\alpha}_\xi({\sigma}_2^\Delta) D^{\alpha}_x({\sigma}_0^{\mathcal{D}}) \Bigr), \\ {\sigma}_{-2}^{\mathcal{D}}&= ({\sigma}_2^\Delta)^{-1}\Bigl({\sigma}_0^{d{\delta}} - {\sigma}_1^\Delta {\sigma}_{-1}^{\mathcal{D}}- {\sigma}_0^\Delta {\sigma}_0^{\mathcal{D}}\\ &\quad - \sum_{|{\alpha}|=1}\left(\del^{\alpha}_\xi({\sigma}_2^\Delta) D^{\alpha}_x({\sigma}_{-1}^{\mathcal{D}}) + \del^{\alpha}_\xi({\sigma}_1^\Delta) D^{\alpha}_x({\sigma}_0^{\mathcal{D}})\right) - \sum_{|{\alpha}|=2} \frac{1}{{\alpha}!} \del^{\alpha}_\xi({\sigma}_2^\Delta)D^{\alpha}_x({\sigma}_0^{\mathcal{D}}) \Bigr), \\ {\sigma}_{-r}^{\mathcal{D}}&= -({\sigma}_2^\Delta)^{-1}\Bigl( {\sigma}_1^\Delta {\sigma}_{-r+1}^{\mathcal{D}}+ {\sigma}_0^\Delta {\sigma}_{-r+2}^{\mathcal{D}}+\sum_{|{\alpha}|=1}\del^{\alpha}_\xi({\sigma}_2^\Delta) D^{\alpha}_x({\sigma}_{-r+1}^{\mathcal{D}}) \\ &\qquad\qquad\quad +\sum_{|{\alpha}|=1}\del^{\alpha}_\xi({\sigma}_1^\Delta) D^{\alpha}_x({\sigma}_{-r+2}^{\mathcal{D}}) + \sum_{|{\alpha}|=2} \frac{1}{{\alpha}!} \del^{\alpha}_\xi({\sigma}_2^\Delta)D^{\alpha}_x({\sigma}_{-r+2}^{\mathcal{D}})\Bigr),\end{aligned}$$ for every $r \geq 3.$ Thus the total symbol of $F$, ${\sigma}^F \sim {\sigma}_0^F + {\sigma}_{-1}^F + \cdots,$ can be recover from the relations ${\sigma}_0^F = 2{\sigma}^{\mathcal{D}}_0 - I$ and ${\sigma}_{-k}^F = 2 {\sigma}_{-k}^{\mathcal{D}}$ for $k\leq 1.$ \[universalityofomega\] The bilinear differential functional $B_n$ on Theorem \[theoremonOmegan\] has a universal expression as a polynomial on ${\nabla}$, $R,$ $df,$ and $dh$ with coefficients rational on $n.$ By choosing the coordinates to be normal coordinates, we can assume the following: $g_{ij}(x)={{\delta}_K}_{ij},$ that the partial derivatives of the metric vanish at $x,$ and that any higher order partial derivative of the metric at $x$ are expressed as polynomials on ${\nabla}$ and $R$ (see for example Corollary 2.9 [@Gray]). By Lemma \[lemmasymbolofcD\] each ${\sigma}_{-k}^F(x,\xi)$ has a polynomial expression on ${\nabla}$ and $R$ at the point $x.$ Every product $\del^{{\alpha}}_\xi({\sigma}^F_{-i}) \del^{{\beta}}_\xi(D^{{\gamma}}_x({\sigma}^F_{-j}))$ will have a polynomial expression on ${\nabla}$ and $R.$ These properties are preserved after integration of $\xi$ on $||\xi||=1$ and hence $$B_n(f,h) := \int_{||\xi||=1} \sum \frac{D^{\beta}_x(f) D^{{\alpha}''+{\delta}}_x(h)}{{\alpha}'!{\alpha}''!{\beta}!{\delta}!} \operatorname{trace}\biggl( \del^{{\alpha}'+{\alpha}''+{\beta}}_\xi({\sigma}^F_{-i}) \del^{{\delta}}_\xi(D^{{\alpha}'}_x({\sigma}^F_{-j})) \biggr)\,d \xi$$ with sum taken over $|{\alpha}'|+|{\alpha}''|+|{\beta}|+|{\delta}|+i+j=n,$ $|{\beta}|\geq 1,$ $|{\delta}|\geq 1$ has an expression as a polynomial in the ingredients ${\nabla}$, $R,$ $df,$ and $dh.$ The explicit expression from Lemma \[lemmasymbolofcD\] for each ${\sigma}_{-k}^F$ in terms of ${\sigma}_i^\Delta$ and ${\sigma}_j^{d{\delta}}$ depends on the metric, the dimension, and the local coordinates since the different ${\sigma}_j$ are not diffeomorphic invariant, but and hence $B_n$ is independent of local representations. We conclude that the expression we obtain for $B_n$ is universal. This independence of local representations allows as to conclude the rationality of the coefficients of $B_n$ as follows. By Lemma \[lemmasymbolofcD\] each ${\sigma}_{-j}^{\mathcal{D}}$ is a polynomial expression in the $\del_x{\sigma}_k^\Delta,$ $\del_x\del_\xi{\sigma}_k^\Delta,$ and $\del_x{\sigma}_k^{d {\delta}},$ with coefficients rational in $n.$ By choosing the local coordinates to be normal coordinates, we can assure that all the expressions of these symbols as polynomials on ${\nabla}$ and $R$ are of coefficients rational in $n.$ A filtration by degree {#afiltrationbydegree} ---------------------- If we fix $h$ (or $f$), $B_n(f,h)$ is a differential operator or order $n-1$ on $f$ (or $h$). It is acting on smooth functions and producing smooth functions. Since $\widehat{B_n} = e^{-n\eta} B_n$ when we transform the metric conformally by $\widehat g =e^{2\eta} g,$ we have that $B_n$ is of level $n,$ (see [@Bra4]) that is, the expression $B_n(f,h)$ is a sum of homogeneous polynomials in the ingredients ${\nabla}^{\alpha}f,$ ${\nabla}^{\beta}h,$ and ${\nabla}^{\gamma}R$ for multi-indices ${\alpha},{\beta},$ and ${\gamma},$ in the following sense, each monomial must satisfies the homogeneity condition given by the rule: $$\mbox{twice~the~appearances~of~$R$~} + \mbox{~number~of~covariant~derivatives} = n$$ where for covariant derivatives we count all of the derivatives on $R,$ $f,$ and $h,$ and any occurrence of $W,$ $Rc,$ ${{\mbox{\sf P}}},$ $Sc,$ or $J$ is counted as an occurrence of $R.$ Furthermore, by the restriction $|{\beta}|\geq 1$ and $|{\delta}|\geq 1$ in Lemma \[lemmas\_nSS\], we know that $B_n(f,h)$ is made of the ingredients ${\nabla}^{\alpha}df,$ ${\nabla}^{\beta}dh,$ and ${\nabla}^{\gamma}R.$ By closing under addition, we denote by ${\mathcal{P}}_n$ the space of these polynomials. For a homogeneous polynomial $p$ in ${\mathcal{P}}_n,$ we denote by $\deg_R$ its degree in $R$ and by $\deg_{\nabla}$ its degree in ${\nabla}.$ In this way, $2\deg_R + \deg_{\nabla}= n,$ with $\deg_{\nabla}\geq 2$ and hence $2\deg_R \leq n-2$ for $B_n(f,h).$ We say that $p$ is in ${\mathcal{P}}_{n,r}$ if $p$ can be written as a sum of monomials with $\deg_R\geq r,$ or equivalently, $\deg_{\nabla}\leq n - 2r.$ We have a filtration by degree: $${\mathcal{P}}_n = {\mathcal{P}}_{n,0} \supseteq {\mathcal{P}}_{n,1} \supseteq {\mathcal{P}}_{n,2} \supseteq\cdots\supseteq {\mathcal{P}}_{n,\tfrac{n-2}{2}},$$ and ${\mathcal{P}}_{n,r} = 0$ for $r> (n-2)/2.$ There is an important observation to make. An expression which a priori appears to be in, say ${\mathcal{P}}_{6,1},$ may actually be in a subspace of it, like ${\mathcal{P}}_{6,2}.$ For example, $$\label{forfiltration} \underbrace{\lcovd{f}{i} \lllcovd{h}{j}{k}{l} W^{ijkl}}_{\in\,{\mathcal{P}}_{6,1}} = \underbrace{\lcovd{f}{i}\lcovd{h}{j} {{\mbox{\sf P}}}_{kl} W^{ikjl} + \frac{1}{2}\lcovd{f}{i}\lcovd{h}{j} W^i{}_{klm}W^{jklm}}_ {\in \,{\mathcal{P}}_{6,2}},$$ by reordering covariant derivatives and making use of the symmetries of the Weyl tensor. In the particular case $n=4,$ $k_R$ can be 0 or 1, hence $B_4$ can be written as $$B_{4}(f,h) = \underbrace{p_4(df,dh)}_{\in\,{\mathcal{P}}_{4,0}} + \underbrace{p_{R,4}(df,dh)}_{\in\,{\mathcal{P}}_{4,1}}$$ where $p_{R,4}(df,dh)$ is a trilinear form on $R,$ $df,$ and $dh.$ Explicitly, $$p_4(df,dh) = B_{4\,{\mathrm{flat}}}(f,h) \hbox{~~and~~} p_{R,4}(df,dh)= 8 \lcovd{f}{i} \ucovd{h}{i} J.$$ In the 6-dimensional case, $k_R \in \{0,1,2\}$ thus $$B_{6}(f,h) = \underbrace{p_6(df,dh)}_{\in {\mathcal{P}}_{6,0} / {\mathcal{P}}_{6,1}} + \underbrace{p_{6,1}(df,dh)}_{\in {\mathcal{P}}_{6,1} / {\mathcal{P}}_{6,2}} + \underbrace{p_{6,2}(df,dh)}_{\in{\mathcal{P}}_{6,2}},$$ with $$p_{6,1} = p_{6,R}+ p_{6,R'} + p_{6,R''},$$ - $p_6(df,dh) = B_{6\,{\mathrm{flat}}}(f,h),$ - $p_{6,R}(df, dh),$ polynomial on ${\nabla}^a df,$ ${\nabla}^b dh,$ and $R,$ - $p_{6,R'}(df, dh),$ polynomial on ${\nabla}^a df,$ ${\nabla}^b dh,$ and ${\nabla}R,$ - $p_{6,R''}(df, dh),$ polynomial on ${\nabla}^a df,$ ${\nabla}^b dh,$ and ${\nabla}{\nabla}R,$ and - $p_{6,2}(df, dh),$ polynomial on ${\nabla}^a df,$ ${\nabla}^b dh,$ and $R R.$ From the previous expressions, it is evident that there exists a sub-filtration inside each ${\mathcal{P}}_{n,l}$ for $l\geq 1.$ Such a filtration is more complicated to describe in higher dimension because of the presence of terms like ${\nabla}^a R {\nabla}^c R\cdots.$ Also it is important to note that $p_n(df,dh)$ is precisely $B_{n,{\mathrm{flat}}}(f,h),$ i.e. the flat version of $B_n$ coincide with the expression in the filtration without curvature terms. $P_n$ and the Wodzicki residue {#PnandWres} ============================== Now we are ready to prove our main result: [**Of Theorem \[theoremonP\_n\]**]{}. Since $B_n(f,h)$ is a differential functional on $f$ and $h$, Stokes’ theorem applied to $\int_M B_n(f,h) \,d x$ leads to the expression $\int_M f P_n(h) \,d x,$ where $P_n$ is a differential operator on $h.$ Uniqueness of $P_n$ follows from the arbitrariness of $f.$ Formally selfadjointness is a consequence of the symmetry of $B_n,$ that is: $$\int_M f P_n(h) \,d x = \int_M B_n(f,h) \,d x = \int_M B_n(h,f) \,d x =\int_M P_n(f) h \,d x.$$ Note that $$\begin{aligned} \int_M f e^{-n\eta} P_n(h) \,d x &= \int_M B_n(fe^{-n\eta},h) \,d x = \int_M \widehat{B_n}(fe^{-n\eta},h) \,\widehat{d x} \\ &= \int_M f e^{-n\eta} \widehat{P_n}(h) \,\widehat{d x} =\int_M f \widehat P_n(h) \,d x\end{aligned}$$ for every $f \in C^\infty(M).$ (ii) follows from the arbitrariness of $f.$ From Lemma \[universalityofomega\] there is a universal expression for $B_n$ and Stoke’s theorem does not affect this universality nor the rationality of its coefficients, thus there is a universal expression for $P_n$ as a polynomial on ${\nabla}$ and $R,$ with coefficients rational on $n,$ and (iii) is proved. Next we prove (iv). Because of the filtration, we can write $B_n(f,h)$ in the form $B_n(f,h) = p_n(df,dh) + p_{n,R}(df,dh)$ where any possible curvature term is in the $p_{n,R}$ part and those terms in $p_n$ are of the form $$\label{fhRfirstkind} f_{;\,j_1}{}^{j_1}\cdots{}_{j_s}{}^{j^s}{}_{i_1 \cdots i_r} \, h_{;\,k_1}{}^{k_1}\cdots{}_{k_t}{}^{k_t i_1 \cdots i_r}$$ with $2r+2s+2t=n.$ Each time we apply Stokes’ theorem to such a term and reorder the indices, we obtain an expression like: $$(\pm)\int f \Delta^{n/2}(h)\,d x + \int f \operatorname{lot}(h,R)\, d x$$ where $\operatorname{lot}(h,R)$ represents a differential operator on $h$ which is polynomial on ${\nabla}$ and $R$ with order on $h$ lower than $n.$ It follows that $P_n(h) = c_n \Delta^{n/2}(h) + \operatorname{lot}$ with $c_n$ a universal constant and $\operatorname{lot}$ a sum of lower order terms. To prove (v) we take a closer look at $B_n(f,h).$ Each time we apply Stokes’ theorem to a term of the form for which $r\geq 1,$ we can reorder the indices in such a way that we have an expression like: $$(\pm)\int f \Delta^{n/2}(h)\,d x + \int f \bigl(\operatorname{lot}(h_i,R)\bigr)_{;\,}{}^i\, d x$$ where now $\operatorname{lot}(h_i,R)$ represents a sum of contractions of products of the form ${\nabla}^{\alpha}h\times$ (some curvature term), for some multi-index ${\alpha}$ with $i$ present in ${\alpha}.$ Each time we apply Stokes’ theorem to a term of the form for which $r=0,$ we can reorder the indices in such a way that we have an expression like: $$\label{fhRthirdkind} (\pm)\int f \Delta^{n/2}(h)\,d x + \int f \bigl(\operatorname{lot}(h,R)\bigr)_{;\,i}{}^i\, d x$$ where $\operatorname{lot}(h,R)$ represents a sum of lower order terms. Each time we apply Stokes’ theorem to any term of $B_n(f,h),$ not of the form , that is to say with some curvature term, we can reorder the indices in such a way that we have an expression of the form: $$\label{fhRfourthkind} \int f\bigl(\operatorname{lot}(h_i,R)\bigr)_{;}{}^i\,d x + \int f\bigl(\operatorname{lot}(h,R_i)\bigr)_{;}{}^i\,d x$$ where $\operatorname{lot}(h,R_i)$ represents a sum of contractions of products of the form ${\nabla}^{\alpha}h\times$ (some curvature term), for some multi-index ${\alpha}$ in such a way that $i$ is present in one of the curvature terms (not in ${\alpha}$). Applying the Leibniz rule to the first appearance of $i$ in the second summand of produces more terms of the same form as . We conclude: $$P_n(h) = c_n\Delta^{n/2}(h) + \sum \bigl(\operatorname{lot}(h_{i},R)\bigr)_{;\,}{}^{i} + \sum \bigl(\operatorname{lot}(h,R_{i})\bigr)_{;\,}{}^{i}.$$ Next, consider the operator $S'_n$ acting on exact 1-forms: $$S'_n :dh = h_{;\,i}\,d x^i \mapsto - \Bigl(\sum \operatorname{lot}(h_{i},R) + \sum \operatorname{lot}(h,R_{i})\Bigr)\,d x^i.$$ It follows that $P_n(h) = c_n\Delta^{n/2}h + \delta S'_n d(h).$ Hence $P_n = \delta S_n d$ where $S_n$ is a constant multiple of $\Delta^{n/2-1} + \operatorname{lot}$ or $(d \delta)^{n/2-1} + \operatorname{lot},$ where any curvature term is absorbed by the $\operatorname{lot}$ part. For (vi), we will prove that for any $k\in C^\infty(M)$ we have $$\label{toprove} \int_M k \bigl(P_n(fh) - fP_n(h) - hP_n(f)\bigr) \,d x = -2\int_M k\,B_n(f,h)\,d x$$ and the result will follow from the arbitrariness of $k.$ By the left hand side of is equal to: $$\begin{aligned} & \int_M B_n(k,fh) - B_n(kf,h) - B_n(kh,f) \,d x \\ &= \operatorname{Wres}\bigl([F,k][F,fh]-[F,kf][F,h]-[F,kh][F,f]\bigr) \\ &= \operatorname{Wres}\bigl(FkFfh - FkfhF - kFFfh + kFfhF \\ &\qquad\quad -FkfFh + FkfhF + kfFFh - kfFhF \\ &\qquad\quad -FhkFf + FhkfF + hkFFf - hkFfF \bigr) \\ &= -2\operatorname{Wres}(kFfFh - kFfhF - kfFFh + kfFhF) \\ &= -2\operatorname{Wres}\bigl(k[F,f][F,h]\bigr) = -2 \int_M k\, B_n(f,h)\,d x,\end{aligned}$$ where we have used the commutativity of the algebra $C^\infty(M),$ the linearity and the trace property of $\operatorname{Wres},$ and the property $F^2=1.$ The value of $c_n$ in Theorem \[theoremonP\_n\] (iii) gets determined in the flat case since no terms with curvature affect it. In the 4 dimensional case $c_4=2$ and in the 6 dimensional case $c_6 =-4.$ The relation between the [GJMS]{} operators in the general even dimensional case and the $P_n$ operators described in this work still unknown to the author at this time. Nevertheless, in the flat case, both operators coincide up to a constant multiple with a power of the Laplacian, $\Delta^{n/2},$ and both operator enjoy the same conformal property $$\widehat P=e^{-n\eta} P$$ for $\widehat g = e^{2\eta} g.$ Hence, they must also coincide, up to a constant multiple, in the conformally flat case. \[lastprop\] In the even dimensional case, inside the conformally flat class of metrics, the critical [GJMS]{} operator and the operator $P_n$ described in this work coincide up to a constant multiple. [25]{} T. P. Branson, *Differential operators canonically associated to a conformal structure*, Math. Scand., **57** (1985), 293–345. T. P. Branson, *Sharp inequalities, the functional determinant, and the complementary series*, Trans. AMS. **347** (1995) 3671-3742. A. Connes, “The action functional in noncommutative-geometry”, Commun. Math. Phys. [**117**]{} (1988) 673-683. A. Connes, *Quantized calculus and applications*, XIth International Congress of Mathematical Physics, Paris 1994, Internatl. Press, Cambridge, MA (1995), 15–36. B. V. Fedosov, F. Golse, E. Leichtman, and E. Schrohe, *The noncommutative residue for manifolds with boundary*, J. Funct. Anal. **142** (1996), 1-31. C. Fefferman and C. R. Graham, *Conformal Invariants*. In *Élie Cartan et les mathématiques d‘aujourd‘hui*, Astérisque, hors série, pages 95-116. Société Mathématique de France, 1985. C. Fefferman and C. R. Graham, *$Q$-Curvature and Poincaré metrics*, Math. Res. Lett. [**9**]{} (2002) 139-151. P. B. Gilkey, *Invariance Theory, the Heat Equation, and the Atiyah–Singer Index Theorem*, 2nd edition, CRC Press, Boca Raton, FL, 1995. A. R. Gover and K. Hirachi, *Conformally invariant powers of the Laplacian – A complete non-existence theorem*, `math.DG/0304082`. A. R. Gover and L. J. Peterson, *Conformally invariant powers of the Laplacian, $Q$-curvature and tractor calculus*, Comm. Math. Phys. 235 (2003), no. 2, 339–378. R. Graham, R. Jenne, L. Mason and G. Sparling, *Conformally invariant powers of the Laplacian, I: Existence*, J. London Math. Soc. (2) **46** (1992), 557–565. C. R. Graham and M. Zworski, *Scattering matrix in conformal geometry*, Invent. Math. 152 (2003), no. 1, 89–118. A. Gray, *The volume of a small geodesic ball of a Riemannian manifold*, Michigan Math. J. **20** (1973), 329–344. W. Kalau and M. Walze, *Gravity, noncommutative geometry and the Wodzicki residue*, J. Geom. Phys. **16** (1995), 327-344. D. Kastler, *The Dirac operator and gravitation*, Commun. Math. Phys. **166** (1995), 633-643. H. Kumano-go, [*Pseudo-Differential Operators*]{}, The MIT Press, Cambridge, Massachusetts, 1981. B. Osgood, R. Phillips, and P. Sarnak, *Extremals of Determinants of Laplacians*, Journal of Functional Analysis [**80**]{}, 148-211, 1988. A. Polyakov, *Quantum geometry of bosonic strings*, Phys. Lett. B **103** (1981), 207–210.—— *Quantum geometry of fermionic strings*, Phys. Lett. B **103** (1981), 211–213. W. J. Ugalde, *Differential forms and the Wodzicki residue*, arXiv:math-DG/0211361. W. J. Ugalde, *Differential forms and the Wodzicki residue*, in *Clifford Algebras: Application to Mathematics, Physics, and Engineering*, R. Ablamowicz, Ed., Progress in Mathematical Physics, Brikhauser, Boston, 2003. W. J. Ugalde, *Automated computations of symbols*, in preparation. J. C. Várilly and J. M. Gracia-Bondía, *Connes’ noncommutative differential geometry and the Standard Model*, J. Geom. Phys. [**12**]{} (1993), 223-301. M. Wodzicki, *Noncommutative residue. Chapter I: Fundamentals*, in *, Arithmetic and Geometry*, Yu. I. Manin, ed., Lecture Notes in Math. [**1289**]{}, Springer, Berlin (1987), pp. 320–399.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We suggest that the recently observed charmed scalar mesons $D_0^{0}(2308)$ (BELLE) and $D_0^{0,+}(2405)$ (FOCUS) are considered as different resonances. Using the QCD sum rule approach we investigate the possible four-quark structure of these mesons and also of the very narrow $D_{sJ}^{+}(2317)$, firstly observed by BABAR. We use diquak-antidiquark currents and work to the order of $m_s$ in full QCD, without relying on $1/m_c$ expansion. Our results indicate that a four-quark structure is acceptable for the resonances observed by BELLE and BABAR: $D_0^{0}(2308)$ and $D_{sJ}^{+}(2317)$ respectively, but not for the resonances observed by FOCUS: $D_0^{0,+}(2405)$.' author: - 'M.E. Bracco$^1$, A. Lozea$^1$, R.D. Matheus$^2$, F. S. Navarra$^2$ and M. Nielsen$^2$' title: 'Disentangling two- and four-quark state pictures of the charmed scalar mesons' --- Recently the first observations of the scalar charmed mesons have been reported. The very narrow $D_{sJ}^+(2317)$ was first discovered in the $D_s^+\pi^0$ channel by the BABAR Collaboration [@babar] and its existence was confirmed by CLEO [@cleo], BELLE [@belle1] and FOCUS [@focus] Collaborations. Its mass was commonly measured as $2317 \MeV$, which is approximately $160 \MeV$ below the prediction of the very successful quark model for the charmed mesons [@god]. The BELLE Collaboration [@belle2] has also reported the observation of a rather broad scalar meson $D_0^{0}(2308)$, and the FOCUS Collaboration [@focus2] reported evidence for broad structures in both neutral and charged final states that, if interpreted as resonances in the $J^P=0^+$ channel, would be the $D_0^{0}(2407)$ and the $D_0^{+}(2403)$ mesons. While the mass of the scalar meson, $D_0^{0}(2308)$, observed by BELLE Collaboration is also bellow the prediction of ref. [@god] (approximately $100 \MeV$), the masses of the states observed by FOCUS Collaboration are in complete agreement with ref. [@god]. Due to its low mass, the structure of the meson $D_{sJ}^+(2317)$ has been extensively debated. It has been interpreted as a $c\bar{s}$ state [@dhlz; @bali; @ukqcd; @ht; @nari], two-meson molecular state [@bcl; @szc], $D-K$- mixing [@br], four-quark states [@ch; @tera; @mppr] or a mixture between two-meson and four-quark states [@bpp]. The same analyses would also apply to the meson $D_0^{0}(2308)$. In the light sector the idea that the scalar mesons could be four-quark bound states is not new [@jaffe] and, therefore, it is natural to consider analogous states in the charm sector. We propose that the resonances observed by BELLE [@belle2] and FOCUS [@focus2] Collaborations be considered as two different resonances. In this work we use the method of QCD sum rules (QCDSR) [@svz] to study the two-point functions of the scalar mesons, $D_{sJ}(2317)$, $D_0(2308)$ and $D_0(2405)$ considered as four-quark states. The use of the QCD sum rules to study the charmed scalar mesons was already done in refs. [@dhlz; @ht; @nari], but in these calculations they were interpreted as two-quark states. In a recent calculation [@sca] some of us have considered that the lowest lying scalar mesons are $S$-wave bound states of a diquark-antidiquark pair. As suggested in ref. [@jawil] the diquark was taken to be a spin zero colour anti-triplet. We extend this prescription to the charm sector and, therefore, the corresponding interpolating fields containing zero, one and two strange quarks are: j\_0&=&\_[abc]{}\_[dec]{}(q\_a\^TC\_5c\_b) (|[u]{}\_d\_5C|[d]{}\_e\^T),\ j\_s&=&[\_[abc]{}\_[dec]{}]{},\ j\_[ss]{}&=&\_[abc]{}\_[dec]{}(s\_a\^TC \_5c\_b)(|[q]{}\_d\_5C|[s]{}\_e\^T), \[int\] where $a,~b,~c,~...$ are colour indices, $C$ is the charge conjugation matrix and $q$ represents the quark $u$ or $d$ according to the charge of the meson. Since $D_{sJ}$ has one $\bar{s}$ quark, we choose the $j_s$ current to have the same quantum numbers of $D_{sJ}$, which is supposed to be an isoscalar. However, since we are working in the SU(2) limit, the isoscalar and isovector states are mass degenerate and, therefore, this particular choice has no relevance here. The QCDSR for the charmed scalar mesons are constructed from the two-point correlation function (q)=id\^4x  e\^[iq.x]{}0 |T\[j\_S(x)j\^\_S(0)\]|0. The coupling of the scalar meson, $S$, to the scalar current, $j_S$, can be parametrized in terms of the meson decay constant $f_S$ as [@sca]: ${\langle}0 | j_S|S{\rangle}=\sqrt{2}f_Sm_S^4$, therefore, the phenomenological side of Eq. (\[2po\]) can be written as \^[phen]{}(q\^2)=[2f\_S\^2m\_S\^8m\_S\^2-q\^2]{}+, where the dots denote higher resonance contributions that will be parametrized, as usual, through the introduction of the continuum threshold parameter $s_0$ [@io1]. In the OPE side we work at leading order and consider condensates up to dimension six. We deal with the strange quark as a light one and consider the diagrams up to order $m_s$. To keep the charm quark mass finite, we use the momentum-space expression for the charm quark propagator. We follow ref. [@su] and calculate the light quark part of the correlation function in the coordinate-space, which is then Fourier transformed to the momentum space in $D$ dimensions. The resulting light-quark part is combined with the charm-quark part before it is dimensionally regularized at $D=4$. We can write the correlation function in the OPE side in terms of a dispersion relation: \^[OPE]{}(q\^2)=\_[m\_c\^2]{}\^ds [(s)s-q\^2]{}, where the spectral density is given by the imaginary part of the correlation function: $\rho(s)={1\over\pi}\mbox{Im}[\Pi^{OPE}(s)]$. After making a Borel transform on both sides, and transferring the continuum contribution to the OPE side, the sum rule for the scalar meson $S$ can be written as 2f\_S\^2m\_S\^8e\^[-m\_S\^2/M\^2]{}=\_[m\_c\^2]{}\^[s\_0]{}ds  e\^[-s/M\^2]{} \_S(s), where $\rho_S(s)=\rho^{pert}(s)+\rh^{m_s}(s)+\rh^{\qq}(s)+\rh^{{\langle}G^2{\rangle}} (s)+\rh^{mix}(s)+\rh^{\qq^2}(s)+\rh^{{\langle}G^3{\rangle}}(s)$, with \^[pert]{}(s)=[12\^[10]{} 3\^6]{}\_\^1 d([1-]{} )\^3(m\_c\^2-s)\^4, \^[G\^2]{}(s)&=&[2\^[10]{}\^6]{}\_\^1 d (m\_c\^2-s) , \^[G\^3]{}(s)=[32\^[12]{} 9\^6]{}\_\^1 d([1-]{})\^3(3m\_c\^2-s), which are common to all three resonances and where the lower limit of the integrations is given by $\La=m_c^2/s$. From $j_0$ we get: $\rh^{m_s}(s)=0$, \^(s)=-[m\_c2\^[6]{}\^4]{}\_\^1 d([1-]{} )\^2(m\_c\^2-s)\^2, \^[mix]{}(s)&=&[m\_c2\^[6]{}\^4]{}, \^[\^2]{}(s)=-[\^212\^2]{}\_\^1 d (m\_c\^2-s). From $j_{s}$ we get: $\rh^{m_s}(s)=0$, \^(s)=[12\^[6]{}\^4]{}\_\^1 d [1-]{} (m\_c\^2-s)\^2 , \^[mix]{}(s)&=&[12\^[6]{}\^4]{}\_\^1 d (m\_c\^2-s)\^[\^2]{}(s)=-[ß12\^2]{}\_\^1 d (m\_c\^2-s). Finally from $j_{ss}$ we get \^[m\_s]{}(s)=-[m\_sm\_c2\^[8]{} 3\^6]{}\_\^1 d([1-]{} )\^3(m\_c\^2-s)\^3, \^(s)=[12\^[6]{}\^4]{}\_\^1 d(m\_c\^2-s)\^2 , \^[mix]{}(s)&=&[12\^[6]{}\^4]{}\_\^1 d (m\_c\^2-s), \^[\^2]{}(s)=-[ß12\^2]{}\_\^1 d (m\_c\^2-s). For the charm quark propagator with two and three gluons attached we use the momentum-space expressions given in ref. [@rry]. In order to get rid of the meson decay constant and extract the resonance mass, $m_S$, we first take the derivative of Eq. (\[sr\]) with respect to $1/M^2$ and then we divide it by Eq. (\[sr\]) to get m\_S\^2=[\_[m\_c\^2]{}\^[s\_0]{}ds  e\^[-s/M\^2]{} s \_S(s)\_[m\_c\^2]{}\^[s\_0]{}ds  e\^[-s/M\^2]{} \_S(s)]{}. In the numerical analysis of the sum rules, the values used for the quark masses and condensates are: $m_s=0.13\,\GeV$, $m_c=1.2\,\GeV$, ${\langle}\bar{q}q{\rangle}=\,-(0.23)^3\,\GeV^3$, $\langle\overline{s}s\rangle\,=0.8{\langle}\bar{q}q{\rangle}$, ${\langle}\bar{q}g\si.Gq{\rangle}=m_0^2 {\langle}\bar{q}q{\rangle}$ with $m_0^2=0.8\,\GeV^2$, ${\langle}g^2G^2{\rangle}=0.5~\GeV^4$ and ${\langle}g^3G^3{\rangle}=0.045~\GeV^6$. The value for the quark condensate was obtained using the Gell-Mann - Oakes - Renner relation, and the mass of the light quarks, $m_u+m_d=14\MeV$, at the renormalization scale of $1\GeV$ [@gale]. Since the charm quark mass introduces a natural scale in the problem, we chose to work at the renormalization scale of $m_c\sim1\GeV$. \[fig1\] We call $D_0^{(0s)}$, $D_{0}^{(1s)}$ and $D_{0}^{(2s)}$ the scalar charmed mesons represented by $j_0$, $j_s$ and $j_{ss}$ (in Eq. (\[int\])) respectively. In Figs. 1 and 2 we show the masses of these three resonances as a function of the Borel mass for different values of the continuum threshold. The Borel window was fixed in such way that the pole contribution is always between 80% and 20% of the total contribution. \[fig2\] Fixing $\sqrt{s_0}=2.7\GeV$ and varying the charm quark and the strange quark masses in the intervals: $1.1\leq m_c\leq 1.3\GeV$ and $0.11\leq m_s\leq 0.15\GeV$, we get results for the resonance masses still between the lower and upper lines in figures 1 and 2. A bigger value for the charm quark mass makes the results more stable as a function of the Borel mass. One can also vary the value of the quark condensate. Keeping the continuum threshold and the quark masses fixed at $\sqrt{s_0}=2.7\GeV$, $m_c= 1.2\GeV$ and $m_s=0.13\GeV$ and varying the quark condensate in the interval: ${\langle}\bar{q}q{\rangle}=\,(-0.23\pm0.01\GeV)^3$, we get a bigger (smaller) result for the resonance masses using a smaller (bigger) value of the condensate. In Fig. 3 we show the the mass of the $D_{0}^{(1s)}$ state, as a function of the Borel mass, for the combination of the values of the continuum threshold and quark condensate that gives the lower and upper limits for the $D_{0}^{(1s)}$ mass. \[fig3\] In ref.[@jala] it was shown that the renormalization scale was an important source of uncertainty, in the analysis of the $B$ meson decay constant. To check how the change of the scale would change our results we also show, through the dashed line in Fig. 3 , the result for the $D_{0}^{(1s)}$ resonance mass using the values of the strange quark mass and quark condensate at the scale $2\GeV$: ${\langle}\bar{q}q{\rangle}(2\GeV)=\,(-0.267\GeV)^3$ and $m_s(2\GeV)=0.10\GeV$ [@jala]. We see that we get a less stable result for the ressonance mass, but it is still compatible with the results at the scale $1\GeV$, considering the variation in the continuum threshold. Therefore, we conclude that it is the variation of the continuum threshold that causes the most significant variations in the resonance masses, and it is our most important source of uncertaintiy. Comparing figures 1 and 2 we see that the $D_{0}^{(1s)}$ and $D_{0}^{(2s)}$ resonance masses are basicaly degenerated, while the mass of $D_{0}^{(0s)}$ is around $100\MeV$ smaller than the others. While it is natural to expect that the inclusion of a strange quark would increase the resonance mass by around the strange quark mass (as was the case when one goes from $D_{0}^{(0s)}$ to $D_{0}^{(1s)}$), it is really interesting to observe that this does not happen when one goes from $D_{0}^{(1s)}$ to $D_{0}^{(2s)}$. In terms of the OPE contributions, we can trace this behavior to the fact that the quark condensate term is smaller in $D_{0}^{(2s)}$ than in $D_{0}^{(1s)}$ (due to the change from $m_c\qq$ to $m_c\ss$), however the inclusion of the term proportional to $m_sm_c$ (which is not present in $D_{0}^{(1s)}$), compensates this decrease. Considering the variations on the quark masses, the quark condensate and on the continuum threshold discussed above, in the Borel window considered here our results for the ressonance masses are given in Table I. \ 0.3cm resonance $D_{0}^{(0s)}$ $D_{0}^{(1s)}$ $D_{0}^{(2s)}$ ------------ ----------------- ----------------- ----------------- mass (GeV) $~2.22\pm0.21~$ $~2.32\pm0.18~$ $~2.30\pm0.20~$ Comparing the results in Table I with the resonance masses given by BABAR, BELLE and FOCUS: $D_{sJ}^+(2317)$, $D_{0}^0(2308)$ and $D_{0}^{0,+}(2405)$, we see that we can identify the four-quark states represented by $D_{0}^{(1s)}$ and $D_{0}^{(2s)}$ with the BABAR and BELLE resonances respectively. However, we do not find a four-quark state whose mass is compatible with the FOCUS resonances, $D_{0}^{0,+}(2405)$. Therefore, we associate the FOCUS resonances, $D_{0}^{0,+}(2405)$, with a scalar $c\bar{q}$ state, since its mass is completly in agreement with the predictions of the quark model in ref. [@god]. It is also interesting to point out that a mass of about $2.4~\GeV$ is also compatible with the the QCD sum rule calculation for a $c\bar{q}$ scalar meson [@ht]. One can still argue that while a pole approximation is justified for the very narrow BABAR resonance, this may not be the case for the rather broad BELLE and FOCUS resonances. To check if the width of the resonances could modify the pattern observed in the masses of the four-quark states, we have modified the phenomenological side of the sum rule, in Eq. (\[sr\]), through the introduction of a Breit-Wigner-type resonance form: \^[phen]{}(M\^2)=2f\_S\^2m\_S\^8\_[(m\_+m\_D)\^2]{}\^[s\_0]{} ds e\^[-s/M\^2]{}\_[BW]{} (s), where \_[BW]{}(s)=[1]{}[(s)m\_S(s-m\_S\^2)\^2+m\_S\^2(s)\^2]{}, with $\Gamma(s)=\Gamma_0{\sqrt{\lambda(s,m_D^2,m_\pi^2)\over\lambda(m_S, m_D^2,m_\pi^2)}}{m_S^2\over s}$, and $\lambda(x,y,z)=x^2+y^2+z^2-2xy-2xz-2yz$. \[fig4\] Of course now we can not obtain an expression for the resonance mass as Eq. (\[m2\]). However, we can still use the resonance mass as a parameter to compare the compatibility between the right-hand side (RHS) and the left-hand side (LHS) of the sum rule in Eq. (\[m2bw\]): =[\_[m\_c\^2]{}\^[s\_0]{}ds e\^[-s/M\^2]{}s \_S(s)\_[m\_c\^2]{}\^[s\_0]{}ds e\^[-s/M\^2]{}\_S(s)]{}. In Fig.4 we show the RHS (solid line) and the LHS of Eq. (\[m2bw\]) for $D_{0}^{(0s)}$, for three different values of the resonance mass, with $\Gamma_0=280~\MeV$ and $\sqrt{s_0}= 2.7~\GeV$. We see that the best agreement is obtained for $m_S\sim2.2~\GeV$, which shows that the inclusion of the width does not change the value of the mass obtained for the resonance. We have presented a QCD sum rule study of the charmed scalar mesons considered as diquark-antidiquark states. We found that the masses of the BABAR, $D_{sJ}^+(2317)$, and BELLE, $D_{0}^0(2308)$, resonances can be reproduced by the four-quark states $(cq)(\bar{q}\bar{s})$ and $(cs)(\bar{u}\bar{s})$ respectively. However, the mass of the FOCUS resonance, $D_{0}^{0,+}(2405)$, which we believe is not the same measured by BELLE, can not be reproduced in the four-quark state picture considered here. Therefore, we interpret it as a normal $c\bar{q}$ state, since its mass is in complete agreement with the predictions of the quark model in ref. [@god]. We also obtain a mass of $\sim 2.2~\GeV$ for a four-quark scalar state $(cq)(\bar{u}\bar{d})$ which was not yet observed, and that should be also rather broad. : We would like to thank I. Bediaga for fruitful discussions. This work has been supported by CNPq and FAPESP. BABAR Coll., B. Auber [*et al.*]{}, Phys. Rev. Lett. [**90**]{}, 242001 (2003); Phys. Rev. [**D69**]{}, 031101 (2004). CLEO Coll., D. Besson [*et al.*]{}, Phys. Rev. [**D68**]{}, 032002 (2003). BELLE Coll., P. Krokovny [*et al.*]{}, Phys. Rev. Lett. [**91**]{}, 262002 (2003). FOCUS Coll., E.W. Vaandering, hep-ex/0406044. S. Godfrey and N. Isgur, Phys. Rev. [**D32**]{}, 189 (1985); S. Godfrey and R. Kokoshi, Phys. Rev. [**D43**]{}, 1679 (1991). BELLE Coll., K. Abe [*et al.*]{}, Phys. Rev. [**D69**]{}, 112002 (2004). FOCUS Coll., J.M. Link [*et al.*]{}, Phys. Lett. [**B586**]{}, 11 (2004). Y.-B. Dai, C.-S. Huang, C. Liu and S.-L. Zhu, Phys. Rev. [**D68**]{}, 114011 (2003). G.S. Bali, Phys. Rev. [**D68**]{}, 071501(R) (2003). A. Dougall, R.D. Kenway, C.M. Maynard and C. Mc-Neile, Phys. Lett. [**B569**]{}, 41 (2003). A. Hayashigaki and K. Terasaki, hep-ph/0411285. S. Narison, Phys. Lett. [**B605**]{}, 319 (2005). T. Barnes, F.E. Close and H.J. Lipkin, Phys. Rev. [**D68**]{}, 054006 (2003). A.P. Szczepaniak, Phys. Lett. [**B567**]{}, 23 (2003). E. van Beveren and G. Rupp, Phys. Rev. Lett. [**91**]{}, 012003 (2003). H.-Y. Cheng and W.-S. Hou, Phys. Lett. [**B566**]{}, 193 (2003). K. Terasaki, Phys. Rev. [**D68**]{}, 011501(R) (2003). L. Maiani, F. Piccinini, A.D. Polosa, V. Riquer, Phys. Rev. [**D71**]{}, 014028 (2005). T. Browder, S. Pakvasa and A.A. Petrov, Phys. Lett. [**B578**]{}, 365 (2004). R.L. Jaffe, [*Phys. Rev.*]{} [**D15**]{}, 267, 281 (1977); [**D17**]{}, 1444 (1978). M.A. Shifman, A.I. and Vainshtein and V.I. Zakharov, [*Nucl. Phys.*]{}, [**B147**]{}, 385 (1979). T.V. Brito, F.S. Navarra, M. Nielsen, M.E. Bracco, Phys. Lett. [**B608**]{}, 69 (2005). R.L. Jaffe and F. Wilczek, [*Phys. Rev. Lett.*]{} [**91**]{}, 232003 (2003). B. L. Ioffe, Nucl. Phys. [**B188**]{}, 317 (1981); [**B191**]{}, 591(E) (1981). H. Kim, S.H. Lee and Y. Oh, Phys. Lett. [**B595**]{}, 293 (2004). L.J. Reinders, H. Rubinstein and S. Yazaky, [*Phys. Rep.*]{} [**127**]{}, 1 (1985). J. Gasser, H. Leutwyler, [*Phys. Rep.*]{} [**87**]{}, 77 (1982). M. Jamin, B. Lange, [*Phys. Rev.*]{} [**D65**]{}, 056003 (2002).
{ "pile_set_name": "ArXiv" }
--- abstract: 'In an earlier work [@GravSer] we showed that the gravitational interaction could be reproduced as a retarded dispersion interaction (Casimir interaction) between particles composed of hypothetical particles having harmonic oscillator interactions. Here we derive the modification of the force when the vacuum modes have a small but finite temperature. The resulting interaction is of finite range. We give the result on analytical form.' author: - 'Bo E. Sernelius' title: 'Gravitation of finite range and the accelerated expansion of the universe.' --- In a previous work [@GravSer] we opened up for the possibility that gravity, considered to be one of the fundamental interactions in nature, is not a fundamental interaction after all. In our model, gravity is a dispersion interaction derived from more fundamental forces between particles having harmonic oscillator interactions. The dispersion forces are obtained from the change in the zero-point energy of the normal modes of the interaction fields in the system when the position of the composite particles are changed. This is in complete analogy with the electromagnetic dispersion forces between atoms [@CasPol; @Ser]. In the electromagnetic case there are two regions: the van der Waals region for moderate distances; the Casimir region for larger distances. We obtain corresponding regions in our model but with the proper choice of characteristic frequency of the composite particle the van der Waals region will never be observed; the range is smaller than the diameter of the composite particle. The dispersion force will have the $r^{ - 2}$ dependence for all distances just as the gravitational force has. The reason we started looking for an alternative origin of gravity is two-fold. One reason is that one has not been able to verify the existence of the graviton, the quantum of the gravitational field. The second is that recent experiments [@Tur] have pointed to an accelerated expansion of the universe. “Vacuum energy is both the most plausible explanation and the most puzzling possibility”, according to Ref. [@Tur]. Another possible explanation could be found in a modified gravity. In the present work we pursue along this second path. Big Bang occurred 13.7 billion years ago. Soon after the formation of atoms, matter and electromagnetic radiation decoupled. Since then the cosmic background radiation has expanded, in the same way as the rest of the universe, with a factor of approximately 11 000. All wave lengths have increased with, and the temperature of the cosmic background radiation has decreased with, this factor. The present temperature has been found to be 2.7 K. In analogy to this the radiation associated with the interaction between the fundamental particles in our model decoupled from matter at around the time when the composite particles were formed. This happened much earlier than the time of atom formation. Consequently the temperature of this cosmic background radiation is much lower than that of the electromagnetic counterpart. Here we will derive the dispersion force at finite temperature and demonstrate that this force has a finite range. There is a temperature correction factor that goes towards zero when the distance goes towards infinity. The normal modes involved in the dispersion forces are of two types. One type is bound to the particles, the atoms in the electromagnetic case and the composite particles in the case of our model; it affects the forces only at close range, in the van der Waals range. The other type is a vacuum mode, modified by the boundary conditions set up by the presence of the atoms or composite particles; this type of mode dominates at large separations, in the Casimir range; it was the only type of mode in the classical work [@Cas] by Casimir on the attraction between two perfectly conducting plates. This second type should be populated in accordance with the temperature of the cosmic background radiation. The first type, which we are not concerned with here, could be populated differently. We found [@GravSer] the interaction potential at zero temperature is $$\begin{array}{l} V\left( r \right) = - \frac{\hbar }{{2\pi }}\int\limits_0^\infty {d\omega \alpha _1 \left( {i\omega } \right)\alpha _2 \left( {i\omega } \right)e^{ - {{2\omega r} \mathord{\left/ {\vphantom {{2\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \left[ {3 - 6\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)} \right.} \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\left. { + \frac{{17}}{2}\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^2 - \frac{9}{2}\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^3 + \frac{3}{2}\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^4 } \right] \\ \,\,\,\,\,\,\,\,\,\,\,\,\, = - \frac{\hbar }{{2\pi }}\alpha _1 \left( 0 \right)\alpha _2 \left( 0 \right)\int\limits_0^\infty {d\omega e^{ - {{2\omega r} \mathord{\left/ {\vphantom {{2\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \left[ {3 - 6\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)} \right.} \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\left. { + \frac{{17}}{2}\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^2 - \frac{9}{2}\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^3 + \frac{3}{2}\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^4 } \right], \\ \end{array}$$ where we have used the fact that the exponential factor drops off so fast that only the static polarizabilities enter and can be brought outside the integral. Making the replacement $\alpha _i \left( 0 \right) = m_i \sqrt {{{32\pi \gamma } \mathord{\left/ {\vphantom {{32\pi \gamma } {25\hbar c}}} \right. \kern-\nulldelimiterspace} {25\hbar c}}}; i=1,2$ of the static polarizabilities gives $$\begin{array}{l} V\left( r \right) = - \gamma m_1 m_2 \left( {\frac{4}{5}} \right)^2 \frac{1}{c}\int\limits_0^\infty {d\omega e^{ - {{2\omega r} \mathord{\left/ {\vphantom {{2\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \left[ {3 - 6\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)} \right.} \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\left. { + \frac{{17}}{2}\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^2 - \frac{9}{2}\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^3 + \frac{3}{2}\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^4 } \right] \\ \,\,\,\,\,\,\,\,\,\,\,\, = - {{\gamma m_1 m_2 } \mathord{\left/ {\vphantom {{\gamma m_1 m_2 } r}} \right. \kern-\nulldelimiterspace} r}, \\ \end{array}$$ where $\gamma$ is the gravitational constant. The force is obtained as $$\begin{array}{l} F\left( r \right) = - \frac{{\partial V}}{{\partial r}} \\ \,\,\,\,\,\,\,\,\,\,\,\,\, = - \gamma m_1 m_2 \left( {\frac{4}{5}} \right)^2 \frac{1}{{cr}}\int\limits_0^\infty {d\omega e^{ - {{2\omega r} \mathord{\left/ {\vphantom {{2\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \left[ { - 12\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)} \right.} \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, + 29\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^2 - \frac{{61}}{2}\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^3 \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\left. {\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, + 15\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^4 - 3\left( {{{\omega r} \mathord{\left/ {\vphantom {{\omega r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^5 } \right] \\ \,\,\,\,\,\,\,\,\,\,\,\,\, = - {{\gamma m_1 m_2 } \mathord{\left/ {\vphantom {{\gamma m_1 m_2 } {r^2 }}} \right. \kern-\nulldelimiterspace} {r^2 }}. \\ \end{array}$$ At finite temperature the integral turns into a summation [@Ser] $$\int\limits_0^\infty {d\omega } \to \frac{{2\pi }}{{\hbar \beta }}\sum\limits_{n = 0}^\infty {} ;\quad \omega _n = \frac{{2\pi n}}{{\hbar \beta }},$$ where the $n=0$ term should be multiplied by a factor of one half. In the present case we do not have to mind this since this term vanishes. We have $$\begin{array}{l} F\left( r \right) = - \gamma m_1 m_2 \left( {\frac{4}{5}} \right)^2 \frac{{2\pi }}{{\hbar \beta cr}}\sum\limits_{n = 0}^\infty {e^{ - {{2\omega _n r} \mathord{\left/ {\vphantom {{2\omega _n r} c}} \right. \kern-\nulldelimiterspace} c}} \left[ { - 12\left( {{{\omega _n r} \mathord{\left/ {\vphantom {{\omega _n r} c}} \right. \kern-\nulldelimiterspace} c}} \right)} \right.} \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, + 29\left( {{{\omega _n r} \mathord{\left/ {\vphantom {{\omega _n r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^2 - \frac{{61}}{2}\left( {{{\omega _n r} \mathord{\left/ {\vphantom {{\omega _n r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^3 \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\left. {\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, + 15\left( {{{\omega _n r} \mathord{\left/ {\vphantom {{\omega _n r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^4 - 3\left( {{{\omega _n r} \mathord{\left/ {\vphantom {{\omega _n r} c}} \right. \kern-\nulldelimiterspace} c}} \right)^5 } \right]. \\ \end{array}$$ This summation may be performed analytically and we find the temperature correction factor is $$\begin{array}{l} G\left( {r,T} \right) = {{F\left( {r,T} \right)} \mathord{\left/ {\vphantom {{F\left( {r,T} \right)} {F\left( {r,0} \right)}}} \right. \kern-\nulldelimiterspace} {F\left( {r,0} \right)}} \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, = \left( {\frac{4}{5}} \right)^2 x\left[ { - 12z^2 } \right.\, + 29z^3 \left( {x + 1} \right) \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, - \frac{{61}}{2}z^4 \left( {x^2 + 4x + 1} \right) \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, + 15z^5 \left( {x^3 + 11x^2 + 11x + 1} \right) \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\left. {\,\,\,\,\,\,\,\,\,\,\, - 3z^6 \left( {x^4 + 26x^3 + 66x^2 + 26x + 1} \right)} \right], \\ \end{array}$$ where $y = {{2\pi r} \mathord{\left/ {\vphantom {{2\pi r} {\hbar \beta c}}} \right. \kern-\nulldelimiterspace} {\hbar \beta c}}$, $x = \exp \left( { - 2y} \right)$, $z = {y \mathord{\left/ {\vphantom {y {\left( {1 - x} \right)}}} \right. \kern-\nulldelimiterspace} {\left( {1 - x} \right)}}$, and $\beta = {1 \mathord{\left/ {\vphantom {1 {k_B T}}} \right. \kern-\nulldelimiterspace} {k_B T}}$. This temperature correction factor is displayed in Fig. (1). ![The temperature correction factor. See the text for details](Fig1.eps){width="8cm"} In summary, we have demonstrated that the gravitational interaction derived from particles having harmonic oscillator interaction potentials has a finite range if the vacuum modes have a non-zero temperature. We have given the temperature correction factor on analytical form. This correction factor should have an effect on the expansion of the universe. To be noted is that the $r$ and $T$ dependences of the correction factor is in the form of the variable $y$. This means that the cut-off distance for gravity increases with the same speed as the expansion of the universe. This research was sponsored by EU within the EC-contract No:012142-NANOCASE and support from the VR Linné Centre LiLi-NFM and from CTS is gratefully acknowledged. [10]{} Bo E. Sernelius, [*Gravitation as a Casimir interaction*]{}, arXiv:0804.3054v1 \[quant-ph\] (2008). H. B. G. Casimir, and D. Polder, [*Phys. Rev.*]{} [**73**]{}, 360 (1948). Bo E. Sernelius, [*Surface Modes in Physics*]{}, Wiley-VCH (2001). M. S. Turner, and D. Huterer, [*J. Phys. Soc. Jap.*]{} [**76**]{}, 1110151 (2007). H. B. G. Casimir, [*Proc. Koninklijke Nederlandse Akad. Wetenschappen*]{} B [**51**]{}, 793-795 (1948).
{ "pile_set_name": "ArXiv" }
--- abstract: 'Recent results from HERA and TevaTron on precision tests of QCD with jets, $W$ and $Z$ bosons and photons associated with jets and heavy flavours are presented. The measurements were used to probe QCD at the highest energies, to provide experimental constraints on SM processes that constitute background to new physics, to extract values of the coupling of the strong interaction and to constrain the proton parton distribution functions. The implications of the results on LHC physics are discussed.' author: - 'C. Glasman' title: Precision tests of QCD with jets and vector bosons at HERA and TevaTron --- Introduction ============ Parton-parton scattering via QCD interactions is the main process at hadron colliders. Besides their intrinsic interest, these processes represent sometimes an irreducible background to other Standard Model (SM) processes and to new physics. The $ep$ collider HERA, which provides a simple hadronic enviroment to test QCD, and the $\pp$ collider TevaTron, which provides the highest available energies, have produced over the years complementary precision measurements of the parameters of the theory and precise tests of its underlying dynamics. All this knowledge will be crucial for understanding any physics at LHC. For instance, the gluon density at low $x$ is a necessary input ingredient for calculating the Higgs production cross section, whereas the predictions for $W$ or $Z$ bosons production cross sections require knowledge of the quark densities. This report summarises the most recent results from jet cross sections, forward jets, photons and $W$ and $Z$ bosons and the underlying event presented by the CDF, DØ, H1 and ZEUS experiments at ICHEP08. Studies of the underlying event =============================== The main experimental uncertainties for jet cross sections arise from the jet energy scale and the underlying event (UE), which contributes with additional energy density on top of the hard interaction but is not related to it; this is a non-perturbative effect which is not included in the calculations. These effects can be simulated with Monte Carlo simulations and are extremely model dependent, so a good understanding of the UE at the available energies is crucial to model its effects at LHC energies. The CDF collaboration has performed a new study of the UE using Drell-Yan processes [@cdf419]. In this analysis, all particles in the final state, except the lepton pair, can be considered as the UE. Thus, the Drell-Yan events constitute a very clean probe of the UE. The study of the UE observables was done for lepton-pair masses around that of the Z boson ($70<M_{l\bar l}<110$ GeV). The transverse plane was separated in four regions: the toward region, which corresponds to the Z-boson direction, the opposite direction, which is called the away region, and the two transverse regions. The transverse regions are most sensitive to the UE. Several observables, such as the charged particle density shown in Fig. \[fig2\]a, were studied as functions of the transverse momentum of the lepton pair. The measurements are approximately constant for the transverse and toward regions and increase with lepton-pair $p_T$ in the away region. The [PYTHIA]{} tune AW MC predictions describe very well the data. Figure \[fig2\]b shows the comparison of the UE observables in Drell-Yan events with the same observable for the leading-jet analysis in the transverse region. The results are in good agreement so the Drell-Yan studies provide insight into the UE in a cleaner environment. (18.0,4.5) (2.0,-0.5) (8.5,-0.5) (7.0,3.0)[(a)]{} (13.5,3.0)[(b)]{} Photoproduction at HERA also allows the study of the UE. At leading order (LO), there are two processes that contribute to jet photoproduction: direct, in which the photon interacts as a point-like particle with the partons from the proton, and resolved, in which the photon interacts via its partonic structure, giving rise to a hadronic-like final state. The observable $\xo=(1/E_{\gamma})({\sum_{\rm jets}\etjet e^{-\etajet}})$, where $\etjet$ is the jet transverse energy, $\etajet$ is the jet pseudorapidity and $E_{\gamma}$ is the energy of the incoming photon, measures the fraction of the photon energy invested in the hard interaction and can be used to separate both contributions since direct processes take high values of $\xo$, whereas resolved processes take smaller values. Measurements of multijet production in photoproduction are directly sensitive to high orders and can be used to test the parton-shower models. Resolved processes allow tests of multiparton interactions and the UE. The UE has been studied at HERA using dijet events in photoproduction. Jet profiles have been measured [@h1849] by the H1 Collaboration in terms of the mean charged multiplicity as a function of the azimuthal angle, separately for $\xo<0.7$, dominated by resolved processes, and $\xo>0.7$, dominated by direct (see Fig. \[fig3\]a and b). For $\xo<0.7$, a significant contribution from the UE is expected. The mean charged multiplicity has also been measured as a function of the leading-jet $\etjet$ for $\xo<0.7$ in four regions of $\phi$ (see Fig. \[fig3\]c, d, e and f). The mean charged multiplicity increases with increasing $\etjet$ in the toward and away regions and decreases in the transverse regions. The inclusion of multiparton interactions (MPIs) in the [PYTHIA]{} MC prediction improves the description of the data in all four regions. The ZEUS Collaboration has measured [@zeus130] the three-jet cross section in photoproduction as a function of $\xo$ for invariant masses below 50 GeV. The measurement is shown in Fig. \[fig5\]a and displays a two-peak structure at low and high $\xo$ values, which can be identified with the resolved and direct processes, respectively. Figure \[fig5\]b shows the four-jet invariant mass cross section. The data are compared with the predictions of [PYTHIA]{} and [HERWIG]{} MC models with and without MPIs; the models without MPIs describe the data at high invariant mass for all $\xo$ values, but fail at low invariant mass and low $\xo$. The models which include MPIs give an improved description of the data at low $\xo$ and low mass. (18.0,7.0) (0.5,3.5) (6.0,3.5) (11.5,3.5) (0.5,-0.5) (6.0,-0.5) (11.5,-0.5) (4.0,6.0)[(a)]{} (9.5,6.0)[(b)]{} (15.0,6.8)[(c)]{} (4.0,2.8)[(d)]{} (9.5,2.8)[(e)]{} (15.0,2.8)[(f)]{} (18.0,5.0) (2.5,-0.5) (8.5,-0.5) (6.5,4.0)[(a)]{} (12.5,3.0)[(b)]{} Probing QCD at the highest energies =================================== The CDF [@cdf452] and the DØ [@d0506] Collaborations have measured the inclusive-jet cross section in $\pp$ collisions at $\sqrt s=1.96$ TeV as a function of $\etjet$ for different regions of rapidity using the mid-point jet algorithm (see Fig. \[fig6\]). These are high precision measurements, especially at high $\etjet$, where new physics might show up. These cross sections are also sensitive to the gluon density at high $x$. The next-to-leading-order (NLO) QCD calculations give a good description of the data. These measurements constitute the most stringent test of pQCD at the highest available energies so far. (18.0,5.5) (2.0,-0.5) (8.0,-0.8) (7.0,3.0)[(a)]{} (14.5,3.0)[(b)]{} Jet production provides the highest energy reach with highest statistics. In particular, dijet production is ideal to test the SM and search for new physics, which might show up as narrow resonances in the dijet invariant mass spectrum. Figures \[fig7\]a and \[fig7\]b show the measurement of the dijet invariant mass from CDF [@cdf452]; the data reach values of up to 1.4 TeV, which constitute the highest energies measured so far. The NLO calculations give a good description of the data in the whole measured range. Since no evidence for new physics is observed, $95\%$ CL limits were set for various models, which include excited quarks, new heavy vector bosons and gravitons. Figure \[fig7\]c and \[fig7\]d show the limits obtained from the data as functions of the new particle mass. (18.0,10.0) (0.7,4.5) (8.7,4.5) (1.0,-0.5) (9.0,-0.5) (6.5,8.0)[(a)]{} (14.5,8.0)[(b)]{} (4.5,2.5)[(c)]{} (12.5,2.5)[(d)]{} Angular correlations in the dijet system are directly sensitive to the underlying parton dynamics and can also be used to search for new physics. The angular correlation, defined as $\chi_{\rm dijets}=\exp(|y_1-y_2|)$, is expected to be constant for Rutherford scattering; QCD processes induce a small deviation from this behaviour, but new physics is expected to have a very different shape at low values of $\chi_{\rm dijets}$. The DØ Collaboration has measured [@d0507] $\chi_{\rm dijets}$ in different regions of dijet invariant mass from $0.25$ to above $1.1$ TeV, as shown in Fig. \[fig8\]a. The NLO predictions give a good description of the data in the whole measured range. Since there is no indication for the presence of new physics, $95\%$ CL limits were set for various models which include quark compositeness and extra dimensions. Figures \[fig8\]b, \[fig8\]c and \[fig8\]d show the $\chi^2$, likelihood and probability as a function of the characteristic parameter of each model. These constitute the most stringent limits for these models from hadron colliders up to date. (18.0,6.0) (0.0,-0.5) (12.7,-0.5) (12.7,0.0) (9.0,-0.5) (9.0,0.0) (5.3,-0.5) (4.0,4.5)[(a)]{} (9.0,3.5)[(b)]{} (12.7,3.5)[(c)]{} (16.5,3.5)[(d)]{} Precision tests of QCD ====================== Recently, very precise measurements of jet cross sections in neutral current (NC) DIS and photoproduction, which are directly sensitive to the gluon content of the proton, have been incorporated in a QCD fit to determine the proton PDFs. The result was an improved determination of the gluon density for mid- to high-$x$ values, a region relevant for new physics searches at LHC. In some regions of phase space the uncertainty in the gluon density decreased by up to a factor of two. Now the H1 and ZEUS Collaborations are making new and more precise jet measurements with full HERA luminosity and extended phase space to take full advantage of this technique. The ZEUS Collaboration has measured double-differential dijet cross sections in NC DIS [@zeus143] and photoproduction [@zeus125] as functions of $\xi=x_{\rm Bj}(1+M_{\rm jj}^2/\q2)$ and $\xpo=(1/2E_p)({\sum_{\rm jets}\etjet e^{\etajet}})$, respectively, which are both estimators of the fractional momentum carried by the struck parton. The measurements are shown in Fig. \[fig9\] and are well described by the NLO calculations. These analyses provide a stringent test of QCD and were optimised to obtain the best sensitivity to the gluon density in the proton. (18.0,5.0) (3.0,-0.6) (9.0,-0.5) Inclusive-jet cross sections in charged current DIS have been measured [@zeus132] by the ZEUS Collaboration. Figure \[fig10\]a shows the cross section as a function of $x$ for electron beams (sensitive to the $u$-quark density) and positron beams (sensitive to the $d$-quark density). The NLO calculations give a good description of the data. Figure \[fig10\]b shows the theoretical uncertainties, clearly dominated by the PDF uncertainty, which is largest for positron beams at high $x$. Therefore, these measurements have the potential to constrain further the valence-quark PDFs if included in global fits. (18.0,5.5) (1.0,-0.5) (8.0,-1.6) (6.0,3.0)[(a)]{} (14.0,5.0)[(b)]{} Inclusive-jet cross sections in NC DIS were measured [@h1845] by the H1 Collaboration as a function of $\etjet$ in different regions of $\q2$ in the range $5<\q2<100$ 2 (see Fig. \[fig11\]a). The NLO predictions give a good description of the data. However, the theoretical uncertainty, which is dominated by terms beyond NLO, is large: it reaches up to $30\%$ at the lowest $\q2$ values. This shows the need for higher-order corrections. These measurements can help to constrain the gluon PDF at low $\q2$ (low $x$) when higher-order calculations become available for NC DIS. (18.0,5.5) (0.0,6.5) (8.0,6.7) (7.5,4.5)[(a)]{} (16.5,4.5)[(b)]{} Precision measurements of $\as$ ------------------------------- The success of pQCD lies on the precise and consistent determinations of $\as$ from many diverse phenomena, such as structure functions, $\tau$ decays, $Z$-line shape, lattice, jets, etc. At HERA, many determinations of $\as$, mostly extracted from jet observables, give values as precise as those from more inclusive measurements. The H1 and ZEUS Collaborations have made new determinations of $\as$, focusing on decreasing the uncertainties further. The H1 Collaboration has determined [@h1845; @h1844] these values of $\as$: $\asz=0.1186\pm 0.0014({\rm exp.})\pm 0.0134({\rm th.})$ and $\asz=0.1196\pm 0.0010({\rm exp.})\pm 0.0053({\rm th.})$ from the inclusive-jet cross sections at low $\q2$ and the normalised inclusive-jet cross sections (see Fig. \[fig11\]b) in the regions $5<\q2<100$ 2 and $150<\q2<15000$ 2, respectively. In this way, a region of phase space was selected in which experimental uncertainties are well under control and also, the use of normalised cross sections yields a cancellation of correlated uncertainties. New determinations of $\as$ were performed by ZEUS from the inclusive-jet cross sections in NC DIS [@zeus121] and photoproduction [@zeus152] at high $\q2$ and high $\etjet$, respectively, where the theoretical uncertainties are minimised. Figure \[fig12\]a shows the cross section in NC DIS as a function of $\q2$ for different values of the jet-radius parameter $R$ in the $\kt$ algorithm. The NLO calculations give a good description of the data for $R=0.5-1.0$ with similar accuracy. Figure \[fig12\]b shows the cross section as a function of $\etjet$ in photoproduction; the NLO calculation also gives a good description of the data for these processes. Values of $\as$ were extracted from these measurements: $\asz=0.1207\pm 0.0014({\rm stat.})_{-0.0033}^{+0.0035}({\rm syst.})_{-0.0023}^{+0.0022}({\rm th.})$ (NC DIS, $\q2>500$ 2) and $\asz=0.1223\pm 0.0001({\rm stat.})_{-0.0021}^{+0.0023}({\rm syst.})\pm 0.0030({\rm th.})$ (photoproduction, $\etjet>17$ GeV). (18.0,6.0) (2.0,-1.0) (9.0,-0.5) (7.0,4.0)[(a)]{} (13.5,3.5)[(b)]{} To reduce the uncertainties even further and to take advantage of the cross-calibration between experiments, a simultaneous fit to the inclusive-jet cross sections in NC DIS from H1 and ZEUS has been performed [@h1zeus628] to give $\asz=0.1198\pm 0.0019\ {\rm (exp.)}\pm 0.0026\ {\rm (th.)}$. The total uncertainty of the combined value, $\pm 2.7\%$, is very competitive with the most recent result from LEP. Probing QCD with vector bosons ============================== Production of isolated photons in $\pp$ collisions at TeVatron are a probe of QCD dynamics. Photons coming directly from the hard interaction are largely independent of hadronisation corrections. The understanding of these processes in QCD is crucial for searches of new particles that decay into photons. The CDF Collaboration has measured [@cdf411] the inclusive cross section for isolated photons as a function of the photon transverse momentum integrated over the photon rapidity range $|\eta^{\gamma}|<1$. Figure \[fig14\]a shows the measurement together with the NLO predictions. The calculations describe the data adequately within the experimental and theoretical uncertainties. (18.0,5.5) (-0.5,-0.5) (7.0,-0.8) (13.0,-0.2) (5.0,2.0)[(a)]{} (10.5,2.5)[(b)]{} (15.0,3.8)[(c)]{} The DØ Collaboration has studied [@d0509] in more detail these processes by measuring the cross section of isolated photons in association with jets. The measurements were done as functions of the photon transverse momentum for different configurations of photon and jet rapidities. These measurements have the potential to constrain the proton PDFs and different angular configurations give access to different regions of $\q2$ and $x$. Figures \[fig14\]b and c show the measurements together with the NLO calculations. The NLO predictions, using different sets of proton PDFs can not describe the shape of the cross sections simultaneously over the entire range measured. Therefore, the theoretical understanding of these processes needs to be improved before these data can be used to constrain the proton PDFs. Inclusive prompt photons and in association with jets have been measured at HERA in photoproduction by the H1 Collaboration [@h1846]. These processes are sensitive to the PDFs both in the proton and the photon. Inclusive prompt photon production has been measured as a function of the transverse energy and pseudorapidity of the photon (see Figs. \[fig16\]a and \[fig16\]b). The NLO calculations are below the data, especially at low $E_T^{\gamma}$ and low $\eta^{\gamma}$. The production of isolated photons in association with jets has been measured as a function of $E_T^{\gamma}$, $\eta^{\gamma}$ and $\xo$ (see Figs. \[fig16\]c to \[fig16\]e). The NLO calculations give a better description of these data than for inclusive photons, except at high $\xo$. (18.0,7.0) (3.0,3.3) (3.0,-0.5) (0.0,-0.5) (12.5,3.3) (12.5,-0.5) (8.3,-0.5) (5.0,6.5)[(a)]{} (8.0,6.5)[(b)]{} (2.0,2.5)[(c)]{} (5.0,2.5)[(d)]{} (8.0,2.5)[(e)]{} (16.0,5.5)[(f)]{} (12.0,2.5)[(g)]{} (16.5,2.3)[(h)]{} Isolated photons have also been measured [@h1797] in NC DIS by the H1 Collaboration. In this case, there are two major contributions to photon emission, from the lepton and from the quark lines. The measurements have been performed inclusively and in association with jets as functions of photon transverse energy and pseudorapidity (see Figs. \[fig16\]f, g and h). The LO calculations describe the shape of the data but underestimate the normalisation by a factor of approximately two, which can be attributed to an underestimation of the quark-line contribution. The NLO calculations, which are only available for photons in association with jets in this process, are higher than the LO predictions, but still below the data. Therefore, the theoretical understanding of these processes needs to be improved. The production of vector bosons in association with jets in $\pp$ collisions is a key channel for studying top production within the SM as well as for searches of Higgs and new physics. The study of the production via QCD processes also provides a stringent test of the theory. The CDF and DØ Collaborations have measured the production of $W$ [@cdf416] and $Z$ [@d0604] bosons in association with jets. Figure \[fig18\]a shows the measurements of $W+$jets from CDF, as a function of $\etjet$ of the first, second and third jets, and Fig. \[fig18\]b shows the measurement of $Z+$jets from DØ, as a function of $\etjet$. The pQCD calculations, which are NLO for up to two jets, are compared with the measurements. The NLO calculations give a good description of the total and differential cross sections. (18.0,6.5) (3.0,-0.3) (9.0,-0.5) (7.0,4.5)[(a)]{} (14.0,4.5)[(b)]{} $W+c$-jet production at TevaTron is dominated by the $s$-gluon fusion channel and so is sensitive to the $V_{\rm cs}$ matrix element. The measurements are also sensitive to the $s$ and $g$ PDFs. This process constitutes a background to top, Higgs and stop production as well as to other searches for new physics. The DØ Collaboration has measured [@d0605] the total fraction of $W+c$-jet to $W+$jets, $\frac{\sigma(W+c-{\rm jet})}{\sigma(W+{\rm jets})}$, to be $0.074\pm 0.019({\rm stat.})_{-0.014}^{+0.012}({\rm syst.})$, as well as differentially as a function of $\etjet$ (see Fig. \[fig19\]a). The CDF Collaboration has measured [@cdf967] the total cross section times the branching ratio of the $W\rightarrow l\nu$ channel, $\sigma_{W+c-{\rm jet}}\times Br(W\rightarrow l\nu)=9.8\pm 3.2$ pb. Figures \[fig19\]b and \[fig19\]c show the distributions of the difference between same-sign and opposite-sign events as a function of the lepton $p_T$, which shows clearly the signal. The predictions are in reasonable agreement with the data. These measurements provide a direct experimental evidence for the signal and constitute an experimental validation of the $W+c$ theoretical prediction for use in searches. (18.0,5.0) (0.5,-0.5) (6.5,-0.5) (5.5,4.0)[(a)]{} (10.5,4.0)[(b)]{} (15.5,4.0)[(c)]{} $W+b$-jet and $Z+b$-jet production are also important backgrounds to Higgs searches and are sensitive to the $b$ parton density needed to predict the production of Higgs, SUSY, top and other new particles. The CDF Collaboration has measured [@cdf415] the total cross sections for $W+b$-jet and $Z+b$-jet production, $\sigma(Z\!+\!b- {\rm jet})=0.86\pm 0.14\pm 0.12$ pb and $\sigma(W\!+\!b- {\rm jet})\!\times\! BR(W\rightarrow l\nu)=2.74\pm 0.27\pm 0.42$ pb. The predictions are 0.53 pb (NLO) and 0.78 pb ([Alpgen]{}), respectively. The measured $\sigma(Z\!+\!b- {\rm jet})$ cross section is somewhat higher than the NLO prediction. The normalised differential cross sections as functions of $\etjet$ and $p_T^Z$ show that the predictions fall below the data at low values but describe the data well at high $\etjet$ and high $p_T^Z$ (see Fig. \[fig20\]). For $W+b$-jet production, the LO prediction is about 3.5 times lower than the data. The calculation has an uncertainty of $30-40\%$, whereas the data has an uncertainty of only $18\%$, so these measurements should be very helpful to constrain the theory. (18.0,3.5) (2.0,-0.5) (9.0,-0.5) The production of photons in association with $b$- or $c$-jets also provides a test of QCD. The DØ Collaboration has measured [@d0526] the cross sections for photons plus $b$- or $c$-jets as functions of the photon transverse energy in different rapidity configurations. Figure \[fig21\] shows the measurements together with the NLO calculations. The predictions are in good agreement with the data for photon$+b$-jets in all the $E_T^{\gamma}$ range and for photon$+c$-jets for $E_T^{\gamma}<50$ GeV. The disagreement between the photon$+c$ measurements and the theory is seen to grow with increasing $E_T^{\gamma}$. The origin of this discrepancy is not yet clear; it could be attributed, for instance, to intrinsic charm in the proton or to uncertainties in the splitting of gluons into heavy-quark pairs. (18.0,10.5) (2.0,5.0) (8.5,5.0) (2.0,-0.5) (8.5,-0.5) Parton dynamics at low $x$ ========================== One of the main channels of Higgs production at LHC is $gg\rightarrow H$ via a top-quark loop. The predictions for this process need information on the parton evolution at low $x$. This information can be obtained from low-$x$ jet data at HERA. At high scales, calculations at NLO using DGLAP evolution give a good description of the data. However, DGLAP evolution is expected to break down at low $x$. Other approaches to parton dynamics at low $x$ include the BFKL and CCFM evolution schemes. One way to study these effects is the one proposed by Müller and Navelet, which consists of analysing the production of jets close to the proton beam direction at HERA or “forward jets”. To search for breakdown of DGLAP evolution, the ZEUS Collaboration has measured [@zeus126] forward-jet production at low $x$. Figure \[fig22\]a shows the cross section as a function of $x$, together with the pQCD predictions; the measured cross section increases as $x$ decreases. The $\oas$ predictions are well below the data, whereas the $\oass$ calculations are closer to the data, but still fall short. The $\oass$ calculation is much larger than the $\oas$ calculation and has large uncertainties; this indicates that higher orders are important. This can be understood by the opening of a new channel (gluon exchange in the $t$-channel) in these calculations, so that the $\oass$ calculation becomes an effective LO estimation. (18.0,6.0) (2.0,-0.5) (8.0,-0.5) (5.0,3.0)[(a)]{} (9.5,4.5)[(b)]{} Multi-jet cross sections are better suited to test parton dynamics at low $x$ since for dijet and three-jet cross sections a “genuine” NLO calculation can be performed. The ZEUS Collaboration has measured [@zeus122] the angular correlation as a function of $\Delta\phi$ for dijets in different regions of $x$ (see Fig. \[fig22\]b). The $\oass$ calculations are one order of magnitude below the data for small jet angular separations, whereas the $\oasss$ calculations describe the data much better; this demonstrates the importance of the higher orders at low $x$. The H1 Collaboration has measured [@h1798] the $x$ distribution for three-jet events and for the configuration of two central jets and one forward jet (see Fig. \[fig24\]). The $\oasss$ calculation describes the data at low $x$ reasonably well. (18.0,4.5) (2.0,-0.5) (8.0,-0.5) Summary and conclusions ======================= A wealth of new measurements from HERA and TevaTron test QCD nowadays with high precision. These data probe the theory up to the highest available energies and down to the lowest possible $x$ values, phase-space regions of special interest at LHC. The exploration of these new regimes may well lead towards a new level of understanding hard processes which will be crucial for interpreting any physics at LHC. In particular, the underlying event has been tested in all possible environments, new high precision data will help to constrain further the proton PDFs, more and more precise determinations of the strong coupling are being obtained (see Fig. \[fig25\]) and successful tests of colour dynamics at the highest available energies and down to the lowest possible $x$ values have been performed. For further progress in understanding QCD and take full advantage of the available data, higher-orders corrections will be extremely useful. (18.0,7.0) (1.0,-0.5) [**Acknowledgments**]{}. I would like to thank the organisers of the ICHEP08 conference for providing me with the opportunity of giving this talk and for a well organised conference. I would like to thank C. Diaconu, G. Dissertori, K. Hatakeyama, G. Hesketh, A. Juste, S. Kluth, C. Pahl, S. Soldner-Rembold, A. Specka, J. Terrn and M. Wobisch for their help in preparing the talk. Special thanks to for a critical reading of this manuscript. [99]{} , abstract 419, CDF/PUB/CDC/PUBLIC/9351. , abstract 849, H1-prelim-08-036. , abstract 130, S. Chekanov , . , abstract 452, CDF/DOC/JET/PUBLIC/8928 and CDF/9246. , abstract 506, hep-ex/0802.2400. , abstract 507, Conference Note 5733-CONF. , abstract 143, ZEUS-prel-07-005. , abstract 125, S. Chekanov , . , abstract 132, S. Chekanov , . , abstract 845, H1-prelim-08-032. , abstract 121, S. Chekanov , . , abstract 152, ZEUS-prel-08-008. , abstract 844, H1-prelim-08-031. H1 and ZEUS Collaborations, abstract 628, H1-prelim-07-132 and ZEUS-prel-07-025. , abstract 411, `` http://www-cdf.fnal.gov/physics/new/qcd/inclpho08/web.html. , abstract 509, hep-ex/0804.1107. , abstract 846, H1-prelim-08-033. , abstract 797, F.D. Aaron , . , abstract 416, T. Aaltonen , and T. Aaltonen , . , abstract 604, hep-ex/0808.1296. , abstract 605, hep-ex/0803.2259. , abstract 967, T. Aaltonen , . , abstract 415, ``http://www-cdf.fnal.gov/physics/new/qcd/zbjet08/index.html and http://www-cdf.fnal.gov/physics/new/qcd/wbjets1900public/index.html. , abstract 526, Conference Note 5724-CONF. , abstract 126, S. Chekanov , . , abstract 122, S. Chekanov , . , abstract 798, F.D. Aaron , .
{ "pile_set_name": "ArXiv" }
--- author: - 'R. Rossi' - 'N. Prokof’ev' - 'B. Svistunov' - 'K. Van Houcke' - 'F. Werner' title: Polynomial complexity despite the fermionic sign --- The notion of fermion sign problem (FSP) was originally formulated in the context of auxiliary-field, path-integral and diffusion quantum Monte Carlo (QMC) methods [@Loh; @CeperleyPIQMC; @CeperleyHouches; @AssadChapter]. There, it was observed that the computational time required for calculating properties of the fermionic system to a given accuracy scales exponentially with the system volume. Later, the notion of FSP was implicitly extended to an arbitrary QMC approach dealing with interacting fermions, referring to the stochastic sampling of a non-sign-definite quantity with a near cancellation between positive and negative contributions. Sign-free Monte Carlo (MC) algorithms were emerging only as exceptions confirming the rule: In each such case, the absence of FSP was due to some special property of the simulated model (see, [*e.g.*]{}, [@WieseMeron; @SachdevSignFree; @Wei; @AletSignFreeFrustratedSpins; @HoneckerSignFreeFrustratedSpins; @troyer_topol_origin; @capponi_honeycomb] and Refs. therein). Nowadays the FSP is generally perceived as one of the most important unsolved problems in the field of numerical studies of interacting fermionic systems in dimensions $d>1$.[^1] The main message of this letter follows from a simple observation. Suppose some quantity $Q$ is computed from a limit $Q = \lim_{n \to \infty} Q_{n} $, with an exponentially fast convergence, $\vert Q-Q_{n} \vert \sim e^{-\# n}$ (where $\#$ denotes some positive constant). Then in order to compute $Q$ up to an error $| Q-Q_{n}|=\epsilon$ it is sufficient to take $n \sim \ln \epsilon^{-1}$. Hence, even if the computational time increases exponentially as a function of $n$, $t \sim e^{\# n}$, the increase of $t$ as a function of $\epsilon^{-1}$ is only polynomial, $\ln t \sim \ln \epsilon^{-1}$. This observation applies to the simulation of interacting fermions by the algorithm introduced in Ref. [@Rossi], denoted hereafter by the acronym CDet, for Connected Determinant Diagrammatic MC. This algorithm works directly in the thermodynamic limit since it evaluates the series of [*connected*]{} Feynman diagrams. It exploits two advantages of the fermionic sign: First, for fermions on a lattice at finite temperature, the series has a finite radius of convergence, so that the convergence as a function of diagram order $n$ is exponential; second, a factorial number of connected Feynman diagrams can be evaluated in exponential time using determinants (and a recursive formula). In general, sign-alternation of observables simulated by MC methods is neither sufficient nor necessary to state that the problem is intractable. One should rather focus on what we will call the “computational complexity problem” (CCP) instead of the FSP. The key question is the one that is most relevant practically[^2]: How easily can one indefinitely increase the accuracy of the computed thermodynamic-limit answer? This leads to the following definition of the CCP that can be applied to any numerical scheme. Let $Q$ be the intensive quantity of interest in the thermodynamic limit. [*A numerical scheme is said to have CCP if the computational time $t$ required to obtain $Q$ with an error $\epsilon$ diverges faster than any polynomial function of $\epsilon^{-1} \to \infty$. The CCP is considered to be solved if*]{} $$t(\epsilon)\; = \; O(\epsilon^{-\alpha}).$$ [Note that we consider unbiased methods, i.e., $\epsilon\to0$ is the difference between computed value and [*exact*]{} value.]{} In what follows, we show in some detail that CDet solves the CCP, [at least at finite temperature and small enough interaction]{}. In parallel, we also discuss the conventional diagrammatic Monte Carlo approach (hereafter denoted by DiagMC) in which the sum over diagram topologies is done stochastically [@diagmchub]. We then show that the conventional FSP leads to a CCP for path-integral and auxiliary-field QMC. In quantum Monte Carlo, one typically generates configurations $\mathcal{C}$ according to a conveniently chosen unnormalised probability distribution $P(\mathcal{C})$ that is positive. Any sign alternation is taken into account when collecting statistics, and is absorbed into the quantity $A(\mathcal{C})$ that is being measured. The average with respect to $P$, $$\langle A \rangle_{P} = \frac{\sum_{\mathcal{C}} P(\mathcal{C}) A(\mathcal{C}) }{\sum_{\mathcal{C}} P(\mathcal{C}) } \; , \label{eq:avA}$$ is estimated through $$\frac{1}{N_{MC}} \sum_{i=1}^{N_{MC}} A(\mathcal{C}_i) \;, \label{eq:avAMC}$$ with $N_{MC}$ the number of MC measurements and $\mathcal{C}_i$ the configuration at the $i$-th measurement. By the central limit theorem, the $1\sigma$ statistical error on (\[eq:avAMC\]) is given by $$\epsilon_{{\rm stat}} = \sigma_A \sqrt{\frac{2~\tau_{\rm auto}+1}{N_{MC}}} \; , \label{eq:epsstat}$$ with $\tau_{\rm auto}$ the integrated autocorrelation time and $\sigma_A^{\phantom{.}2} = \langle A^2 \rangle_P - \langle A \rangle_P^{\phantom{.}2}$ the variance on individual measurements. We now specialize to CDet and DiagMC. We consider the computation of an observable ([*e.g.*]{} density or double occupancy). For convenience, we make two simplifications regarding the Monte Carlo algorithm. We expect that this does not change the final CCP scaling. The first simplification is that a separate simulation is performed for each order, while the normalisation factors $z_n$ (see below) are known. The second simplification is that in DiagMC, rather than sampling the self-energy diagrams and then obtaining observables from the Dyson equation, we consider here sampling the diagrams for the observable (including one-particle reducible diagrams), so that external variables are simply fixed (to zero in space and imaginary-time representation). A DiagMC configuration is then defined by a Feynman diagram topology together with values of the internal variables $X$. In CDet a configuration is defined only by the internal variables $X$ (the space and imaginary-time coordinates of the interaction vertices), while the weight of a configuration is given by the sum over all possible connected diagram topologies connecting the internal and external vertices. Let us denote the contribution of a diagram of topology $\mathcal{T}$ for fixed internal variables $X$ by $\mathcal{D}(\mathcal{T},X)$. Let $a_n$ be the sum of all Feynman diagrams of order $n$: $$\begin{aligned} a_n = \int dX \sum_{\mathcal{T} \in \mathcal{S}_n} \mathcal{D}(\mathcal{T},X) \; ,\end{aligned}$$ with $\mathcal{S}_n$ the set of all diagram topologies at order $n$. This can be rewritten in the form of Eq. (\[eq:avA\]): $$a_n = \langle A_n \rangle_{P_n} \; ,$$ with the unnormalised distribution to be sampled chosen to be $$\left\{ \begin{array}{ll} \displaystyle P_n(\mathcal{T},X) = \left| \mathcal{D}(\mathcal{T},X) \right|\quad & {\rm (DiagMC)} \\ \displaystyle P_n(X) = \left| \sum_{\mathcal{T}\in \mathcal{S}_n} \mathcal{D}(\mathcal{T},X) \right| \quad & {\rm (CDet)} \end{array} \right.$$ and $$A_n =\left\{ \begin{array}{ll} \displaystyle z_n ~ {\rm sign}[\mathcal{D}(\mathcal{T},X)] \quad & {\rm (DiagMC)} \\ \displaystyle z_n ~ {\rm sign}\left[\sum_{\mathcal{T}\in \mathcal{S}_n} \mathcal{D}(\mathcal{T},X)\right] \quad & {\rm (CDet)} \end{array} \right.$$ with the normalization factors $$z_n =\left\{ \begin{array}{ll} \displaystyle \int dX \sum_{\mathcal{T}\in \mathcal{S}_n} |\mathcal{D}(\mathcal{T}, X)| \quad & {\rm (DiagMC)} \\ \displaystyle \int dX \left|\sum_{\mathcal{T}\in \mathcal{S}_n} \mathcal{D}(\mathcal{T}, X) \right| \quad & {\rm (CDet)} \; . \end{array} \right.$$ So, in DiagMC the diagrams are sampled according to the distribution $ \left| \mathcal{D}(\mathcal{T},X) \right|$, while in CDet diagrams are grouped together via determinants and in the MC part of the algorithm one samples $X$ according to the distribution $\left| \sum_{\mathcal{T}\in \mathcal{S}_n} \mathcal{D}(\mathcal{T},X) \right|$. In what follows we neglect the statistical error on the normalisation factors $z_n$ since they are obtained by sampling a sign-positive quantity. [Here we consider fermions on a lattice at finite temperature, so that]{} the radius of convergence of the diagrammatic series is finite [[@Mastropietro; @diagmchub; @Rossi]]{}. Assuming that we are inside the radius of convergence, the convergence is exponential, $$|a_n| \underset{n \to \infty}{=} O( R^{-n}) \label{eq:an_Rn}$$ with $R>1$ a constant. Here and in what follows we omit multiplicative constants and power laws which do not affect the dominant scaling behavior. The number of diagrams scales factorially with the order $n$. For CDet, however, one takes into account cancellations between different diagram topologies. More specifically, we expect that for fermions on a lattice at finite temperature, $$\begin{aligned} \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! z_n &\underset{n\to \infty}{\sim}& R_D^{-n} ~n! \quad\quad\quad\quad\quad\quad\quad {\rm (DiagMC)} \label{eq:RD} \\ z_n &\underset{n\to \infty}{\sim}& R_C^{-n} \quad \quad \quad\quad \quad \quad \quad \quad \quad {\rm (CDet)} \label{eq:RTD} \end{aligned}$$ with $R_D$ and $R_C$ positive constants.[^3] Let us discuss the behavor of the average sign, $\langle{\rm sign}\rangle := \langle {\rm sign}\, A_n\rangle_{P_n}=a_n/z_n $, as a function of the order $n$. For DiagMC, we see that $\langle{\rm sign}\rangle$ tends to zero factorially, a manifestation of the near-compensation between different diagrams. For CDet, $\langle{\rm sign}\rangle$ tends to zero exponentially in the generic case where $R_C<R$. As a result, the variance on individual measurements behaves as $$\begin{aligned} \sigma_{A_n} = z_n \sqrt{1 - \langle {\rm sign} \rangle^2_{P_n}} \underset{n\to \infty}{\sim} z_n \; , \label{eq:sigmaAn}\end{aligned}$$ which together with Eq. (\[eq:epsstat\]) gives for the statistical error bar on the $n$-th order contribution $a_n$, $$\epsilon_{{\rm stat}}(n) \underset{n\to\infty}{\sim} z_n\, \sqrt{\frac{2~\tau_{\rm auto}(n)+1}{N_{MC}(n)}}. \label{eq:estatN}$$ Here $\tau_{\rm auto}(n)$ is expected to increase at most polynomially with $n$, which we have checked numerically for CDet; we will neglect this $n$-dependence of $\tau_{\rm auto}$ since it will not affect the final scalings. An appropriate dependence $N_{MC}(n)$ of the number of MC steps on order will be specified below. Note that the relative statistical error $\epsilon_{\rm stat}(n)/a_n \propto z_n / a_n = 1/\langle{\rm sign}\rangle_{P_n}$ diverges factorially for DiagMC and exponentially for CDet. This can be viewed as a sign problem for diagrammatic Monte Carlo methods, limiting the order that can be reached. On the other hand, the exponential convergence (\[eq:an\_Rn\]), which is only possible thanks to the fermionic sign, implies that reaching very high orders is not necessary, as we now quantify. The assumption of exponential convergence implies that the systematic error due to the finite diagram-order cut-off $N$, $$\epsilon_{\rm sys}(N) = \sum_{n=N+1}^{\infty} a_n,$$ decreases exponentially, $$\epsilon_{\rm sys}(N) \underset{N \to \infty}{=} O(R^{-N}) \; . \label{eq:esysN}$$ To achieve a final error $\sim\epsilon$, it is then natural to work in a regime where systematic and statistical errors are on the same order. We thus choose $N$ such that $R^{-N} \sim \epsilon$, and we take a computational time $t$ such that the total statistical error is $\epsilon_{\rm stat}\sim\epsilon$. Neglecting correlations between different orders, we have $\epsilon_{\rm stat}^2 \simeq \sum_{n=0}^N \epsilon_{\rm stat}(n)^2$, which leads us to choose $N_{MC}(n)$ such that $\epsilon_{\rm stat}(n)$ is $n$-independent. Equation (\[eq:estatN\]) together with Eqs. (\[eq:RD\],\[eq:RTD\]) then yield $$t_n \underset{n\to\infty}{\sim}\left\{ \begin{array}{ll} \displaystyle \frac{1}{\epsilon^2}\,\frac{(n!)^2}{R_D^{2n}} \quad & {\rm (DiagMC)} \\ \displaystyle \frac{1}{\epsilon^2}\,\left(\frac{3}{R_C^2}\right)^n \quad & {\rm (CDet)} \; , \end{array} \right.$$ where the factor $3$ for CDet comes from the fact that the computational time per MC-step is $\sim 3^n$, because of the recursive formula that needs to be evaluated in order to eliminate disconnected diagrams [@Rossi]. As a result, for DiagMC most time is spent sampling the highest order, while for CDet this is the case only for $R_C<\sqrt{3}$. Finally, we get $$\begin{aligned} t(\epsilon) &\sim& \epsilon^{- \# \ln (\ln \epsilon^{-1})} \quad\quad\quad\quad\quad\quad {\rm (DiagMC)} \\ t(\epsilon) &\sim& \epsilon^{-\alpha} \quad\quad\quad\quad\quad\quad\quad\quad\quad\quad {\rm (CDet)} \label{eq:CCP_CDet}\end{aligned}$$ Hence polynomial scaling is nearly reached with DiagMC, and is achieved with CDet. The exponent for CDet is given by $$\alpha = 2 + \frac{ \ln(3/{R_C}^2) }{\ln R}$$ if $R_C < \sqrt{3}$. The case $R_C > \sqrt{3}$ is particularly instructive. Here, most time is spent sampling low diagram orders, and one has $\alpha=2$, which is the best scaling one can achieve in any Monte Carlo computation. We thus conclude that fermionic sign—all by itself—does not necessarily leads to [*any*]{} qualitative effect on the scaling of computational time with $\epsilon$. ![(a) Absolute value of the sum of all order-$n$ diagrams $|a_n|$, and (b) weight $z_n$ of the order-$n$ configuration space of Connected Determinant Diagrammatic Monte Carlo, for the pressure of the Fermi-Hubbard model (at $U=2$, $\beta=8$, $n \simeq 0.875$). The lines are linear fits to the data at large $n$. \[Fig1\]](an.eps "fig:"){width="\columnwidth"} ![(a) Absolute value of the sum of all order-$n$ diagrams $|a_n|$, and (b) weight $z_n$ of the order-$n$ configuration space of Connected Determinant Diagrammatic Monte Carlo, for the pressure of the Fermi-Hubbard model (at $U=2$, $\beta=8$, $n \simeq 0.875$). The lines are linear fits to the data at large $n$. \[Fig1\]](zn.eps "fig:"){width="\columnwidth"} The above scalings are not purely academic considerations, as we illustrate with an example for CDet. We analyse the computation, reported in Ref. [@Rossi], of the pressure of Fermi-Hubbard model in two dimensions. The diagrammatic scheme is a bare series with bare tadpoles taken into account through a shift of the chemical potential. The Hubbard parameters are: interaction $U = 2$, chemical potential $\mu=0.55978$ and inverse temperature $\beta=8$ (with hopping $=1$); this corresponds to a density $n = 0.87500(2)$. Figure \[Fig1\](a) shows that $|a_n|$ approaches an exponential behavior $R^{-n}$ with $R=2.5(1)$, while Figure \[Fig1\](b) shows that $z_n$ approaches $R_C^{-n}$ with $R_C=0.75(3)$.[^4] We can make three important observations. First, the exponent $\alpha=3.8(2)$ is not too large. Second, we clearly reach the asymptotic regime where Eqs. (\[eq:RTD\],\[eq:esysN\]), and therefore also Eq. (\[eq:CCP\_CDet\]), are valid. Third, $|a_n|$ at low $n$ is ${\sim} 100$ times larger than the extrapolation to low orders of the large-order behavior shown by the straight line in Fig. 1(a). These three observations explain why it was possible to obtain a ${\sim}10^{-6}$ relative accuracy for the pressure in Ref. [@Rossi]. Interestingly, the second and third observations hold independently of $U$ (with $\mu(U)=\mu_0+U n_0/2$ as in [@Rossi]). In contrast, $\alpha$ diverges when $U$ tends to the critical value $U_c = 2\,R(U=2)\simeq 5.1$ such that $R(U_c)=1$. Divergent-series summation methods may allow to approach this point and even to go beyond it; we leave this for future study. To avoid possible confusion, let us remark that we do not claim any connection between our results and the computational complexity theory of computer science. In this theory, a ‘problem instance’ is defined by $\mathcal{N}$ parameters, and the P complexity-class is defined by polynomial scaling of computational time [*with respect to $\mathcal{N}$ for $\mathcal{N}\to\infty$*]{} (in the worst case with respect to all possible instances). In the spin-glass problem discussed in [@TroyerWieseNP], an instance is a disorder realisation, and $\mathcal{N}$ is the number of random couplings (which happens to coincide with the system volume). In contrast, here we consider problems defined by a [*fixed*]{} (usually small) number of model-parameters (e.g. $T$, $U$ and $\mu$ for the Hubbard model in the thermodynamic limit). We turn to ‘traditional’ QMC, by which we mean here path-integral or auxiliary-field QMC. We consider the computation of an intensive quantity at finite temperature. Generically the FSP leads to an exponential scaling with spatial volume and inverse temperature of the average sign, and hence, due to Eq. (\[eq:epsstat\]), of the statistical error (see, e.g., [@Loh; @CeperleyPIQMC; @AssadChapter; @troyer_topol_origin]): $$\epsilon_{\rm stat}(L) \underset{L\to\infty}{\sim} \frac{e^{\# \beta L^d}}{\sqrt{t}} \; , \label{eq:errL}$$ where $L$ is the linear system size. Beside the statistical error, we also need to take into account the systematic error $\epsilon_{\rm sys}(L)$ coming from the finite size $L$. The total error $\epsilon$ entering the CCP is $\epsilon \sim \epsilon_{\rm stat}(L)+\epsilon_{\rm sys}(L)$. We assume that finite-size corrections decrease exponentially, $$\epsilon_{\rm sys}(L) \underset{ L \to \infty}{\sim} e^{-\# L}, \label{eq:Lerr}$$ which is expected generically (away from second-order phase transitions and at finite temperature). For a given computational-time $t$, the optimal strategy is to choose $L$ so that $\epsilon_{\rm sys}\sim \epsilon_{\rm stat}$, which yields $$t(\epsilon) \sim e^{ \# \beta(\ln \epsilon^{-1})^d}. \label{eq:ccppi}$$ So for $d>1$ the scaling of $t$ with $\epsilon^{-1}$ is quasi-polynomial and there is a CCP. In one dimension there is no CCP, which is another illustration of the simple observation presented in the introduction. Apart from these asymptotic scalings, there are also practical advantages of diagrammatic methods over traditional QMC. For traditional QMC, the condition for getting close to the thermodynamic limit is typically $L$ much larger than the correlation length. Equation (\[eq:errL\]) then yields a computational time $t\propto e^{\# 2\beta L^d}$ which is often prohibitive, meaning that one cannot get close to the thermodynamic limit, and that one cannot even reach the asymptotic scaling regimes (\[eq:Lerr\],\[eq:ccppi\]). The situation is very different in diagrammatic expansions, where as shown by the above example, the asymptotic regime is accessible, and moreover the lowest orders typically set the scale while higher-order contributions are merely corrections. In conclusion, [unbiased]{} numerical methods for solving quantum many-fermion problems should be evaluated on the basis of their scaling of computational time with respect to the final error bar on thermodynamic-limit quantities. The presence of a fluctuating sign does not suffice to say that a problem is intractable by Monte Carlo. Nothing prevents in principle a polynomial scaling of the CPU-time versus the inverse error bar. We demonstrated that such polynomial complexity is indeed achieved by the recently proposed CDet method when inside the radius of convergence of the Feynman diagrammatic series. Since this method offers the possibility to calculate properties of many-fermion systems in polynomial time, it is fair to say that the sign problem has become irrelevant here and a numerical solution to the many-fermion problem is available, at least in some region of parameter space. We acknowledge support from ERC Grant [*Thermodynamix*]{} (NP and FW), the Simons Foundation’s [*Many Electron Collaboration*]{}, the National Science Foundation under the grant PHY-1314735, and the MURI Program [*New Quantum Phases of Matter*]{} from AFOSR (NP and BS). Some of us are members of [*Paris Sciences et Lettres*]{}, [*Sorbonne Universités*]{} (RR, KVH and FW) and [*Sorbonne Paris-Cité*]{} (RR and KVH). [99]{} E. Y. Loh, J. E. Gubernatis, R. T. Scalettar, S. R. White, D. J. Scalapino, R. L. Sugar, Phys. Rev. B [**41**]{}, 9301 (1990). D. M. Ceperley, [*Path integral Monte Carlo methods for fermions*]{}, in [*Monte Carlo and Molecular Dynamics of Condensed Matter Systems*]{}, Ed. K. Binder and G. Ciccotti, Editrice Compositori, Bologna, Italy (1996). D. M. Ceperley, [*Quantum Monte Carlo Methods for Fermions*]{}, Proceedings of the Les Houches Summer School, Session 56, [*Strongly Interacting Fermions and High $T_{c}$ Superconductivity*]{}, eds. B. Doucot and J. Zinn-Justin, Elsevier (1995). F.F. Assaad and H.G. Evertz, [*World-line and Determinantal Quantum Monte Carlo Methods for Spins, Phonons and Electrons*]{}, in [*Computational Many-Particle Physics*]{}, H. Fehske, R. Schneider and A. Wei[ß]{}e (Eds.), Springer, Lect. Notes Phys. [**739**]{}, 277-356 (2008). S. Chandrasekharan and U.-J. Wiese, Phys. Rev. Lett. [**83**]{}, 3116 (1999). E. Berg, M. A. Metlitski, and S. Sachdev, Science [**338**]{}, 1606 (2012). Z. C. Wei, C. Wu, Yi Li, S. Zhang, and T. Xiang, Phys. Rev. Lett. [**116**]{}, 250601 (2016). F. Alet, K. Damle, S. Pujari, Phys. Rev. Lett. [**117**]{}, 197203 (2016). A. Honecker, S. Wessel, R. Kerkdyk, T. Pruschke, F. Mila, B. Normand, Phys. Rev. B [**93**]{}, 054408 (2016) M. Iazzi, A. A. Soluyanov, and M. Troyer, Phys. Rev. B [**93**]{}, 115102 (2016). S. Capponi, J. Phys.: Condens. Matter [**29**]{}, 043002 (2017). R. Rossi, arxiv:1612.05184. K. Van Houcke, E. Kozik, N. Prokof’ev and B. Svistunov, [*Diagrammatic Monte Carlo*]{}, in “Computer Simulation Studies in Condensed Matter Physics XXI. CSP-2008” (Eds. D.P. Landau, S.P. Lewis, and H.B. Sch[ü]{}ttler), Physics Procedia [**6**]{}, 95 (2010). [G. Benfatto, A. Giuliani, and V. Mastropietro, Annales H. Poincar[é]{} [**7**]{}, 809 (2006).]{} A. Abdesselam and V. Rivasseau, Lett. Math. Phys. [**44**]{}, 77 (1998). [J. Zinn-Justin, Phys. Rep. [**70**]{}, 109 (1981).]{} M. Troyer and U.-J. Wiese, Phys. Rev. Lett. [**94**]{}, 170201 (2005). [^1]: Frustrated spin and frustrated bosonic (with restricted on-site Hilbert space) lattice models can be mapped to a system of interacting fermions and thus are part of the present discussion. [^2]: We closely follow ideas expressed by D. Ceperley at the Meeting of the [*Simons collaboration on the many-electron problem*]{}, New York, Nov. 19-20, 2015. [^3]: [ Equation (\[eq:RTD\]) is a natural conjecture given Eq. (\[eq:an\_Rn\]); its]{} rigorous proof may be obtained using techniques similar to those of Ref. [@abdes] (J. Magnen, private communication). [Equation (\[eq:RD\]) is plausible since this is the generic large-order behavior for bosonic theories [@ZinnRevue].]{} [^4]: For the density and the kinetic energy, we find the same value of $R_C$ within our error bars.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Through multi-agent competition, the simple objective of *hide-and-seek*, and standard reinforcement learning algorithms at scale, we find that agents create a self-supervised autocurriculum inducing multiple distinct rounds of emergent strategy, many of which require sophisticated tool use and coordination. We find clear evidence of six emergent phases in agent strategy in our environment, each of which creates a new pressure for the opposing team to adapt; for instance, agents learn to build multi-object shelters using moveable boxes which in turn leads to agents discovering that they can overcome obstacles using ramps. We further provide evidence that multi-agent competition may scale better with increasing environment complexity and leads to behavior that centers around far more human-relevant skills than other self-supervised reinforcement learning methods such as intrinsic motivation. Finally, we propose transfer and fine-tuning as a way to quantitatively evaluate targeted capabilities, and we compare hide-and-seek agents to both intrinsic motivation and random initialization baselines in a suite of domain-specific intelligence tests.' author: - | Bowen Baker[^1]\ OpenAI\ `[email protected]`\ Ingmar Kanitscheider\ OpenAI\ `[email protected]`\ Todor Markov\ OpenAI\ `[email protected]`\ Yi Wu\ OpenAI\ `[email protected]`\ Glenn Powell\ OpenAI\ `[email protected]`\ Bob McGrew\ OpenAI\ `[email protected]`\ Igor Mordatch[^2]\ Google Brain\ `[email protected]`\ bibliography: - 'paper.bib' title: 'Emergent Tool Use From Multi-AgentAutocurricula' --- Introduction ============ Related Work ============ Hide and Seek ============= Policy Optimization =================== Auto-Curricula and Emergent Behavior ==================================== Evaluation ========== Discussion and Future Work ========================== ### Acknowledgments {#acknowledgments .unnumbered} We thank Pieter Abbeel, Rewon Child, Jeff Clune, Harri Edwards, Jessica Hamrick, Joel Liebo, John Schulman and Peter Welinder for their insightful comments on this manuscript. We also thank Alex Ray for writing parts of our open sourced code. [^1]: This was a large project and many people made significant contributions. Bowen, Bob, and Igor conceived the project and provided guidance through all stages of the work. Bowen created the initial environment, infrastructure and models, and obtained the first results of sequential skill progression. Ingmar obtained the first results of tool use, contributed to environment variants, created domain-specific statistics, and with Bowen created the final environment. Todor created the manipulation tasks in the transfer suite, helped Yi with the RND baseline, and prepared code for open-sourcing. Yi created the navigation tasks in the transfer suite, intrinsic motivation comparisons, and contributed to environment variants. Glenn contributed to designing the final environment and created final renderings and project video. Igor provided research supervision and team leadership. [^2]: Work performed while at OpenAI
{ "pile_set_name": "ArXiv" }
--- abstract: '“Anarchy” is the hypothesis that there is no fundamental distinction among the three flavors of neutrinos. It describes the mixing angles as random variables, drawn from well defined probability distributions dictated by the group Haar measure. We perform a Kolmogorov–Smirnov (KS) statistical test to verify whether anarchy is consistent with all neutrino data, including the new result presented by KamLAND. We find a KS probability for Nature’s choice of mixing angles equal to 64%, quite consistent with the anarchical hypothesis. In turn, assuming that anarchy is indeed correct, we compute lower bounds on $|U_{e3}|^2$, the remaining unknown “angle” of the leptonic mixing matrix.' author: - 'André de Gouvêa$^{1}$ and Hitoshi Murayama$^{2,3}$' title: Statistical Test of Anarchy --- All fermions in the Standard Model of particle physics (SM) seem to come in threes. The three copies of each fundamental matter particle have in common all properties except one – the mass. It is common to say that there are three families, generations, or [*flavors*]{} of each matter particle in the SM. Currently we do not know the reason behind the number three, nor why the matter particles should “repeat” at all. Therefore, it is important to look for any information that may shed light into the origin of flavor. Within the SM, it has been known for quite some time that different quark flavors can mix quantum mechanically, and that the weak interactions can turn one flavor into another. The “amount” of mixing is summarized by the so-called Cabibbo–Kobayashi–Maskawa (CKM) unitary matrix. The CKM matrix, in turn, can be parameterized by three mixing angles $\theta_{12}, \theta_{13}, \theta_{23}$ and one complex phase $\delta$ (throughout, we use the “PDG parameterization” [@PDG] for the mixing matrices). A non-vanishing phase $\delta$ indicates that SM processes can violate CP invariance, distinguishing matter from anti-matter in a subtle manner. With the beautiful data from the $B$-factory experiments, we have been able to confirm the CKM framework, and measure all angles and the CP-odd phase with $O(10)\%$ accuracy. A noteworthy feature of the CKM matrix is that it is rather well approximated by the unit matrix, meaning that the quark mixing angles are all small. This fact, combined with the fact that the quark masses are quite distinct (the ratio of the lightest to heaviest quark mass is $O(10^{-5})$), is interpreted as evidence for the existence of some underlying symmetry or physical mechanism that differentiates the quark families and hence explains the hierarchy in the quark masses and the small mixing angles. In the SM, all neutrinos are exactly massless. This being the case, one can always choose a basis where the Maki–Nakagawa–Sakata (MNS) unitary matrix, the leptonic analog of the CKM matrix, is the unit matrix without loss of generality. This means that there are no SM processes through which one lepton flavor can turn into another. This hypothesis has been indeed confirmed by all experimental searches for charged lepton flavor violation to date [@PDG]. If neutrinos have masses, and these masses are distinct, there is no reason to expect that the MNS matrix is trivial, and lepton flavor transitions are observable in principle. In this case, the most sensitive probes for lepton flavor transitions are neutrino oscillation processes, through which a neutrino produced in a well-defined flavor state $\nu_{\alpha}$ is detected in a different flavor state $\nu_{\beta}$ after propagating over a macroscopic distance $L$. The transition probabilities depend on the mixing angles and the CP-odd phase of the MNS matrix, plus the difference of the neutrino masses-squared, $\Delta m^2_{ij}\equiv m_i^2-m_j^2$. Since 1998, there is compelling evidence that neutrino flavor transitions do occur when the neutrinos traverse macroscopic distances. Atmospheric [@atm], solar [@solar], and, very recently, reactor neutrino experiments [@KamLAND] have all observed data consistent with the neutrino oscillation hypothesis. In light of all the experimental evidence, it appears that neutrinos have masses, and that leptonic flavors mix. There are two striking features regarding the values of the oscillation parameters which are extracted from the current neutrino data. One is that the neutrino masses are extremely small. Neutrino oscillation experiments have determined that the neutrino mass-squared differences are [^1] $|\Delta m^2_{23}|=(2-7)\times10^{-3}$ eV$^2$ [@atm] and $\Delta m^2_{12}=(4-20)\times 10^{-5}$ eV$^2$ [@KamLAND]. These results, combined with direct searches for neutrino masses [@PDG], yield that the heaviest neutrino mass is less than $O(1)$ eV, over six orders of magnitude smaller than the smallest charged fermion mass of which we know (the electron mass). The other is that, of the mixing angles, two ($\theta_{12},\theta_{23}$) are known with some precision, and are both large: $\sin^22\theta_{23}\gtrsim0.9$ [@atm] and $\sin^22\theta_{12}\gtrsim0.4$ [@KamLAND]. Assuming a three family mixing scenario, there are two more parameters in the MNS mixing matrix that are still unknown: $\theta_{13}$ and $\delta$. In particular, if $\delta\neq 0$ neutrino oscillation processes need not conserve CP. Leptogenesis models [@leptogenesis], on the other hand, try to relate the existence of matter but no anti-matter in the Universe to the CP violation present in the neutrino sector, making its observation of the utmost interest. CP-violating effects parameterized by the CP-odd phase $\delta$ of the MNS matrix can be probed in accelerator-based long-baseline neutrino oscillation experiments if, for example, one compares the flavor transformation probabilities of neutrinos and anti-neutrinos (written here in vacuum), $$\begin{aligned} & & P(\nu_e \rightarrow \nu_\mu) - P(\bar{\nu}_e \rightarrow \bar{\nu}_\mu) = -16 s_{12} c_{12} s_{13} c^2_{13} s_{23} c_{23} \nonumber \\ & & \qquad \sin \delta \sin \frac{\Delta m^2_{12}L}{4E} \sin \frac{\Delta m^2_{13}L}{4E} \sin \frac{\Delta m^2_{23}L}{4E}\ ,\end{aligned}$$ where $s_{ij} = \sin \theta_{ij}$, $c_{ij} = \cos \theta_{ij}$. It is well known that the observation of CP violation in neutrino oscillations is possible only if $\theta_{12}$ and $\Delta m^2_{12}$ are “large enough” (and the atmospheric parameters are also large, as has been established by the atmospheric data). The KamLAND result has shown that this is the case. The remaining question, therefore, is whether $\theta_{13}$ is also large enough to render the experimental search for CP violation possible. The only information we currently have is that $\theta_{13}$ is relatively small: $\sin^2\theta_{13}\lesssim 0.05$, constrained by the CHOOZ experiment [@CHOOZ]. The purpose of this letter is two-fold. First, we examine if the current data “requires” new symmetry principles in order to control the structure of the MNS matrix, analogous to the situation in the quark sector. Saying that there is no symmetry principle behind the MNS matrix means there is no fundamental distinction among the three flavors of neutrinos. If this is the case, the MNS matrix is distributed (statistically) according to the bi-invariant Haar measure of group theory, which dictates the probability distribution of the mixing angles. The hypothesis here is that Nature has chosen one point according to this probability distribution. This is the concept of “anarchy” in neutrinos [@anarchy1; @anarchy2]. We would like to examine if the data are consistent with anarchy by performing a Kolmogorov–Smirnov (KS) statistical test. We find that they are perfectly consistent. Second, given the empirical success of anarchy, we study what it has to say about $\theta_{13}$. Anarchy prefers large values for $\theta_{13}$, meaning that a small $\theta_{13}$ would be inconsistent with the anarchical hypothesis. By turning this argument around, we can place a [*lower limit on*]{} $\theta_{13}$ at various confidence levels, again using the KS test. Consider the following situation: there is a model that “predicts” that a certain quantity is described by a probability distribution. For example, one may construct a model that predicts that a given quantity $x$ may have any value from 0 to 1, with equal probability. This means that the probability density $f(x)$ is [^2] $$\label{f(x)} f(x) = \left\{ \begin{array}{ll} 1 & {\rm if}~x\in [0,1], \\ 0 & {\rm otherwise.} \end{array} \right.$$ Let us assume that the value of $x$ is known: $x=x_0$. The question to be addressed is how well does the result $x=x_0$ agree with the model presented above (that the probability density for $x$ is given by Eq. (\[f(x)\]))? This question can be answered using the KS test. Given that we have drawn the specific value $x=x_0$, we would like to test the hypothesis ${\cal H}_f$ that the probability distribution associated with the random variable $x$ is $f(x)$. In order to do so we define the distribution function [^3] $F(x)\equiv\int_{-\infty}^x f(x^{\prime}){\rm d}x^{\prime}$. For Eq. (\[f(x)\]), $$\label{F(x)} F(x) = \left\{ \begin{array}{ll} 0 & {\rm if}~x\le 0, \\ x & {\rm if}~x\in [0,1], \\ 1 & {\rm if}~x\ge 1. \end{array} \right.$$ We then compare $F(x)$ with the best possible guess for a distribution function $F_{\rm guess}(x)$ that can be obtained given that $x=x_0$ has been “drawn,” namely, $$F_{\rm guess}(x)=\theta(x-x_0).$$ Note that it is very easy to generalize this to $N$ random drawings of $x$, which yield, say, $x_0,x_1,\ldots,x_{N-1}$ [@KS_ref]. The (two-sided) KS statistic (“$D$-function”) is defined by [@KS_ref] $$\label{D} D={\rm sup}_x[|F_{\rm guess}(x)-F(x)|].$$ In the example we have been discussing, $D_0=x_0$ if $x_0\ge 0.5$ or $D_0=1-x_0$ if $x_0\le 0.5$ (note that the two expressions agree at $x_0=0.5$, and we assume that $x_0\in[0,1]$). If the hypothesis ${\cal H}_f$ is correct, the probability that a larger value of $D$ ([*i.e.*]{} a “worse fit”) would be computed from a different random drawing of $x$ is [@KS_ref] $$\label{P_KS} P(D\ge D_0)=2(1-D_0),$$ which is, in the example we have been discussing, $$P(x_0) = \left\{ \begin{array}{ll} 2x_0 & {\rm if}~x_0\le 0.5, \\ 2(1-x_0) & {\rm if}~x_0\ge 0.5. \end{array} \right.$$ The smaller the value of $P(x_0)$, the less likely it is that ${\cal H}_f$ is correct. In this context, we allow statements such as ${\cal H}_f$ is only allowed at the $[1-P(x_0)]$ confidence level. We wish to apply the test described above to the MNS and CKM mixing matrices for leptons and quarks, respectively. Our model is that the mixing matrices are random variables drawn from a “flat” distribution of unitary $3\times 3$ matrices. Following the PDG convention, we define the three mixing angles as in Table \[table\]. “angle” CKM \[90% expt.\] MNS \[3$\sigma$ expt.\] --------------------- ---------------------------------------- --------------------------------------- $\sin^2\theta_{13}$ $|V_{ub}|^2~~[(6.2-23)\times10^{-6}]$ $|U_{e3}|^2~~[0-0.05]$ $\sin^2\theta_{12}$ $\sin^2\theta_C~~[0.048-0.051]$ $\sin^2\theta_{\rm sol}~~[0.2-0.5]$ $\sin^2\theta_{23}$ $|V_{cb}|^2~~[(1.4-1.9)\times10^{-3}]$ $\sin^2\theta_{\rm atm}~~[0.35-0.65]$ : $\sin^2\theta_{ij}$ in the MNS and CKM mixing matrices, according to the PDG parameterization [@PDG]. In square brackets we quote the currently allowed experimental values for the CKM (MNS) entries at the 90% (three sigma) confidence level.[]{data-label="table"} Within this convention, the hypothesis is that the marginalized probability density function is given by (see [@anarchy2] for a detailed discussion of this point) $$\int f({U(3)}){\rm d(phases)}= f(\cos^4\theta_{13},\sin^2\theta_{12},\sin^2\theta_{12})=1,$$ where we have integrated over all (both physical and unphysical) complex phases. The mixing angles are defined such that $\theta_{ij}\in[0,\pi/2],~\forall i,j$. The probability distribution is flat in $\sin^2\theta_{12}$, $\sin^2 \theta_{23}$, and $\cos^4 \theta_{13}$. It is clear that $f=1$ is correctly normalized, $$\int f~ {\rm d}(\cos^4\theta_{13}){\rm d}(\sin^2\theta_{12}) {\rm d}(\sin^2\theta_{23})=1, \label{eq:f}$$ as it should be. Since anarchy implies that the three mixing angles are distributed as [*uncorrelated*]{} random variables according to Eq. (\[eq:f\]), we are allowed to perform a [*separate*]{} KS test for each of the three mixing angles. The three distinct $D$-functions are (from Eq. (\[D\]) and the line that succeeds it), $$\begin{aligned} D_{\theta^0_{13}}&=&(1-\sin^2\theta^0_{13})^2, \\ D_{\theta^0_{12}}&=&1-\sin^2\theta^0_{12}, \\ D_{\theta^0_{23}}&=&1-\sin^2\theta^0_{23}.\end{aligned}$$ The superscript $0$ refers to the randomly picked value ([*i.e.,*]{} the physical value, “drawn” by Nature) of the corresponding mixing angle. We have assumed that $\sin^2\theta^0_{12,23}<1/2$, $\cos^4\theta^0_{13}>1/2$. The generalization for all values of $\theta_{ij}^0$ is trivial and does not add to our discussion [^4]. Because the three “random variables” are not correlated, we calculate the probability that a different random draw would yield a worse result by computing the area where the product of three one-variable probabilities $$\begin{aligned} \epsilon^0&=&P(\theta^0_{12}) \times P(\theta^0_{13}) \times P(\theta^0_{23}) \nonumber \\ &=& 8\sin^2\theta^0_{12}(2\sin^2\theta^0_{13}-\sin^4\theta^0_{13}) \sin^2\theta^0_{23}\end{aligned}$$ is worse than the data. Here, $P(\theta^0_{ij})=2(1-D_{\theta^0_{ij}})$, as in Eq. (\[P\_KS\]). Therefore $$P(KS) = \int {\rm d}(\cos^4 \theta_{13}) {\rm d}(\sin^2 \theta_{12}) {\rm d}(\sin^2 \theta_{23}) \theta(\epsilon-\epsilon^0),$$ where $\epsilon$ is given by the same expression as $\epsilon^0$, evaluated at the observed values of the mixing parameters, and $\theta(\epsilon-\epsilon^0)$ is the usual step function. We obtain $$P(KS) = \epsilon\left(1-\log\epsilon+\frac{1}{2}\log^2\epsilon\right).$$ By using the best fit values $\sin^2\theta_{12} = 0.3$ [@Fogli:2002pb] and $\sin^2\theta_{23} = 0.5$ [@atm] for the MNS matrix, we find $$\label{eq:2} \epsilon = 2.4 (\sin^2\theta_{13} - \frac{1}{2} \sin^4\theta_{13}).$$ Given the bound $\sin^2\theta_{13} \lesssim 0.05$, the anarchical hypothesis is consistent with the current data, with probability $64\%$. One can also check whether anarchy works in the quark sector. Using the values tabulated in Table \[table\], one obtains a probability smaller than $6\times 10^{-6}$, implying that the hypothesis that the CKM matrix is a random unitary $3\times3$ matrix is safely discarded. Hence, a fundamental distinction among the three flavors of quarks seems to be required. Once we have established as consistent the hypothesis that the MNS matrix is a matrix drawn from a random sample of unitary $3\times 3$ matrices, we now turn the argument around, and try to place a lower limit on $\theta_{13}$. What we require is that $P(KS) > 1-P_0$, where $P_0$ is defined to be the confidence level of the limit. Fig. \[KS\_prob\] depicts $P(KS)$ for the MNS matrix as a function of $\sin^2\theta_{13}\equiv |U_{e3}|^2$ within the three sigma bounds allowed experimentally for $\sin^2\theta_{12} \equiv\sin^2\theta_{\rm sol}$ and $\sin^2\theta_{23} \equiv\sin^2\theta_{\rm atm}$, as tabulated in Table \[table\]. For the best fit values of $\sin^2\theta_{\rm atm}$ and $\sin^2\theta_{\rm sol}$, one is able to “rule out” $|U_{e3}|^2<0.0007$ at the two sigma level and $|U_{e3}|^2<0.00002$ (which is, curiously, the upper bound for $|V_{ub}|^2$) at the three sigma level. Fig. \[KS\_prob\] also depicts $P(KS)$ as a function of $\sin^2\theta_{13}\equiv |V_{ub}|^2$ for the CKM matrix within the 90% experimentally allowed ranges defined in Table \[table\]. ![$P(KS)$ for the MNS matrix as a function of $\sin^2\theta_{13}\equiv|U_{e3}|^2$, see text for details. The top dashed curve corresponds to $\sin^2\theta_{\rm sol}=\sin^2\theta_{\rm atm}=0.5$, while the bottom dashed curve corresponds to $\sin^2\theta_{\rm sol}=0.2$ and $\sin^2\theta_{\rm atm}=0.35$. The solid curve corresponds to the best fit values $\sin^2\theta_{\rm sol}=0.3$ and $\sin^2\theta_{\rm atm}=0.5$. The hatched region is currently excluded by the neutrino data. In the bottom left corner, $P(KS)$ for the CKM matrix as a function of $\sin^2\theta_{13}\equiv|V_{ub}|^2$ is also depicted within the experimentally allowed range for $|V_{ub}|^2$, assuming that the values of $|V_{cb}|^2$ and $\sin^2\theta_C$ vary within the range indicated in Table \[table\].[]{data-label="KS_prob"}](ksprobnew2.eps){width="\columnwidth"} Note, however, that these bounds are obtained [*a posteriori*]{}, and turn out to be rather weak. This is due to the fact that the observed values of the angles $\theta_{12}$ and $\theta_{23}$ agree “too well” with the anarchical hypothesis, hence allowing a larger-than-usual fluctuation for $\theta_{13}$. We believe that $P(KS)$ is a good tool for testing the anarchical hypothesis against the data, but not as useful a tool for studying what values of $|U_{e3}|^2$ are preferred by the hypothesis. For this reason, we choose to make use of another method of obtaining lower bounds on $|U_{e3}|^2$ assuming anarchy. This method can be thought of as yielding an [*a priori*]{} prediction for $|U_{e3}|^2$, which, we believe, is more appropriate, and will be described promptly. Let us compute the marginalized probability density function of $\theta_{13}$ [*only.*]{} From Eq. (\[eq:f\]), $$\int f~{\rm d}(\sin^2\theta_{12}) {\rm d}(\sin^2\theta_{23})\equiv g(\cos^4\theta_{13})=1. \label{prob_1d}$$ This probability distribution ($g$) is rather similar to $f$, (see Eq. (\[eq:f\])) but is to be interpreted in a slightly different way. Remember that $f$ is the probability distribution function derived for $U(3)$ marginalized over all CP-odd phases, including: three unphysical phases that can be removed by redefining fermionic SM fields, two “Majorana phases” which may or may not be physical, depending on whether or not the neutrinos are Majorana fermions, and one “Dirac phase.” Marginalyzing over unphysical phases is the only reasonable procedure to follow, and we have chosen to also marginalize over the physical phase(s) in order to address whether the current information we have on mixing angles fits the anarchical hypothesis. This [*choice*]{} only makes sense if the anarchical hypothesis for the “whole” MNS matrix is consistent, which is the case as we have observed above. Similarly, $g$ is the probability distribution function derived for $U(3)$ marginalized over all CP-odd phases [ *and*]{} the “solar” and “atmospheric” mixing angles. The single variable probability is $$P(KS)_{1d}=P(\theta_{13}^0) =4(\sin^2\theta_{13} - \frac{1}{2} \sin^4\theta_{13}). \label{P(KS)_1d}$$ Fig. \[KS\_prob\_1d\] depicts $P(KS)_{1d}$ as a function of $\sin^2\theta_{13}\equiv|U_{e3}|^2$. Again, we can interpret $P(KS)_{1d}$ as the probability that a “worse fit” is obtained assuming that $\theta_{13}$ is a random variable drawn from the probability distribution Eq. (\[prob\_1d\]). These bounds are [*a priori*]{} predictions of the anarchy unlike the previous ones, and should be taken more seriously. We exclude $|U_{e3}|^2<0.011$ (0.0007) at the two (three) sigma confidence level. Note that the CKM-equivalent $P(KS)_{1d}=P(V_{ub})$ is less than $10^{-4}$ (this can be easily read off from Fig. \[KS\_prob\_1d\]), again indicating that the anarchical hypothesis in the quark sector can be safely discarded. ![$P(KS)_{1d}$ as a function of $\sin^2\theta_{13}\equiv|U_{e3}|^2$, see text for details. The hatched region is currently excluded by the neutrino data.[]{data-label="KS_prob_1d"}](bound_1d.eps){width="\columnwidth"} Finally, It is worth recalling that anarchy predicts a flat probability distribution for the CP-violating phase $\delta$ [@anarchy2], and hence the distribution in $\sin\delta$ is $1/|\cos\delta|$, peaked at $\sin\delta=\pm 1$. If anarchy is correct, chances are that the observation of CP violation in long-baseline oscillation experiments is indeed within reach! We now summarize our results, with more discussions to follow. We have statistically tested the hypothesis that the MNS matrix is a matrix drawn from a random “flat” sample of unitary $3\times 3$ matrices. According to the KS test performed, this “anarchical hypothesis” is consistent with the data. The anarchical hypothesis fails the KS test when it is performed with the CKM matrix. Our result is different from other attempts to statistically “test” anarchy. For example, the authors of [@no_anarchy] have claimed that the neutrino sector prefers the existence of some symmetry behind neutrino masses and mixing angles to completely random entries. We have not attempted to perform such a “compartive test,” which is, at least, hard to interpret in a well defined way. We do not believe that such tests are capable of indicating whether one hypothesis is favored with respect to the other. Our test has a well defined statistical interpretation, and directly probes whether anarchy in the neutrino sector is a good hypothesis. Having checked that anarchy is consistent with our current understanding of the MNS matrix, we were able to use the anarchical hypothesis to “predict” the value of the still unobserved mixing angle $\theta_{13}$. At the two sigma level, anarchy requires that $\sin^2\theta_{13}>0.011$, for example (bound obtained from $P(KS)_{1d}$, see Fig. \[KS\_prob\_1d\]). If there is indeed no structure in the leptonic mixing matrix, it seems very likely that one should be able to observe CP-violation in long-baseline neutrino oscillation experiments, as not only are all angles large, but the CP-odd parameter $\sin\delta$ is also “predicted” to be large. We have nothing to say about the value of the neutrino masses. The hypothesis we tested is that the MNS matrix is “random,” independent of whether the masses are degenerate, partially degenerate or hierarchical [@anarchy2]. Even in the case of non-LMA solutions to the solar neutrino puzzle (currently ruled out at 99.95% C.L. [@KamLAND]), one can obtain random mixing matrices [@min_model]. Incidently, it is interesting to note that the the neutrino masses seem to be “less hierarchical” than the charged fermion masses. Assuming that the neutrino masses are not degenerate, it turns out that $m_3/m_2\simeq \sqrt{\Delta m^2_{23}/\Delta m^2_{12}}= 3-13$, not too far away from unity (of course, we do not know $m_2/m_1\ldots$). This is consistent with random mass matrices generated via the seesaw mechanism [@anarchy1]. We would like to underline important assumptions and limitations of our result. By hypothesis, the probability distributions for the mixing angles are uncorrelated. Our discriminatory procedure does not include information regarding whether the different variables are more likely to be correlated than not. Given the minimal statistics (provided by the fact that we live in only one Universe), adding this sort of information would not lead to different conclusions, although one should start to worry if, say, it turns out that $\sin^22\theta_{23}=\sin^22\theta_{12}=1$. One should also be warned that the KS test performed here need not be the most powerful test for the anarchical hypothesis, statistically speaking [@KS_ref]. Finally, we emphasize what our result [*does not*]{} imply. Although the anarchical hypothesis is consistent with the data, neutrino mass models which rely on flavor symmetries and nontrivial “textures” are not disfavored in any well defined way. Some are perfectly justified by top-down arguments, including, say, grand unification of matter fields. We would like to point out, however, that the “burden of proof” is with the models that assume that there is structure in the leptonic mixing matrix. The anarchical hypothesis may be viewed as the simplest of flavor models – a model of flavor without flavor. In light of our long experience with quark masses and mixing angles, it is remarkable that, in the neutrino sector, one can do without new symmetry principles in order to appreciate the entries of the MNS mixing matrix. [**Note Added –**]{} After the first version of the this manuscript became publicly available, a preprint discussing our results [@Espinosa] appeared. All of the comments contained there apply to this version of our manuscript as well. While we appreciate most of the arguments contained in [@Espinosa], we disagree with its author in a few key points. Most importantly, we do not agree with the claim that $\sin^2\theta_{12},\sin^2\theta_{23},\cos^4\theta_{13}\in [0,1]$ are “angular variables.” Note that the Haar measure is flat in these variables rather than in the mixing angles, which are convention dependent. Therefore we stand by our claim that a KS test can be used to test the anarchical hypothesis. Furthermore, [@Espinosa] contains an alternative statistical test of the anarchical hypothesis (a Kuiper’s test, see [@Espinosa] for details), and the result that maximal mixing is “preferred” is obtained, in qualitative agreement with the results presented here. The author of [@Espinosa], however, dismisses the result of the Kuiper’s test, claiming that the statistical sample is too small. While we appreciate that some are uneasy about the statistics of very small data samples, namely one chosen by Mother Nature, we point out that the dismissal of such results is not mathematically justified. We thank Paolo Creminelli for very useful questions and discussions, and José Ramon Espinosa for corresponce regarding [@Espinosa]. The work of AdG is supported by the US Department of Energy Contract DE-AC02-76CHO3000. The work of HM is supported in part by the Director, Office of Science, Office of High Energy and Nuclear Physics, Division of High Energy Physics of the U.S. Department of Energy under Contract DE-AC03-76SF00098 and in part by the National Science Foundation under grant PHY-00-98840. [99]{} Particle Data Group (K. Hagiwara [*et al.*]{}), Phys Rev. [ **D 66**]{}, 010001 (2002). Y. Fukuda [*et al.*]{} \[Super-Kamiokande Collaboration\], Phys. Rev. Lett.  [**81**]{}, 1562 (1998). Q. R. Ahmad [*et al.*]{} \[SNO Collaboration\], Phys. Rev. Lett. [**89**]{}, 011301 (2002); 011302 (2002). K. Eguchi [*et al.*]{} \[KamLAND Collaboration\], hep-ex/0212021. M. Fukugita and T. Yanagida, Phys. Lett. [**B174**]{}, 45 (1986). M. Apollonio [*et al.*]{} \[CHOOZ Collaboration\], Phys. Lett. [**B466**]{}, 415 (1999). L.J. Hall, H. Murayama, and N. Weiner, Phys. Rev. Lett. [**84**]{}, 2572 (2000). N. Haba and H. Murayama, Phys. Rev. [**D 63**]{}, 053010 (2001). See, [*e.g.*]{}, V.K. Rohatgi and A.K. Md. Ehsanes Saleh, “An Introduction to Probability and Statistics,” New York, 1976. See, [*e.g.*]{}, G. L. Fogli, G. Lettera, E. Lisi, A. Marrone, A. Palazzo and A. Rotunno, Phys. Rev. D [**66**]{}, 093008 (2002) \[arXiv:hep-ph/0208026\]. V. Antonelli, F. Caravaglios, R. Ferrari, and M. Picariello, Phys. Lett. [**B549**]{}, 325 (2002); G. Altarelli, F. Feruglio, and I. Masina, JHEP [**0301**]{}, 035 (2003). See, for example, A. de Gouvêa and J.W. Valle, Phys. Lett. [**B501**]{}, 115 (2001). J. R. Espinosa, hep-ph/0306019. [^1]: We define the neutrino mass eigenvalues such that $m_2^2>m_1^2$, and $\Delta m^2_{12}<|\Delta m^2_{13,23}|$. [^2]: The probability density function is defined in such a way the probability that $x$ has a value between $x_0$ and $x_0+{\rm d}x$ is given by $f(x_0){\rm d}x$. [^3]: The distribution function is defined in such a way the the probability that $x\in [a,b]$ is $F(b)-F(a)$. [^4]: Of course, all angles in the quark sector satisfy $\sin^2\theta_{ij}\ll 1/2$. The same is true of $|U_{e3}|^2$, while the preferred values of $\sin^2\theta_{\rm sol}$ are smaller than 1/2 at the three sigma level. We don’t know whether $\sin^2\theta_{\rm atm}$ is less than or greater than 1/2. The experimental information we do have is such, however, that $\theta_{\rm atm}$ and $\pi/2-\theta_{\rm atm}$ cannot be discriminated. These “degenerate” solutions lead to the same $D_{\theta^0_{23}}$.
{ "pile_set_name": "ArXiv" }
--- abstract: '**Self-assembled, epitaxially-grown InAs/GaAs quantum dots are promising semiconductor quantum emitters that can be integrated on a chip for a variety of photonic quantum information science applications. However, self-assembled growth results in an essentially random in-plane spatial distribution of quantum dots, presenting a challenge in creating devices that exploit the strong interaction of single quantum dots with highly confined optical modes. Here, we present a photoluminescence imaging approach for locating single quantum dots with respect to alignment features with an average position uncertainty $<30$ nm ($<10$ nm when using a solid immersion lens), which represents an enabling technology for the creation of optimized single quantum dot devices. To that end, we create quantum dot single-photon sources, based on a circular Bragg grating geometry, that simultaneously exhibit high collection efficiency (48 $\%~\pm$ 5 $\%$ into a 0.4 numerical aperture lens, close to the theoretically predicted value of 50 $\%$), low multiphoton probability ($g^{(2)}(0)<$1 $\%$), and a significant Purcell enhancement factor ($\approx$ 3)**.' author: - Luca Sapienza - Marcelo Davanço - Antonio Badolato - Kartik Srinivasan title: 'Nanoscale optical positioning of single quantum dots for bright and pure single-photon emission' --- Single InAs/GaAs quantum dots are one of the more promising solid-state quantum emitters for applications such as quantum light generation and single-photon level nonlinear optics [@ref:Michler_book_2009]. Critical to many such applications is the incorporation of the quantum dot within an engineered photonic environment so that the quantum dot interacts with only specific optical modes. A variety of geometries such as photonic crystal devices and whispering gallery mode resonators have been employed to achieve such behavior for bright single-photon sources and strongly coupled quantum dot-cavity systems [@ref:Santori_Book]. The optical field in many such geometries varies significantly over distances of $\approx$ 100 nm, setting a scale for how accurately the quantum dot position should be controlled within the device for optimal interaction. While site-controlled growth of quantum dots presents one attractive option [@ref:Juska_site_controlled_QD_entangled], the properties of such quantum dots (in terms of homogeneous linewidth, for example) have not yet matched those of quantum dots grown by strain-mediated self-assembly (Stranski-Krastanow growth) [@ref:Huggenberger_Forchel_site_controlled]. However, the in-plane location, polarization, and emission wavelength of such self-assembled quantum dots are not accurately controlled in a deterministic fashion, and thus techniques are required to determine these properties prior to device fabrication, in order to create optimally performing systems. Several techniques for location of self-assembled InAs/GaAs quantum dots prior to device fabrication have been reported, including atomic force microscopy (AFM) [@ref:Hennessy3], scanning confocal photoluminescence microscopy [@ref:Thon_APL_09] (including in-situ, cryogenic photolithography [@ref:Lee_Taylor_cryo_litho; @ref:Dousse_Senellart_QD_in_situ_litho]), photoluminescence imaging [@ref:Kojima_Noda_positioning], and scanning cathodoluminescence [@ref:Gschrey_Reitzenstein_CL_positioning]. Of these approaches, photoluminescence imaging is particularly attractive given its potential to combine high throughput sub-50 nm positioning accuracy, spectral information, and compatibility with high-resolution electron-beam lithography that is typically used to pattern small features such as those used in photonic crystals. Localization of single molecules to 10 nm scale accuracy by imaging their fluorescence onto a sensitive camera has proven to be a powerful technique in the biological sciences [@ref:Thompson_Webb_localization]. Here, we present a two-color photoluminescence imaging technique to determine the position of single quantum dots with respect to fiducial alignment marks, with an average position uncertainty $<30$ nm obtained for an image acquisition time of 120 s (the average position uncertainty is reduced to $<10$ nm when using a solid immersion lens). This wide-field technique is combined with confocal measurements within the same experimental setup to determine emission wavelength and polarization. We use this information to fabricate and demonstrate quantum dot single-photon sources in a circular Bragg grating geometry that simultaneously exhibit high collection efficiency (48 $\%~\pm$ 5 $\%$ into a lens with numerical aperture of 0.4), low multiphoton probability at this collection efficiency ($g^{(2)}(0)<$1 $\%$), and a significant Purcell enhancement factor ($\approx$ 3). Our results constitute an important step forward for both the general creation of nanophotonic devices using positioned quantum dots, and the specific performance of quantum dot single-photon sources. **** **Quantum dot location via photoluminescence imaging** ![image](Figure1_submit.pdf){width="0.75\linewidth"} Prior to sample interrogation, an array of metal alignment marks is fabricated on quantum-dot-containing material through a standard lift-off process (see Methods). The samples are then placed on a stack of piezo-electric stages to allow motion along three orthogonal axes (x,y,z) within a closed-cycle cryostat that reaches temperatures as low as 6 K. The simplest photoluminescence imaging configuration we use is a subset of Fig. \[fig:Fig1\](a), and starts with excitation by a 630 nm LED, which is sent through a 90/10 (reflection/transmission percentage) beamsplitter and through a 20x infinity-corrected objective (0.4 numerical aperture) to produce an $\approx$ 200 $\mu$m diameter spot on the sample. Reflected light and fluorescence from the sample goes back through the 90/10 beamsplitter and is imaged onto an Electron Multiplied Charged Couple Device (EMCCD) using a variable zoom system. When imaging the fluorescence from the quantum dots, the 630 nm LED power is set to its maximum power ($\approx$ 40 mW, corresponding to an intensity of $\approx$ 130 W/cm$^2$), and a 900 nm long-pass filter (LPF) is inserted in front of the EMCCD camera to remove reflected 630 nm light. Imaging of the alignment marks is done by reducing the LED power to 0.8 mW, turning off the EMCCD gain, and removing the 900 nm LPF. ![image](Figure2_submit.pdf){width="0.7\linewidth"} Representative images of the quantum dot photoluminescence and alignment marks are shown in Fig. \[fig:Fig1\](b) and Fig. \[fig:Fig1\](d). In Fig. \[fig:Fig1\](b), circular bright spots surrounded by Airy rings - a signature of optimally focused collection - are clearly visible and represent the emission from single quantum dots excited within an $\approx$ 56 $\mu$m x 56 $\mu$m field of view. Orthogonal linescans of the bright spots (Fig. \[fig:Fig1\](c)) are fit with Gaussian functions using a nonlinear least squares approach (see Supplementary Note 1), with the extracted peak positions showing one standard deviation uncertainties as low as $\approx$ 9 nm. A similar analysis of orthogonal linescans of the alignment marks (Fig. \[fig:Fig1\](d)-(e)) shows their center positions to be known with an uncertainty that is typically $\approx$ 15 nm. Figure \[fig:Fig1\](f) shows how this uncertainty changes as a function of system magnification (and hence field of view), which is adjusted using the variable zoom system. We see that the quantum dot uncertainty values show a decreasing trend with higher magnification, and values as low as $\approx$ 5 nm are measured. This can be understood because the increased magnification spreads the quantum dot emission over a larger number of pixels on the EMCCD camera, resulting in a smaller fit uncertainty, provided that the collected fluorescence level produces an adequate per pixel signal-to-noise level. On the other hand, the uncertainty in the alignment mark center position shows no obvious trend with changing magnification. Ultimately, we have found that the alignment mark uncertainties are limited by the blur induced by the two intermediate fused silica cryostat windows (vacuum and radiation shield, 2 mm and 1 mm thick, respectively) between the objective and sample, which has been confirmed by measurements in ambient conditions with the windows removed. While the 630 nm LED can thus be used for imaging both the quantum dots and alignment marks, it requires the acquisition of two separate images, with insertion of a filter needed when collecting the quantum dot photoluminescence. As filter insertion can result in beam shifts that will be manifested as an uncontrolled error in determination of the separation between quantum dot and alignment mark, we implement a modified setup (Fig. \[fig:Fig1\](a)) in which a second, infrared LED at 940 nm is combined with the 630 nm LED when illuminating the sample. Unlike the 630 nm LED, the 940 nm LED does not excite the quantum dots, but instead serves only to illuminate the alignment marks, with the wavelength chosen to approximately match the expected wavelength of the quantum dot emission. By adjusting the 940 nm LED power appropriately, both the quantum dots and alignment marks can be observed in a single image with the 900 nm LPF in place. Figure \[fig:Fig2\](a) shows an image taken when the sample is co-illuminated by both 630 nm and 940 nm LEDs, with the 940 nm power set to be $\approx$ 4 $\mu$W, about four orders of magnitude smaller than that of the 630 nm LED power. Orthogonal line scans through the quantum dot and alignment marks under this co-illumination scheme are shown in Fig. \[fig:Fig2\](b). As expected, the uncertainty values determined for quantum dot and alignment mark positions are larger than those obtained when acquiring two separate images (Fig. \[fig:Fig1\](c),(e)), for which the LED power can be optimized independently to maximize the image contrast and minimize each uncertainty. However, we have favoured the co-illumination approach due to its ability to reduce some potential uncertainties, like sample drift, that may occur during schemes requiring multiple images to be acquired. Ultimately, one might envision time-multiplexing and drift compensation techniques being employed to correct for such factors. After carrying out a systematic study of the position uncertainties as a function of magnification, integration time, and EMCCD gain, we have found optimized settings for image acquisition (under 40x magnification), in terms of the combined quantum dot and alignment mark uncertainty: an integration time of 120 s, an EMCCD gain of 200, and the aforementioned LED powers. Under these conditions, we have studied the uncertainties in the quantum dot position, alignment mark position, and quantum dot-alignment mark separation for a number of different quantum dots on our sample. Histograms of the measured values are reported in Fig. \[fig:Fig2\](c), and show that the mean uncertainty in the quantum dot-alignment mark separation is $\approx$ 28 nm. Finally, we note that in the present setup, the available 630 nm LED power is below that required to saturate the quantum dot emission (a comparison with the saturation counts obtained under laser excitation shows that it is about half the value required). Higher 630 nm LED power would increase the collected photoluminescence and reduce the uncertainty values that we have reported. This pre-eminent role of collected photon flux is well-established in the single emitter localization literature [@ref:Thompson_Webb_localization]. We have confirmed it in our experiments by using a solid-immersion lens [@ref:Zwiller_Bjork_JAP; @ref:Serrels_SIL], which can both increase the LED intensity at the quantum dot and the fraction of quantum dot emission that is collected by the microscope objective. Placing a hemispherical lens with refractive index $n$ = 2 on the surface of the sample yields individual quantum dot and alignment mark position uncertainties of $\approx$ 5 nm (Fig. \[fig:Fig2\](d)-(e)), so that the overall uncertainty in locating the quantum dot with respect to the alignment mark is $<10$ nm (more details provided in Supplementary Note 2). In total, we note that the positioning uncertainties that we obtain are 2$\times$ to 5$\times$ smaller than previously reported [@ref:Dousse_Senellart_QD_in_situ_litho; @ref:Thon_APL_09; @ref:Kojima_Noda_positioning], and are obtained with a single image, acquired over a 120 s acquisition time, and spanning an area of the sample greater than 100 $\mu$m $\times$ 100 $\mu$m. **Realization of circular Bragg grating bullseye cavities** ![**Circular dielectric gratings tailored to specific quantum dot emitters.** (a) Normalized cavity mode electric field intensity $|E|^2$ superimposed on a scanning electron microscope image of the center of one of the cavities. Scale bar represents 200 nm. (b) Experimental central wavelength of 50 circular grating cavities with varying period and central radius, plotted as a function of the simulated central wavelength. When only one peak is observed in the spectrum, black squares are used to denote the peak wavelength. When two peaks are observed, red circles and blue triangles are used. Such two-peak behavior is also seen in simulations depending on the device parameters, and is due to coupling to a second cavity mode. Top inset: Atomic Force Microscope image of a circular grating cavity and a linecut (along the dashed line) showing the etch depth of the trenches. Bottom inset: Examples of photoluminescence spectra of circular grating cavities, measured from a high-quantum dot density region.[]{data-label="fig:Fig3"}](Figure3_submit.pdf){width="\linewidth"} We now use the optical positioning technique to fabricate nanophotonic structures tailored for the properties of a specific quantum dot and engineered to enhance the collection efficiency of single photons in free space. First, we obtain information about the quantum dot emission wavelength by spatially selecting one quantum dot and collecting its emission into a single-mode fiber that is coupled into a grating spectrometer (a half waveplate and polarizing beamsplitter are used to switch the collection path between the EMCCD camera and single-mode fiber). Spatial selection is achieved by exciting individual quantum dots with a 780 nm laser, incorporated into the same micro-photoluminescence setup (Fig. \[fig:Fig1\](a)), and producing a focused spot size of $\approx$ 2 $\mu$m on the sample surface. The half waveplate and polarizing beamsplitter also enable determination of the quantum dot polarization. Having thus obtained emission wavelength to go along with the spatial position obtained from the imaging setup, a properly calibrated fabrication process can enable the creation of nanophotonic structures that are tailored to the specific emitter properties. This allows one to minimise (and potentially avoid altogether) the need for mutual spectral tuning of the emitter with respect to the optical resonance of the cavity, which is a clear limitation of the scalability of these sources. The specific nanophotonic structure we focus on is a circular Bragg grating ‘bullseye’ geometry, which has been developed as a planar structure in which quantum dot photons are funneled into a near-Gaussian far-field pattern over a moderate spectral bandwidth (few nm) with high efficiency (theoretical efficiency of 50 $\%$ into a 0.4 numerical aperture) and with the potential for Purcell enhancement of the radiative rate [@ref:Davanco_BE; @ref:Ates_JSTQE]. The cavity mode of interest is tightly confined, and optimal performance requires the quantum dot to be within a couple hundred nanometres of the centre of the bullseye structure. This is illustrated in Fig. \[fig:Fig3\](a), which plots the normalized electric field intensity superimposed on a scanning electron microscope image of the center of a fabricated device. An important parameter in the fabrication of these devices is the etch depth of the asymmetric grating, as this determines the fraction of emission in the upwards direction (towards our collection optics) compared to the downwards direction (towards the substrate). Furthermore, given the high refractive index difference between GaAs and air, a change in etch depth of 1 nm results in a shift of the optical resonances of about 1 nm. We use AFM to determine the GaAs dry etch rate within the grating grooves (Fig. \[fig:Fig3\](b), top inset), and based on this calibration, we fabricate (see Methods) 50 circular gratings whose parameters (pitch and central diameter) have been adjusted so that the cavity resonances cover the 930 nm to 1000 nm range of wavelengths. These samples were fabricated in a region of the wafer with a high density of quantum dots, so that the resulting emission under high power excitation is broad enough to feed the cavity modes. Example spectra collected from different circular grating cavities are shown in the bottom inset of Fig. \[fig:Fig3\](b). These measurements allow us to calibrate the experimental cavity resonances with respect to simulations, as shown in the main panel of Fig. \[fig:Fig3\](b), and tailor the design to match the specific quantum dot emission wavelength. **Optimized quantum dot single-photon source** Using the quantum dot positions with respect to alignment marks as determined by photoluminescence imaging, emission wavelengths as determined by grating spectrometer measurements, and the aforementioned calibration of the circular grating geometry to match target wavelengths, we fabricate (see Methods) a series of circular grating cavities containing single quantum dots. Photoluminescence imaging of the devices after fabrication, as shown in Fig. \[fig:Fig4\](a) for a representative device excited by the 630 nm LED, qualitatively indicates that the quantum dot emission originates from the centre of the bullseye structure, as intended. A measurement of the far-field emission from the device on the EMCCD, as shown in Fig. \[fig:Fig4\](b), shows that it is close to a circular Gaussian function, as confirmed by a nonlinear least squares fit. As the overlap with a perfect circular Gaussian is $\approx$ 70 $\%$, this far-field patten is expected to mode match well to a single-mode fiber, an important consideration for long-distance transmission of single photons for quantum information applications. ![image](Figure4_submit.pdf){width="0.8\linewidth"} We now characterise the emission produced by the optically positioned quantum dots within the circular grating cavities, in terms of collection efficiency, single-photon purity, and spontaneous emission rate. For these measurements, a second cryostat and photoluminescence setup was used, as it provides direct free-space in-coupling to a grating spectrometer that is also used for spectral isolation of the quantum dot excitonic state (Supplementary Fig. 1). First, we determine the collection efficiency by pumping the devices with a 780 nm wavelength, 50 MHz repetition rate pulsed laser (50 ps pulse width), and varying the laser power until the emission from the quantum dot saturates (Fig. \[fig:Fig4\](c)). Assuming a quantum dot radiative efficiency of unity, and taking into account the losses within the optical setup (see Supplementary Note 3), we measure a collection efficiency as high as 48.5 $\%$ $\pm$ 5.0 $\%$ into a 0.4 numerical aperture objective, where the uncertainty is due to fluctuations in power measurements done to calibrate losses in the optical setup, and represents a one standard deviation value. This collection efficiency is close to the theoretical value of 50 $\%$ expected for a centrally located quantum dot, and is more than two orders of magnitude larger than the collection efficiency for a quantum dot in unpatterned GaAs, as shown in Fig. \[fig:Fig4\](c). We note that a 80 $\%$ collection efficiency is theoretically expected if a higher numerical aperture optic (e.g., NA = 0.7) is used. In previous studies of quantum dots in circular grating cavities [@ref:Davanco_BE; @ref:Ates_JSTQE], where no optical positioning was used, device fabrication in a material containing a higher density of quantum dots was performed, to ensure that some non-negligible fraction (which turned out to be a few percent) of devices would have a quantum dot spectrally and spatially overlapped with the desired cavity mode (see Supplementary Note 4). In comparison, the optical positioning used here allows us to work with a much lower density of quantum dots ($\lesssim$ 1 per 1000 $\mu$m$^2$). One consequence of this is the comparatively clean emission spectra we observe, even when exciting with pump powers that completely saturate the quantum dot emission (Fig. \[fig:Fig4\](d)). Such clean spectra might be expected to correspond to clean (low multi-photon probability) single-photon emission, and to test this, the spectrally filtered emission from the bright quantum dot exciton line is measured in a standard Hanbury-Brown and Twiss setup. Under non-resonant, 780 nm pulsed excitation, we measure $g^{(2)}(0) = 0.15 \pm 0.03$ when the quantum dot emission is saturated (Supplementary Fig. 2(a)). When the system is excited above quantum dot saturation non-resonantly, we observe emission from the bullseye cavity modes superimposed with the quantum dot emission (data in Supplementary Fig. 2(b), collected under continuous wave 780 nm excitation). Together, this suggests that quasi-continuum states, originating from the combined single quantum dot - wetting layer system, feed the optical cavity mode [@ref:Winger2009; @ref:Chauvin_Fiore; @ref:Laucht_Finely_2010] and limit the device’s single-photon purity. We next consider pumping the device on an excited state of the quantum dot, as such excitation (sometimes referred to as quasi-resonant or p-shell pumping) has been shown to reduce $g^{(2)}(0)$ [@ref:Santori_NJP]. Measurement of the quantum dot emission under pulsed 857 nm excitation shows that, at the saturation pump intensity (where the collection efficiency is maximized), the spectrum is nearly identical to that under 780 nm excitation (Supplementary Fig. 2(d)). Moreover, increased excitation power above saturation (achieved using a 857 nm continuous wave laser) yields far less cavity mode feeding than in the corresponding 780 nm case (Supplementary Fig. 2(c)), suggesting that improved single-photon purity should be observed. This is confirmed by intensity autocorrelation measurements, which indicate that on-demand single-photon emission with a purity of 99.1$\%$ ($g^{(2)}(0) = 0.009 \pm 0.005$) is achieved at quantum dot saturation. We note that the $g^{(2)}(0)$ levels are determined from raw coincidences, without any background subtraction, and with an uncertainty value given by the standard deviation in the area of the peaks away from time zero. We also measure the spontaneous decay rate of the quantum dot emission under 780 nm pulsed excitation (measurements at 857 nm have also been performed and yield unchanged results). The spontaneous emission decay of a quantum dot in bulk and a quantum dot in a circular grating cavity are shown in Fig. \[fig:Fig4\](f). The exponential fit of the decay curve allows us to extract a lifetime of $\approx$ 520 ps for the quantum dot in the bullseye cavity, corresponding to a Purcell enhancement of the spontaneous emission rate by a factor of $\approx$ 3. A Purcell factor as high as 4 is measured in other devices that have a smaller detuning with respect to the cavity mode (the detuning is 1.6 nm for the device we focus on here). Theoretically, Purcell factors as high as $\approx$ 11 are expected [@ref:Davanco_BE; @ref:Ates_JSTQE] for quantum dots with perfect spectral and spatial alignment with respect to the cavity mode. Different methods to achieve such precise spectral resonance are currently under consideration; preliminary measurements indicate that in-situ $N_2$ deposition is ill-suited to the circular grating geometry, as the cavity mode degrades before a significant wavelength shift is observed. Going forward, it would be relevant to determine the location of the optically positioned quantum dots within fabricated devices, in order to understand sources of error within our overall fabrication approach (which combines optical positioning with aligned electron-beam lithography). Supplementary Note 5 presents a detailed discussion on the results of finite-difference time-domain simulations examining the Purcell factor, collection efficiency, and degree of polarization in the collected far-field as a function of dipole position and orientation within the cavity. Our calculations indicate that the Purcell enhancement, in particular, very sensitively depends on the dipole location, while the collection efficiency is not as sensitive. For the devices we have focused on in the main text, we find that a simulated offset between 50 nm and 250 nm with respect to the cavity center produces results that are consistent with our measurements. **** There has been much progress in the development of bright quantum dot single-photon sources in recent years, including micropillar [@ref:Strauf_NPhot; @ref:Gazzano_Senellart_QD_SPS], vertical nanowire waveguide [@ref:Claudon; @ref:Reimer_Zwiller_nanowire_SPS; @ref:Kremer_Gerardot_QD_nanowire_SPS], fiber-coupled microdisk [@ref:Ates_Srinivasan_Sci_Rep], and photonic crystal cavity [@ref:Madsen_Lodahl_QD_PC_SPS] geometries. Many metrics are needed to characterize these sources, and the choice of which ones are of particular importance is largely determined by the intended application. Within the landscape of these sources, the results presented here are unique in terms of simultaneously exhibiting high collection efficiency, nearly perfect single-photon purity at the highest measured collection efficiency, and Purcell enhancement of the spontaneous emission rate. For example, previous bright, Purcell-enhanced microcavity single-photon sources have shown significant non-zero $g^{(2)}(0)$ values (e.g $\gtrsim0.1$) at their highest collection efficiencies [@ref:Strauf_NPhot; @ref:Gazzano_Senellart_QD_SPS; @ref:Ates_Srinivasan_Sci_Rep; @ref:Madsen_Lodahl_QD_PC_SPS], while bright nanowire sources show $g^{(2)}(0)\approx0$ but do not exhibit Purcell enhancement [@ref:Claudon; @ref:Reimer_Zwiller_nanowire_SPS; @ref:Kremer_Gerardot_QD_nanowire_SPS]. For some applications, the metrics demonstrated thus far should be combined with a high degree of photon indistinguishability [@ref:Gazzano_Senellart_QD_SPS], which is limited in our work by the coherence time of the quantum dots in this sample ($<300$ ps, as confirmed by measurements with a scanning Fabry Perot interferometer; other emitters on the same wafer show coherence times as long as 500 ps). Future work will focus on resonant excitation [@ref:Muller; @ref:Ates_PRL09; @ref:He_indistinguishable_SPS] to improve the coherence time and fine control of the cavity-quantum dot detuning to achieve shorter radiative lifetimes [@ref:Santori2; @ref:Varoutsis_PRB05]. Together, these advances may provide a route to a source that simultaneously provides bright, pure, and indistinguishable single-photons. In conclusion, we have developed a photoluminescence imaging technique that enables the location of single quantum dots with respect to alignment markers with an average position uncertainty $<$ 30 nm and reaching values as low as $<10$ nm. We have combined this technique with systematic calibration of our fabrication process to create single-photon sources based on a circular Bragg grating geometry that simultaneously exhibit high brightness, purity, and Purcell enhancement of the spontaneous emission rate. More generally, this technique is an important step forward in the ability to create functional single quantum dot nanodevices, including quantum light sources, strongly-coupled quantum dot - microcavity systems for achieving single photon nonlinearities [@ref:Fushman_Vuckovic_Science_phase_shift; @ref:Reinhard_Imamoglu_photon_blocakde; @ref:Kim_Waks_QD_logic_gate] coupled quantum dot - nanomechanical structures [@ref:Wilson_Rae_Imamoglu_QD_mechanics_theory; @ref:Metcalfe_Lawall_QD_SAW; @ref:Yeo_strain_QD], and integrated systems involving multiple quantum dot nodes. Devices are fabricated in a wafer grown by molecular beam epitaxy, consisting of a single layer of InAs quantum dots (QDs) embedded in a 190nm thick layer of GaAs, which in turn is grown on top of a 1$\mu$m thick layer of Al$_x$Ga$_{1-x}$As with an average $x$=0.65. The s-shell peak of the QD ensemble is located near 940 nm, and a gradient in the QD density is grown along one axis of the wafer. Low-temperature photoluminescence imaging of portions of the wafer is performed prior to any device definition to determine the appropriate location on the wafer (in terms of QD density) to fabricate devices. Alignment marks are fabricated using positive tone electron-beam lithography and a lift-off process. Polymethyl methacrylate (PMMA) with a molecular weight of 495,000 is spin coated onto the sample, and 2 $\mu$m wide, 50 $\mu$m long crosses are patterned in the resist using a 100 keV electron-beam lithography tool. After exposure, the resist is developed in a 1:3 (by volume) solution of methyl isobutyl ketone (MIBK) and isopropanol, and 20 nm of Cr and 100 nm of Au are deposited on the sample using an electron-beam evaporator. Microposit remover 1165 is used for lift-off, with gentle ultra-sonication applied if necessary. After location of quantum dots with respect to the alignment marks through photoluminescence imaging, circular Bragg grating ‘bullseye’ microcavities are fabricated as follows. First, the sample is spin-coated with a positive tone electron-beam resist (ZEP 520A), and aligned electron-beam lithography with a 100 keV tool and four mark detection is performed. Next, the pattern is transferred into the GaAs layer using an Ar-Cl$_2$ inductively-coupled plasma reactive ion etch. After removal of the electron beam resist, the sample is undercut in hydrofluoric acid. Atomic force microscopy (AFM) was used in the calibration of the etch rate, with the samples scanned in tapping mode using a commercial, etched silicon probe whose backside is coated with Al. The AFM probe cantilever has a vendor-specified spring constant of 42 N/m, frequency of 300 kHz, and probe tip radius and height of 8 nm and 10 $\mu$m, respectively. [10]{} url \#1[[\#1]{}]{}urlprefix P. Michler, [*[Single Semiconductor Quantum Dots]{}*]{} (Springer Verlag, Berlin, 2009). C. Santori, D. Fattal, and Y. Yamamoto, [*[Single-photon Devices and Applications]{}*]{} (Wiley-VCH, Leipzig, 2010). G. Juska, V. Dimastrodonato, L. O. Mereni, A. Gocalinska, and E. Pelucchi, “Towards quantum-dot arrays of entangled photon emitters,” Nature Photonics [**7**]{}, 527–531 (2013). A. [Huggenberger]{}, S. [Heckelmann]{}, C. [Schneider]{}, S. [H[ö]{}fling]{}, S. [Reitzenstein]{}, L. [Worschech]{}, M. [Kamp]{}, and A. [Forchel]{}, “[Narrow spectral linewidth from single site-controlled In(Ga)As quantum dots with high uniformity]{},” Appl. Phys. Lett. [**98**]{}, 131104 (2011). K. Hennessy, A. Badolato, M. Winger, D. Gerace, M. Atature, S. Guide, S. Falt, E. Hu, and A. Imamoglu, “[Quantum nature of a strongly coupled single quantum dot-cavity system]{},” Nature (London) [**445**]{}, 896–899 (2007). S. M. Thon, M. T. Rakher, H. Kim, J. Gudat, W. T. M. Irvine, P. M. Petroff, and D. Bouwmeester, “Strong coupling through optical positioning of a quantum dot in a photonic crystal cavity,” Appl. Phys. Lett. [**94**]{}, 111115 (2009). K. H. [Lee]{}, A. M. [Green]{}, R. A. [Taylor]{}, D. N. [Sharp]{}, J. [Scrimgeour]{}, O. M. [Roche]{}, J. H. [Na]{}, A. F. [Jarjour]{}, A. J. [Turberfield]{}, F. S. F. [Brossard]{}, D. A. [Williams]{}, and G. A. D. [Briggs]{}, “[Registration of single quantum dots using cryogenic laser photolithography]{},” Appl. Phys. Lett. [**88**]{}, 193106 (2006). A. [Dousse]{}, L. [Lanco]{}, J. [Suffczy[ń]{}ski]{}, E. [Semenova]{}, A. [Miard]{}, A. [Lema[î]{}tre]{}, I. [Sagnes]{}, C. [Roblin]{}, J. [Bloch]{}, and P. [Senellart]{}, “[Controlled light-matter coupling for a single quantum dot embedded in a pillar microcavity using far-field optical lithography]{},” Phys. Rev. Lett. [**101**]{}, 267404 (2008). T. Kojima, K. Kojima, T. Asano, and S. Noda, “[Accurate alignment of a photonic crystal nanocavity with an embedded quantum dot based on optical microscopic photoluminescence imaging]{},” Appl. Phys. Lett. [**[102]{}**]{}, 011110 ([2013]{}). M. Gschrey, F. Gericke, A. Schuessler, R. Schmidt, J. H. Schulze, T. Heindel, S. Rodt, A. Strittmatter, and S. Reitzenstein, “[In situ electron-beam lithography of deterministic single-quantum-dot mesa-structures using low-temperature cathodoluminescence spectroscopy]{},” Appl. Phys. Lett. [**[102]{}**]{}, 251113 ([2013]{}). R. Thompson, D. Larson, and W. Webb, “[Precise nanometer localization analysis for individual fluorescent probes]{},” Biophysical Journal [**82**]{}, 2775–2783 (2002). V. [Zwiller]{} and G. [Bjork]{}, “[Improved light extraction from emitters in high refractive index materials using solid immersion lenses]{},” J. Appl. Phys. [**92**]{}, 660–665 (2002). K. A. Serrels, E. Ramsay, P. A. Dalgarno, B. Gerardot, J. O’Connor, R. H. Hadfield, R. Warburton, and D. Reid, “Solid immersion lens applications for nanophotonic devices,” Journal of Nanophotonics [**2**]{}, 021854 (2008). M. [Davanço]{}, M. T. Rakher, D. Schuh, A. Badolato, and K. Srinivasan, “[A circular dielectric grating for vertical extraction of single quantum dot emission]{},” Appl. Phys. Lett. [**99**]{}, 041102 (2011). S. Ates, L. Sapienza, M. Davanco, A. Badolato, and K. Srinivasan, “Bright single-photon emission from a quantum dot in a circular Bragg grating microcavity,” IEEE Journal of Selected Topics in Quantum Electronics [**18**]{}, 1711–1721 (2012). M. Winger, T. Volz, G. Tarel, S. Portolan, A. Badolato, K. Hennessy, E. L. Hu, A. Beveratos, J. Finley, V. Savona, and A. Imamoglu, “[Explanation of photon correlations in the far-off-resonance optical emission from a quantum - dot – cavity system]{},” Phys. Rev. Lett. [**103**]{}, 207403 (2009). N. Chauvin, C. Zinoni, M. Francardi, A. Gerardino, L. Balet, B. Alloing, L. H. Li, and A. Fiore, “Controlling the charge environment of single quantum dots in a photonic-crystal cavity,” Phys. Rev. B [**80**]{}, 241306 (2009). A. Laucht, M. Kaniber, A. Mohtashami, N. Hauke, M. Michler, and J. Finley, “Temporal monitoring of nonresonant feeding of semiconductor nanocavity modes by quantum dot multiexciton transitions,” Phys. Rev. B [ **81**]{}, 241302 (2010). C. [Santori]{}, D. [Fattal]{}, J. [Vuckovic]{}, G. S. [Solomon]{}, and Y. [Yamamoto]{}, “[Single-photon generation with InAs quantum dots]{},” New J. Phys. [**6**]{}, 89 (2004). S. [Strauf]{}, N. G. [Stoltz]{}, M. T. [Rakher]{}, L. A. [Coldren]{}, P. M. [Petroff]{}, and D. [Bouwmeester]{}, “[High-frequency single-photon source with polarization control]{},” Nature Photonics [**1**]{}, 704–708 (2007). O. Gazzano, S. M. de Vasconcellos, C. Arnold, A. Nowak, E. Galopin, I. Sagnes, L. Lanco, A. Lema[î]{}tre, and P. Senellart, “Bright solid-state sources of indistinguishable single photons,” Nature Communications [**4**]{}, 1425 (2013). J. [Claudon]{}, J. [Bleuse]{}, N. S. [Malik]{}, M. [Bazin]{}, P. [Jaffrennou]{}, N. [Gregersen]{}, C. [Sauvan]{}, P. [Lalanne]{}, and J. [G[é]{}rard]{}, “[A highly efficient single-photon source based on a quantum dot in a photonic nanowire]{},” Nature Photonics [**4**]{}, 174–177 (2010). M. E. Reimer, G. Bulgarini, N. Akopian, M. Hocevar, M. B. Bavinck, M. A. Verheijen, E. P. Bakkers, L. P. Kouwenhoven, and V. Zwiller, “Bright single-photon sources in bottom-up tailored nanowires,” Nature Communications [**3**]{}, 737 (2012). P. E. Kremer, A. C. Dada, P. Kumar, Y. Ma, S. Kumar, E. Clarke, and B. D. Gerardot, “Strain-tunable quantum dot embedded in a nanowire antenna,” Phys. Rev. B [**90**]{}, 201408 (2014). S. Ates, I. Agha, A. Gulinatti, I. Rech, A. Badolato, and K. Srinivasan, “Improving the performance of bright quantum dot single photon sources using temporal filtering via amplitude modulation,” Scientific Reports [**3**]{}, 1397 (2013). K. H. Madsen, S. Ates, J. Liu, A. Javadi, S. M. Albrecht, I. Yeo, S. Stobbe, and P. Lodahl, “Efficient out-coupling of high-purity single photons from a coherent quantum dot in a photonic-crystal cavity,” Phys. Rev. B [ **90**]{}, 155303 (2014). A. Muller, E. Flagg, P. Bianucci, X. Wang, D. Deppe, W. Ma, J. Zhang, G. Salamo, M. Xiao, and C. Shih, “[Resonance fluorescence from a coherently driven semiconductor quantum dot in a cavity]{},” Phys. Rev. Lett. [**99**]{}, 187402 (2007). S. Ates, S. M. Ulrich, S. Reitzenstein, A. Löffler, A. Forchel, and P. Michler, “Post-selected indistinguishable photons from the resonance fluorescence of a single quantum dot in a microcavity,” Phys. Rev. Lett. [**103**]{}, 167402 (2009). Y.-M. He, Y. He, Y.-J. Wei, D. Wu, M. Atat[ü]{}re, C. Schneider, S. H[ö]{}fling, M. Kamp, C.-Y. Lu, and J.-W. Pan, “On-demand semiconductor single-photon source with near-unity indistinguishability,” Nature Nanotechnology [**8**]{}, 213–217 (2013). C. Santori, D. Fattal, J. Vuckovic, G. Solomon, and Y. Yamamoto, “[Indistinguishable photons from a single-photon device]{},” Nature [**419**]{}, 594–597 (2002). S. Varoutsis, S. Laurent, P. Kramper, A. Lemaître, I. Sagnes, I. Robert-Philip, and I. Abram, “Restoration of photon indistinguishability in the emission of a semiconductor quantum dot,” Phys. Rev. B [**72**]{}, 041303 (2005). I. [Fushman]{}, D. [Englund]{}, A. [Faraon]{}, N. [Stoltz]{}, P. [Petroff]{}, and J. [Vu[č]{}kovi[ć]{}]{}, “[Controlled phase shifts with a single quantum dot]{},” Science [**320**]{}, 769–772 (2008). A. [Reinhard]{}, T. [Volz]{}, M. [Winger]{}, A. [Badolato]{}, K. J. [Hennessy]{}, E. L. [Hu]{}, and A. [Imamo[ğ]{}lu]{}, “[Strongly correlated photons on a chip]{},” Nature Photonics [**6**]{}, 93–96 (2012). H. [Kim]{}, R. [Bose]{}, T. C. [Shen]{}, G. S. [Solomon]{}, and E. [Waks]{}, “[A quantum logic gate between a solid-state quantum bit and a photon]{},” Nature Photonics [**7**]{}, 373–377 (2013). I. [Wilson-Rae]{}, P. [Zoller]{}, and A. [Imamo[ǧ]{}lu]{}, “[Laser cooling of a nanomechanical resonator mode to its quantum ground state]{},” Phys. Rev. Lett. [**92**]{}, 075507 (2004). M. [Metcalfe]{}, S. M. [Carr]{}, A. [Muller]{}, G. S. [Solomon]{}, and J. [Lawall]{}, “[Resolved sideband emission of InAs/GaAs quantum dots strained by surface acoustic waves]{},” Phys. Rev. Lett. [**105**]{}, 037401 (2010). I. Yeo, P.-L. de Assis, A. Gloppe, E. Dupont-Ferrier, P. Verlot, N. S. Malik, E. Dupuy, J. Claudon, J.-M. G[é]{}rard, A. Auff[è]{}ves, et al., “Strain-mediated coupling in a quantum dot-mechanical oscillator hybrid system,” Nature Nanotechnology, [**[9]{}**]{}, 106–110 (2013). **Acknowledgements** L.S. acknowledges support under the Cooperative Research Agreement between the University of Maryland and NIST-CNST, Award 70NANB10H193. The authors thank Serkan Ates and Krishna Coimbatore Balram for useful discussions and early contributions to this work. They also thank Christopher Long and Santiago Solares for helpful advice regarding atomic force microscopy. **Author Contributions** L.S. and K.S. fabricated the devices, performed the measurements and analyzed the experimental data, and wrote the manuscript. A.B. grew the quantum dot material, M.D. and L.S. performed the electromagnetic simulations, and K.S. supervised the project.\ The identification of any commercial product or trade name is used to foster understanding. Such identification does not imply recommendation or endorsement or by the National Institute of Standards and Technology, nor does it imply that the materials or equipment identified are necessarily the best available for the purpose. **Additional Information** Correspondence and requests for materials should be addressed to L.S. and K.S. **Competing financial interests** The authors declare no competing financial interests. [[**SUPPLEMENTARY INFORMATION**]{}]{} **Supplementary Note 1: Quantum dot positioning setup and measurements** In this note, we provide additional details on the quantum dot positioning setup shown schematically in Fig. 1 of the main text. The samples are housed within a cryogen-free cryostat with a base temperature as low as 6 K. Sample motion is achieved using a three-axis cryogenic piezo-positioning stage system. A confocal micro-photoluminescence geometry is utilized, in which a microscope objective (20x magnification and 0.4 numerical aperture) both focuses excitation light on the sample and collects light emitted and reflected by the sample. As described in the main text, photoluminescence imaging is done with co-illumination by 630 nm and 940 nm LEDs, where the former is used to excite the quantum dots and the latter is used to image alignment marks. Excitation of single quantum dots for spectroscopy is performed by focusing a 780 nm laser on the sample. A 90/10 (reflection/transmission percentage) beamsplitter followed by a 900 nm long-pass filter is used to send the light emitted and reflected by the sample towards the imaging and spectroscopic characterisation paths. Selection between the two paths is accomplished with a half waveplate and polarizing beasmplitter. For photoluminescence imaging, the collected light is coupled into a variable zoom system and Electron Multiplied Charged Couple Device (EMCCD), while for spectroscopy, it is coupled to a single mode fiber whose output is sent to a grating spectrometer equipped with a silicon Charged Coupled Device (CCD). In the photoluminescence imaging measurements, the 900 nm longpass filter serves to reject 630 nm excitation light, while allowing both the quantum dot emission and reflected 940 nm LED light to pass. A total system magnification of 40x (20x from the objective, and 2x from the zoom barrel) is used, corresponding to a field of view of $\approx$ 200 $\mu$m x 200 $\mu$m. EMCCD images are acquired with an integration time of 120 s and gain of 200. While the 630 nm LED power is always set at its maximum ($\approx$ 40 mW, corresponding to an intensity of $\approx$ 130 W/cm$^2$) to generate as much fluorescence from the quantum dot as possible, the 940 nm LED power is set to achieve a reflected signal from the alignment marks that is approximately equal to the intensity of the quantum dot emission. This choice of 940 nm LED power is a tradeoff between the improved alignment mark position uncertainty produced at higher powers, and the degraded quantum dot position uncertainty that results if the reflected 940 nm LED signal swamps the quantum dot emission. A typical 940 nm LED power is $\approx$ 4 $\mu$W. The linecuts of the images taken by the EMCCD camera are analyzed using a commercial software and fitted by Gaussian functions to determine the location of the quantum dot and centers of the alignment marks. The fit is optimized using a Levenberg Marquardt iteration algorithm. The central position of the Gaussian function and its error are then translated from a pixel value on the camera to a distance on the sample by using a calibration obtained by imaging, under the same magnification conditions, a microscope calibration target presenting etched features with known separations. **Supplementary Note 2: Solid immersion lenses for reduced positioning uncertainties** Solid immersion lenses have been used to increase the collection of light emitted by semiconductor quantum dots by increasing the effective numerical aperture of the collection optics [@ref:Serrels_SIL_SI]. Moreover, because the solid immersion lens reduces the focused excitation spot size, it also increases the excitation intensity at the sample. This can lead to an increased photon flux from the quantum dot, because (as was noted in the main text), without the solid immersion lens, the maximum 630 nm LED excitation intensity at the sample is not enough to saturate the quantum dot emission. Taken together, the increased emission signal from the quantum dot should lead to a lower uncertainty in its position. We also expect that the solid immersion lens can improve the alignment mark uncertainty, since the amount of 940 nm LED power used to image the mark will be increased (see discussion above) to match the increased quantum dot emission level. We test the above experimentally using a 2 mm diameter, high refractive index ($n~\approx~2$) half-ball lens placed directly on the sample surface, with a thin layer of cryogenic grease applied between the sample and lens, obtaining the results shown in Fig. 2(d)-(e) from the main text (the x-axis scans are similar). We measure uncertainties in the quantum dot-marker distance as low as 7.6 nm, a reduction of about a factor 4 compared to the average error measured without the lens (and a factor of 2 compared to the best error measured without the lens). However, there are some considerations to take into account other than the reduced positioning error possible with a solid immersion lens. First, the solid immersion lenses are generally 1 mm or 2 mm in diameter. Therefore, when using them for imaging the quantum dot emission and alignment marks, the area of the sample that can be probed within a single measurement session is highly reduced, unless multiple lenses are used. Second, given that the lenses are hemispherical, care must be taken in optimally focusing the imaging and excitation light on the apex of the solid immersion lens, in order to avoid distortions of the image that would affect the inferred distance between the alignment mark and the quantum dot. **Supplementary Note 3: Quantum dot single-photon source characterization** A schematic of the experimental setup used to evaluate the collection efficiency and single-photon purity of the quantum dot emission is shown in Supplementary Fig. \[fig:Janis\_setup\], and is similar to that used in previous work [@ref:Ates_JSTQE_SI]. The sample is mounted on the cold finger of a liquid helium flow cryostat that sits on a two-axis nano-positioning stage. Spectral properties of the quantum dot emission are investigated via low-temperature micro-photoluminescence, where a 20x microscope objective (numerical aperture of 0.4) is used for both the illumination of the sample and the collection of the emission. Four different excitation sources are available for use. The first is a continuous wave 780 nm diode laser for basic spectroscopy. The second is a continuous wave Ti:sapphire laser, tunable beteween 780 nm and 1000 nm, that can be used to excite the quantum dot on its different excited state transitions. The third is a gain-switched, 780 nm pulsed laser diode (50 ps pulse width; 50 MHz repetition rate) for photon counting, lifetime, and correlation measurements. The final source is a 820 nm to 950 nm tunable fiber laser ($<10$ ps pulse width; 80 MHz repetition rate) used for counting, lifetime, and correlation measurements under excitation of a quantum dot’s excited state. ![Schematic of the experimental setup: (a) confocal micro-photoluminescence; (b) time-resolved photoluminescence; (c) Hanbury Brown and Twiss photon correlation setup. SMF: single-mode fiber, LPF: long-pass filter, CCD: charge-coupled device, PL: photoluminescence, SPAD: single-photon avalanche diode, QE: quantum efficiency.[]{data-label="fig:Janis_setup"}](Figure1_SI_submit.pdf){width="0.6\linewidth"} The collected signal is directed to a spectrometer either to record an emission spectrum with a Si CCD camera, or to filter a single emission line for further investigation (Supplementary Fig. \[fig:Janis\_setup\](a)). The spectrally filtered emission line is coupled into a single mode optical fiber to enable measurements using fiber-coupled single-photon avalanche diodes (SPADs). Single quantum dot fluorescence decay dynamics are measured through time-correlated single-photon counting, which relies on measuring the time delay between an excitation pulse and detection of an emitted photon by a SPAD (Supplementary Fig. \[fig:Janis\_setup\](b)). We use a thin Si SPAD whose timing jitter is $<50$ ps to enable measurement of fast quantum dot decay dynamics. For the second-order correlation function $g^{(2)}(\tau)$ measurements, the spectrally filtered emission is directed to a Hanbury-Brown and Twiss interferometer that consists of a fiber-coupled, 50/50 non-polarizing beam-splitter and two fiber-coupled single-photon avalanche diodes (SPADs), as shown in Supplementary Fig. \[fig:Janis\_setup\](c). These SPADs have a timing jitter of $\approx$ 700 ps, and their outputs are connected to a time-correlated single-photon counting board. A time bin width of 512 ps is chosen for the $g^{(2)}(\tau)$ measurements. Calibration of the quantum dot single-photon source collection efficiency into the 20x (0.4 numerical aperture) objective proceeds as follows. First, the transmission of the optical path from the QD source to the detector is determined. The emitted light escapes the cryostat by traveling through two fused silica windows (total transmission $\approx$ 87 $\%$), it is then collected by a microscope objective (transmission of $\approx$ 70 $\%$), goes through a 90/10 beamsplitter (transmission of $\approx$ 89 $\%$), reflects off four dielectric mirrors and travels through a polarizer (total transmission of $\approx$ 78 $\%$) before being focused through the slit of the grating spectrometer. The total transmission of the optical path up to the spectrometer is 42 $\%$ $\pm$ 4 $\%$, where the uncertainty is based on the spread of transmission values measured for the optical components, and represents a one standard deviation value. Next, a known laser power (22.6 $\mu$W) is sent into the spectrometer after being attenuated by an independently measured attenuation (64.21 dB) using a variable attenuator, and the detected counts on the Si CCD coupled to the spectrometer are recorded. The counts measured from a quantum dot, excited with a 50 MHz repetition rate source, are then recorded and compared to the laser counts (taking into account the transmission of the optical path) in order to extract the emitter’s single-photon collection efficiency. Figure 4 and the accompanying discussion in the main text present data characterizing single-photon source performance for an optically positioned quantum dot within a circular grating ‘bullseye’ cavity. Here, we present supplementary data referred to in the main text discussion. Supplementary Fig. \[fig:positioned\_QD\_SPS\_SI\_data\](a) shows an intensity autocorrelation measurement ($g^{(2)}(\tau)$) under pulsed excitation at 780 nm when the quantum dot emission is saturated. Despite what appears to be a relatively clean emission spectrum (in terms of an absence of spectral features other than the quantum dot emission) in Fig. 4(d) of the main text, the relatively significant multi-photon component measured ($g^{(2)}(0) = 0.15 \pm 0.03$) indicates the presence of emission that is spectrally resonant with the emission from the quantum dot excitonic line. Quasi-continuum states, generated by hybridization of states of the single quantum dot with those of the wetting layer [@ref:Winger2009_SI; @ref:Chauvin_Fiore_SI; @ref:Laucht_Finely_2010_SI], are thought to be a potential source of such multi-photon emission, particularly in Purcell-enhanced (e.g., microcavity) geometries. This is consistent with our measurements, as more intense excitation (through a 780 nm continuous wave laser that provides more output power than the pulsed 780 nm laser) yields a spectrum in which the cavity mode emission is clearly visible (Supplementary Fig. \[fig:positioned\_QD\_SPS\_SI\_data\](b)). Given that the cavity mode linewidths are many nanometers wide, while the quantum dot excitonic states are orders of magnitude narrower, a spectrally broad emission source such as a quasi-continuum state is needed to reconcile the presence of the cavity mode within the measured spectrum. ![Supplementary data for the single photon source characterization of Figure 4 from the main text. (a) $g^{(2)}(\tau)$ under 780 nm pulsed excitation, with the quantum dot emission saturated. (b) Spectrum under intense 780 nm continuous wave excitation (far above quantum dot saturation), for which the bullseye cavity modes are visible. The quantum dot line of interest is +1.6 nm detuned with respect to the shorter wavelength cavity mode. The red solid line is a fit of the data to the sum of two Gaussians, which is used to determine the center wavelengths of the two cavity modes. The two Gaussians making up the sum are shown as black dashed lines. (c) Spectrum under intense 857 nm continuous wave excitation (far above quantum dot saturation), on resonance with a quantum dot excited state. Compared to 780 nm excitation far above saturation, significantly reduced cavity mode emission is observed. (d) Comparison of the quantum dot spectra under saturation conditions (the conditions under which the $g^{(2)}(\tau)$ data in part (a) and Fig. 4(e) were taken) for both 780 nm and 857 nm pulsed excitation.[]{data-label="fig:positioned_QD_SPS_SI_data"}](Figure2_SI_submit.pdf){width="\linewidth"} The contribution of such quasi-continuum states should be limited if the system is pumped on an excited state of the quantum dot, which would prevent the generation of high energy carriers that could fill those states. Using a narrow linewidth ($<1$ MHz) continuous wave Ti:sapphire laser, we have identified 857.0 nm and 876.4 nm as wavelengths that are resonant with quantum dot excited states. Under intense excitation at these wavelengths, the cavity mode feeding is significantly reduced relative to the 780 nm case, as shown in Supplementary Fig. \[fig:positioned\_QD\_SPS\_SI\_data\](c) for 857 nm excitation. Switching to pulsed 857 nm excitation, we find that the resulting spectrum at saturation of the quantum dot emission is nearly identical to that observed under pulsed 780 nm excitation in the main text, as shown in Supplementary Fig. \[fig:positioned\_QD\_SPS\_SI\_data\](d). In contrast, the measured $g^{(2)}(\tau)$ (Fig. 4(e) in the main text) is markedly different, with $g^{(2)}(0) = 0.009 \pm 0.005$. Overall, these results indicate the importance of excited state pumping to achieving pure single photon emission, even in situations in which there is only one quantum dot that can interact with the cavity mode. Finally, we note that in the photon antibunching experiments, the grating spectrometer was used as a monochromator to spectrally isolate the quantum dot emission, and had a throughput of $\approx$ 11 $\%$. The output of the monochromator was coupled into single mode fiber and sent into the Hanbury-Brown and Twiss setup as described above, and the detected count rates on each of the two SPADs was $\approx~2{\times}10^4$ counts/s in the measurements from Fig. 4(e) in the main text. Overall, this detected count rate includes the collection efficiency of quantum dot emission into the NA = 0.4 lens ($\approx$ 48 $\%$), the transmission of the photoluminescence setup ($\approx$ 42 $\%$), the throughput of the monochromator ($\approx$ 11 $\%$), coupling from the monochromator output into single mode fiber and throughput of the single-mode-fiber-based Hanbury-Brown and Twiss setup ($\approx$ 12 $\%$), and the SPAD quantum efficiency ($\approx$ 20 $\%$). **Supplementary Note 4: Comparison to single-photon sources created without optical positioning** For the purposes of comparison, in this section we present data from quantum dot single-photon sources in which quantum dot positioning was not employed (so that the position of the quantum dot with respect to bullseye cavity center was uncontrolled). The investigation of these devices was described in detail in Ref. , where spectroscopy, lifetime, and photon correlation measurements were presented. In Supplementary Fig. \[fig:unpositioned\_BE\_data\](a), we show an EMCCD image of a subset of the array of cavities investigated in Ref. , where the array has been illuminated by the 630 nm red LED. This EMCCD image reveals two new pieces of information. First, only one of twelve displayed devices shows an emission lobe near the center of the cavity, for which the collection efficiency is expected to be maximized. For this unpositioned sample, the maximum collection efficiency measured was $\approx$ 10 $\%$, and the fraction of devices producing this efficiency was a couple of percent. Next, the quantum dot density in this sample is significantly higher than that studied in the current manuscript. While the density is still low enough so that only a single quantum dot can spatially and spectrally interact with a mode of the cavity, it is about two orders of magnitude larger than what we use in the positioned quantum dot devices. The background emission caused by these quantum dots, and in particular, their potential for supporting quasi-continuum states with broad emission bandwidths [@ref:Winger2009_SI], may limit the purity of single-photon emission. Given that the yield for this sample is only a couple of percent, reducing the quantum dot density without locating the quantum dots prior to fabrication is impractical. ![Representative data from quantum dot - bullseye cavity devices fabricated without using the quantum dot positioning technique, as in Ref. . (a) EMCCD image of an array of cavities, illuminated by a 630 nm LED. Only one of the devices (within the dashed box) shows an emission lobe centered with respect to the cavity. Scale bar represents 50 $\mu$m. (b) Photoluminescence spectrum from a device exhibiting collection efficiency $\approx$ 10 $\%$. (c) $g^{(2)}(\tau)$ from the same device. Note that parts (b) and (c) are re-displayed data from Ref. .[]{data-label="fig:unpositioned_BE_data"}](Figure3_SI_submit.pdf){width="0.65\linewidth"} Spectroscopy and photon counting measurements from Ref.  further address these points. A typical photoluminescence spectrum under non-resonant pulsed excitation is shown in Supplementary Fig. \[fig:unpositioned\_BE\_data\](b). In contrast to the clean spectrum shown in the main text in Fig. 4(d), the spectrum of Supplementary Fig. \[fig:unpositioned\_BE\_data\](b) shows significant background emission attributed to feeding of the cavity mode by multi-excitonic states of nearby quantum dots. This emission can be expected to limit the purity of the single-photon source produced by spectrally isolating a single excitonic state, and indeed, the $g^{(2)}(\tau)$ measurement in Supplementary Fig. \[fig:unpositioned\_BE\_data\](c) shows a significant departure from $g^{(2)}(0)=0$. While this measurement is of a device with a particularly high $g^{(2)}(0)$ value, in general, unpositioned devices studied in Ref.  showed $g^{(2)}(0)\gtrsim15~\%$. That being said, the discussion from the previous section indicates that even the drastically reduced quantum dot density used in the current work most likely needs to be supplemented by excited state pumping of the quantum dot in order to achieve $g^{(2)}(0)~\approx~0$. **Supplementary Note 5: Electromagnetic simulations** As discussed in the main text and in Ref. , the bullseye cavity supports dipole-like resonant modes (shown in Supplementary Figs. \[fig:dipole\](b) and (c)) that are well-suited for the creation of bright single-photon sources - a combination of relatively high Purcell-type radiative enhancement, efficient vertical light extraction from the semiconductor, and near-Gassian far-field for efficient collection into an optical fiber. These modes are strongly localized at the center of the cavity (the central intensity peak has a full-width at half-maximum of $\approx~100$ nm), and a sequence of satellite peaks along the radial direction. Because the electric dipole coupling to a cavity mode is proportional to the squared electric field magnitude [*at the dipole location*]{} [@ref:Xu_99], we expect that the Purcell enhancement factor $F_p$, coupling efficiency $\eta$, and emitted polarization state will vary significantly with dipole position. An understanding of these parameters is not just important from a device performance perspective, but also provides information about the actual quantum dot location. We employ full-wave numerical electromagnetic simulations to investigate the sensitivity of the emission properties of our single-photon source to the location of the quantum dot within the bullseye cavity. Purcell Factor and Collection Efficiency ---------------------------------------- Following Ref. , we use finite-difference time domain (FDTD) simulations to model the system as an electric dipole radiating inside a suspended bullseye cavity. The dipole is allowed to radiate with a short Gaussian pulse time dependence, and the electromagnetic field is allowed to evolve over a long time span. The steady-state electromagnetic field is recorded at all edges of the computational window, so that the total dipole radiated power $P_{rad}$ can be determined. The Purcell factor can then be obtained as $F_p=P_{rad}/P_{hom}$, where $P_{hom}$ is the dipole radiated power in a homogeneous medium [@ref:Xu_99]. We also record the power $P_z$ emitted upwards in the $+z$ direction, which in a real setting is partially collected with a microscope objective with numerical aperture $NA$. The steady-state field recorded at a parallel plane above the bullseye cavity is used to calculate the emitted far-field, which is then integrated within an angular cone corresponding to a numerical aperture $NA$ to yield the collection efficiency $\eta_{NA}$. Perfectly matched layers are used to simulate free-space above and below the cavity, so that effects related to the substrate are not taken into account. The dipole is assumed to be on the $z=0$ plane (which corresponds to the center of the semiconductor membrane) as defined by the quantum dot growth process, and to have no $z$-components. The latter assumption is appropriate for epitaxially grown InAs dots, given their few nanometer vertical size, negligible compared to the membrane thickness [@ref:Michler_book_2009_SI]. ![(a) Schematic of the simulation geometry, showing a dipole (green arrows) inside a bullseye cavity. Due to the circular symmetry, a dipole located anywhere within the cavity and with any orientation can be represented by a dipole on the $x$-axis with $d_x$ and $d_y$ components equivalent to the radial and azimuthal ones. (b) Electric field amplitude squared profile for a ’h’-type bullseye cavity mode. The $yz$-plane boundary conditions satisfied by the mode are given in the figure. This mode is only excited by $x$-dipoles. (c) Electric field amplitude squared profile for a ’v’-type bullseye cavity mode. The $yz$-plane boundary conditions satisfied by the mode are given in the figure. This mode is only excited by $y$-dipoles.[]{data-label="fig:dipole"}](Figure4_SI_submit.pdf) In the circular geometry of the cavity, an electric dipole with moment $\mathbf{d}$ with arbitrary orientation placed anywhere in the cavity is equivalent to a dipole located on the $x$-axis with components along the $x$ and $y$ directions, corresponding to the radial and azimuthal components. This is illustrated in Supplementary Fig. \[fig:dipole\](a). The symmetry of the problem allows a description of the electromagnetic fields supported by the cavity in terms of orthogonal, symmetric and anti-symmetric cavity eigenmodes with respect to the $y=0$ plane (’$h$’-modes) and degenerate modes of the 90-degree rotated geometry (’$v$’-modes), as illustrated in Supplementary Figs. \[fig:dipole\](b) and (c). An $x$-dipole will however only excite symmetric $h$-modes and anti-symmetric $v$-modes, and vice-versa is valid for a $y$-dipole; in other words, $\mathbf{d}\cdot\mathbf{E^h}=d_x\cdot E^h_x$ and $\mathbf{d}\cdot\mathbf{E^v}=d_y\cdot E^v_y$. From ref. , the total power emitted by the dipole in the cavity is $P_{rad}\propto\sum_n{\left|\mathbf{d}\cdot\mathbf{E^n}\right|^2}$, where $\mathbf{d}$ is the dipole moment and $\mathbf{E^n}$ is the (normalized) electric field for mode $n$, evaluated at the dipole location. With the symmetry considerations above, we can write $$P_{rad}\propto\sum_{n}\left|d_x\cdot E^{h,n}_x\right|^2+\left|d_y\cdot E^{v,n}_y\right|^2=\left|\mathbf{d}\right|^2\left\{\sum_{n}\left|\cdot E^{h,n}_x\right|^2\cos^2\phi+\left|\cdot E^{v,n}_y\right|^2\sin^2\phi\right\}, \label{eq:P_rad}$$ where $\phi$ is the dipole orientation - the angle the dipole makes with respect to the x-axis - which we assume to be unknown. Equation (\[eq:P\_rad\]) thus allows us to determine the dipole emitted power $P_{rad}$ for a dipole positioned anywhere on the $x$ axis, with arbitrary angle given by $\phi$, just based on the $E^h$ and $E^v$ modes which are respectively excited by the $x$ and $y$ dipole components. As such, we proceed to calculate $P_{rad}$ and the collection efficiency $\eta_{0.4}$ for $NA=0.4$ separately for $x$- and $y$-oriented dipoles located on the $x$-axis at varying distance $x_0$ from the cavity center. We then use eq.(\[eq:P\_rad\]) to determine the range within which the quantities of interest can vary due to the (unknown) azimuthal dipole orientation $\phi$. To verify that this procedure is valid, we also simulate the case $\phi=45^\circ$, and compare the collection efficiency with that obtained through eq. (\[eq:P\_rad\]) and the $x-$ and $y-$ dipole solutions. As shown in Supplementary Fig. \[fig:dipole45\_comp\], the difference between the two types of calculations is $ \lesssim1~\%$ almost everywhere. ![Collection efficiency $\eta_{0.4}$ into a $NA=0.4$ objective as a function of wavelength and dipole position along the $x$-axis inside the bullseye cavity. The dipole is on the $xy$-plane and is oriented at an azimuthal angle $\phi=45^\circ$. (a) Results obtained using eq.(\[eq:P\_rad\]) with separate simulations for $x$ and $y$ dipoles individually. (b) Results obtained by simulating a dipole at $\phi=45^\circ$. (c) Absolute value of difference between results shown in panels (a) and (b).[]{data-label="fig:dipole45_comp"}](Figure5_SI_submit.pdf) In Supplementary Figs. \[fig:Fp\](a) and (b), we show the Purcell Factor $F_p$ as a function of wavelength for $x$- and $y$-dipoles, respectively, located at varying positions along $x$. At a wavelength of $948$ nm, the $x$ dipole couples to the ’$h$’ resonance shown in Supplementary Fig. \[fig:dipole\](b), while the $y$-dipole couples to the degenerate ’$v$’ mode in Supplementary Fig. \[fig:dipole\](c), and the Purcell Factor $F_p$ peaks for dipoles at the cavity center. For y-dipoles displaced from the center, $F_{p}$ shows a sequence of satellite peaks observed for increasing distances, which contrasts with the $x$-dipole case. This can be understood based on the variation of of the $v$ and $h$ field profiles along the $x$-axis (Supplementary Figs. \[fig:dipole\](b)-(c)), as $F_{p}\propto\left|E\right|^2$. A second resonance centered at 957 nm exists that is also excited by dipoles in both orientations, however displays considerably lower Purcell enhancement and collection efficiency (shown later). Supplementary Figs. \[fig:Fp\](c) and (d) show the overall maximum and minimum achievable $F_p$, and the shaded areas in Supplementary Fig. \[fig:Fp\](e) correspond to overall allowed values of $F_p$ as a function of dipole displacement, for three wavelengths around the resonance center. Essentially, these ranges correspond to the uncertainty in our knowledge of $F_p$ due to lack of knowledge of the in-plane dipole orientation. Dotted white lines, on the other hand, correspond to the case $\phi=45^\circ$, which corresponds to an in-plane isotropic dipole. ![Purcell Factor $F_p$ as a function of wavelength and dipole position along the $x$-axis, inside the bullseye cavity. (a) Results for an $x$-oriented dipole; (b) Results for a $y$-oriented dipole; (c) maximum achievable $F_p$; (d) minimum achievable $F_p$; (e) Purcell Factor as a function of dipole displacement from the bullseye cavity center, at resonance ($\lambda=948.02$ nm) and at $\pm$ 1.6 nm away \[shown as dashed lines in (a)-(d)\]. Shaded areas correspond to the uncertainty in $F_p$ due to lack of knowledge of the dipole azimuthal angle $\phi$. The white dotted line corresponds to the case $\phi=45^\circ$.[]{data-label="fig:Fp"}](Figure6_SI_submit.pdf) Supplementary Figs. \[fig:eff\](a) and (b) show the overall maximum and minimum achievable collection efficiency $\eta_{0.4}$, and the shaded areas in Supplementary Fig. \[fig:Fp\](c) correspond to overall allowed values of $\eta_{0.4}$ as a function of dipole displacement, for three wavelengths around the resonance center. White dotted lines are for the $\phi=45^\circ$ case. It is evident that the collection efficiency is a much slower function of both wavelength and dipole displacement than the Purcell factor. As a result, for the QD-cavity wavelength detuning of the device we focus on in the main text (1.6 nm), there is a $\approx~\pm~250$ nm range of dipole positions consistent with the experimentally observed collection efficiency ($48~\%\pm5~\%$) and Purcell Factor ($\approx3$). The lack of knowledge about the QD dipole orientation $\phi$ prevents us from more precisely estimating the location of the emitter within the cavity. For example, if $\phi=45^\circ$, from Supplementary Figs. \[fig:Fp\](e) and Supplementary Fig. \[fig:eff\](c) we can estimate that the dipole is located within 50 nm of the cavity center, in order to display the experimentally observed $F_{p}$ and $\eta_{0.4}$. ![Collection efficiency $\eta_{0.4}$ into a $NA=0.4$ optic as a function of wavelength and dipole position along the $x$-axis inside the bullseye cavity. (a) Maximum achievable $\eta_{0.4}$; (b) minimum achievable $\eta_{0.4}$; (c) Collection efficiency as a function of dipole displacement from the bullseye cavity center, at resonance ($\lambda=948.02$ nm) and $\pm$ 1.6 nm away \[shown as dashed lines in (a) and (b)\]. Shaded areas correspond to the uncertainty in $\eta_{0.4}$ due to lack of knowledge of the dipole azimuthal angle $\phi$. The white dotted line corresponds to the case $\phi=45^\circ$.[]{data-label="fig:eff"}](Figure7_SI_submit.pdf) We note however that the collection efficiency maximum is shifted with respect to the resonance center by approximately -5 nm, as can be seen in Supplementary Figs. \[fig:Fp\](a)-(d) and Supplementary Figs. \[fig:eff\](a)-(b). This is due to far-field collection efficiency, which is actually asymmetric with respect to the resonance center, being higher by approximately $0.5~\%$ at a maximizing blue-shifted wavelength. While this information is still not sufficient to pinpoint the quantum dot location based on our experimental data, it further corroborates our explanation that the relatively low observed Purcell factors can still exist with high collection efficiencies. Polarization of the light emitted by a dipole embedded within a bullseye cavity ------------------------------------------------------------------------------- We now study the polarization properties of the light emitted by a dipole in the bullseye cavity. In particular, our goal is to understand the degree to which polarization-resolved measurements of the far-field intensity can be used to identify the dipole orientation, which in turn would enable more precise determination of the dipole location from Purcell enhancement and collection efficiency measurements. The ’$h$’ and ’$v$’ bullseye cavity modes overall display major electric field components oriented in the $x$ and $y$ directions, respectively. This can be verified in two ways: plots of $|E_x|^2$ and $|E_y|^2$ for the ’$v$’ mode in Supplementary Figure \[fig:polar1\](a) show the former to be overall at least an order of magnitude larger than the latter; and the ratio $R_{xy}=\int_{S_{NA}}\mathbf{dS}|E_x|^2/\int_{S_{NA}}\mathbf{dS}|E_y|^2$, where $S_{NA}$ is the spherical surface corresponding to a $NA=0.4$ cone, is calculated to be $R_{xy}=3.47$. As such, we expect the far-field produced by a dipole at an arbitrary orientation characterized by the azimuthal angle $\phi$ to display some degree of polarization. This degree of polarization can in principle be resolved by introducing of a linear polarizer above the cavity and determining the variation of the transmitted power (collected into a $0.4~NA$ optic) with respect to the polarizer orientation. ![(a) Magnitude-square, (b) Real part, and (c) Imaginary part of the $E_x$ far-field for a ’$v$’ mode. d) Magnitude-square, (e) Real part, and (f) Imaginary part of the $E_y$ far-field for a ’$v$’ mode. []{data-label="fig:polar1"}](Figure8_SI_submit.pdf) To perform this calculation, we first note that the radial component of the far electric field is much smaller than the azimuthal and polar ones ($|\mathbf{E_\rho}|\ll|\mathbf{E_\phi}|,|\mathbf{E_\theta}|$ in spherical coordinates). We then assume that the collection cone is narrow enough that the field at the entrance of the collecting lens can be well represented as $\mathbf{E}=E_x\mathbf{\hat{x}}+E_y\mathbf{\hat{y}}$, where $E_x$ and $E_y$ are the $x$- and $y$-components of the far-field (in other words, we take $E_x$ and $E_y$ to be the transverse components of the far-field). This allows us to use Jones matrix formalism to estimate the power transmitted through the polarizer. We represent a polarizer oriented at an angle $\theta_p$ with respect to the $x$-axis with the Jones matrix $$\mathbf{M} = \left[ \begin{array}{cc} \cos\theta_p & \sin\theta_p \\ -\sin\theta_p & \cos\theta_p \end{array}\right] \left[\begin{array}{cc} 1 & 0 \\ 0 & 0 \end{array}\right] \left[ \begin{array}{cc} \cos\theta_p & -\sin\theta_p \\ \sin\theta_p & \cos\theta_p \end{array} \right] = \left[ \begin{array}{cc} \cos^2\theta_p & -\sin\theta_p\cos\theta_p \\ -\cos\theta_p\sin\theta_p & \sin^2\theta_p \end{array}\right]$$ The transmitted electric field $\mathbf{E_{out}}=\mathbf{ME}$ is, then, $$\left[ \begin{array}{c} E_x \\ E_y \end{array}\right]_{out}= \left[\begin{array}{c} E_x\cos^2\theta_p - E_y\sin\theta_p\cos\theta_p \\ E_y\sin^2\theta_p - E_x\sin\theta_p\cos\theta_p \end{array}\right].$$ The transmitted power is proportional to $|\mathbf{E}|^2=|E_x|^2+|E_y|^2$. If the emitting dipole is at an arbitrary orientation, both ’$h$’ and ’$v$’ modes are produced in the cavity, so that, in the far-field, $\mathbf{E}=\alpha_h\mathbf{E^h}+\alpha_v\mathbf{E^v}$ ($\alpha_{h,v}$ represent the dipole coupling strength to the $h$ and $v$ modes). In this case, the resulting expression for the transmitted power consists of a sum of terms $E^i_kE^{j*}_l$, where $i,j\in\{h,v\}$ and $k,l\in\{x,y\}$. In determining the transmitted power, all of these terms are integrated over a portion of a spherical surface which represents the acceptance cone of the collection lens. Because of the cylindrical symmetry of the cavity, the $x$ and $y$ components of the ’$h$’ and ’$v$’ fields obey the following symmetry relations (as seen in Supplementary Figs. \[fig:polar1\](b),(c),(d) and (f)): $E^h_x$ and $E^v_y$ are even in $x$ and $y$; $E^v_x$ and $E^h_y$ are odd in $x$ and $y$. Because the integration is performed symmetrically in the $xy$ plane, any cross-term $E^i_kE^{j*}_l$ that results odd in $x$ and $y$ has no contribution to the power; these are cross terms with $\{i\neq j,k=l\}$ and $\{i=j,k\neq l\}$. We can thus write the integrands $$\left|E_x\right|^2=\cos^4\theta_p\left(\left|E_x^h\right|^2+\left|E_x^v\right|^2\right)+ \cos^2\theta_p\sin^2\theta_p\left(\left|E_y^h\right|^2+\left|E_y^v\right|^2\right)- 2\cos\theta_p\sin^3\theta_p\Re\left\{E_x^hE_y^{v*}+E_x^vE_y^{h*}\right\} \label{eq:P_pol_Ex2}$$ and $$\left|E_y\right|^2=\sin^4\theta_p\left(\left|E_y^h\right|^2+\left|E_y^v\right|^2\right)+ \cos^2\theta_p\sin^2\theta_p\left(\left|E_x^h\right|^2+\left|E_x^v\right|^2\right)- 2\cos^3\theta_p\sin\theta_p\Re\left\{E_x^{h*}E_y^v+E_x^{v*}E_y^h\right\}, \label{eq:P_pol_Ey2}$$ where the substitutions $\mathbf{E^{h,v}\leftarrow\alpha_{h,v}\mathbf{E^{h,v}}}$ were done for simplicity. We use equations (\[eq:P\_pol\_Ex2\]) and (\[eq:P\_pol\_Ey2\]) to estimate the collected power that is transmitted through the polarizer, at any polarizer orientation angle . The visibility $V$ can be determined from the maximum and minimum intensities with respect to the polarizer angle as $$V = \frac{I_{max}-I_{min}} {I_{max}+I_{min}},$$ where $I = \int_{S_{NA=0.4}}\mathbf{dS}\left|\mathbf{E}\right|^2$, where $S_{NA=0.4}$ is the spherical surface corresponding to the $NA=0.4$ collection cone. Once again, the power radiated into $\mathbf{E^h}$ and $\mathbf{E^v}$ by a dipole located at an arbitrary position in the cavity depends on the dipole’s orientation; and because the dipole orientation is not known, we can only determine the possible ranges of $V$ at each dipole location. As such, we can only determine the range of achievable visibilities $V$ at each dipole location. This is shown as a function of wavelength in Supplementary Fig. \[fig:V\](s). These plots indicate the non-monotonic dependence of the visibility on dipole location and orientation. As a result, a measurement of the visibility, taken together with measurements of the Purcell enhancement and collection efficiency, usually does not provide an unambiguous estimate of the dipole location. For example, we have measured $V=0.8$ for the device described in detail in the main text. Based on the quantum dot cavity detuning at the time of this measurement ($-1.6$ nm; lower panel in Supplementary Fig. \[fig:V\](c)), we find that this visibility is consistent with multiple different locations for the dipole, including $~\approx100$ nm away from the cavity center. While this is consistent with our estimate of dipole location based on the Purcell enhancement and collection efficiency measurements ($<250$ nm from the bullseye center), it does not provide a significantly improved estimate of the dipole location. Additional measurement techniques (for example, spatially-resolved polarization-dependent far-field measurements) may be required to achieve a better estimate. ![Visibility $V$ as a function of wavelength and dipole position along the $x$-axis inside the bullseye cavity. (a) Maximum $V$; (b) minimum $V$; (c) Visibility as a function of dipole displacement from the bullseye cavity center, at resonance ($\lambda=948.02$ nm) and $\pm$ 1.6 nm away \[shown as dashed lines in (a) and (b)\]. Shaded areas correspond to the uncertainty in $V$ due to lack of knowledge of the dipole azimuthal angle $\phi$. The white dotted line corresponds to the case $\phi=45^\circ$.[]{data-label="fig:V"}](Figure9_SI_submit.pdf) **** [1]{} url \#1[[\#1]{}]{}urlprefix K. A. Serrels, E. Ramsay, P. A. Dalgarno, B. Gerardot, J. O’Connor, R. H. Hadfield, R. Warburton, and D. Reid, “Solid immersion lens applications for nanophotonic devices,” Journal of Nanophotonics [**2**]{}, 021854 (2008). S. Ates, L. Sapienza, M. Davanco, A. Badolato, and K. Srinivasan, “Bright single-photon emission from a quantum dot in a circular Bragg grating microcavity,” IEEE Journal of Selected Topics in Quantum Electronics [**18**]{}, 1711–1721 (2012). M. Winger, T. Volz, G. Tarel, S. Portolan, A. Badolato, K. Hennessy, E. L. Hu, A. Beveratos, J. Finley, V. Savona, and A. Imamoglu, “[Explanation of photon correlations in the far-off-resonance optical emission from a quantum - dot – cavity system]{},” Phys. Rev. Lett. [**103**]{}, 207403 (2009). N. Chauvin, C. Zinoni, M. Francardi, A. Gerardino, L. Balet, B. Alloing, L. H. Li, and A. Fiore, “Controlling the charge environment of single quantum dots in a photonic-crystal cavity,” Phys. Rev. B [**80**]{}, 241306 (2009). A. Laucht, M. Kaniber, A. Mohtashami, N. Hauke, M. Michler, and J. Finley, “Temporal monitoring of nonresonant feeding of semiconductor nanocavity modes by quantum dot multiexciton transitions,” Phys. Rev. B [ **81**]{}, 241302 (2010). Y. Xu, J. S. Vučković, R. K. Lee, O. J. Painter, A. Scherer, and A. Yariv, “Finite-difference time-domain calculation of spontaneous emission lifetime in a microcavity,” J. Opt. Soc. Am. B [**16**]{}, 465–474 (1999). P. Michler, [*[Single Semiconductor Quantum Dots]{}*]{} (Springer Verlag, Berlin, 2009).
{ "pile_set_name": "ArXiv" }
--- abstract: 'We compute the diffusion coefficient and the Lyapunov exponent for a diffusive intermittent map by means of cycle expansion of dynamical zeta functions. The asymptotic power law decay of the coefficients of the relevant power series are known analytically. This information is used to resum these power series into generalized power series around the algebraic branch point whose immediate vicinity determines the desired quantities. In particular we consider a realistic situation where all orbits with instability up to a certain cutoff are known. This implies that only a few of the power series coefficients are known exactly and a lot of them are only approximately given. We develop methods to extract information from these [*stability ordered*]{} cycle expansions and compute accurate values for the diffusion coefficient and the Lyapunov exponent. The method works successfully all the way up to a phase transition of the map, beyond which the diffusion coefficient and Lyapunov exponent are both zero.' author: - | Carl P. Dettmann\ Center for Chaos and Turbulence Studies\ Niels Bohr Institute\ Blegdamsvej 17, DK-2100 Copenhagen, Denmark\ and\ Per Dahlqvist\ Royal Institute of Technology\ S-100 44 Stockholm, Sweden\ title: | Computing the diffusion coefficient for intermittent maps:\ Resummation of stability ordered cycle expansions --- =4.5cm Introduction ============ Given the task of computing an average, such as a Lyapunov exponent or diffusion coefficient of a chaotic system, one can take two different approaches. Firstly, simulation is usually simple and it provides an answer without bothering to understand the topology of the flow, but it may suffer from severe convergence problems. Secondly, these averages can be extracted from dynamical zeta functions and their expansions, known as [*cycle expansions*]{}. The basic advantage of expanding the average over cycles is that the asymptotic limit $t\rightarrow \infty$ is already taken from the beginning. Longer cycles provide corrections to the results obtained from shorter ones[@DasBuch]. Real success applying zeta functions has so far only been demonstrated for quite a restricted class of dynamical systems[@DasBuch; @AAC; @rugh]. The topology of the flow should be Markovian — symbolic dynamics may be introduced and this symbolic dynamics is of finite subshift type (meaning that there is only a finite number of forbidden substrings). In addition the system need to be hyperbolic — the stability of cycles is exponentially bounded with length. The class of systems complying with these two properties is called Axiom-A. This class is far too restricted to have any major relevance in applications. Success in expanding a zeta function depends on its analytic structure. Convergence is hampered by singularities close to the zero being studied. However, if the nature of a disturbing singularity is known, one can utilize this knowledge in a resummation scheme. If the singularity is solely due to intermittency the convergence problem is thus tamed to a large extent[@PDresum]. To appreciate the relevance of stability ordering of cycle expansions we imagine a fairly generic system, given by some set of differential equations. The problem of finding periodic orbits in a systematic way is largely facilitated if one has some symbolic dynamics. For a few potentials this is possible, for example the $x^2y^2$ model[@PDGR], the Helium atom[@he], the diamagnetic Kepler problem[@dia] and the anisotropic Kepler problem[@akp1; @akp2]. For generic flows it is often not clear what Poincaré section should be used, and how it should be partitioned to generate a symbolic dynamics. Cycles can be detected numerically by searching a long trajectory for near recurrences. The long trajectory method for finding cycles discussed in[@ACEGP; @MR] preferentially finds the least unstable cycles, regardless of their topological length. If you can find all cycles with stability $\Lambda_p$ less than a certain cutoff you can use [*stability*]{} ordered cycle expansions. Stability ordering was introduced in [@PDGR; @PDreson]. It has later been studied more systematically in [@DM97; @DC]. It is much easier to implement for a generic dynamical system than the curvature expansions which rely on finite subshift approximations to a given flow. A general stability ordered cycle expansion looks like $\sum_{i=0}^{N_{\rm max}} a_i \exp(-s l_i)$, where $a_i$ is a monotonically decreasing sequence but $l_i$ is not monotonic. In this paper we will restrict our attention to maps. (It would then be relevant to speak of stability truncation rather than stability ordering.) The expansion looks like $\sum_{i=0}^{N_{\rm max}} a_i z^i$ where a few of the coefficients may be exact whereas the rest are only approximate. In particular if the system is intermittent the number of approximate coefficients greatly exceeds the number of exact ones and the main task of this paper is to extract the information they carry. Moreover, we will make use of our a priori knowledge of the power law decay of the exact coefficients and employ the resummation technique of ref [@PDresum] to improve convergence. We believe that the idea of stability ordering has its biggest potential for systems which cannot be described by a symbolic dynamics of finite subshift type. But in order to identify the problems due only to intermittency, we will study a map with complete symbolic dynamics. Theory ====== Averages and zeta functions {#s:theory} --------------------------- A nice introduction to chaotic averages is found with proper references in [@PCLA]; we will take a slightly different approach. The reason for this is that the key step in [@PCLA] assumes that the leading zero of a zeta function is isolated. We will try to avoid this assumption by starting from an expression for the invariant density in terms of periodic orbits. The price we pay is that this formula is not rigorously proven for the intermittent systems we will study. The aim is to compute averages like $$\langle w(x,n)\rangle = \int \rho(x) w(x,n) dz$$ where $\rho(x)$ is the invariant density of the ergodic map $x \mapsto f(x)$. This density can be expressed in terms of periodic orbits via $$\rho(x)=\lim_{n \rightarrow \infty} \sum_{p} \sum_{r=1}^{\infty} \frac{\delta_{n,rn_p}}{|\Lambda_p|^r} \sum_{x_i \in p} \delta(x-x_i)$$ where $r$ is the number of repetitions of primitive orbit $p$, having period $n_{p}$, and stability $\Lambda_p=\frac{df^{n_p}}{dx}|_{x=x_i}$ with $x_i$ being any point along $p$. The weight $w(x_0,n)$ is associated with the trajectory starting at $x_0$ and evolving during $n$ iterations in such a way that it is multiplicative along the flow: $w(x_0,n_1+n_2)=w(x_0,n_1)\cdot w(f^{n_1}(x_0),n_2)$. As we are dealing with maps, it is simply $w(x_0,n)=w(x_0,1)\cdot w(f(x_0),1)\cdot w(f^2(x_0),1)\ldots w(f^{n-1},1)$. The phase space average of $w(x_0,n)$ may now be expanded in terms of periodic orbits as $$\lim_{n \rightarrow \infty} \langle w(x_0,n)\rangle = \lim_{n \rightarrow \infty} \sum_p n_p \sum_{r=1}^{\infty} w_p^r \frac{\delta_{n,rn_p}} {{|\Lambda_p^r|}} \ \ , \label{eqn:tracedef}$$ $w_p$ is the weight along with cycle $p$. Zeta functions are introduced by observing that the average [(\[eqn:tracedef\])]{} may be written as $$\lim_{n \rightarrow \infty}\langle w(x_0,n)\rangle = \lim_{n \rightarrow \infty} \frac{1}{2\pi i} \int_C z^{-n} \frac{d}{dz} \log \zeta_w^{-1}(z) dz \label{eqn:intlogder} \ \ ,$$ with the zeta function $$1/\zeta_w(z)=\prod_{p} \left(1-w_p \frac{z^{n_{p}}} {{|\Lambda_{p}|} }\right) \ \ . \label{eqn:zetaw}$$ $C$ is a small contour encircling the origin in clockwise direction. Eq. (\[eqn:intlogder\]) may be verified by inserting the zeta function (\[eqn:zetaw\]) and let the integral pick up the residues from $z=0$. The result can be recast into a sum over residues outside $C$, that is, it may be related to the analytic structure of the zeta function. The [*Lyapunov exponent*]{} can be expressed in terms of a generating function $$\lambda\equiv \lim_{n \rightarrow \infty} \frac{1}{n} \langle \log |\Lambda(x_0,n)| \rangle =\lim_{n \rightarrow \infty}\frac{1}{n}\frac{d}{d\beta} \langle \Lambda(x_0,n)^\beta\rangle\mid_{\beta=0}$$ One therefore introduce the multiplicative weight $$w (x_0,n)=\Lambda(x_0,n )^\beta \label{eqn:wlam} \ \ ,$$ One can now express the Lyapunov exponent in terms of the associated zeta function $$\lambda= \lim_{n \rightarrow \infty} \frac{1}{n}\frac{1}{2\pi i} \int_C z^{-n} \frac{d}{d \beta} \frac{d}{d z} \log \zeta^{-1}_\lambda(z) \mid_{\beta=0} dz \ \ . \label{eqn:lamC}$$ For a diffusive map $\hat{f}:\bf R \mapsto R$, the diffusion coefficient can also expressed in terms of a generating function $$\begin{aligned} D&=&\lim_{n \rightarrow \infty}\frac{1}{2n} \langle(\hat{f}^n(\hat{x}_0) -\hat{x}_0)^2\rangle\nonumber\\ &=&\lim_{n \rightarrow \infty}\frac{1}{2n}\frac{d^2}{d\beta^2} \langle e^{\beta (\hat{f}^n(\hat{x}_0)-\hat{x}_0)}\rangle\mid_{\beta=0} \label{7.5}\end{aligned}$$ motivating the introduction of the weight $$w_{D}(x_0,t)=e^{\beta (\hat{f}^n(\hat{x}_0) )-\hat{x}_0)} \ \ . \label{eqn:wD}$$ If $\hat{f}(\hat{x}+nL)=\hat{f}(\hat{x})+nL$ where $\hat{x} \in I$ ($I$ is some interval of length L) then the map can be reduced to a map $f: I \mapsto I$ on the [*elementary cell*]{}. This may be expressed in terms of a zeta function with the weight $w_p$ along cycle $p$ on the elementary cell given by $$w_p=e^{\beta \sigma_p}$$ where $$\sigma_p=\sum_{x_i \in p} \left( \hat{f}(x_i)-f(x_i) \right)$$ is the corresponding drift in the full system. The diffusion coefficient may now be expressed in terms of the associated zeta function $$D= \lim_{n \rightarrow \infty}\frac{1}{2n} \frac{1}{2\pi i} \int_C z^{-n} \frac{d^2}{d\beta^2} \frac{d}{d z} \log \zeta^{-1}_D(z) \mid_{\beta=0} dz \ \ . \label{eqn:DC}$$ In both [(\[eqn:lamC\])]{} and [(\[eqn:DC\])]{} the asymptotic behavior will determined by the leading singularity of the integrand. Since the integrand is evaluated at $\beta=0$, the singularity is located at $z=1$. If this singularity is isolated the asymptotic result is obtained by simply integrating around it. For the intermittent system we are going to consider there is a complication. The singularity is not isolated. The zeta function has a branch cut along $\mbox{Re}(z)\geq 1$ and $\mbox{Im}(z)=0$. To extract the asymptotic behavior of these integrals we need to integrate around this cut. Let us return to the Lyapunov exponent and assume that $$\begin{aligned} 1/\zeta_\lambda(z,\beta)&=&[ a_1 (1-z) + O((1-z)^{\gamma}) ]\nonumber\\ & &+ \beta [ b_0 + O(1-z) ] + O(\beta^2) \label{e:lyapansatz}\end{aligned}$$ where $\gamma >1$. This particular assumption will be motivated later in this paper. We need to evaluate $$\begin{aligned} & &\frac{1}{2\pi i} \int_{\Gamma_0} (1-s)^{-n} \frac{d}{d \beta} \frac{d}{d s} \log \zeta^{-1}_\lambda(z(s)) \mid_{\beta=0} ds \nonumber\\ &=&\frac{1}{2\pi i} \int_{\Gamma_0} (1-s)^{-n} \left(-\frac{b_0}{a_1}\right) \frac{1+O(s^{\gamma-1})}{s^2} ds\nonumber\\ &=&-\frac{b_0}{a_1}n+O(n^{2-\gamma})\end{aligned}$$ where we have changed variable to $s=1-z$. $\Gamma_0$ is a contour encircling the negative real $s$-axis in an anti-clockwise direction. When evaluating these integrals the following formula is useful $$\frac{1}{2\pi i} \int_{\Gamma_0} \frac{1}{s^\rho} e^{st}ds= \frac{t^{\rho -1}}{\Gamma(\rho)}$$ The Lyapunov exponent is thus found to be $$\lambda=-\frac{b_0}{a_1}$$ For the diffusion case we assume that $$\begin{aligned} 1/\zeta_D(z,\beta)&=&[ a_1 (1-z) + O((1-z)^{\gamma}) ]\nonumber\\ & &+ \beta^2 [ c_0 + O(1-z) ] + O(\beta^4) \label{e:diffansatz}\end{aligned}$$ We now need to evaluate $$\begin{aligned} & &\frac{1}{2\pi i} \int_{\Gamma_0} (1-s)^{-n} \frac{d^2}{d \beta^2} \frac{d}{d s} \log \zeta^{-1}_D(z(s)) \mid_{\beta=0} ds\nonumber\\ &=&-2\frac{c_0}{a_1}n+O(n^{2-\gamma})\end{aligned}$$ and $$D=-\frac{c_0}{a_1}$$ To obtain the $a$, $b$ and $c$ coefficients of this section we need to expand the zeta function (\[eqn:zetaw\]) in powers of $z$ around $z=0$ which will be discussed in the next section (\[s:expand\]) Then we will resum the series around $z=1$ in section \[s:resum\]. Expanding zeta functions {#s:expand} ------------------------ For the Lyapunov exponent calculation we expand the zeta function $$1/\zeta_\lambda(z)=\prod_p \left(1-\frac{z^{n_p}}{|\Lambda|^{1-\beta}}\right)$$ $$=\prod_p \left(1-\frac{z^{n_p}}{|\Lambda|}- \beta \frac{z^{n_p}\log |\Lambda| }{|\Lambda|}+O(\beta^2)\right)$$ $$=\prod_p \left(1-\frac{z^{n_p}}{|\Lambda|}- \beta \frac{z^{n_p}\log |\Lambda| }{|\Lambda|}\right)+O(\beta^2)$$ $$\equiv \sum_{j=0}^{\infty} \hat{a}_j z^j +\beta \left( \sum_{j=0}^{\infty} \hat{b}_j z^j \right)+O(\beta^2)$$ resulting in two power series. Similarly for the diffusion calculation we expand $$1/\zeta_D(z)=\prod_p \left(1-\frac{z^{n_p}e^{\beta \sigma_p}}{|\Lambda|}\right)$$ $$\begin{aligned} 1/\zeta_D(z)&=&\prod_p \left(1-\frac{z^{n_p}}{|\Lambda|} -\beta \frac{z^{n_p}\sigma_p}{|\Lambda|} -\beta^2 \frac{z^{n_p}\sigma_p^2}{2|\Lambda|}\right.\nonumber\\ & &\left.-\beta^3 \frac{z^{n_p}\sigma_p^3}{6|\Lambda|} \right) +O(\beta^4) \label{e:evenbeta}\end{aligned}$$ $$\equiv \sum_{j=0}^{\infty} \hat{a}_j z^j +\beta^2 \left( \sum_{j=0}^{\infty} \hat{c}_j z^j \right)+O(\beta^4)$$ We restrict our attention to systems with no net drift, that is $$\lim_{n\rightarrow\infty}\frac{1}{n}\langle \hat{f}^n(\hat{x}_0)-\hat{x}_0\rangle=0\;\; . \label{e:nodrift}$$ Therefore only even powers of $\beta$ appear in Eq. (\[e:evenbeta\]). The set of coefficients we obtain in this way depends on the truncation used in the expansion of the infinite product. For truncation by topological length, we count cycles up to a given length $N_{\rm top}$. For maps with a few branches this number is limited to roughly of order $\sim 10^1$, due to the exponential growth in the number of cycles with the topological length. All combinations of cycles with total length less than or equal to $N_{\rm top}$ are also included, as these contribute to the first $N_{\rm top}$ coefficients in each series. Thus we obtain $N_{\rm top}$ exact coefficients in each series by topological length truncation. For truncation by stability, we count cycles up to a given stability $\Lambda_{\rm max}$, and combinations where the product of stabilities is less than $\Lambda_{\rm max}$. They have lengths up to $N_{\rm max}$, and so contribute to all of the first $N_{\rm max}$ coefficients in each series, but are not the only contributions to such coefficients. They give us an approximation to the zeta function which for the intermittent case is more accurate than that obtained from the length truncation, but the values of the coefficients themselves are not exact beyond some $N_{\rm exact}(\Lambda_{\rm max})$, a quantity growing logarithmically: $N_{\rm exact}\sim \log \Lambda_{\rm max}$. For intermittent maps, as the one we will consider, $N_{\rm max}$ increases as a power of $\Lambda_{\rm max}$ and $N_{\rm max} \gg N_{\rm exact}$. Often it is found that stability ordered cycle expansions lead to noisy results as a function of $\Lambda_{\rm max}$. This is due to the breaking of shadowing pairs. For example a cycle $AB$ usually gives a contribution roughly equal to and of the opposite sign as the combination of cycles $A$ and $B$ (we will refer to such a combination as a [*pseudocycle*]{}). This means the total contribution is quite small. The phenomenon is called shadowing, and is the main mechanism for the rapid convergence of cycle expansions in hyperbolic systems. It is still present to some degree in intermittent systems. However, if one such term is included but the other is excluded because they lie on opposite sides of $\Lambda_{\rm max}$, there may be a substantial error generated. Partial shadowing which may be present can be (partially) restored by smoothing the stability ordered cycle expansions by replacing each term with inverse pseudocycle stability $\Lambda^{-1}= (\Lambda_{p_1}\cdots\Lambda_{p_k})^{-1}$ by $S(\Lambda)\Lambda^{-1}$. Here, $S(\Lambda)$ is a monotonically decreasing function, with $S(0)=1$ and $S(\Lambda > \Lambda_{\rm max})=0$. A typical “shadowing error” induced by the cutoff is due to two pseudocycles of stability $\Lambda$ separated by $\Delta\Lambda$, and whose contribution is of opposite signs. Ignoring possible weighting factors the magnitude of the resulting term is of order $\Lambda^{-1}-(\Lambda+\Delta\Lambda)^{-1} \approx\Delta\Lambda/\Lambda^2$. With smoothing there is an extra term of the form $S'(\Lambda)\Delta\Lambda/\Lambda$, which we want to minimize. A reasonable guess might be to keep $S'(\Lambda)/\Lambda$ constant and as small as possible, that is $$S(\Lambda)=\left[1-\left(\frac{\Lambda}{\Lambda_{\rm max}}\right)^2\right] \Theta(\Lambda_{\rm max}-\Lambda)$$ This function still contains a non-analytic point at $\Lambda=\Lambda_{\rm max}$, however the discontinuity is now in the derivative, not in the original function, so a smoothing error estimated by $S'(\Lambda)/\Lambda$ ($\Lambda<\Lambda_{\rm max}$) is finite. We use this smoothing function below when evaluating the zeta coefficients, and demonstrate the improvement numerically. Resumming zeta functions {#s:resum} ------------------------ The result of the [*cycle expansions*]{} in sec 2.2 is a set of power series of the form $\sum_i \hat{a}_i z^i$ around $z=0$. And, according to section 2.1, what we need are coefficients from some kind of (resummed) series around $z=1$. We now describe a method of obtaining such a series, along the lines of Ref. [@PDresum]. Suppose for a moment that the series $\sum_{i=0}^{\infty} \hat{a}_i z^i$ has a radius of convergence exceeding unity. In a practical calculation we have only a finite number $n$ (say, $N_{\rm top}$ or $N_{\rm max}$) of coefficients $\hat{a}_i$ at our disposal. We assume them to be exact, the treatment of the approximate coefficients from stability ordered expansions are discussed in sec. \[ss:stab\]. Then we can in principle expand it into another truncated (resummed) Taylor series around $z=1$. $$\sum_{i=0}^{n} \hat{a}_i z^i =\sum_{i=0}^{n} a_i (z-1)^i$$ This leads to a linear systems of equations which is trivially invertible $$a_i=\sum_{j=i}^{n} \left( \begin{array}{c}j\\i\end{array}\right)\; \hat{a}_j \label{eqn:aninf}$$ In this way one obtains the standard formulae [@DasBuch] $$\begin{aligned} \lambda&=&\frac{\sum(-1)^k\frac{\log\Lambda_1+\ldots+\log\Lambda_k} {|\Lambda_1\ldots\Lambda_k|}} {\sum(-1)^k\frac{n_1+\ldots+n_k}{|\Lambda_1\ldots\Lambda_k|}}\\ D&=&\frac{1}{2}\frac{\sum(-1)^k\frac{(\sigma_1+\ldots+\sigma_k)^2} {|\Lambda_1\ldots\Lambda_k|}} {\sum(-1)^k\frac{n_1+\ldots+n_k}{|\Lambda_1\ldots\Lambda_k|}} \label{e:Preddiff}\end{aligned}$$ where the sums run over all distinct pseudocycles. This approach is particularly cumbersome for intermittent systems where $\hat{a}_i$ (as well as $\hat{b}_i$ and $\hat{c}_i$) decays according to some power law. Then the coefficients either diverges or converges slowly as $n\rightarrow \infty$. So, for intermittent systems the resummed series cannot be a Taylor series, it has to be some generalized power series. Assume that the asymptotic behavior of the coefficients is a power law $$\hat{a}_i \sim n^{-(\gamma+1)}$$ Then the leading singularity is of the form $(1-z)^\gamma$, and the simplest possible expansion would be $$\begin{aligned} & &\sum_{i=1}^{\infty} a_i (1-z)^i + (1-z)^{\gamma} \sum_{i=0}^{\infty} \bar{a}_i (1-z)^i\nonumber\\ &=&\sum_{i=0}^\infty \hat{a}_i z^i \end{aligned}$$ Having only a finite number $n$ of coefficients $\hat{a}_i$ we propose the following resummation[@PDresum] $$\begin{aligned} & &\sum_{i=1}^{n_a} a_i (1-z)^i + (1-z)^{\gamma} \sum_{i=0}^{\bar{n}_a} \bar{a}_i (1-z)^i\nonumber\\ &=&\sum_{i=0}^n \hat{a}_i z^i +O(z^{n+1})\end{aligned}$$ If $n_a +\bar{n}_a+2=n+1$ we just get a linear system of equations to solve in order to to determine the coefficients $a_i$ and $\bar{n}_a$ from the coefficients $\hat{a}_i$. It also natural to require that $|n_a +\gamma -\bar{n}_a|<1$. The basic philosophy is to build in as much as information as possible into the ansatz. If the original power series correspond to the unweighted zeta function we know that $a_0=0$. The ansatz is thus accordingly modified, we fix $a_0=0$ and modify $n_a$ or $\bar{n}_a$ so we still get a solvable system of equations. Numerical studies of an intermittent diffusive map ================================================== The map ------- In the interval $\hat{x}\in[-1/2,1/2)$, which we call the elementary cell, our model map, following Ref. [@DC] takes the form $$\hat{f}(\hat{x})=\hat{x}(1+2|2\hat{x}|^\alpha)\;\;, \label{e:map}$$ The parameter range we consider here is $\alpha\in(0,1)$, where the Lyapunov exponent and diffusion coefficient are both nonzero. For any value of $\alpha$, this maps the interval $\hat{x}\in[-1/2,1/2)$ monotonically to $[-3/2,3/2)$. Outside the elementary cell, the map is defined to have a discrete translational symmetry, $$\hat{f}(\hat{x}+n)=\hat{f}(\hat{x})+n\qquad n\in\cal{Z}\;\;.$$ See Fig. \[f:map\]. A typical initial $\hat{x}$ in the elementary cell diffuses, wandering over the real line. The map is parity symmetric, $ \hat{f}(-\hat{x})=-\hat{f}(\hat{x}), $ so the average value of $\hat{x}_{n+1}-\hat{x}_n$ is zero, and there is no mean drift, as expressed in (\[e:nodrift\]). We now restrict the dynamics to the elementary cell, that is, we define $$x=\hat{x}-[\hat{x}+1/2]\;\;,$$ where $[z]$ is the greatest integer less than or equal to $z$, so that $x$ is restricted to the range $[-1/2,1/2)$. The reduced map is $$f(x)=\hat{f}(x)-[\hat{f}(x)+1/2]\;\;. \label{e:RedMap}$$ As discussed in Ref. [@DC], the intermittency of this map appears in the form of long cycles near the marginal point with power law stabilities. This is in contrast to Axiom-A systems for which $\Lambda$ may be bounded by exponentials of the topological length. The map has three complete branches in the elementary cell. Symbolic dynamics is introduced by labeling the branches $\{ -,0,+\}$. Due to the completeness of the symbolic dynamics the zeta functions are approximated by [@ACL] $$1/\zeta_\lambda(z)\approx 1-\sum_{n=0}^{\infty}\frac{z^{n+1}} {|\Lambda_{\overline{-0^n}}|^{1-\beta}} -\sum_{n=0}^{\infty}\frac{z^{n+1}} {|\Lambda_{\overline{+0^n}}|^{1-\beta}} \label{eqn:zeta_appr_lambda}$$ and $$1/\zeta_D(z)\approx 1-e^{-\beta}\sum_{n=0}^{\infty}\frac{z^{n+1}} {|\Lambda_{\overline{-0^n}}|} -e^{+\beta}\sum_{n=0}^{\infty}\frac{z^{n+1}} {|\Lambda_{\overline{+0^n}}|} \label{eqn:zeta_appr_diff}$$ This approximation may seem crude. For instance, the zeta functions (\[eqn:zeta\_appr\_lambda\], \[eqn:zeta\_appr\_diff\]) fail to preserve flow conservation. However in [@PDresum] we presented evidence that they capture the leading singularity structure correctly. This was obtained by comparing coefficients of the piecewise linear approximation of the intermittent map (sharing the singularity structure with the approximation above) by the exact cycle expansion. The asymptotic behavior of the fundamental cycles is given by $$\Lambda_{\overline{-0^n}}= \Lambda_{\overline{+0^n}} \sim n^{1+1/\alpha} \ \ , \label{eqn:powerlaw}$$ see [*eg.*]{}[@PDresum; @DC] for a derivation. We obtain immediately from (\[eqn:zeta\_appr\_lambda\],\[eqn:zeta\_appr\_diff\],\[eqn:powerlaw\]) $$\begin{aligned} \hat{a}_n&\sim&n^{-1-1/\alpha}\\ \hat{b}_n&\sim&n^{-1-1/\alpha}\log n\\ \hat{c}_n&\sim&n^{-1-1/\alpha}\end{aligned}$$ This leads to the forms (\[e:lyapansatz\],\[e:diffansatz\]) with $\gamma=1/\alpha$ as long as $\alpha<1$. For a general orbit we can only bound the stability in the range $$Cn_p^{1+1/\alpha} < |\Lambda_p| \leq (\mbox{max} |f'|)^{n_p} =(3+2\alpha)^{n_p}$$ so when using stability cutoff we get for the parameters $N_{\rm max}$ and $N_{\rm exact}$ discussed in section \[s:expand\]: $$N{\rm max}\sim \Lambda_{\rm max}^{\frac{\alpha}{1+\alpha}}$$ and $$N_{\rm exact} > \frac{\log \Lambda_{\rm max}}{\log (3+2\alpha)} \ \ .$$ Resumming topologically ordered cycle expansions {#ss:top} ------------------------------------------------ We will most of the time concentrate on the diffusion coefficient, but a similar analysis holds for the Lyapunov exponent, to which we return at the very end. We calculated the diffusion coefficient from resummed cycle expansions obtained using topological ordering as described in Sect. \[s:resum\] with the number of coefficients $n$ determined by the maximum topological length, up to $10$. We also used the direct formula (\[e:Preddiff\]), and performed direct simulations with roughly the same amount of computer time. The results are shown in Fig. \[f:Dvsn\], showing that the resummation gives much improvement, and is consistent with direct simulation. Resumming stability ordered cycle expansions {#ss:stab} -------------------------------------------- Now we come to the central part of our numerical work: the resummation of stability ordered cycle expansions. First we calculate the $\hat{a}_i$ as described in Sec. \[s:expand\]. The coefficients are all negative except $\hat{a}_0=1$, and their magnitudes are plotted in Fig. \[f:smooth\] where we have used $\Lambda_{\rm max} =10^5$ which corresponds to $N_{\rm exact}=8$ and $N_{\rm max}=81$. The unsmoothed coefficients are thus exact for $n\leq N_{\rm exact}$. The smoothed are not exact but are still quite accurate. For $n> N_{\rm exact}$ we clearly see how the unsmoothed begins to oscillate in an irregular fashion where as the smoothed ones are stable for much larger $n$. The next issue is how to best make use of the information contained in the $\hat{a}_n$ coefficients. As pointed out in Sect. \[s:expand\] these coefficients are not exact, but they give a better representation of the zeta function than the limited number of exact coefficients obtained from topological ordering. In order to match the series at $z=1$, we must again solve a linear set of equations, but the number of coefficients ($N_{\rm max}$) for intermittent systems is much larger than for the topological ordering. We cannot match such a large number of coefficients in both series, because the solution would be unstable to the errors in the coefficients, so we must represent the information contained in the $\hat{a}_n$ in the (fewer) number of degrees of freedom that the expansion really contains. There may be more than one solution to this problem; the solution we use here is to perform [*two*]{} resummations, the first from $z=0$ to an intermediate $0<z^{'}<1$, and the second from $z=z^{'}$ to $z=1$. $$\sum_{i=0}^{N_{\rm max}}\hat{a}_iz^i= \sum_{i=0}^{N_{\rm max}}a^{'}_i(z-z^{'})^i$$ which can be explicitly inverted $$a^{'}_n=\sum_{i=n}^{N_{\rm max}}\left( \begin{array}{c}i\\n\end{array}\right)\hat{a}_iz^{'(i-n)}$$ With $z^{'}$ suitably chosen, we have thus used the information available in the $\hat{a}_n$ approximately in proportion to their reliability. That is, the accurate low order coefficients appear with large weights in the first few $a^{'}_n$, while the less accurate high order coefficients appear with small weights. As we will see, this approach is better than one which simply ignores the higher order coefficients (this corresponds to putting $z^{'}=0$ below). As for the topological length truncation, the resummation from $z=z^{'}$ to $z=1$ leads to a set of linear equations obtained by equating coefficients in $$\begin{aligned} & &\sum_{i=1}^{n_a} a_i (1-z)^i + (1-z)^{\gamma} \sum_{i=0}^{\bar{n}_a}\bar{a}_i (1-z)^i\nonumber\\ &=&\sum_{i=0}^{n^{'}}a^{'}_i(z-z^{'})^i +O(z^{n+1})\end{aligned}$$ Again, we adjust $n_a$ and $\bar{n}_a$ so as to obtain a consistent series in powers of $z-1$ and a consistent set of linear equations. We have two parameters in the double resummation scheme, $n^{'}$ and $z^{'}$. The idea is to calculate $\lambda$ or $D$ for a range of both parameters and look for the most consistent solution. The general behavior is shown in Fig. \[f:var\]. For each stability cutoff $\Lambda_{\rm max}$, small values of $n^{'}$ lead to a variation of $D$ with $z^{'}$ which has a single maximum. Larger values of $n^{'}$ lead to functions that are either monotonically decreasing or oscillatory. We estimate the diffusion coefficient by finding the maximum for the largest value of $n^{'}$ before monotonic or oscillatory behavior sets in. The convergence of this method with the stability cutoff is shown in Fig. \[f:stabresum\]. This figure also contains the direct simulation results, obtained by estimating the left side of Eq. (\[7.5\]) for $3\times 10^3$ iterations over a sample of $3\times 10^3$ trajectories, similar to the computer time required to find the cycles with $\Lambda<10^5$. The errors were obtained by looking at the scatter in this statistical sample of trajectories; for intermittent maps the diffusion coefficient always tends to be too high because long intermittent episodes are not sampled sufficiently. Close to the phase transition at $\alpha=1$ convergence is practically logarithmic in the number of iterations, with exponentially long times required to achieve convergence. For example, with the numerical procedure described above we find $D=0.0524\pm0.0005$ at $\alpha=1$ where we know $D=0$. Even at $\alpha=0.7$, a reasonable distance from the transition, the resummed cycle expansion result is more accurate than direct simulation. Our final value for the diffusion coefficient at $\alpha=0.7$ is $D=0.1267\pm0.0003$ with the resummation method. It is then quite compatible with the topological ordering discussed in Sect \[ss:top\], which yielded the result $D=1.262\pm0.0003$. In this example, it is clear that resummed cycle expansions, whether ordered by topological length or stability provide an accurate method of analyzing intermittent systems. Stability ordering is most important in more complicated systems where topological ordering is not a realistic alternative. There is an additional advantage with the stability ordered expansion in that it provides a large number of approximate coefficients, thus facilitating a numerical estimate of the power law if it is not known analytically. Recall that this power law is used in the resummation ansatz, and was absolutely essential for the good result in Sect \[ss:top\]. The phase transition -------------------- Having gained confidence in the resummation method for $\alpha=0.7$, far from the phase transition at $\alpha=1$, we now vary $\alpha$, including values for which direct simulation is totally impractical, due to logarithmically slow convergence. At $\alpha=0.99$ we obtain Fig. \[f:0.99\] for the diffusion coefficient, showing a consistent value of $D=0.0066\pm0.0001$. Plotting $D$ vs $\alpha$ (Fig. \[f:phasediff\]) we find a linear dependence near the phase transition at $\alpha=1$. Finally we performed the same analysis for the Lyapunov exponent, which has a similar dependence on $\alpha$, shown in Fig. \[f:phaselyap\] Conclusion ========== We have demonstrated that resummed stability ordered cycle expansions can provide accurate estimates of dynamical averages for intermittent maps, even close to a phase transition. This analysis could equally apply to maps with uncontrolled symbolic dynamics, as long as a reliable method exists for locating the cycles. Our methods can also be applied without much modification to flows. Then the variable $z$ is replaced by $\exp(-s)$ and the cycle expansion is actually a Dirichlet series, $\sum_i a_i \exp(-sl_i)$, where the lengths of the pseudo orbits $l_i$ are not restricted to integer values. With an additional resummation step at $s'$ (corresponding to $z'$ in this paper), the zeta function may be represented as a standard power series, thus allowing it to be matched to a generalized power series at $s=0$. For intermittent systems $s=0$ is again a branchpoint, and information about it can be obtained from the methods described in [@PDreson; @PDsin] or numerically from the stability ordered expansion. C. D.  is supported by the Danish Research Academy. P. D.  is supported by the Swedish Natural Science Research Council (NFR) under contract no. F-AA/FU 06420-311. We thank the Göran Gustafsson foundation and NORDITA for support. [99]{} P. Cvitanović et.al. [*Classical and Quantum Chaos: A Cyclist Treatise*]{}, http://www.nbi.dk/ChaosBook/, Niels Bohr Institute (Copenhagen 1997). R. Artuso, E. Aurell and P. Cvitanović, Nonlinearity [**3**]{}, 325 and 361, (1990). F. Christansen, P. Cvitanović and H. H. Rugh, Phys. Rev. Lett. [**65**]{}, 2087 (1990). P. Dahlqvist, J. Phys. A [**30**]{}, L351 (1997). P. Dahlqvist and G. Russberg, J. Phys. A [**24**]{}, 4763 (1991). D. Wintgen, K. Richter and G. Tanner, Chaos [**2**]{}, 19 (1992). B. Eckhardt and D. Wintgen, J. Phys. [**B 23**]{}, 355 (1990). M. C. Gutzwiller, J. Math. Phys. ,[**18**]{}, 806 (1977). R. Devaney, J. Diff. Eq. [**29**]{}, 253 (1978). D. Auerbach, P. Cvitanović, J.-P. Eckmann, G. H. Gunaratne and I. Procaccia, Phys. Rev. Lett. [**58**]{}, 2387 (1987). G. P. Morriss and L. Rondoni, J. Stat. Phys. [**75**]{}, 553 (1994). P. Dahlqvist, J. Phys. A [**27**]{}, 763 (1994). C. P. Dettmann and G. P. Morriss, Phys. Rev. Lett., 4201 (1997). C. P. Dettmann and P. Cvitanović, Phys. Rev. E, in press (1997). P. Cvitanović, Physica [**D83**]{}, 109 (1995). R. Artuso, G. Casati and R. Lombardi, Phys. Rev. Lett. [**71**]{} (1993) 62. P. Dahlqvist, Nonlinearity [**8**]{},11 (1995).
{ "pile_set_name": "ArXiv" }
--- abstract: '[ I prove near-optimal frequentist regret guarantees for the finite-horizon Gittins index strategy for multi-armed bandits with Gaussian noise and prior. Along the way I derive finite-time bounds on the Gittins index that are asymptotically exact and may be of independent interest. I also discuss computational issues and present experimental results suggesting that a particular version of the Gittins index strategy is an improvement on existing algorithms with finite-time regret guarantees such as UCB and Thompson sampling. ]{}' bibliography: - 'all.bib' title: '[[Regret Analysis of the Finite-Horizon Gittins Index Strategy for Multi-Armed Bandits]{}]{}' --- Introduction ============ The stochastic multi-armed bandit is a classical problem in sequential optimisation that captures a particularly interesting aspect of the dilemma faced by learning agents. How to explore an uncertain world, while simultaneously exploiting available information? Since [@Rob52] popularised the problem there have been two main solution concepts. The first being the Bayesian approach developed by [@BJK56], [@Git79] and others, where research has primarily focussed on characterising optimal solutions. The second approach is frequentist, with the objective of designing policies that minimise various forms of regret [@LR85]. The purpose of this article is to prove frequentist regret guarantees for popular Bayesian or near-Bayesian algorithms, which explicates the strong empirical performance of these approaches observed by [@KCOG12] and others. In each round the learner chooses an arm $I_t \in {\left\{1,\ldots,d\right\}}$ based on past observations and receives a Gaussian reward $X_t \sim \mathcal N(\mu_{I_t}, 1)$ where $\mu \in {\mathbb R}^d$ is the vector of unknown means. A strategy is a method for choosing $I_t$ and is denoted by $\pi$. The performance of a particular strategy $\pi$ will be measured in terms of the expected regret, which measures the difference between the expected cumulative reward of the optimal strategy that knows $\mu$ and the expected rewards of $\pi$. Let $\mu^* = \max_i \mu_i$ be the mean reward for the optimal arm, then the expected regret up to horizon $n$ is defined by [$$\begin{aligned} \label{eq:regret} R^\pi_\mu(n) = {\mathbb E}\left[\sum_{t=1}^n (\mu^* - \mu_{I_t})\right]\,,\end{aligned}$$]{} where the expectation is taken with respect to the uncertainty in the rewards and any randomness in the strategy. If $Q$ is a probability measure on ${\mathbb R}^d$, then the Bayesian regret is the expectation of \[eq:regret\] with respect to the prior $Q$. [$$\begin{aligned} \label{eq:bayes} {B\!R}^\pi_Q(n) = {\mathbb E}_{\theta \sim Q} \! \left[{\mathbb E}\left[\sum_{t=1}^n (\mu^* - \mu_{I_t}) \Bigg| \mu = \theta\right]\right]\,.\end{aligned}$$]{} I assume that $Q$ is Gaussian with diagonal covariance matrix $\Sigma = {\operatorname{diag}}(\sigma_1^2,\ldots,\sigma_d^2)$. A famous non-Bayesian strategy is UCB [@KR95; @Agr95; @ACF02], which chooses $I_t = t$ for rounds $t \in {\left\{1,\ldots,d\right\}}$ and subsequently maximises an upper confidence bound. [$$\begin{aligned} I_t = {\operatornamewithlimits{arg\,max}}_i \hat \mu_i(t-1) + \sqrt{\frac{\alpha \log t}{T_i(t-1)}}\,,\end{aligned}$$]{} where $\hat \mu_i(t-1)$ is the empirical estimate of the return of arm $i$ based on samples from the first $t-1$ rounds and $T_i(t-1)$ is the number of times that arm has been chosen. For $\alpha > 2$ and any choice of $\mu$ it can be shown that [$$\begin{aligned} \label{eq:ucb} R^{{\text{\scalebox{0.8}{UCB}}}}_\mu(n) = O\left( \sum_{i: \Delta_i > 0} \frac{\log(n)}{\Delta_i} + \Delta_i\right)\,,\end{aligned}$$]{} where $\Delta_i = \mu^* - \mu_i$ is the regret incurred by choosing arm $i$ rather than the optimal arm. No strategy enjoying sub-polynomial regret for all mean vectors can achieve smaller asymptotic regret than \[eq:ucb\], so in this sense the UCB strategy is optimal [@LR85]. The Bayesian strategy minimises \[eq:bayes\], which appears to be a hopeless optimisation problem. A special case where it can be solved efficiently is called the one-armed bandit, which occurs when there are two arms ($d = 2$) and the expected return of the second arm is known ($\sigma^2_2 = 0$). [@BJK56] showed that the Bayesian optimal strategy involves choosing the first arm as long as its “index” is larger than the return of the second arm and thereafter choosing the second arm. The index depends only on the number of rounds remaining and the posterior distribution of the first arm and is computed by solving an optimal stopping problem. Another situation when \[eq:bayes\] can be solved efficiently is when the horizon is infinite ($n = \infty$) and the rewards are discounted geometrically. Then [@Git79] was able to show that the Bayesian strategy chooses in each round the arm with maximal index. Gittins’ index is defined in the same fashion as the index of [@BJK56] but with obvious modifications to incorporate the discounting. The index has a variety of interpretations. For example, it is equal to the price per round that a rational learner should be willing to pay in order to play the arm until either the horizon or their choice of stopping time [@Web92]. Gittins’ result is remarkable because it reduces the seemingly intractable problem of finding the Bayesian optimal strategy to solving an optimal stopping problem for each arm separately. In our setting, however, the horizon is finite and the rewards are not discounted, which means that Gittins’ result does not apply and the Bayesian optimality of Gittins index strategy is not preserved.[^1] Nevertheless, the finite-horizon Gittins index strategy has been suggested as a tractable heuristic for the Bayesian optimal policy by [@Nin11] and [@Kau16], a claim that I support with empirical and theoretical results in the special case that $d = 2$ (with both means unknown). A brief chronological summary of the literature on multi-armed bandits as relevant for this article is given in \[sec:history\]. I make a number of contributions. Upper and lower bounds on the finite-horizon Gittins index for the Gaussian prior/noise model that match asymptotically and are near-optimal in finite time. Asymptotic approximations are known for the discounted case via an elegant embedding of the discrete stopping problem into continuous time and solving the heat equation as a model for Brownian motion [@Bat83; @Yao06]. In the finite-horizon setting there are also known approximations, but again they are asymptotic in the horizon and are not suitable for regret analysis [@CR65; @BK97]. A proof that the Gittins index strategy with the improper flat Gaussian prior enjoys finite-time regret guarantees comparable to those of UCB. There exists a universal constant $c > 0$ such that [$$\begin{aligned} \label{eq:log-regret} R^{{\text{\scalebox{0.8}{Gittins}}}}_\mu(n) \leq c\cdot \left( \sum_{i : \Delta_i > 0} \frac{\log(n)}{\Delta_i} + \Delta_i \right)\,.\end{aligned}$$]{} I also show that the Gittins index strategy is asymptotically order-optimal for arbitrary Gaussian priors. There exists a universal constant $c > 0$ such that for all non-degenerate Gaussian priors the Bayesian strategy satisfies [$$\begin{aligned} \limsup_{n\to\infty} \frac{R^{{\text{\scalebox{0.8}{Gittins}}}}_\mu(n)}{\log (n)} \leq c \cdot \left(\sum_{i:\Delta_i > 0} \frac{1}{\Delta_i}\right) \,.\end{aligned}$$]{} While \[eq:log-regret\] depends on a particular choice of prior, the above result is asymptotic, but independent of the prior. This is yet another example of a common property of (near) Bayesian methods, which is that the effect of the prior is diluted in the limit of infinite data. A unique aspect of the bandit setting is that the strategy is collecting its own data, so although the result is asymptotic, one has to prove that the strategy is choosing the optimal arm sufficiently often to get started. For the special case that there are two arms ($d = 2$) I show by a comparison to the Gittins index strategy that the fully Bayesian strategy (minimising \[eq:bayes\]) also enjoys the two regret guarantees above. I present a method of computing the index in the Gaussian case to arbitrary precision and with sufficient efficiency to simulate the Gittins index policy for horizons of order $10^4$. This is used to demonstrate empirically that the Gittins index strategy with a flat prior is competitive with state-of-the-art algorithms including UCB, Thompson sampling [@Tho33], OCUCB [@Lat15-ucb] and the fully Bayesian strategy. I propose an efficient approximation of the Gittins index that (a) comes with theoretical guarantees and (b) is empirically superior to Thompson sampling and UCB, and only marginally worse than OCUCB. The approximation may also lead to improvements for index-based algorithms in other settings. Bounds on the Gittins Index {#sec:approx} =========================== As previously remarked, the Gittins index depends only on the posterior mean and variance for the relevant arm and also on the number of rounds remaining in round $t$, which is denoted by $m = n - t + 1$. Let $\nu \in {\mathbb R}$ and $\sigma^2 \geq 0$ be the posterior mean and variance for some arm in a given round. Let $\mu \sim \mathcal N(\nu, \sigma^2)$ and $Y_1, Y_2, \ldots, Y_m$ be a sequence of random variables with $Y_t \sim \mathcal N(\mu, 1)$. Assume ${\left\{Y_t\right\}}_{t=1}^m$ are independent after conditioning on $\mu$. The Gittins index for an arm with posterior mean $\nu$ and variance $\sigma^2$ and horizon $m = n-t+1$ is given by [$$\begin{aligned} \label{eq:gittins} \gamma(\nu, \sigma^2, m) = \max{\left\{\gamma : \sup_{1 \leq \tau \leq m} {\mathbb E}\left[(\mu - \gamma) \tau\right] \geq 0\right\}}\,,\end{aligned}$$]{} where the supremum is taken over stopping times with respect to the filtration generated by random variables $Y_1,\ldots, Y_m$. Equivalently, ${\mathds{1}\!\!{\left\{\tau = t\right\}}}$ must be measurable with respect to $\sigma$-algebra generated by $Y_1,\ldots,Y_t$. I denote $\tau(\nu, \sigma^2, m)$ to be the maximising stopping time in \[eq:gittins\] for $\gamma = \gamma(\nu,\sigma^2, m)$. The following upper and lower bound on $\gamma$ is the main theorem of this section. \[thm:gittins\] If $\beta \geq 1$ is chosen such that $\gamma(\nu, \sigma^2, m) = \nu + \sqrt{2 \sigma^2 \log \beta}$. Then there exists a universal constant $c > 0$ such that [$$\begin{aligned} c \min{\left\{\frac{m}{\log_{+}^{\frac{3}{2}}(m)},\, \frac{m \sigma^2}{\log_{+}^{\frac{1}{2}}(m\sigma^2) }\right\}} \leq \beta \leq \frac{m}{\log^{\frac{3}{2}}(m)}\,,\end{aligned}$$]{} where ${\log_{+}\!}(x) = \max{\left\{1, \log(x)\right\}}$. I use $c, c'$ and $c''$ as *temporary* positive constants that change from theorem to theorem and proof to proof. A table of notation may be found in \[sec:notation\].\ The upper bound in \[thm:gittins\] is rather trivial while the lower bound relies on a carefully constructed stopping time. As far as the asymptotics are concerned, as $m$ tends to infinity the upper and lower bounds on $\beta$ converge up to constant factors, which leads to [$$\begin{aligned} \label{eq:asy} \exp\left(\left(\frac{\gamma(\nu, \sigma^2, m) - \nu}{\sqrt{2\sigma^2}}\right)^2\right) = \Theta\left(\frac{m}{\log^{\frac{3}{2}} m}\right)\,.\end{aligned}$$]{} The $\log^{\frac{3}{2}}(m)$ looks bizarre, but is surprisingly well-justified as shall be seen in the next section. In order to obtain logarithmic regret guarantees it is necessary that the approximation be accurate in the exponential scale as in \[eq:asy\]. Merely showing that $\log \beta / \log(m)$ tends to unity as $m$ grows would not be sufficient. Empirically choosing $c = 1/4$ in the lower bound leads to an excellent approximation of the Gittins index (see \[fig:approx-index\] in Appendix \[sec:exp\]). The proof of \[thm:gittins\] relies on a carefully chosen stopping time. [$$\begin{aligned} \label{eq:stopping} \tau = \min{\left\{m,\,\, \min{\left\{t \geq \frac{1}{\nu^2} : \hat \mu_t + \sqrt{\frac{4}{t} \log\left(4t \nu^2 \right)} \leq 0\right\}}\right\}}\,,\end{aligned}$$]{} where $\hat \mu_t = \frac{1}{t} \sum_{s=1}^t Y_s$. Note that ${\mathds{1}\!\!{\left\{\tau = t\right\}}}$ depends only on $Y_1,Y_2,\ldots, Y_t$ so this is a stopping time with respect to the right filtration. \[lem:bandit\] If $\nu < 0$, then there exists a universal constant $c' \geq 2e$ such that: 1. If $\theta \in (-\infty, \nu]$, then $\displaystyle {\mathbb E}[\tau|\mu = \theta] \leq 1 + \frac{c'}{\nu^2}$. 2. If $\theta \in (\nu, 0)$, then $\displaystyle {\mathbb E}[\tau|\mu = \theta] \leq 1 + \frac{c'}{\theta^2} \log\left(\frac{e\nu^2}{\theta^2}\right)$. 3. If $\theta \in [0,\infty)$, then ${\mathbb E}[\tau|\mu = \theta] \geq m/2$. The proof of Lemma \[lem:bandit\] may be found in \[sec:lem:bandit\]. Before the proof of \[thm:gittins\] we need two more results. The proofs follow directly from the definition of the Gittins index and are omitted. \[lem:shift\] For all $m \geq 1$ and $\sigma^2 \geq 0$ and $\nu, \nu' \in {\mathbb R}$ we have $\gamma(\nu, \sigma^2, m) - \nu = \gamma(\nu',\sigma^2, m) - \nu'$. \[lem:positive\] $\gamma(\nu, \sigma^2, m) \geq \nu$ for all $\nu$, $\sigma^2$ and $m$. [\[thm:gittins\]]{} Starting with the lower bound, by Lemma \[lem:shift\] it is enough to bound the Gittins index for any choice of $\nu$. Let $\nu = -\sqrt{2\sigma^2 \log \beta}$ where $\beta \geq 1$ will be defined subsequently. I will shortly show that there exists a stopping time $\tau$ adapted to the filtration generated by $Y_1,\ldots,Y_m$ for which [$$\begin{aligned} \label{eq:zero} {\mathbb E}\left[\mu\tau\right] \geq 0\,.\end{aligned}$$]{} Therefore by the definition of the Gittins index we have $\gamma(\nu, \sigma^2, m) \geq 0$ and so by Lemma \[lem:shift\] [$$\begin{aligned} (\forall \nu') \qquad \gamma(\nu', \sigma^2, m) \geq \nu' + \sqrt{2\sigma^2 \log \beta}\,.\end{aligned}$$]{} I now prove \[eq:zero\] holds for the stopping time given in \[eq:stopping\]. First note that if $\beta = 1$, then $\nu = 0$ and ${\mathbb E}\left[\mu \tau\right] = 0$ for the stopping time with $\tau = 1$ surely. Now assume $\beta > 1$ and define $P = \mathcal N(\nu, \sigma^2)$ to be a Gaussian measure on ${\mathbb R}$. Then [$$\begin{aligned} {\mathbb E}\left[\mu\tau\right] = {\mathbb E}_{\theta \sim P} \left[\theta {\mathbb E}\left[\tau|\mu = \theta\right]\right] &= \int^\infty_0 \theta {\mathbb E}\left[\tau|\mu = \theta\right] dP(\theta) + \int^0_{-\infty} \theta {\mathbb E}\left[\tau|\mu = \theta\right] dP(\theta)\,. \label{eq:split}\end{aligned}$$]{} The integrals will be bounded seperately. Define [$$\begin{aligned} {\operatorname{Erfc}}(x) &= \frac{2}{\sqrt{\pi}} \int^\infty_x \exp\left(-y^2\right)dy & f(\beta) &= \frac{1}{\beta} \sqrt{\frac{1}{2\pi}} - \sqrt{\frac{\log \beta}{2}}{\operatorname{Erfc}}\left(\sqrt{\log \beta}\right)\,.\end{aligned}$$]{} A straightforward computation combined with Lemma \[lem:bandit\] shows that [$$\begin{aligned} \label{eq:first} &\int^\infty_0 \theta {\mathbb E}\left[\tau|\mu = \theta\right] dP(\theta) \geq \frac{m}{2} \int^\infty_0 \theta dP(\theta) = \frac{m\sigma}{2} f(\beta)\,.\end{aligned}$$]{} The function $f$ satisfies $f(\beta) \beta {\log_{+}\!}(\beta) = \Theta(1)$ with the precise statement given in Lemma \[lem:beta\] in \[sec:algebra\]. Let $W:[0,\infty) \to {\mathbb R}$ be the product logarithm, defined implicitly such that $x = W(x) \exp(W(x))$. For the second integral in \[eq:split\] we can use Lemma \[lem:bandit\] again, which for $\theta \in (\nu, 0)$ gives [$$\begin{aligned} \theta\left({\mathbb E}[\tau|\mu = \theta] - 1\right) &\geq \theta \min{\left\{m,\, \frac{c'}{\theta^2} \log\left(\frac{e\nu^2}{\theta^2}\right)\right\}} \geq -\sqrt{c' m {\operatorname{W\!}}\left(\frac{e m\nu^2}{c'}\right)}\,,\end{aligned}$$]{} where in the first inequality we have exploited the fact that $\tau \leq m$ occurs surely and the second follows by noting that $\theta$ is negative and that $\theta m$ is increasing in $\theta$ and $1/\theta \log(e \nu^2 \theta^{-2})$ is decreasing. Let $\epsilon = 1 - \sqrt{7/8}$, which is chosen such that by tail bounds on the Gaussian integral (Lemma \[lem:gaussian\]): [$$\begin{aligned} \int^0_{\epsilon \nu} dP(\theta) \leq \exp\left(-\frac{(1 - \epsilon)^2 \nu^2}{2\sigma^2}\right) = \left(\frac{1}{\beta}\right)^{\frac{7}{8}} \quad\text{ and }\quad \int^{0}_{-\infty} \theta dP(\theta) \geq \nu - \frac{\sigma}{\sqrt{2\pi}}\,. \end{aligned}$$]{} Combining the above two displays with Lemma \[lem:bandit\] we have for some universal constant $c'' > 0$ [$$\begin{aligned} \int^0_{-\infty} \theta &{\mathbb E}[\tau|\mu = \theta] dP(\theta) \geq \left(1 + \frac{c'}{\epsilon^2 \nu^2} \log\left(\frac{e}{\epsilon^2}\right)\right) \int^{0}_{-\infty} \theta dP(\theta) - \int^0_{\epsilon\nu} \sqrt{c'm {\operatorname{W\!}}\left(\frac{e m\nu^2}{c'} \right)} dP(\theta) \\ &{\stackrel{(a)}}\geq c''\left(\left(1 + \frac{1}{\nu^2}\right) \left(\nu - \sigma\right) - \sqrt{m {\operatorname{W\!}}\left(\frac{e m\nu^2}{c'} \right)} \left(\frac{1}{\beta}\right)^{\frac{7}{8}}\right) \\ &{\stackrel{(b)}}\geq -c''\left(\left(1 + \frac{1}{2\sigma^2 \log(\beta)}\right) \left(\sqrt{2\sigma^2 \log(\beta)} + \sigma\right) + \sqrt{m {\operatorname{W\!}}\left(m\sigma^2 \log(\beta) \right)} \left(\frac{1}{\beta}\right)^{\frac{7}{8}}\right)\,,\end{aligned}$$]{} where in (a) I have substituted the bounds on the integrals in the previous two displays and naively chosen a large constant $c''$ to simplify the expression and in (b) I substitute $\nu=-\sqrt{2\sigma^2 \log(\beta)}$ and exploited the assumption that $c' \geq 2e$. Therefore by \[eq:first\] and \[eq:split\] we have [$$\begin{aligned} {\mathbb E}[\mu\tau] \geq \frac{m\sigma}{2} f(\beta) - c''\left(\left(1 \!+\! \frac{1}{2\sigma^2 \log(\beta)}\right) \left(\sqrt{2\sigma^2 \log(\beta)} \!+\! \sigma\right) + \sqrt{m {\operatorname{W\!}}\left(m\sigma^2 \log(\beta) \right)} \beta^{-\frac{7}{8}}\right)\,.\end{aligned}$$]{} Therefore by expanding the bracketed product, dropping the non-dominant $1/(\sigma \log(\beta))$ term and upper bounding the sum by the max there exists a universal constant $c''' \geq 1$ such that the following implication holds: [$$\begin{aligned} m\sigma f(\beta) \geq c'''\max{\left\{\sigma,\, \frac{1}{\sqrt{\sigma^2 \log(\beta)}},\, \sqrt{\sigma^2 \log(\beta)},\, \sqrt{mW(m\sigma^2 \log(\beta))} \beta^{-\frac{7}{8}}\right\}} \!\implies {\mathbb E}[\mu\tau] \geq 0\,.\end{aligned}$$]{} Let $c \geq 1$ be a sufficiently large universal constant and define $\beta$ by [$$\begin{aligned} \beta_1 &=\frac{m}{c \log_+^{\frac{3}{2}}(m)} & \beta_2 &=\frac{m\sigma^2}{c \log_+^{\frac{1}{2}}(m\sigma^2)} & \beta &= \begin{cases} \min{\left\{\beta_1,\, \beta_2\right\}} & \text{if } \min{\left\{\beta_1,\, \beta_2\right\}} \geq 3 \\ 1 & \text{otherwise} \end{cases}\end{aligned}$$]{} If $\beta \geq 3$, then the inequality leading to the implication is shown tediously by applying Lemma \[lem:beta\] to approximate $f(\beta)$ (see \[sec:algebra\] for the gory algebraic details). If $\beta = 1$, then $\nu = 0$ and ${\mathbb E}[\mu\tau] \geq 0$ is trivial for the stopping time $\tau = 1$ surely. Finally we have the desired lower bound by Lemma \[lem:shift\], [$$\begin{aligned} \gamma(\nu', \sigma^2, m) \geq \nu' + \sqrt{2\sigma^2 \log(\beta)} \geq \nu' + \sqrt{2\sigma^2 \log\left(\frac{1}{3c}\min{\left\{\frac{m}{\log_+^{\frac{3}{2}}(m)},\, \frac{m\sigma^2}{\log_+^{\frac{1}{2}}(m\sigma^2)}\right\}}\right)}\,.\end{aligned}$$]{} For the upper bound we can proceed naively and exploit the fact that the stopping time must be chosen such that $1 \leq \tau \leq m$ surely. This time choose $\nu$ such that $\gamma(\nu, \sigma^2, m) = 0$ and let $\beta \geq 1$ be such that $\nu = -\sqrt{2\sigma^2 \log(\beta)}$, which is always possible by Lemma \[lem:positive\]. Then [$$\begin{aligned} 0 &= \sup_\tau \int^\infty_{-\infty} \theta {\mathbb E}[\tau|\mu = \theta] dP(\theta) \leq m\int^\infty_0 \theta dP(\theta) + \int^\nu_{-\infty} \theta dP(\theta) =m\sigma f(\beta) - \frac{\sigma}{\sqrt{2\pi}} + \frac{\nu}{2}\,.\end{aligned}$$]{} Rearranging and substituting the definition of $\nu$ and applying Lemma \[lem:beta\] in the appendix leads to [$$\begin{aligned} \frac{m}{\sqrt{8\pi} \beta \log \beta} \geq m f(\beta) \geq \frac{1}{\sqrt{2\pi}} + \sqrt{\frac{1}{2} \log \beta} \geq \sqrt{\frac{1}{2} \log \beta}\,.\end{aligned}$$]{} Naive simplification leads to $\displaystyle \beta \leq m / \log^{\frac{3}{2}}(m)$ as required. Regret Bounds for the Gittins Index Strategy {#sec:finite} ============================================ I start with the finite-time guarantees. Before presenting the algorithm and analysis we need some additional notation. The empirical mean of arm $i$ after round $t$ is denoted by [$$\begin{aligned} \hat \mu_i(t) &= \frac{1}{T_i(t)} \sum_{s=1}^t {\mathds{1}\!\!{\left\{I_s = i\right\}}} X_{s} & T_i(t) &= \sum_{s \leq t} {\mathds{1}\!\!{\left\{I_s = i\right\}}}\,.\end{aligned}$$]{} I will avoid using $\hat \mu_i(t)$ for rounds $t$ when $T_i(t) = 0$, so this quantity will always be well-defined. Let $\Delta_{\max} = \max_i \Delta_i$ and $\Delta_{\min} = \min {\left\{\Delta_i : \Delta_i > 0\right\}}$. The Gittins index strategy for a flat Gaussian prior is summarised in \[alg:flat\]. Since the flat prior is improper, the Gittins index is not initially defined. For this reason the algorithm chooses $I_t = t$ in rounds $t \in {\left\{1,\ldots,d\right\}}$ after which the posterior has unit variance for all arms and the posterior mean is $\hat \mu_i(d)$. An alternative interpretation of this strategy is the limiting strategy as the prior variance tends to infinity for all arms. Choose each arm once\ \[thm:finite\] Let $\pi$ be the strategy given in \[alg:flat\]. Then there exist universal constants $c, c' > 0$ such that [$$\begin{aligned} R^\pi_\mu(n) &\leq c \sum_{i: \Delta_i > 0} \left(\frac{\log(n)}{\Delta_i} + \Delta_i \right) & R^\pi_\mu(n) &\leq c' \left(\sqrt{dn \log(n)} + \sum_{i=1}^d \Delta_i\right)\end{aligned}$$]{} While the problem dependent bound is asymptotically optimal up to constant factors, the problem independent bound is sub-optimal by a factor of $\sqrt{\log(n)}$ with both the MOSS algorithm by [@AB09] and OCUCB by [@Lat15-ucb] matching the lower bound of $\Omega(\sqrt{dn})$ given by [@ACFS95]. The main difficulty in proving Theorem \[thm:finite\] comes from the fact that the Gittins index is smaller than the upper confidence bound used by UCB. This is especially true as $t$ approaches the horizon when the Gittins index tends towards the empirical mean, while the UCB index actually grows. The solution is (a) to use very refined concentration guarantees and (b) to show that the Gittins strategy chooses near-optimal arms sufficiently often for $t \leq n/2$, ensuring that the empirical mean of these arms is large enough that bad arms are not chosen too often in the second half. Before the proof we require some additional definitions and lemmas. Assume for the remainder of this section that $\mu_1 \geq \mu_2 \geq \ldots \geq \mu_d$, which is non-restrictive, since if this is not the case, then the arms can simply be re-ordered. Let $F$ be the event that there exists a $t \in {\left\{1,\ldots,n\right\}}$ and $i \in {\left\{1,\ldots, d\right\}}$ such that [$$\begin{aligned} \left|\hat \mu_i(t) - \mu_i\right| \geq \sqrt{\frac{2}{T_i(t)} \log (dn^2)}\,.\end{aligned}$$]{} The index of the $i$th arm in round $t$ is abbreviated to $\gamma_i(t) = \gamma(\hat \mu_i(t-1), T_i(t-1)^{-1}, n - t + 1)$, which means that for rounds $t > d$ \[alg:flat\] is choosing $I_t = {\operatornamewithlimits{arg\,max}}_i \gamma_i(t)$. Define random variable $Z = \mu_1 - \min_{1 \leq t \leq n/2} \gamma_1(t)$, which measures how far below $\mu_1$ the index of the first arm may fall some time in the first $n/2$ rounds. For each $i \in {\left\{1,\ldots, d\right\}}$ define [$$\begin{aligned} \label{def:u} u_i = {\left \lceil {\frac{32}{\Delta_i^2} \log(2dn^2)} \right\rceil}\,.\end{aligned}$$]{} Now we are ready for the lemmas. First is the key concentration inequality that controls the probability that the Gittins index of the optimal arm drops far below the true mean. It is in the proof of \[lem:conc\] that the odd-looking powers of the logarithmic terms in the bounds on the Gittins index are justified. Any higher power would lead to super-logarithmic regret, while a lower power could potentially lead to a sub-optimal trade-off between failure probability and exploration. \[lem:conc\] Let $c > 0$ be a universal constant, and $\Delta > 0$ and $Y_1,Y_2,\ldots$ be a sequence of i.i.d. random variables with $Y_t \sim \mathcal N(0, 1)$ and $S_t = \sum_{s=1}^t Y_s$. Then there exist universal constants $c' > 0$ and $n_0 \in {\mathbb N}$ such that whenever $n \geq n_0$. [$$\begin{aligned} &{\mathbb{P}\left\{\exists t : S_t \geq t\Delta + \max{\left\{0,\,\, \sqrt{2t \log\left(\frac{cn}{2\log_{+}^{\frac{3}{2}}(n/2)}\right)}\right\}}\right\}} \leq c' \cdot \frac{\log(n)}{n\Delta^2} \\ &{\mathbb{P}\left\{\exists t : S_t \geq t\Delta + \max{\left\{0,\,\, \sqrt{2t \log\left(\frac{cn}{2t \log_{+}^{\frac{1}{2}}\!\!\left(\frac{n}{2t}\right)}\right)}\right\}}\right\}} \leq c' \cdot \frac{{\log_{+}\!}(n\Delta^2)}{n\Delta^2}\,.\end{aligned}$$]{} The proof of Lemma \[lem:conc\] follows by applying a peeling device and may be found in \[sec:lem:conc\]. Note that similar results exist in the literature. For example, by [@AB09; @PR13] and presumably others. What is unique here is that the improved concentration guarantees for Gaussians are also being exploited. Assume that $n \geq n_0$, which is non-restrictive since $n_0$ is a universal constant and \[thm:finite\] holds trivially with $c = n_0$ for $n \leq n_0$. \[lem:Z\] There exists a universal constant $c > 0$ such that for all $\Delta > 0$ we have [$$\begin{aligned} {\mathbb{P}\left\{Z \geq \Delta\right\}} \leq c\cdot \frac{\log(n) + {\log_{+}\!}(n\Delta^2)}{n\Delta^2}\,.\end{aligned}$$]{} Apply \[thm:gittins\] and \[lem:conc\] and the fact that $m = n - t+1 \geq n/2$ for $t \leq n/2$. \[lem:symmetry\] $\displaystyle {\mathbb E}[T_i(n)] \leq n / i$. The result follows from the assumption that $\mu_1 \geq\ldots \geq \mu_d$, the definition of the algorithm, the symmetry of the Gaussian density and because the exploration bonus due to Gittins index is shift invariant (Lemma \[lem:shift\]). \[lem:F\] $\displaystyle {\mathbb{P}\left\{F\right\}} \leq 1/n$. For fixed $T_i(t) = u$ apply the standard tail inequalities for the Gaussian (Lemma \[lem:gaussian\]) to bound $\mathbb{P}\{|\hat \mu_i(t) - \mu_i| \geq \sqrt{2 \log(2dn^2)/u}\} \leq 1/(dn^2)$. The result is completed by applying the union bound over all arms and values of $u \in {\left\{1,\ldots,n\right\}}$. \[lem:half\] If $F$ does not hold and $i$ is an arm such that $Z < \Delta_i / 2$, then $T_i(n/2) \leq u_i$. Let $t \leq n/2$ be some round such that $T_i(t - 1) = u_i$. Then by \[thm:gittins\] and the definitions of $F$ and $u_i$ we have [$$\begin{aligned} \gamma_i(t) &\leq \hat \mu_i(t) + \sqrt{\frac{2}{u_i} \log(n)} \leq \mu_i + \sqrt{\frac{2}{u_i} \log(2dn^2)} + \sqrt{\frac{2}{u_i} \log(n)} \\ &\leq \mu_i + \frac{\Delta_i}{2} = \mu_1 - \frac{\Delta_i}{2} < \mu_1 - Z \leq \gamma_1(t)\,.\end{aligned}$$]{} Therefore $I_t \neq i$ and so $T_i(n/2) \leq u_i$. [\[thm:finite\]]{} The regret may be re-written as [$$\begin{aligned} \label{eq:decomp} R^\pi_\mu(n) = \sum_{i=1}^d \Delta_i {\mathbb E}[T_i(n)]\,.\end{aligned}$$]{} To begin, let us naively bound the regret for nearly-optimal arms. Define a set $C \subset {\left\{1,\ldots, d\right\}}$ by [$$\begin{aligned} C = {\left\{i : 0 < \Delta_i \leq 4\sqrt{\frac{8i}{n} \log(2dn^2)}\right\}} = {\left\{i : 0 < \Delta_i \leq \frac{128i}{n\Delta_i} \log(2dn^2)\right\}}\,.\end{aligned}$$]{} Therefore by Lemma \[lem:symmetry\] we have [$$\begin{aligned} \label{eq:C} \sum_{i \in C} \Delta_i {\mathbb E}[T_i(n)] \leq \sum_{i \in C} \frac{128}{\Delta_i} \log(2dn^2)\,.\end{aligned}$$]{} For arms not in $C$ we need to do some extra work. By Lemma \[lem:F\] [$$\begin{aligned} \sum_{i \notin C} \Delta_i {\mathbb E}[T_i(n)] &\leq {\mathbb{P}\left\{F\right\}} n \Delta_{\max} + {\mathbb E}\left[{\mathds{1}\!\!{\left\{\neg F\right\}}} \sum_{i \notin C} \Delta_i T_i(n)\right] \nonumber \\ &\leq \Delta_{\max} + {\mathbb E}\left[{\mathds{1}\!\!{\left\{\neg F\right\}}} \sum_{i \notin C} \Delta_i T_i(n)\right]\,. \label{eq:de}\end{aligned}$$]{} From now on assume $F$ does not hold while bounding the second term. By Lemma \[lem:half\], if $Z < \Delta_i / 2$, then $T_i(n/2) \leq u_i$. Define disjoint (random) sets $A, B \subseteq {\left\{1,\ldots,d\right\}}$ by [$$\begin{aligned} B = {\left\{i : Z < \Delta_i / 2 \text{ and } \sum_{j \geq i} u_j \leq \frac{n}{4}\right\}} \quad \text{and} \quad A = {\left\{1,\ldots,d\right\}} - B\,.\end{aligned}$$]{} The set $A$ is non-empty because $u_1 = \infty$, which implies that $1 \in A$. Let $i = \max A$, which satisfies either either $Z \geq \Delta_i / 2$ or $n \leq 4\sum_{j \geq i} u_j$. Therefore [$$\begin{aligned} \sum_{k \in A} \Delta_k T_k(n) &\leq n \Delta_i \leq \max{\left\{2nZ {\mathds{1}\!\!{\left\{\Delta_{\min}/2 \leq Z\right\}}}, 4\Delta_i \sum_{j \geq i} u_j\right\}} \nonumber \\ &\leq 2nZ {\mathds{1}\!\!{\left\{\Delta_{\min}/2 \leq Z\right\}}} + 4\sum_{j : \Delta_j > 0} \Delta_j u_j\,. \label{eq:A}\end{aligned}$$]{} The next step is to show that there exists an arm in $A$ that is both nearly optimal and has been chosen sufficiently often that its empirical estimate is reasonably accurate. This arm can then be used to show that bad arms are not chosen too often. From the definition of $B$ and because $F$ does not hold we have $\sum_{i \in A} T_i(n/2) = n/2 - \sum_{i \in B} T_i(n/2) \geq n/2 - \sum_{i \in B} u_i \geq n/4$. Therefore there exists an $i \in A$ such that $T_i(n/2) \geq n/(4|A|)$. Suppose $j \notin A \cup C$ and $\Delta_j \geq 4\Delta_i$ and $T_j(t) = u_j$. Then [$$\begin{aligned} \gamma_j(t) &\leq \mu_j + \frac{\Delta_j}{2} = \mu_i - \frac{\Delta_j}{2} + \Delta_i \leq \mu_i - \frac{\Delta_j}{4} < \mu_i - \sqrt{\frac{8j}{n} \log(dn^2)} \nonumber \\ &\leq \mu_i - \sqrt{\frac{8|A|}{n} \log(dn^2)} \leq \hat \mu_i(t) \leq \gamma_i(t)\,. \label{eq:A2}\end{aligned}$$]{} Therefore $I_t \neq j$ and so $T_j(n) \leq u_j$. Now we consider arms with $\Delta_j < 4\Delta_i$. By the same reasoning as in \[eq:A\] we have [$$\begin{aligned} \sum_{j : \Delta_j < 4\Delta_i} \Delta_j T_j(n) \leq 4n\Delta_i \leq 8nZ {\mathds{1}\!\!{\left\{\Delta_{\min}/2 \leq Z\right\}}} + 16 \sum_{j : \Delta_j > 0} \Delta_j u_j\,.\end{aligned}$$]{} Finally we have done enough to bound the regret due to arms not in $C$. By the above display and \[eq:A\] and the sentence after \[eq:A2\] we have [$$\begin{aligned} \sum_{i \notin C} {\mathds{1}\!\!{\left\{\neg F\right\}}} \Delta_i T_i(n) \leq 10n Z {\mathds{1}\!\!{\left\{\Delta_{\min}/2 \leq Z\right\}}} + 20 \sum_{j : \Delta_j > 0} \Delta_j u_j + \sum_{j \notin C} \Delta_j u_j\,.\end{aligned}$$]{} Combining this with \[eq:C\] and the regret decomposition \[eq:decomp\] and \[eq:de\] we have [$$\begin{aligned} \label{eq:rnear} R^\pi_\mu(n) \leq \Delta_{\max} + \sum_{j \in C} \frac{128}{\Delta_j} \log(dn^2) + 21 \sum_{j : \Delta_j > 0} \Delta_j u_j + 10n{\mathbb E}\left[Z {\mathds{1}\!\!{\left\{\Delta_{\min}/2 \leq Z\right\}}}\right]\,.\end{aligned}$$]{} From Lemma \[lem:Z\] there exists a universal constant $c'' > 0$ such that [$$\begin{aligned} {\mathbb E}Z {\mathds{1}\!\!{\left\{\Delta_{\min} / 2 \leq Z\right\}}} &\leq \int^\infty_{\Delta_{\min}/2} {\mathbb{P}\left\{Z \geq z\right\}} dz + \frac{\Delta_{\min}}{2} {\mathbb{P}\left\{Z \geq \frac{\Delta_{\min}}{2}\right\}} \\ &\leq \frac{c'' (\log(n) + {\log_{+}\!}(n\Delta_{\min}^2))}{n\Delta_{\min}}\,.\end{aligned}$$]{} Substituting into \[eq:rnear\] and inserting the definition of $u_i$ and naively simplifying completes the proof of the problem dependent regret bound. To prove the second bound in \[thm:finite\] it suffices to note that the total regret due to arms with $\Delta_i \leq \sqrt{d/n \log(n)}$ is at most $\sqrt{nd \log(n)}$. The previous results relied heavily on the choice of a flat prior. For other (Gaussian) priors it is still possible to prove regret guarantees, but now with a dependence on the prior. For the sake of simplicity I switch to asymptotic analysis and show that any negative effects of a poorly chosen prior wash out for sufficiently large horizons. Let $\nu_i$ and $\sigma_i^2$ be the prior mean and variance for arm $i$ and let $\nu_i(t)$ and $\sigma_i^2(t)$ denote the posterior mean and variance at the end of round $t$. A simple computation shows that [$$\begin{aligned} \nu_i(t) &= \left(\frac{\nu_i}{\sigma^2_i} + \sum_{s=1}^t {\mathds{1}\!\!{\left\{I_s = i\right\}}} X_s\right) \!\!\!\Bigg/ \!\!\! \left(\frac{1}{\sigma^2_i} + T_i(t)\right) & \label{eq:update} \sigma_i^2(t) &= \left(\frac{1}{\sigma^2_i} + T_i(t)\right)^{-1}\,.\end{aligned}$$]{} \[eq:nonflat\] I\_t = \_i (\_i(t-1), \_i\^2(t-1), n - t + 1). && \[thm:asymptotic\] Assume that $\sigma_i^2 > 0$ for all $i$. Let $\pi_n$ be the strategy given in \[eq:nonflat\], then there exists a universal $c > 0$ such that for $\mu \in {\mathbb R}^d$, $\limsup_{n\to\infty} R^{\pi_n}_\mu(n) / \log n \leq c \sum_{i : \Delta_i > 0} \Delta_i^{-1}$. The proof may be found in \[sec:thm:asymptotic\]. It should be emphasised that the limit in \[thm:asymptotic\] is taken over a sequence of strategies and an increasing horizon. This is in contrast to similar results for UCB where the strategy is fixed and only the horizon is increasing. Discussion {#sec:conc} ========== I have shown that the Gittins strategy enjoys finite-time regret guarantees, which explicates the excellent practical performance. The proof relies on developing tight bounds on the index, which are asymptotically exact. If the prior is mis-specified, then the resulting Gittins index strategy is order-optimal asymptotically, but its finite-time regret may be significantly worse. Experimental results show the Gittins strategy with a flat improper Gaussian prior is never much worse and often much better than the best frequentist algorithms (please see \[sec:exp\]). The index is non-trivial to compute, but reasonable accuracy is possible for horizons of $O(10^4)$. For larger horizons I propose a new index inspired by the theoretical results that is efficient and competitive with the state-of-the-art, at least in the worst case regime for which experiments were performed. There are a variety of open problems some of which are described below. The Gaussian noise/prior was chosen for its simplicity, but bounded or Bernoulli rewards are also interesting. In the latter, the Gittins index can be computed by using a Beta prior and dynamic programming. Asymptotic approximations by [@BK97] suggest that one might expect to derive the KL-UCB algorithm in this manner [@CGMMS13], but finite-time bounds would be required to obtain regret guarantees. Many of the concentration inequalities used in the Gaussian case have information-theoretic analogues, so in principle I expect that substituting these results into the current technique should lead to good results [@Gar13]. For non-flat priors I only gave asymptotic bounds on the regret, showing that the Gittins index strategy will eventually recover from even the most poorly mis-specified prior. Of course it would be nice to fully understand how long the algorithms takes to recover by analysing its finite-time regret. Some preliminary work has been done on this question in a simplified setting for Thompson sampling by [@LL15]. Unfortunately the results will necessarily be quite negative. If a non-flat prior is chosen in such a way that the resulting algorithm achieves unusually small regret with respect to a particular arm, then the regret it incurs on the remaining arms must be significantly larger [@Lat15-unfair]. In short, there is a large price to pay for favouring one arm over another and the Bayesian algorithm cannot save you. This is in contrast to predictive settings where a poorly chosen prior is quickly washed away by data. \[thm:asymptotic\] shows that the Gittins index strategy is eventually *order-optimal* for any choice of prior, but the leading constant does not match optimal rate. The reason is that the failure probability for which the algorithm suffers linear regret appears to be $O(\log(n)/n)$ and so this term is not asymptotically insignificant. In contrast, the UCB confidence level is chosen such that the failure probability is $O(1/n)$, which is insignificant for large $n$. Recent bandit algorithms including MOSS [@AB09] and OCUCB [@Lat15-ucb] exploit the knowledge of the horizon. The Gittins strategy also exploits this knowledge, and it is not clear how it could be defined without a horizon (the index tends to infinity as $n$ increases). The most natural approach would be to choose a prior on the unknown horizon, but what prior should you choose and is there hope to compute the index in that case?[^2] You can also ask what is the benefit of knowing the horizon? The exploration bonus of the Gittins strategy (and MOSS and OCUCB) tend to zero as the horizon approaches, which makes these algorithms more aggressive than anytime algorithms such as UCB and Thompson sampling and improves practical performance. For finite-armed bandits and simple priors the index may be approximated in polynomial time, with practical computation possible to horizons of $\sim\!\!\! 10^4$. It is natural to ask whether or not the computation techniques can be improved to compute the index for larger horizons. I am also curious to know if the classical results can be extended to more complicated settings such as linear bandits or partial monitoring. This has already been done to some extent (eg., restless bandits), but there are many open problems. The recent book by [@GGW11] is a broad reference for existing extensions. Acknowledgements {#acknowledgements .unnumbered} ---------------- My thanks to Marcus Hutter for several useful suggestions and to Dan Russo for pointing out that the finite-horizon Gittins strategy is not generally Bayesian optimal. The experimental component of this research was enabled by the support provided by WestGrid ([www.westgrid.ca](www.westgrid.ca)) and Compute Canada ([www.computecanada.ca](www.computecanada.ca)). Comparison to Bayes {#sec:bayes} =================== As remarked in the introduction, it turns out that both geometric discounting and an infinite horizon are crucial for the interchange argument used in all proofs of the Gittins index theorem and indeed the result is known to be false in general for non-geometric discounting as shown by [@BF85]. Thus the observation that the Gittins index strategy is not Bayesian optimal in the setting considered here should not come as a surprise, but is included for completeness. Let $n = 2$ and $\nu_1 = 0$ and $\sigma^2_1 = 1$ and $\sigma^2_2 = 1/2$. Then the Gittins indices are [$$\begin{aligned} \gamma(\nu_1, \sigma^2_1, 2) &\approx 0.195183 & \gamma(\nu_2, \sigma^2_2, 2) &\approx \nu_2 + 0.112689\,.\end{aligned}$$]{} Therefore the strategy based on Gittins index will choose the second action if $\nu_2 \gtrsim 0.082494$. Computing the Bayesian value of each choice is possible analytically [$$\begin{aligned} \sup_\pi {\mathbb E}\left[\sum_{t=1}^n X_t \Bigg| I_1 = 1\right] &= \int^\infty_{-\infty} \max{\left\{\nu_2, \delta\right\}} \frac{1}{\sqrt{\pi}} \exp\left(-\delta^2 \right) d\delta \\ &= \frac{\exp\left(-\nu_2^2\right) }{2\sqrt{\pi}} + \frac{\nu_2 + \nu_2 {\operatorname{Erf}}(\nu_2)}{2}\,. \\ \sup_\pi {\mathbb E}\left[\sum_{t=1}^n X_t \Bigg| I_1 = 2\right] &= \nu_2 + \int^\infty_{-\infty} \max{\left\{0, \nu_2 + \delta\right\}} \frac{1}{\sqrt{\pi/3}} \exp\left(-3\delta^2\right) d\delta \\ &= \nu_2 + \frac{\exp\left(-3\nu_2^2\right)}{\sqrt{2\pi}} + \frac{\nu_2 + \nu_2 {\operatorname{Erf}}\left(\sqrt{3} \nu_2\right)}{2}\,.\end{aligned}$$]{} Solving leads to the conclusion that $I_2 = 2$ is optimal only if $\nu_2 \gtrsim 0.116462$ and hence the Gittins strategy is not Bayesian optimal (it does not minimise \[eq:bayes\]). Despite this, the following result shows that the regret analysis for the Gittins index strategy can also be applied to the intractable fully Bayesian algorithm when the number of arms is $d = 2$. The idea is to show that the Bayesian algorithm will never choose an arm for which a UCB-like upper confidence bound is smaller than the largest Gittins index. Then the analysis in Section \[sec:finite\] may be repeated to show that that Theorems \[thm:finite\] and \[thm:asymptotic\] also hold for the Bayesian algorithm when $d = 2$. \[thm:bayes-arm\] Assume $d = 2$ and the horizon is $n$. Let $Q$ be the multivariate Gaussian prior measure with mean $\nu \in {\mathbb R}^d$ and covariance matrix $\Sigma = {\operatorname{diag}}(\sigma^2)$. If the Bayesian optimal action is $I_1 = 1$, then [$$\begin{aligned} \nu_1 + \sqrt{2c\sigma^2_1 \log n} \geq \gamma(\nu_2, \sigma^2_2, n)\,,\end{aligned}$$]{} where $c > 0$ is a universal constant. The proof of \[thm:bayes-arm\] is surprisingly tricky and may be found in \[sec:thm:bayes-arm\]. Empirically the behaviour of the Bayesian algorithm and the Gittins index strategy is almost indistinguishable, at least for two arms and small horizons (see \[sec:exp\]). Computing the Gittins Index {#sec:compute} =========================== I briefly describe a method of computing the Gittins index in the Gaussian case. A variety of authors have proposed sophisticated methods for computing the Gittins index, both in the discounted and finite horizon settings [@Nin11; @CM13 and references therein]. The noise model used here is Gaussian and so continuous, which seems to make prior work inapplicable. Fortunately the Gaussian model is rather special, mostly due to the shift invariance (Lemma \[lem:shift\]), which can be exploited to compute and store the index quite efficiently. Let $\nu \in {\mathbb R}$ and $\sigma^2 > 0$ be the prior mean and variance respectively and $m$ be the number of rounds remaining. For $t \in {\left\{1,\ldots,m\right\}}$ define independent random variables [$$\begin{aligned} \eta_t \sim \mathcal N\left(0, \frac{\sigma^2}{1 + (t-1) \sigma^2} \cdot \frac{\sigma^2}{1 + t\sigma^2}\right)\,.\end{aligned}$$]{} Let $\nu_1 = \nu$ and $\nu_t = \nu_{t-1} + \eta_{t-1}$ for $t > 1$. Then [$$\begin{aligned} \label{eq:git-alt} \gamma(\nu, \sigma^2, m) &= \max{\left\{\gamma : \sup_{1\leq \tau\leq m} {\mathbb E}\left[\sum_{t=1}^\tau (\nu_t - \gamma)\right] = 0\right\}} \end{aligned}$$]{} where the stopping time is with respect to the filtration generated by $\eta_1,\ldots,\eta_m$. \[eq:git-alt\] is the Bayesian view of \[eq:gittins\] with the integral over $\mu$ in that equation incorporated into the posterior means. As in the proof of \[thm:gittins\], let $\nu$ be such that $\gamma(\nu, \sigma^2, m) = 0$. Then [$$\begin{aligned} 0 = \gamma(\nu, \sigma^2, m) = \sup_{1 \leq \tau \leq m} {\mathbb E}\left[\sum_{t=1}^\tau \nu_t\right]\,.\end{aligned}$$]{} The optimal stopping problem above can be solved by finding the root of the following Bellman equation for $t = m$. [$$\begin{aligned} \label{eq:bellman} {\mathcal V}_t(x, \sigma^2) = \begin{cases} x + {\mathbb E}_{\eta \sim \mathcal N(0, \sigma^4/(1 + \sigma^2))} \left[\max{\left\{0, {\mathcal V}_{t-1}\left(x + \eta, \frac{\sigma^2}{1+\sigma^2}\right)\right\}}\right] & \text{if } t \geq 1 \\ 0 & \text{otherwise}\,. \end{cases}\end{aligned}$$]{} Then by Lemma \[lem:shift\] the Gittins index satisfies [$$\begin{aligned} \label{eq:root} \gamma(0, \sigma^2, m) = \gamma(\nu, \sigma^2, m) - \nu = -\nu = -\max{\left\{x : {\mathcal V}_m(x, \sigma^2) = 0\right\}}\,.\end{aligned}$$]{} Now \[eq:bellman\] is a Bellman equation and conveniently there is a whole field devoted to solving such equations [@BT95 and references therein]. An efficient algorithm that computes the Gittins index to arbitrary precision by solving \[eq:bellman\] using backwards induction and approximating $\max{\left\{0, {\mathcal V}_t(x, \sigma^2)\right\}}$ using quadratic splines may be found at . The motivation for choosing the quadratic is because the expectation in \[eq:bellman\] can be computed explicitly in terms of the error function and because ${\mathcal V}_t(x, \sigma^2)$ is convex in $x$, so the quadratic leads to a relatively good fit with only a few splines. The convexity of ${\mathcal V}$ also means that \[eq:root\] can be computed efficiently from a sufficiently good approximation of ${\mathcal V}_m$. Finally, in order to implement \[alg:flat\] up to a horizon of $n$ we may need $\gamma(\nu, 1/T, m) = \nu + \gamma(0,1/T, m)$ for all $T$ and $m$ satisfying $T + m \leq n$. This can computed by solving $n$ copies of \[eq:bellman\] with $\sigma^2 = 1$ and $m \in {\left\{1,\ldots,n-1\right\}}$. The total running time is $O(n^2 N)$ where $N$ is the maximum number of splines required for sufficient accuracy. The results can be stored in a lookup table of size $O(n^2)$, which makes the actual simulation of \[alg:flat\] extremely fast. Computation time for $n = 10^4$ was approximately 17 hours using 8 cores of a Core-i7 machine. Experiments {#sec:exp} =========== I compare the Gittins strategy given in \[alg:flat\] with UCB, OCUCB [@Lat15-ucb] and Thompson sampling (TS) with a flat Gaussian prior [@Tho33; @AG12]. Due to the computational difficulties in calculating the Gittins index we are limited to modest horizons. For longer horizons the Gittins index may be approximation by the lower bound in \[thm:gittins\]. I also compare the Gittins index strategy to the Bayesian optimal strategy that can be computed with reasonable accuracy for the two-armed case and a horizon of $n = 2000$. Error bars are omitted from all plots because they are too small to see. All code is available from . [ll]{} & [**Index**]{} \ UCB & $\hat \mu_i(t-1) + \sqrt{\frac{2}{T_i(t-1)} \log t}$\ OCUCB & $\hat \mu_i(t-1) + \sqrt{\frac{3}{T_i(t-1)} \log \left(\frac{2 n}{t}\right)}$\ TS & $\sim \mathcal N(\hat \mu_i(t-1), (T_i(t-1)+1)^{-1})$\ I start with the worst case regret for horizons $n \in {\left\{10^3, 10^4\right\}}$. In all experiments the first arm is optimal and has mean $\mu_1 = 0$. The remaining arms have $\mu_i = -\Delta$. In \[fig:worst\] I plot the expected regret of various algorithms with $d \in {\left\{2,5,10\right\}}$ and varying $\Delta$. The results demonstrate that the Gittins strategy significantly outperforms both UCB and Thompson sampling, and is a modest improvement on OCUCB in most cases. +\[\] table\[x index=0,y index=1\] ; +\[\] table\[x index=0,y index=3\] ; +\[\] table\[x index=0,y index=2\] ; +\[\] table\[x index=0,y index=4\] ; +\[\] table\[x index=0,y index=1\] ; +\[\] table\[x index=0,y index=3\] ; +\[\] table\[x index=0,y index=2\] ; +\[\] table\[x index=0,y index=4\] ; +\[\] table\[x index=0,y index=1\] ; +\[\] table\[x index=0,y index=3\] ; +\[\] table\[x index=0,y index=2\] ; +\[\] table\[x index=0,y index=4\] ; +\[\] table\[x index=0,y index=1\] ; +\[\] table\[x index=0,y index=3\] ; +\[\] table\[x index=0,y index=2\] ; +\[\] table\[x index=0,y index=4\] ; +\[\] table\[x index=0,y index=1\] ; +\[\] table\[x index=0,y index=3\] ; +\[\] table\[x index=0,y index=2\] ; +\[\] table\[x index=0,y index=4\] ; +\[\] table\[x index=0,y index=1\] ; +\[\] table\[x index=0,y index=3\] ; +\[\] table\[x index=0,y index=2\] ; +\[\] table\[x index=0,y index=4\] ; Current methods for computing the Gittins index are only practical for horizons up to $O(10^4)$. For longer horizons it seems worthwhile to find a closed form approximation. Taking inspiration from \[thm:gittins\], define $\tilde \gamma(\nu, \sigma^2, m) \approx \gamma(\nu, \sigma^2, m)$ by [$$\begin{aligned} \tilde \gamma(\nu, \sigma^2, m) &= \nu + \sqrt{2\sigma^2 \log \beta(\sigma^2, m)} \qquad \text{where} \qquad \\ \beta(\sigma^2, m) &= \max{\left\{1,\,\, \frac{1}{4} \min{\left\{\frac{m}{\log^{\frac{3}{2}}(m)},\,\, \frac{m\sigma^2}{\log^{\frac{1}{2}}(m\sigma^2)}\right\}}\right\}}\,.\end{aligned}$$]{} This is exactly the lower bound in \[thm:gittins\], but with the leading constant chosen (empirically) to be $1/4$. I compare the indices in two key regimes. First fix $\nu = 0$ and $\sigma^2 = 1$ and vary the horizon. In the second regime the horizon is fixed to $m = 10^3$ and $\sigma^2 = 1/T$ is varied. The results (\[fig:approx-index\]) suggest a relatively good fit. +\[solid\] table\[x index=0,y index=2\] ; ; +\[domain=100:10000\] [sqrt(2 \* min(ln(x / 4 / ln(x)\^(1/2)), ln(x / 4 / ln(x)\^(3/2))))]{}; +\[solid\] table\[x index=1,y index=2\] ; ; +\[domain=1:100\] [sqrt(2 / x \* min(ln(1000 / 4 / x / ln(1000/x)\^(1/2)), ln(1000 / 4 / ln(1000)\^(3/2))))]{}; ; The approximated index strategy is reasonably competitive with the Gittins strategy (\[fig:approx\]). For longer horizons the Gittins index cannot be computed in reasonable time, but the approximation can be compared to other efficient algorithms such as OCUCB, Thompson sampling and UCB (\[fig:approx\]). The approximated version Gittins index is performing well compared to Thompson sampling and UCB, and is only marginally worse than OCUCB. As an added bonus, by cloning the proofs of Theorems \[thm:finite\] and \[thm:asymptotic\] it can be shown that the approximate index enjoys the same regret guarantees as the Gittins strategy/UCB. +\[black,solid\] table\[x index=0,y index=1\] ; ; +\[black,dashed\] table\[x index=0,y index=2\] ; ; +\[\] table\[x index=0,y index=2\] ; ; +\[\] table\[x index=0,y index=4\] ; ; +\[\] table\[x index=0,y index=3\] ; ; +\[\] table\[x index=0,y index=1\] ; ; Unfortunately the Bayesian optimal strategy is not typically an index strategy. The two natural exceptions occur when the rewards are discounted geometrically or when there are only two arms and the return of the second arm is known. Despite this, the Bayesian optimal strategy can be computed for small horizons and a few arms using a similar method as the Gittins index. Below I compare the Gittins index against the Bayesian optimal and OCUCB for $n = 2000$ and $d = 2$ and $\mu_1 = 0$ and $\mu_2 = -\Delta$. The results show the similar behaviour of the Gittins index strategy and the Bayesian optimal strategy, while OCUCB is making a different trade-off (better minimax regret at the cost of worse long-term regret). +\[blue,dashed\] table\[x index=0,y index=1\] ; ; +\[black,solid\] table\[x index=0,y index=2\] ; ; +\[red,dashed\] table\[x index=0,y index=3\] ; ; Proof of Lemma \[lem:bandit\] {#sec:lem:bandit} ============================= For parts (1) and (2) we have $\theta \in (-\infty, 0)$. Define $t_\theta \in {\left\{1,2,\ldots\right\}}$ by [$$\begin{aligned} t_\theta &= \min{\left\{t \geq \frac{1}{\nu^2} : \sqrt{\frac{4}{t} \log (4t\nu^2)} \leq -\theta/2\right\}}\,. \end{aligned}$$]{} Then [$$\begin{aligned} {\mathbb E}[\tau|\mu= \theta] &= \sum_{t=1}^m {\mathbb{P}\left\{\tau \geq t| \mu = \theta\right\}} \leq t_\theta + \sum_{t=t_\theta+1}^m {\mathbb{P}\left\{\tau \geq t | \mu = \theta\right\}} \\ &\leq t_\theta + \sum_{t=t_\theta}^{m-1} {\mathbb{P}\left\{\hat \mu_t + \sqrt{\frac{4}{t} \log\left(4t\nu^2\right)} > 0\Bigg| \mu = \theta\right\}} \\ &\leq t_\theta + \sum_{t=t_\theta}^{m-1} {\mathbb{P}\left\{\hat \mu_t \geq \theta / 2 | \mu= \theta\right\}} \leq t_\theta + \sum_{t=t_\theta}^\infty \exp\left(-\frac{t \theta^2}{8}\right) \leq t_\theta + \frac{8}{\theta^2}\,.\end{aligned}$$]{} The results follow by naively bounding $t_\theta$. For part 3, assume $\mu \geq 0$ and define $S_t = t \hat \mu_t$ and $t_k = {\left \lceil {1/\nu^2} \right\rceil} 2^k$. By using the peeling device we obtain [$$\begin{aligned} {\mathbb{P}\left\{\tau < m|\mu = \theta\right\}} &= {\mathbb{P}\left\{\exists t \geq \frac{1}{\nu^2} : \hat \mu_t + \sqrt{\frac{4}{t} \log(4t\nu^2)} \leq 0 \Bigg|\mu = \theta\right\}} \\ &\leq \sum_{k=0}^\infty {\mathbb{P}\left\{\exists t \leq t_{k+1} : S_t + \sqrt{2 \cdot t_{k+1} \log (4t_k \nu^2)} \leq 0 \Big|\mu = \theta\right\}} \\ &\leq \sum_{k=0}^\infty \frac{1}{4t_k \nu^2} \leq \sum_{k=0}^\infty \frac{1}{4 \cdot 2^k} = \frac{1}{2}\,.\end{aligned}$$]{} Therefore ${\mathbb E}[\tau|\mu = \theta] \geq m/2$. Concentration of Gaussian Random Variables {#sec:gaussian} ========================================== These lemmas are not new, but are collected here with proofs for the sake of completeness. \[lem:gaussian\] Let $\sigma \geq 0$ and $\nu < 0$ and $X \sim \mathcal N(\nu, \sigma^2)$, then for $x \geq \nu$ 1. $\displaystyle {\mathbb{P}\left\{X \geq x\right\}} \leq \exp\left(-\frac{(\nu - x)^2}{2\sigma^2}\right)$. 2. $\displaystyle {\mathbb E}[X {\mathds{1}\!\!{\left\{X \leq 0\right\}}}] \geq \nu - \sigma / \sqrt{2\pi}$. The first is well known. The second follows since [$$\begin{aligned} {\mathbb E}[X {\mathds{1}\!\!{\left\{X \leq 0\right\}}}] &= \int^0_{-\infty} \frac{x}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\nu - x)^2}{2\sigma^2}\right) dx \\ &= \frac{\nu}{2} {\operatorname{Erfc}}\left(\frac{\nu}{\sqrt{2\sigma^2}}\right) - \sqrt{\frac{\sigma^2}{2\pi}}\exp\left(-\frac{\nu^2}{2\sigma^2}\right) \\ &\geq \nu - \sigma / \sqrt{2\pi} \end{aligned}$$]{} as required. \[lem:maximal\] Let $\delta \in (0,1)$ and $\Delta \geq 0$ and $X_1,\ldots,X_n$ be a sequence of i.i.d. random variables with $X_t \sim \mathcal N(0, 1)$ and $S_t = \sum_{s=1}^t X_s$. Then [$$\begin{aligned} {\mathbb{P}\left\{\exists t \leq n : S_t \geq n\Delta + \sqrt{2 n \log\left(\frac{1}{\delta}\right)}\right\}} \leq \frac{\delta}{\sqrt{\pi \log(1/\delta)}} \exp\left(-\frac{n\Delta^2}{2}\right)\,.\end{aligned}$$]{} Let $\epsilon > 0$. By the reflection principle we have ${\mathbb{P}\left\{\exists t \leq n : S_t \geq \epsilon\right\}} \leq 2 {\mathbb{P}\left\{S_n \geq \epsilon\right\}}$ and [$$\begin{aligned} 2{\mathbb{P}\left\{S_n \geq \epsilon\right\}} &= 2\int^\infty_\epsilon \frac{1}{\sqrt{2\pi n}} \exp\left(-\frac{x^2}{2n}\right) dx \\ &\leq \frac{2}{\epsilon} \int^\infty_\epsilon \frac{x}{\sqrt{2\pi n}} \exp\left(-\frac{x^2}{2n}\right) dx \\ &= \frac{1}{\epsilon} \sqrt{\frac{2n}{\pi}} \exp\left(-\frac{\epsilon^2}{2n}\right)\,.\end{aligned}$$]{} There result is completed by substituting $\epsilon = n\Delta + \sqrt{2n \log(1/\delta)}$ and naive simplification. Proof of Lemma \[lem:conc\] {#sec:lem:conc} =========================== For the first part we can apply Lemma \[lem:maximal\] and the union bound. [$$\begin{aligned} &{\mathbb{P}\left\{\exists t \leq n : S_t \geq t\Delta + \sqrt{2t \log\left(\frac{cn}{2\log_+^{\frac{3}{2}}(n/2)}\right)}\right\}} \\ &\qquad\qquad\leq \sum_{t=1}^{n} {\mathbb{P}\left\{S_t \geq t\Delta + \sqrt{2t \log\left(\frac{cn}{2\log_+^{\frac{3}{2}}(n/2)}\right)}\right\}} \\ &\qquad\qquad\leq \frac{2\log_+^{\frac{3}{2}}(n/2)}{cn \sqrt{\pi \log\left(\frac{cn}{2\log_+^{\frac{3}{2}}(n/2)}\right)}} \sum_{t=1}^\infty \exp\left(-\frac{t\Delta^2}{2}\right) \\ &\qquad\qquad\leq \frac{\log(n)}{n\Delta^2} \cdot \frac{4\log_+^{\frac{3}{2}}(n/2)}{c \log(n) \sqrt{\pi \log\left(\frac{cn}{2\log_+^{\frac{3}{2}}(n/2)}\right)}}\,. \end{aligned}$$]{} The result follows by noting that [$$\begin{aligned} \lim_{n\to\infty} \frac{4\log_+^{\frac{3}{2}}(n/2)}{c \log(n) \sqrt{\pi \log\left(\frac{cn}{2\log_+^{\frac{3}{2}}(n/2)}\right)}} = \frac{4}{c \sqrt{\pi}}\end{aligned}$$]{} and then choosing $n_0$ sufficiently large. Moving to the second part. Let $\eta \in (0, 1]$ be a constant to be chosen later and $t_k = (1 + \eta)^k$ and [$$\begin{aligned} \alpha_k &= \frac{2\log_+^{\frac{1}{2}}\left(\frac{n}{2t_k}\right)}{c} & K &= \max{\left\{k : t_k \leq 2en \quad \text{and} \quad \sqrt{\frac{n}{t_k}} \geq \alpha_k \geq 1\right\}}\,.\end{aligned}$$]{} An easy computation shows that there exists a universal constant $c' > 0$ such that $t_K \geq c' n$. Then [$$\begin{aligned} &{\mathbb{P}\left\{\exists t \leq n : S_t \geq t\Delta + \max{\left\{0, \sqrt{2t \log\left(\frac{c n}{2t \log_+^{\frac{1}{2}}(n/(2t))}\right)}\right\}}\right\}} \nonumber \\ &\leq \sum_{k=0}^K {\mathbb{P}\left\{\exists t \leq t_k : S_t \geq \frac{t_k \Delta}{(1 + \eta)} + \sqrt{\frac{2t_k}{1+\eta} \log\left(\frac{n}{t_k\alpha_k}\right)}\right\}} + {\mathbb{P}\left\{\exists t_K \leq t \leq n : S_t \geq t\Delta\right\}}\,. \label{eq:conc1}\end{aligned}$$]{} The second term in \[eq:conc1\] is easily bounded by a peeling argument. [$$\begin{aligned} {\mathbb{P}\left\{\exists t_K \leq t \leq n: S_t \geq t\Delta\right\}} &\leq \sum_{i=1}^\infty {\mathbb{P}\left\{\exists t \leq (i+1)t_K : S_t \geq i t_K \Delta\right\}} \nonumber \\ &\leq \sum_{i=1}^\infty \exp\left(-\frac{i^2 t_K^2 \Delta^2}{2(i+1)t_K}\right) \nonumber \\ &\leq \frac{c''}{n\Delta^2}\,, \label{eq:conc2}\end{aligned}$$]{} where $c'' > 0$ is some sufficiently large universal constant and we have used the fact that $t_K \geq c'n$ and naive algebra. The first term in \[eq:conc1\] is slightly more involved. Using peeling and Lemmas \[lem:maximal\] and \[lem:tech1\] we have universal constants $c'$ and $c''$ such that [$$\begin{aligned} &\sum_{k=0}^K {\mathbb{P}\left\{\exists t \leq t_k : S_t \geq \frac{t_k \Delta}{(1 + \eta)} + \sqrt{\frac{2t_k}{1+\eta} \log\left(\frac{n}{t_k\alpha_k}\right)}\right\}} \\ &{\stackrel{(a)}}\leq \sum_{k=0}^K \sqrt{\frac{1 + \eta}{\pi \log\left(\frac{n}{t_k \alpha_k}\right)}} \left(\frac{t_k \alpha_k}{n}\right)^{1/(1+\eta)} \exp\left(-\frac{t_k \Delta^2}{2(1 +\eta)^2}\right) \\ &{\stackrel{(b)}}\leq \sum_{k=0}^K \alpha_k \sqrt{\frac{2}{\pi \log\left(\frac{n}{t_k\alpha_k}\right)}} \left(\frac{t_k}{n}\right)^{1/(1+\eta)} \exp\left(-\frac{t_k \Delta^2}{2(1 +\eta)^2}\right) \\ &{\stackrel{(c)}}\leq \sum_{k=0}^K \frac{2\log_+^{\frac{1}{2}}\left(\frac{n}{t_k}\right)}{c} \sqrt{\frac{4}{\pi \log\left(\frac{n}{t_k}\right)}} \left(\frac{t_k}{n}\right)^{1/(1+\eta)} \exp\left(-\frac{t_k \Delta^2}{2(1 +\eta)^2}\right) \\ &{\stackrel{(d)}}= \frac{4}{c\sqrt{\pi}} \sum_{k=0}^K \left(\frac{t_k}{n}\right)^{1/(1+\eta)} \exp\left(-\frac{t_k \Delta^2}{2(1 +\eta)^2}\right) \\ &{\stackrel{(e)}}\leq \frac{3\cdot 4}{c\eta} \left(\frac{2(1 + \eta)}{n\Delta^2}\right)^{1/(1 + \eta)} \\ &{\stackrel{(f)}}\leq \frac{4 \cdot 3\cdot 4}{c\eta} \left(\frac{1}{n\Delta^2}\right)^{1/(1 + \eta)}\,.\end{aligned}$$]{} where (a) follows from Lemma \[lem:maximal\], (b) since $\eta \in (0,1]$ and $\alpha_k \geq 1$, (c) by substituting the value of $\alpha_k$ and because $\sqrt{n/t_k} \geq \alpha_k$, (d) and (f) are trivial while (e) follows from Lemma \[lem:tech1\]. Then choose [$$\begin{aligned} \eta = \frac{1}{{\log_{+}\!}(n\Delta^2)} \in (0, 1]\,,\end{aligned}$$]{} which satisfies [$$\begin{aligned} \frac{1}{\eta} \left(\frac{1}{n\Delta^2}\right)^{1/(1+\eta)} \leq \frac{e}{n\Delta^2} {\log_{+}\!}(n\Delta^2)\,. \end{aligned}$$]{} The result by combining the above reasoning with \[eq:conc2\] and substituting into \[eq:conc1\]. Proof of \[thm:asymptotic\] {#sec:thm:asymptotic} =========================== As before I assume for convenience that $\mu_1 \geq \mu_2 \geq \ldots \geq \mu_d$. Let $n$ be fixed and abbreviate $\gamma_i(t) = \gamma(\nu_i(t-1), \sigma^2_i(t-1), n - t + 1)$. Like the proof of \[thm:finite\] there are two main difficulties. First, showing for suboptimal arms $i$ that $\gamma_i(t) < \mu_i + \Delta_i / 2$ occurs with sufficiently high probability once $T_i(t)$ is sufficiently large. Second, showing that $\gamma_1(t) \geq \mu_1 - \Delta_i / 2$ occurs with sufficiently high probability. Let $F$ be the event that there exists an arm $i$ and round $t \in {\left\{1,\ldots,n\right\}}$ such that [$$\begin{aligned} \left|\hat \mu_i(t) - \mu_i\right| \geq \sqrt{\frac{2}{T_i(t)} \log (2dn^2)}\,.\end{aligned}$$]{} As in the proof of \[thm:finite\] we have ${\mathbb{P}\left\{F\right\}} \leq 1/n$. The regret due to $F$ occurring is negligible, so from now on assume that $F$ does not hold. By \[eq:update\] we have [$$\begin{aligned} \nu_i(t) = \eta_i(t) \nu_i + (1 - \eta_i(t)) \hat \mu_i(t)\,,\end{aligned}$$]{} where [$$\begin{aligned} \eta_i(t) = \frac{1}{1 + \sigma_i^2 T_i(t)} = O\left(\frac{1}{T_i(t)}\right)\,.\end{aligned}$$]{} Therefore for sufficiently large $n$ and $u_i = {\left \lceil {\frac{32}{\Delta^2} \log (dn^2)} \right\rceil}$ and $T_i(t-1) = u_i$ [$$\begin{aligned} \gamma_i(t) &\leq \eta_i(t-1) \nu_i + (1 - \eta_i(t-1)) \hat \mu_i(t-1) + \sqrt{2 \sigma^2_i(t-1) \log (2n)} \\ &\leq \eta_i(t-1) (\nu_i - \mu_i) + \mu_i + \sqrt{\frac{2}{T_i(t-1)} \log (2n)} + \sqrt{\frac{2}{T_i(t-1)} \log (2dn^2)} \\ &\leq \eta_i(t-1) (\nu_i - \mu_i) + \mu_i + \Delta_i / 2 \leq \mu_i + 3\Delta_i / 4\,.\end{aligned}$$]{} For the first arm we have for $t \leq n /2$ that [$$\begin{aligned} \gamma_1(t) \geq \eta_1(t-1) \nu_1 + (1 - \eta_1(t-1)) \hat \mu_1(t-1) + \sqrt{2\sigma_1^2(t-1) \log \beta_t}\,.\end{aligned}$$]{} where by \[thm:gittins\] [$$\begin{aligned} \beta_t \geq c \min{\left\{\frac{n/2}{\log_{+}^{\frac{3}{2}}(n/2)}, \,\, \frac{n\sigma_1^2(t-1)/2}{\log_{+}^{\frac{1}{2}}(n\sigma_1^2(t-1)/2)}\right\}}\,.\end{aligned}$$]{} For any $\Delta > 0$ define $F_\Delta$ to be the event that there exists a $t \leq n/2$ for which [$$\begin{aligned} \hat \mu_1(t-1) + \sqrt{\frac{2}{T_1(t-1)} \log \beta_t} \leq \mu_1 - \Delta\,.\end{aligned}$$]{} By Lemma \[lem:conc\] there is a universal constant $c'>0$ such that [$$\begin{aligned} {\mathbb{P}\left\{F_\Delta\right\}} \leq c'(\log(n)+ {\log_{+}\!}(n\Delta^2))/(n\Delta^2)\end{aligned}$$]{} and when $F_\Delta$ does not occur we have for $t \leq n/2$ that [$$\begin{aligned} \gamma_1(t) &\geq \eta_1(t-1) \nu_1 + (1 - \eta_1(t-1)) \left(\mu_1 - \Delta - \sqrt{\frac{2}{T_1(t-1)} \log \beta_t}\right) \\ &\qquad + \sqrt{2\sigma_1^2(t-1) \log \beta_t} \\ &\geq \mu_1 - \Delta - O\left(\frac{1}{T_1(t-1)}\right) - (1 - \eta_1(t-1)) \sqrt{\frac{2}{T_1(t-1)} \log \beta_t} \\ &\qquad + \sqrt{2\sigma_1^2(t-1) \log \beta_t}\,. \end{aligned}$$]{} Comparing the ratio of the last two terms gives [$$\begin{aligned} \frac{(1 - \eta_1(t-1)) \sqrt{\frac{2}{T_1(t-1)} \log \beta_t}}{\sqrt{2\sigma_1^2(t-1)} \log \beta_t} &= \sqrt{\frac{\sigma_1^2 T_1(t-1)}{1 + \sigma_1^2 T_1(t-1)}} < 1\,.\end{aligned}$$]{} Since $\log \beta_t = \Theta(\log n)$ for $t \leq n/2$, if $F_\Delta$ does not hold and $n$ is sufficiently large, then [$$\begin{aligned} \gamma_1(t) \geq \mu_1 - 2\Delta \qquad \text{for all } t \leq n/2\,.\end{aligned}$$]{} The proof is completed by duplicating the argument given in the proof of \[thm:finite\] and taking the limit as $n$ tends to infinity. Proof of \[thm:bayes-arm\] {#sec:thm:bayes-arm} ========================== \[lem:cgaussian\] Let $X \sim \mathcal N(0, \sigma^2)$ and $A$ be an arbitrary event. Then [$$\begin{aligned} {\mathbb E}{\mathds{1}\!\!{\left\{A\right\}}} X \leq {\mathbb{P}\left\{A\right\}} \sqrt{2\sigma^2 \log \left(\frac{1}{{\mathbb{P}\left\{A\right\}}}\right)}\,.\end{aligned}$$]{} \[lem:brownian\] Let $B_t$ be the standard Brownian motion. Let $\epsilon > 0$ and define stopping time $\tau = \min{\left\{t : B_t = -\epsilon\right\}}$. Then [$$\begin{aligned} {\mathbb{P}\left\{\tau > t\right\}} = \Phi\left(\frac{\epsilon}{\sqrt{t}}\right) - \Phi\left(\frac{-\epsilon}{\sqrt{t}}\right)\,. \end{aligned}$$]{} The proofs of Lemmas \[lem:cgaussian\] and \[lem:brownian\] are omitted, but the former follows by letting $\delta = {\mathbb{P}\left\{A\right\}}$ and noting that the worst-case occurs when $A = {\mathds{1}\!\!{\left\{X \geq \alpha\right\}}}$ where $\alpha$ is such that [$$\begin{aligned} \int^\infty_\alpha \frac{\exp\left(-x^2/(2\sigma^2)\right)}{\sqrt{2\pi\sigma^2}} dx = \delta\,.\end{aligned}$$]{} The second lemma is well known [@Ler86 for example]. For what follows we need an alternative view of the Gittins index. Let $\nu \in {\mathbb R}$ and $1 \geq \sigma^2 \geq 0$. Furthermore define [$$\begin{aligned} \sigma^2_t = \frac{\sigma^2}{1 + (t-1)\sigma^2} \cdot \frac{\sigma^2}{1 + t\sigma^2}\,,\end{aligned}$$]{} which satisfies $\sum_{t=1}^\infty \sigma^2_t = \sigma^2$. I abbreviate $\sigma^2_{\leq t} = \sum_{s=1}^t \sigma^2_t$. Let $\nu_1 = \nu$ and $\nu_{t+1} = \nu_t + \eta_t$ where $\eta_t \sim \mathcal N(0, \sigma^2_t)$. Then the Gittins index is given by [$$\begin{aligned} \gamma(\nu, \sigma^2, n) = \sup_{1 \leq \tau \leq n} \frac{{\mathbb E}\left[\sum_{t=1}^\tau \nu_t\right]}{{\mathbb E}[\tau]}\,,\end{aligned}$$]{} where ${\mathds{1}\!\!{\left\{\tau = t\right\}}}$ is measurable with respect to the $\sigma$-algebra generated by $\nu_1,\nu_2,\ldots,\nu_{t+1}$. This is equivalent to the definition in \[eq:gittins\], but written in terms of the evolution of the posterior mean rather than the observed reward sequences (these can be computed from each other in a deterministic way). \[lem:lower-weak\] There exists a universal constant $1 \geq c > 0$ such that $\displaystyle \gamma(\nu, \sigma^2, n) \geq c \sqrt{2\sigma^2_{\leq n}}$. The proof is left as an exercise (hint: choose a stopping time $\tau = {\left \lceil {n/2} \right\rceil}$ unless $\nu_{{\left \lceil {n/2} \right\rceil}}$ is large enough when $\tau = n$). Recall that $\tau = \tau(\nu, \sigma^2, n)$ is the stopping rule defining the Gittins index for mean $\nu$, variance $\sigma^2$ and horizon $n$. The following lemma shows that $\tau = n$ with reasonable probability and is crucial for the proof of \[thm:bayes-arm\]. \[lem:tau-bound\] Let $\nu \in {\mathbb R}$ and $\sigma^2 > 0$ and let $\tau = \tau(\nu, \sigma^2, n)$. Then there exists a universal constant $c' > 0$ such that [$$\begin{aligned} {\mathbb{P}\left\{\tau = n\right\}} \geq \frac{c'}{n^2 \log(n)}\,.\end{aligned}$$]{} Assume without loss of generality (by translation and Lemma \[lem:shift\]) that $\gamma(\nu, \sigma^2, n) = 0$ and let $\mathcal E = -\nu \geq c\sqrt{2\sigma^2_{\leq n}}$ by Lemma \[lem:lower-weak\]. By definition of the Gittins index we have [$$\begin{aligned} {\mathbb E}\left[\sum_{t=1}^\tau \nu_t\right] = 0\,.\end{aligned}$$]{} Let $\delta_t = {\mathbb{P}\left\{\tau \geq t \text{ and } \nu_t \geq \mathcal E/(2n)\right\}}$ and $\alpha_t = {\mathbb E}[\nu_t|\tau \geq t \text{ and } \nu_t \geq \mathcal E/(2n)]$. By Lemma \[lem:cgaussian\] we have $\alpha_t \leq \sqrt{2\sigma_{\leq t}^2 \log(1/\delta_t)} - \mathcal E \leq \sqrt{2\sigma_{\leq t}^2 \log(1/\delta_t)}$. Therefore [$$\begin{aligned} 0 = {\mathbb E}\left[\sum_{t=1}^\tau \nu_t\right] \leq -\mathcal E/2 + \sum_{t=2}^n \delta_t \alpha_t\,.\end{aligned}$$]{} Therefore there exists a $t$ such that $\delta_t \alpha_t \geq \mathcal E / (2n)$ and so [$$\begin{aligned} \delta_t \sqrt{2\sigma^2_{\leq t} \log\frac{1}{\delta_t}} \geq \delta_t \alpha_t \geq \frac{\mathcal E}{2n} \geq \frac{c \sqrt{2\sigma^2_{\leq n}}}{2n}\,.\end{aligned}$$]{} Straightforward analysis shows that [$$\begin{aligned} \delta_t \geq \frac{c}{2n \log\left(\frac{2n}{c}\right)}\,.\end{aligned}$$]{} Now we apply Lemma \[lem:brownian\]. First note that if $\tau \geq t$ and $\nu_s \geq 0$ for all $s \geq t$, then $\tau = n$. Therefore [$$\begin{aligned} {\mathbb{P}\left\{\tau = n\right\}} &\geq {\mathbb{P}\left\{\tau \geq t \text{ and } \nu_s \geq 0 \text{ for all } s \geq t\right\}} \\ &\geq \frac{c}{2n \log\left(\frac{2n}{c}\right)} \left(\Phi\left(\frac{\mathcal E}{2n \sigma_{\leq n}}\right) - \Phi\left(-\frac{\mathcal E}{2n \sigma_{\leq n}}\right) \right) \\ &\geq \frac{c}{2n \log\left(\frac{2n}{c}\right)} \left(\Phi\left(\frac{c}{2n}\right) - \Phi\left(-\frac{c}{2n}\right) \right) \\ &\geq \frac{c^2}{8n^2 \log\left(\frac{2n}{c}\right)}\end{aligned}$$]{} as required. It is rather clear that the above proof is quite weak. Empirically it appears that ${\mathbb{P}\left\{\tau = n\right\}} = \Omega(1/n)$, but the result above is sufficient for our requirements. [\[thm:bayes-arm\]]{} Let $\pi^*$ be the Bayesian optimal policy, which by assumption chooses $I_1 = 1$. Define $\pi$ be the policy that chooses arm $2$ until $t$ such that $\gamma_2(t) < \gamma_2(1)$ and thereafter follows $\pi^*$ as if no samples had been observed and substituting the observed values for the second arm as it is chosen. Let $\tau = \tau(\nu_2, \sigma^2_2, n)$, then $\pi$ chooses the second time until at least $t = \tau$. For any policy $\pi$ let $T^\pi_i(n)$ be the number of plays of arm $i$ by policy $\pi$, which is a random variable. Let $\nu_{i,s}$ be the posterior mean of the $i$th arm after $s-1$ plays. Then the value of policy $\pi$ is [$$\begin{aligned} V^\pi = {\mathbb E}\left[\sum_{i=1}^d \sum_{s=1}^{T^\pi_i(n)} \nu_{i,s}\right]\,,\end{aligned}$$]{} which is maximised by the Bayesian optimal policy $\pi^*$. Therefore if we abbreviate $T^{\pi^*}_i(n)$ by $T_i(n)$ and $\gamma_2(1)$ by $\gamma$, then [$$\begin{aligned} \nonumber 0 \nonumber &\geq V^\pi - V^{\pi^*} \\ \nonumber &\geq {\mathbb E}\left[{\mathds{1}\!\!{\left\{\tau > T_2(n)\right\}}} \sum_{s=T_1(n) - \tau + T_2(n)}^{T_1(n)} (\gamma - \nu_{1,s})\right] \\ \nonumber &\geq \sum_{s=1}^n {\mathbb E}\left[{\mathds{1}\!\!{\left\{\tau - T_2(n) \geq s\right\}}} (\gamma - \max_{s \leq n} \nu_{1,n})\right] \\ \nonumber &\geq \sum_{s=1}^n {\mathbb{P}\left\{\tau - T_2(n) \geq s\right\}} \left(\gamma - \nu_1 - \sqrt{2\sigma_1^2 \log \left(\frac{1}{{\mathbb{P}\left\{\tau - T_2(n) \geq s\right\}}}\right)}\right) \\ &= \sum_{s=1}^n \delta_s \left(\gamma - \nu_1 - \sqrt{2\sigma_1^2 \log \left(\frac{1}{\delta_s}\right)}\right)\,, \label{eq:bayes-arm} \end{aligned}$$]{} where $\delta_s = {\mathbb{P}\left\{\tau - T_2(n) \geq s\right\}}$. Since the optimal policy chooses action $1$ in the first round by assumption, if $\tau = n$, then $\tau - T_2(n) \geq 1$ is guaranteed. Therefore by Lemma \[lem:tau-bound\] we have [$$\begin{aligned} \delta_1 \geq \frac{c'}{n^2 \log(n)}\,.\end{aligned}$$]{} By rearranging \[eq:bayes-arm\] we have for any $\delta \leq \delta_1$ that [$$\begin{aligned} \gamma &\leq \nu_1 + \frac{\sum_{s=1}^n \delta_s \sqrt{2\sigma_1^2 \log \left(\frac{1}{\delta_s}\right)}}{\sum_{s=1}^n \delta_s} \\ &\leq \nu_1 + \sqrt{2\sigma_1^2 \log\left(\frac{1}{\delta}\right)} + \frac{\sum_{s:\delta_s < \delta} \delta_s \sqrt{-2\sigma_1^2 \log \delta_s}}{\delta_1} \\ &\leq \nu_1 + \sqrt{2\sigma_1^2 \log\left(\frac{1}{\delta}\right)} + \frac{\sum_{s:\delta_s < \delta} \sqrt{2\sigma_1^2 \delta_s}}{\delta_1} \\ &\leq \nu_1 + \sqrt{2\sigma_1^2 \log\left(\frac{1}{\delta}\right)} + \frac{n\sqrt{2\sigma_1^2 \delta}}{\delta_1}\,.\end{aligned}$$]{} The result is completed by choosing $\delta = \delta_1^2 n^{-2}$. Technical Lemmas {#sec:tech} ================ \[lem:tech1\] Let $\eta \in (0,1)$ and $\Delta > 0$. Then [$$\begin{aligned} \sum_{k=0}^\infty (1 + \eta)^{\frac{k}{1 + \eta}} \exp\left(-(1 + \eta)^k \Delta^2\right) \leq \frac{3}{\eta} \left(\frac{1}{\Delta^2}\right)^{1/(1 + \eta)}\,.\end{aligned}$$]{} Let $L = \min{\left\{L : (1 + \eta)^k \geq \frac{1}{\Delta^2}\right\}}$. Then [$$\begin{aligned} \sum_{k=0}^\infty &(1 + \eta)^{\frac{k}{1 + \eta}} \exp\left(-(1 + \eta)^k \Delta^2\right) \\ &=\sum_{k=0}^{L-1} (1 + \eta)^{\frac{k}{1 + \eta}} \exp\left(-(1 + \eta)^k \Delta^2\right) + \sum_{k=L}^{\infty} (1 + \eta)^{\frac{k}{1 + \eta}} \exp\left(-(1 + \eta)^k \Delta^2\right) \\ &\leq\sum_{k=0}^{L-1} (1 + \eta)^{\frac{k}{1 + \eta}} + \sum_{k=L}^{\infty} (1 + \eta)^{\frac{k}{1 + \eta}} \exp\left(-(1 + \eta)^k \Delta^2\right) \\ &\leq \frac{1}{\eta} \left(\frac{1}{\Delta^2}\right)^{1/(1 + \eta)} + \sum_{k=L}^{\infty} (1 + \eta)^{\frac{k}{1 + \eta}} \exp\left(-(1 + \eta)^k \Delta^2\right) \\ &\leq \frac{1}{\eta} \left(\frac{1}{\Delta^2}\right)^{1/(1 + \eta)} + \left(\frac{1+\eta}{\Delta^2}\right)^{1/(1+\eta)} \sum_{k=0}^{\infty} (1 + \eta)^{\frac{k}{1 + \eta}} \exp\left(-(1 + \eta)^k \right) \\ &\leq \frac{1}{\eta} \left(\frac{1}{\Delta^2}\right)^{1/(1 + \eta)} + \left(\frac{1+\eta}{\Delta^2}\right)^{1/(1+\eta)} \sum_{k=0}^{\infty} (1 + \eta)^k \exp\left(-(1 + \eta)^k \right) \\ &\leq \frac{1}{\eta} \left(\frac{1}{\Delta^2}\right)^{1/(1 + \eta)} + \left(\frac{1+\eta}{\Delta^2}\right)^{1/(1+\eta)} \left(\frac{1}{e} + \int^\infty_0 (1 + \eta)^k \exp\left(-(1 + \eta)^k \right) dk\right) \\ &\leq \frac{1}{\eta} \left(\frac{1}{\Delta^2}\right)^{1/(1 + \eta)} + \left(\frac{1+\eta}{\Delta^2}\right)^{1/(1+\eta)} \frac{1}{e}\left(1 + \frac{1}{\log (1 + \eta)}\right) \\ &\leq \frac{2}{\eta} \left(\frac{1+\eta}{\Delta^2}\right)^{1/(1 + \eta)} \\ &\leq \frac{3}{\eta} \left(\frac{1}{\Delta^2}\right)^{1/(1 + \eta)}\end{aligned}$$]{} as required. Completing Proof of \[thm:gittins\] {#sec:algebra} =================================== First we need and easy lemma. \[lem:beta\] $f(\beta) = \frac{1}{\beta} \sqrt{\frac{1}{2\pi}} - \sqrt{\frac{\log \beta}{2}} {\operatorname{Erfc}}\left(\sqrt{\log \beta}\right)$ satisfies: 1. $\lim_{\beta \to \infty} f(\beta) \beta \log(\beta) = \frac{1}{\sqrt{8\pi}}$. 2. $f(\beta) \geq \frac{1}{10\beta \log(\beta)}$ for all $\beta \geq 3$. 3. $f(\beta) \leq \frac{1}{\sqrt{8\pi}} \cdot \frac{1}{\beta \log \beta}$. To begin we choose $c = (40c''')^4$. Recall that we need to show that [$$\begin{aligned} m\sigma f(\beta) &\geq c''' \max{\left\{\sigma, \sqrt{\sigma^2 \log(\beta)}, \frac{1}{\sqrt{\sigma^2 \log(\beta)}}, \sqrt{mW\left(m\sigma^2 \log(\beta)\right)} \beta^{-\frac{7}{8}}\right\}}\,,\end{aligned}$$]{} where [$$\begin{aligned} \beta_1 &= \frac{m}{c \log_+^{\frac{3}{2}}(m)} & \beta_2 &=\frac{m\sigma^2}{c \log_+^{\frac{1}{2}}(m\sigma^2)} & \beta &= \min{\left\{\beta_1,\, \beta_2\right\}}\end{aligned}$$]{} and $\beta \geq 3$ is assumed. This is equivalent to showing: [$$\begin{aligned} (1) \qquad m\sigma f(\beta) &\geq c'''\sigma && \text{and}\\ (2) \qquad m\sigma f(\beta) &\geq c'''\sqrt{\sigma^2 \log(\beta)} && \text{and} \\ (3) \qquad m\sigma f(\beta) &\geq c'''/\sqrt{\sigma^2 \log(\beta)} && \text{and} \\ (4) \qquad m\sigma f(\beta) &\geq c''' \sqrt{mW(m\sigma^2 \log(\beta))} \beta^{-\frac{7}{8}}\,.\end{aligned}$$]{} By \[lem:beta\] we have [$$\begin{aligned} m\sigma f(\beta) &\geq \frac{m\sigma}{10 \beta {\log_{+}\!}(\beta)} \geq \frac{m \sigma}{10 \cdot {\frac{m}{c\log^{\frac{3}{2}}_+(m)}}{\log_{+}\!}\left({\frac{m}{c\log^{\frac{3}{2}}_+(m)}}\right)} \\ &\geq c''' \sigma \sqrt{{\log_{+}\!}(m)} \geq c''' \sqrt{\sigma^2 \log(\beta)} \geq c''' \sigma\,.\end{aligned}$$]{} Therefore (1) and (2) hold. For (3) we have [$$\begin{aligned} m\sigma f(\beta) &\geq \frac{m\sigma}{10\beta {\log_{+}\!}(\beta)} \geq \frac{m \sigma}{10\cdot {\frac{m\sigma^2}{c \log^{\frac{1}{2}}_+(m\sigma^2)}}\log^{\frac{1}{2}}_+\left({\frac{m\sigma^2}{c \log^{\frac{1}{2}}_+(m\sigma^2)}}\right) \sqrt{{\log_{+}\!}(\beta)}} \\ &\geq \frac{c'''}{\sqrt{\sigma^2 \log(\beta)}}\,.\end{aligned}$$]{} Therefore (3) holds. Finally [$$\begin{aligned} m\sigma f(\beta) \beta^{\frac{7}{8}} &{\stackrel{(a)}}\geq \frac{m\sigma}{10\beta^{\frac{1}{8}} {\log_{+}\!}(\beta)} {\stackrel{(b)}}\geq \frac{m\sigma}{40\beta^{\frac{1}{4}}} {\stackrel{(c)}}\geq \frac{m\sigma}{40 \left(\frac{m\sigma^2}{(40 c''')^4}\right)^{\frac{1}{4}}} {\stackrel{(d)}}= c'''\sqrt{m} \left(m\sigma^2\right)^{\frac{1}{4}} \\ &{\stackrel{(e)}}\geq c'''\sqrt{m W(m\sigma^2 \log(m\sigma^2))} {\stackrel{(f)}}\geq c''' \sqrt{m W(m\sigma^2 \log(\beta))} \,,\end{aligned}$$]{} where (a) follows from Lemma \[lem:beta\]. (b) since $\beta^{\frac{1}{8}} \log(\beta) \leq 4 \beta^{\frac{1}{4}}$ for $\beta \geq 3$. (c) by substituting and naively upper bounding $\beta \leq \beta_2$. (d) is trivial. (e) since $x^{\frac{1}{4}} \geq \sqrt{W(x \log(x))}$ for all $x \geq 0$. (f) is true because $\beta \leq \beta_2 \leq m\sigma^2$. A Brief History of Multi-Armed Bandits {#sec:history} ====================================== This list is necessarily not exhaustive. It contains in (almost) chronological order the papers that are most relevant for the present work, with a special focus on earlier papers that are maybe less well known. Please contact me if there is a serious omission in this list. - [@Tho33] proposes a sampling approach for finite-horizon undiscounted Bernoulli bandits now known as Thompson sampling. No theoretical results are given. - [@Rob52] sets out the general finite-horizon bandit problem and gives some Hannan consistent algorithms and general discussion. Robbins seems to have been unaware of Thompson’s work. - [@BJK56] present the optimal solution for the Bayesian one-armed bandit in the finite-horizon undiscounted setting. They also prove many of the counter-intuitive properties of the Bayesian optimal solution for multi-armed bandits. The index used in the present article could just as well be called the Bradt–Johnson–Karlin index. - [@Git79] presents the optimal solution for the Bayesian multi-armed bandit in the infinite-horizon geometrically discounted setting. Gittins seems to have been unaware of the work by [@BJK56] who defined a similar index. Gittins’ result does not apply in the finite-horizon setting. - [@Whi80; @Web92; @Tsi94] all give alternative proofs and/or characterisations of the Gittins index for the infinite-horizon discounted case. - [@Bat83] gives an asymptotic approximation of the infinite-horizon Gittins index (see also the work by [@Yao06]). - The textbook by [@BF85] summarises many of the earlier results on Bayesian multi-armed bandits. They also prove that geometric discounting is essentially necessary for the Gittins index theorem to hold and give counter-examples in the Bernoulli case. - [@LR85] develop asymptotically optimal frequentist algorithms for the multi-armed bandits and prove the first lower bounds. - [@Agr95] and [@KR95] independently develop the UCB algorithm. - [@BK97] give asymptotic approximations for the finite-horizon Gittins index (more specific results in a similar vein are given by [@CR65]). - [@ACF02] proves finite-time guarantees for UCB. - [@Nin11] presents some methods for computing the finite-horizon Gittins index in the discrete case, and also suggest that index algorithm is a good approximation for the intractable Bayesian solution (no reference given). - [@KKM12] gave promising empirical results for the finite-horizon Gittins index strategy for the Bernoulli case, but did not study its theoretical properties. They also incorrectly claim that the Gittins strategy is Bayesian optimal. Their article played a large role in motivating the present work. Table of Notation {#sec:notation} ================= ------------------------------------------------------- ------------------------------------------------------------------------------------------------------ $d$ number of arms $n$ horizon $t$ current round $\mu_i$ expected return of arm $i$ $\mu^*$ maximum mean reward, $\mu^* = \max_i \mu_i$ $\hat \mu_i(t)$ empirical estimate of return of arm $i$ after round $t$ $T_i(t)$ number of times arm $i$ has been chosen after round $t$ $I_t$ arm chosen in round $t$ $X_t$ reward in round $t$ $\nu_i$ prior mean of arm $i$ $\sigma_i^2$ prior variance for mean of arm $i$ $\Delta_i$ gap between the expected returns of the best arm and the $i$th arm $\Delta_{\max}$ maximum gap, $\max_i \Delta_i$ $\Delta_{\min}$ minimum non-zero gap, $\min {\left\{\Delta_i : \Delta_i > 0\right\}}$ ${\log_{+}\!}(x)$ $\max{\left\{1, \log(x)\right\}}$ ${\operatorname{W\!}}(x)$ product logarithm $x = W(x) exp(W(x))$ ${\operatorname{diag}}(\sigma_1^2,\ldots,\sigma_d^2)$ diagonal matrix with entries $\sigma_1^2,\ldots,\sigma_d^2$ $\gamma(\nu, \sigma^2, m)$ Gittins index for mean/variance $(\nu, \sigma^2)$ and horizon $m$ $\tau(\nu, \sigma^2, m)$ stopping time that determines the Gittins index ${\operatorname{Erfc}}(x)$ complementary error function [$$\begin{aligned} {\operatorname{Erfc}}(x) = \frac{2}{\sqrt{\pi}} \int^\infty_x \exp(-y^2) dy\end{aligned}$$]{} ${\operatorname{Erf}}(x)$ error function ${\operatorname{Erf}}(x) = 1 - {\operatorname{Erfc}}(x)$ $\mathcal N(\mu, \sigma^2)$ Gaussian with mean $\mu$ and variance $\sigma^2$ $\Phi(x)$ standard Gaussian cdf [$$\begin{aligned} \Phi(x) = \frac{1}{\sqrt{2\pi}}\int^x_{-\infty} \exp\left(-\frac{t^2}{2}\right) dt\end{aligned}$$]{} ------------------------------------------------------- ------------------------------------------------------------------------------------------------------ [^1]: This contradicts the opposite claim made without proof by [@KCOG12] but counter-examples for Bernoulli noise have been known for some time and are given here for Gaussian noise in \[sec:bayes\]. In fact, the Gittins index strategy is only Bayesian optimal in all generality for geometric discounting [@BF85 §6]. [^2]: An exponential prior leads to the discounted setting for which anytime regret guarantees are unlikely to exist, but where computation is efficient. A power-law would be a more natural choice, but the analysis becomes very non-trivial in that case.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We present new high resolution spectroscopic observations of the Herbig-Haro object HH 32 from System Verification observations made with the GMOS IFU at Gemini North Observatory. The 3D spectral data covers a $8''''.7\times 5''''.85$ spatial field and 4820 - 7040 Å spectral region centered on the HH 32 A knot complex. We show the position-dependent line profiles and radial velocity channel maps of the H$\alpha$ line, as well as line ratio velocity channel maps of \[O III\] 5007/H$\alpha$, \[O I\] 6300/H$\alpha$, \[N II\] 6583/H$\alpha$, \[S II\] (6716+6730)/H$\alpha$ and \[S II\] 6716/6730. We find that the line emission and the line ratios vary significantly on spatial scales of $\sim$1$''''$ and over velocities of $\sim$50 km/s. A “3/2-D” bow shock model is qualitatively successful at reproducing the general features of the radial velocity channel maps, but it does not show the same complexity as the data and it fails to reproduce the line ratios in our high spatial resolution maps. The observations of HH 32 A show two or three superimposed bow shocks with separations of $\sim 3''''$, which we interpret as evidence of a line of sight superposition of two or three working surfaces located along the redshifted body of the HH 32 outflow.' author: - 'Tracy L. Beck, A. Riera, A. C. Raga, C. Aspin,' title: 'The 3-Dimensional Structure of HH 32 from GMOS IFU Spectroscopy' --- Introduction ============ HH 32 is one of the brightest sources in the original catalog of Herbig-Haro objects (Herbig 1974). Spectroscopic studies have shown that HH 32 has a high excitation spectrum with significant emission from H, He and a wide range of metals in the 3700 to 10800 Å region (Dopita 1978; Brugel, Böhm & Mannery 1981; Raga, Böhm & Cantó 1996). HH 32 also has an unusually large extinction (E(B-V)$\approx 0.7$; Brugel, Böhm & Mannery 1981), hence it is surprising that this object was detected in the ultraviolet with IUE (Böhm & Böhm-Vitense 1984; Lee et al. 1988; Moro-Martín et al. 1996). The UV detection results from of the strong emission in the high exitation \[C III\] 1909 Å and \[C IV\] 1550 Å ultraviolet lines. Interestingly, this high excitation object also shows H$_2$ [*v*]{} = 1-0 S(1) emission in the infrared which is excited in shocks as the flow encounters ambient cloud material. (Zealey et al. 1986; Zinnecker et al. 1989; Davis, Eislöffel & Smith 1996). The HH 32 outflow is clearly associated with the T Tauri star AS 353A, with condensations A, B and D at $\sim 20''$ to the west, and condensation C at $\sim 10''$ to the east of the star (see Figure 1 from Curiel et al. 1997). AS 353A has a rich emission line spectrum (Eislöffel, Solf & Böhm 1990), but shows only two forbidden emission lines: \[O I\] 5577, 6300 Å. High resolution optical spectroscopy of the HH condensations has shown that HH 32 A, B and D have very broad, redshifted line profiles, with full widths of $\sim 400$ km s$^{-1}$ (Herbig & Jones 1983; Hartigan, Mundt & Stocke 1986; Solf, Böhm & Raga 1986; Hartigan, Raymond & Hartmann 1987). The faint knot C to the east of the exciting source is blueshifted, indicating that the HH 32 emission knots are part of a bipolar outflow from AS 353A. The very large line widths, the shape of the line profiles, and the position-velocity diagrams have been interpreted successfully in terms of “3/2-D” bow shock models (Solf et al. 1986; Raga, Böhm & Solf 1986; Raga & Böhm 1986; Hartigan et al. 1987). In such models, the bow shock is modeled with a parametrized shape (i.e., Hartigan et al. 1987) and the emission is computed by assuming that the post-bow shock recombination region resembles the structure of a 1D stationary shock. Comparisons with the models show a $v_{bs}\sim 300-350$ km s$^{-1}$ velocity for the bow shock (Solf et al. 1986; Hartigan et al. 1987). By comparing the proper motions with the radial velocities we know that HH 32A is moving away from the observer at a $\phi\approx 70^\circ$ angle with respect to the plane of the sky (Herbig & Jones 1983). This angle is consistent with the orientation necessary to model the emission line profiles (Solf et al. 1986; Hartigan et al. 1987). The detailed morphological structure and the proper motions of the sub-condensations of HH 32A described by Curiel et al. (1997) have been interpreted in terms of 3/2-D bow shock models to obtain the same $\phi=70^\circ$ and $v_{bs}\sim 300$ km s$^{-1}$ values that are deduced from the emission line profiles (Raga et al. 1997). In this paper we present and discuss new spectroscopic observations with 2D spatial coverage of a field centered on the HH 32A condensation, which give new information on the kinematics and structure of the emission knots. The previous spatially resolved spectroscopy of Solf et al. (1986) covered all of HH 32 with several different long-slit positions centered on AS 353A, but did not have a comparable angular resolution nor as high a signal-to-noise ratio as the spectra we present here. Also, the HST images of Curiel et al. (1997) show the high resolution spatial structure of HH 32, but lack any kinematic information. In $\S$2, we present the observations and discuss briefly the method of data reduction. We show the H$\alpha$ line profiles and velocity channel maps obtained from the spectra, and the velocity channel maps of different line ratios in $\S$3 and $\S$4. In $\S$5 we interpret the observations in terms of a 3/2-D bow shock model and discuss its successes and failures in modeling the data. Finally, we present our conclusions and a discussion of the implications in $\S$6. The observations ================ The data for this project were obtained on 2002 Aug 03 with the Gemini Multi-Object Spectrograph (GMOS; Hook et al. 2002) Integral Field Unit (IFU; Allington-Smith et al. 2002) at the Gemini North Fredrick C. Gillett Telescope. These data were taken as System Verification for dithered observations on an emission line source, and are available to the public at: http://www.gemini.edu/sciops/instruments/gmos/SVobs/gmosSVobs73.html under Gemini program ID GN-2002B-SV-73. The observations were made in cloudy conditions (0.5 to 1 magnitude of extinction) with $\sim$0.$''$5 average seeing. The source star of the HH 32 outflow, AS 353A, was used to provide guiding and tip-tilt corrections using the GMOS on-instrument wavefront sensor (OIWFS). These data were obtained with the IFU in 1-slit mode, which provides a spatial field of $3\farcs5$ x 5$''$ for each resulting science datacube. With this observing configuration, the GMOS IFU is comprised of 750 fibers; each spans a $0\farcs2$ hexagonal region on the sky. 500 fibers make up the $3\farcs5$ x 5$''$ science field of view, and 250 fibers make up a smaller, dedicated sky field which is fixed at a 1$'$ distance from the science position (Allington-Smith et al. 2002). A total of 12 900 second frames were obtained; two steps through a 3 x 2 dither pattern resulting in total integration time of 30 minutes at each of the 6 dither positions. We used the R831 grating in GMOS, which has a 0.034 nm/pixel scale and results in $\approx 15$ km s$^{-1}$ spectral resolution. We were able to obtain spectra over the 4920 - 7040 Å region, which includes the \[O III\] 4969/5007, \[O I\] 6300/63, \[N II\] 6548/83, H$\alpha$, and \[S II\] 6716/31 lines. We used the GMOS IFU scripts available in the Gemini IRAF package for reduction of the data. The [gfreduce]{} task was used to prepare the frames for reduction, subtract off the bias and overscan levels, reject cosmic rays by calling the [gscrrej]{} routine, and call the additional reduction routines [gfextract]{} and [gfskysub]{}. [gfextract]{} traces and extracts each IFU fiber to 1-D spectral output in the flat field image, applies this trace to the science data, extracts the science spectra using a $\pm$2.5 pixel aperture on each side of the traces, and applies the flat field correction. [gfskysub]{} subtracts off a sky spectrum from the science data using averaged spectra from the dedicated sky field inherent in the IFU. The [gswavelength]{} task then establishes the wavelength calibration for the IFU arc lamp images, which is thus applied to the science frames using the [gftransform]{} routine. These 2-Dimensional data images were reformatted into a 3-D datacube format (x, y, $\lambda$) and resampled to square pixels with a 0.$''$05 spatial resolution using the [gfcube]{} routine. The dithered data were mosaiced together to form a larger cube in the spatial dimension by manually offsetting and coadding the individual cubes using IDL. The final datacube has a spatial extent of 8$''$.7 by 5.$''$85. At the forementioned webpage we have made available the raw data, the final IRAF datacubes, the mosaiced cube constructed in IDL, an IRAF cl-reduction script, and a README file that describes our data reduction process. In Figure 1 we present a 30 second exposure r-band image of the HH 32 system that was obtained for setup on the IFU field; the HH 32 A1, A2, B and D condensations are labeled (using the notation of Curiel et al. 1997). Overplotted on the image is the position and orientation in the spatial dimension of the final mosaiced datacube, and a vector indicating the direction to the outflow source, AS 353A. In Figure 2 we show a spectrum of the HH 32 A2 knot averaged over the inner $1\farcs5$ in the spatial dimension; the \[O I\] 6300/63, \[N II\] 6548/83, H$\alpha$ and \[S II\] 6716/31 lines are identified. All of the lines show a pronounced double-peaked structure, consistent with previous observations (Hartigan, Mundt & Stocke 1986). H$\alpha$ line profiles and channel maps ======================================== Figure 3 shows a contour map of the H$\alpha$ emission constructed by summing the line over its wavelength extent in the final mosaiced datacube. Overplotted in Figure 3 are the H$\alpha$ line profiles obtained by integrating the emission over boxes of $10\times 10$ pixels (corresponding to $0\farcs5 \times 0\farcs5$). Many of the observed line profiles are double peaked, with a stronger peak at a $v_r\approx 90$ km s$^{-1}$ heliocentric radial velocity and a weaker high velocity peak at $v_r\approx 240\to 276$ km s$^{-1}$. Several of the line profiles in the emission joining the maxima of the A1 and A2 condensations show a dominant high velocity peak, and a few of the profiles in the periphery of A1 show three components. In many of the line profiles along the edges of the observed field, only the low velocity component is visible. In Figure 4 we present the H$\alpha$ radial velocity channel maps derived from the IFU observations. The H$\alpha$ emission is detected from heliocentric radial velocities ranging from $v_r\approx -20$ to $+400$ km s$^{-1}$. The surface brightness of the condensations grows with increasing $v_r$, reaching a maximum emission level in the $+72$ km s$^{-1}$ map. It then decreases, reaching a minimum at $+165$ km s$^{-1}$, and increases again reaching a second maximum at $v_r=+241$ km s$^{-1}$. For the larger radial velocities, the surface brightness decreases monotonically. At $v_r=10$ km s$^{-1}$, there are two condensations with a separation of $\approx 2\farcs40$. The western condensation (A2) has a structure of two superimposed arcs and the eastern condensation (A1) also has an arc-like structure. For both condensations, this basic morphology is preserved to velocities of $v_r\approx 100$ km s$^{-1}$. We find a pronounced arc of emission that connects the A1 and A2 condensations in the $v_r \approx 25$ km s$^{-1}$ to $v_r \approx 87$ km s$^{-1}$ maps. A1 and A2 no longer have their arc-like shapes and have a more compact morphology for $v_r> 118$ km s$^{-1}$. They approach each other in the spatial direction for increasing values of $v_r$, and have a separation of only $\approx 1\farcs15$ in the $v_r=304$ km s$^{-1}$ map. Condensation A1 dominates the total intensity for $v_r> 300$ km s$^{-1}$, and again develops a compact, arc-like shape for higher values of $v_r$. Although not shown, the channel maps for the \[O I\], \[N II\], and \[S II\] lines show qualitatively similar morphologies as those for H$\alpha$ . However, the channel maps for the \[O III\] line show significant differences. The surface brightness of the two condensations in \[O III\] grows with increasing radial velocity, reaching a maximum level at +168 km s$^{-1}$, and then the surface brightness decreases for larger velocities. No second maximum is detected. At the lowest radial velocities, subcondensation A2 shows an elongated structure that, despite its low signal-to-noise, resembles the arc-shaped structure also visible in the H$\alpha$ maps. At higher radial velocities, subcondensation A1 has a more compact morphology. Subcondensation A1 is detected at $v_r$ from +26 to +350 km s$^{-1}$, and it dominates the total intensity for $v_r$ $>$ +188 km s$^{-1}$. Subcondensation A1 shows a compact morphology at these radial velocities; no arc-like structure is detected in this subcondensation in the \[O III\] emission. As reported for the H$\alpha$ maps, both subcondenstions approach each other in the spatial direction as the radial velocity increases. The line ratios =============== The \[S II\] 6716/6731 ratio depends only weakly on temperature, so it is a direct diagnostic of the electron density of the emitting plasma. Unfortunately, the other ratios between the observed lines do not have such a direct interpretation. However, ratios such as \[S II\] (6716+6731)/H$\alpha$ and \[O III\] 5007/H$\alpha$ do show clear trends as a function of shock velocity in models of HH jets (see, e. g., Hartigan et al. 1987), and can therefore be used to estimate the velocities of the shocks which dominate the emission along a given line of sight. Our spectra are not flux calibrated, so we are not able to give direct estimates of the values of the shock velocities. Therefore, our results can only be used to obtain a qualitative picture of the variations of the shock properties across the emitting region of HH 32. From the position-velocity (PV) cubes of the \[O I\] 6300, \[N II\] 6583, \[S II\] 6716, 6731, \[O III\] 5007 and the H$\alpha$ lines, we have generated seven channel maps centered at heliocentric radial velocities of $v_r=-5$, 48, 110, 172, 234, 296 and 358 km s$^{-1}$ (each map has a velocity width of 47 km s$^{-1}$). We have binned the maps spatially over $5\times 5$ pixels of the original spectrum to decrease the effective pixel size to $0\farcs25$. Figure 5 shows the \[O I\] 6300 and \[N II\] 6583 channel maps divided by the corresponding H$\alpha$ maps, Figure 6 shows the \[S II\] (6716+6731)/H$\alpha$ and the \[S II\] 6716/6731 line ratio channel maps and Figure 7 shows the \[O III\] 5007 Å channel maps divided by the corresponding H$\alpha$ channel maps. The overall spatial size of the line ratio maps is the same as the H$\alpha$ channel maps, but the pixels have been binned by 5 in the x and y directions. The line ratio maps appear to cover a wider area than the H$\alpha$ emission maps because there is non-negligible flux in the lines at distances spatially extended from the condensations seen in H$\alpha$. As a function of radial velocity, the behavior of the line ratio maps can be described as: -  6583/H$\alpha$ : at the position of knot A2, this line ratio has low values of $\approx 0.4$ at low radial velocities, and the ratio grows with increasing $v_r$. At the position of knot A1, the maximum values of $\approx 0.7$ for the line ratio occur at $v_r=48$ km s$^{-1}$, and the ratios decrease for increasing $v_r$, reaching values of $\approx 0.4$ for $v_r=296$ km s$^{-1}$. Interestingly, the region with the highest line ratio lies to the South of knot A1, with values of $\approx 1$ for $v_r=48$ km s$^{-1}$. -  6300/H$\alpha$ : knot A2 shows an initial increase in this line ratio, from $\approx 0.2$ for $v_r=-5$ km s$^{-1}$ to $\approx 0.4$ for $v_r=110$ km s$^{-1}$. The line ratio has a $\approx 0.3$ value in the $v_r=172$ km s$^{-1}$ velocity channel and then it increases for higher radial velocities, reaching a value of $\approx 0.5$ for $v_r=296$ km s$^{-1}$. Knot A1 shows high values of $\approx 0.7$ at $v_r=48$ km s$^{-1}$, decreasing to values of $\approx 0.1$ at $v_r=296$ km s$^{-1}$. The highest values in this line ratio are found to the East of knot A1 in the $v_r=110$ km s$^{-1}$ channel map. -  (6716+6731)/H$\alpha$ : this line ratio shows a general trend of decreasing values with increasing radial velocity for $v_r=-5\to 172$ km s$^{-1}$, and then has similar values for all higher radial velocity channel maps. Knot A2 has line ratios of $\approx 1$ for $v_r=-5$ km s$^{-1}$, dropping to $\approx 0.4$ for $v_r=110$ km s$^{-1}$, and then keeping this value for the higher radial velocity channels. Knot A1 has a more drastic drop in the line ratio, with values of $\approx 2$ at low velocities, decreasing to $\approx 0.2$ for $v_r=234$ km s$^{-1}$. -  6716/6731 : for both A1 and A2, this line ratio shows values of $\approx 0.7$ (corresponding to $n_e\approx 230$ cm$^{-3}$) in the $v_r=-5$ km s$^{-1}$ map. The line ratio then decreases with increasing radial velocities, with values of $\approx 0.6$ ($n_e\approx 420$ cm$^{-3}$) for $v_r=48$ km s$^{-1}$ and $\approx 0.55$ ($n_e\approx 650$ cm$^{-3}$) for $v_r=110$ km s$^{-1}$. The line ratio at the positions of knots A1 and A2 remains close to the high density limit for all of the radial velocity channel maps with $v_r> 110$ km s$^{-1}$. It is clear in the lower heliocentric radial velocity channel maps (with $v_r\leq 110$ km s$^{-1}$) that knots A1 and A2 are surrounded by a diffuse emitting region (filling most of the observed field in the $v_r=-5$ km s$^{-1}$ map) with line ratios of $\approx 1$ ($n_e\approx 150$ cm$^{-3}$). - 5007/H$\alpha$: at the position of condensation A1, this emission line ratio grows continually with radial velocities from -5 km s$^{-1}$ up to +300 km s$^{-1}$. At the condensation A2, this line ratio is roughly constant for radial velocities from +110 to +238 km s$^{-1}$ and increases for larger radial velocities. This line ratio is clearly larger at the position of condensation A1 than at the knot A2. These results show that the line ratios of HH 32 have strong spatial variations at arcsecond scales, as well as variations as a function of radial velocity. The ratio maps for the \[S II\] lines show that the brighter A1 and A2 condensates are enclosed within a diffuse emitting region with lower electron density than the knots. This diffuse low electron density region surrounding knots A1 and A2 has low H$\alpha$ emission; it has significant values for both the high excitation \[N II\] 6583/H$\alpha$ and the low excitation \[O I\] 6300/H$\alpha$ line ratios (see Figure 5). Indeed, the peak \[N II\] 6583/H$\alpha$ and \[O I\] 6300/H$\alpha$ values are found in this diffuse region rather than centered on the positions of the A1 or A2 condensations. Although the behavior of the line ratios is very complex, some of the observed trends are consistent with a simple, plane shock wave interpretation. For example, the \[O III\]/H$\alpha$ ratio in knot A1 grows monotonically with radial velocity, consistent with the fact that this line ratio increases with increasing shock velocities. Also, the \[S II\] (6716+6731)/H$\alpha$ ratio in the A1 condensation first decreases and then stabilizes as a function of increasing $v_r$, which qualitatively follows the line ratio vs. shock velocity trend predicted from plane-parallel shock models (see Hartigan et al. 1987). However, different regions of HH 32 show different trends in the line ratio versus radial velocity which cannot be easily interpreted with the present plane shock models. As an example, knot A1 shows a decrease in the \[O I\] 6300/H$\alpha$ ratio as a function of increasing radial velocity, while knot A2 has the opposite trend. Interestingly, for this oxygen line ratio the plane shock models predict a decrease in intensity as a function of shock velocities (from 20 to 90 km/s), followed by a strong increase for larger velocities (from 100 to 300 km/s, see Hartigan et al. 1987). The observed line ratios of \[O I\] 6300/H$\alpha$ in the two knots are not consistent with the present models. The behavior as a function of $v_r$ of the line emission of knots A1 and A2 can be seen in a more quantitative way in Figure 8. In this figure, we have plotted the ratios between the H$\alpha$, \[O III\] 5007, \[O I\] 6300, \[N II\] 6583 and \[S II\] (6716+31) line emission of A1 and A2 as a function of $v_r$. From this graph we see that knot A1 is brighter than A2 for heliocentric radial velocities $v_r<150$ km s$^{-1}$ in the \[O I\] 6300 and \[S II\] (6716+31) lines, and A2 is brighter than A1 for $150<v_r<350$ km s$^{-1}$. For $v_r>350$ km s$^{-1}$, knot A1 is again brighter than A2. The relative intensities between A1 and A2 in the H$\alpha$ and \[N II\] 6583 lines show a similar behavior to \[O I\] and \[S II\], but have a much shallower dependence on radial velocity for $v_r<300$ km s$^{-1}$. The H$\alpha$ fluxes of A1 and A2 differ by at most 20 % in this radial velocity range. In the H$\alpha$ and \[N II\] lines, condensation A1 brightens quite steeply relative to A2 for $v_r>300$ km s$^{-1}$, becoming twice as bright for $v_r=390$ km s$^{-1}$. The relative \[O III\] intensity between A1 and A2 shows a different behavior than the other emission lines plotted (the \[O III\] channel maps are shown in Figure 7). The \[O III\] emission seen in A1 brightens quite steeply relative to A2 for $v_r$ from 0 to +120 km s$^{-1}$. The relative intensities remain more or less constant with a mean value of 1.75 for radial velocities in the range from +120 to +300 km s$^{-1}$, and decrease again for larger velocities ($v_r$ $>$ 300 km s$^{-1}$). A 3/2-D bow shock model ======================= In order to interpret the observations of HH 32, we consider a traditional “3/2-D” bow shock model (Hartigan et al. 1987). It is assumed that the shock velocity is equal to the component of the pre-bow shock flow velocity normal to the shock surface. Such models have been used to interpret line profiles of several HH objects (see, e. g., Hartigan et al. 1987), with 1D or 2D spatial resolution (Raga & Böhm 1985; 1986; Solf et al. 1991; Morse et al. 1992). To date, one of the most successful applications of the 3/2-D bow shock models comes from the line profile fits to the HH 32 outflow (Solf et al. 1986; Hartigan et al. 1987). We consider a bow shock with a functional form: $${z\over a}=\left({r\over a}\right)^p\,, \label{zr}$$ where $z$ is measured along the symmetry axis and $r$ is the cylindrical radius. The constants $a$ and $p$ are the free parameters of the model, with $p$ determining the shape of the bow shock, and $a$ its physical size. The bow shock is assumed to be moving at a velocity $v_{bs}$ with respect to a uniform pre-shock medium, directed at an angle $\phi$ with respect to the plane of the sky. Without any loss of generality, we can set $a=1$, and scale the intensity maps predicted from the model in an arbitrary way to produce spatial scales comparable to the ones of HH 32A. In order for the model to fit closely with the observations, we have assumed that the wings of the bow shock (with the shape given by equation \[zr\]) suddenly end at a distance $z_{max}$, as measured along the symmetry axis from the head of the bow shock. We then use the geometry of the system to calculate: 1) the shock velocity, which is the velocity component of the bow shock normal to the surface of the shock, 2) the velocity of the post-bow shock flow which is approximately equal to the normal velocity for the compressions found in radiative shocks, and 3) the intensity per unit area of the bow shock derived by interpolating between the line emissions predicted from a series of plane-parallel shock models with differing shock velocities. The necessary geometrical construction we have used is described by Raga & Böhm (1986) and in more detail by Hartigan et al. (1987). We computed the emission using the tabulation of plane-parallel models with a “self consistent” pre-ionization and pre-shock density of $n_{pre}=100$ cm$^{-3}$ as described by Hartigan et al. (1987). In the models, we fixed the bow shock velocity to a $v_{bs}=350$ km s$^{-1}$ value necessary to obtain line widths comparable to the observed ones, and we adopt an orientation angle of $\phi=70^\circ$ moving away from the observer, as derived by the observed line profiles, radial velocities and proper motions (Raga et al. 1986; Hartigan et al. 1987; Curiel et al. 1997). We further assume that the pre-shock medium is at rest with respect to the outflow source. At this point, the power law index $p$ is the only free parameter in the model. If the shock wings extend to large distances from the head of the bow shock, then we can set $z_{max} \to \infty$ (see equation \[zr\]). In this case, the best fits for the model are obtained for $p\sim 2-3$. However, if we allow a finite value for $z_{max}$, the models more closely resemble the observations. In particular, we choose a model with $p=2$ and $z_{max}=1.5a$. From this bow shock model, we have computed velocity channel maps which are used to compare with the observations of HH 32 (shown in Figure 9). From comparison of the velocity channel maps shown in Figures 4 and 9, we find that the shock model predicts structures that are in qualitative agreement with the HH 32 observations. At low radial velocities ($v_r=15\to 90$ km s$^{-1}$), the model shows strong H$\alpha$ emission distributed in a single arc-like filament. The observations of HH 32 also show strong H$\alpha$ emission but it is distributed in two or three arc-like structures. At $v_r=100\to 200$ km s$^{-1}$, the bow shock model shows fainter H$\alpha$ emission with a circular structure, which becomes more compact for increasing radial velocities. In this velocity range, the HH 32 data also shows fainter, more compact emission at increasing radial velocities, but with a more complex morphology of knots (see Figure 4). For $v_r>200$ km s$^{-1}$, both the model and HH 32 data show an increase in H$\alpha$ emission, which is then followed by an intensity decrease, with tighter knot-like structures at the higher radial velocities. There is a lack of quantitative agreement because the high velocity H$\alpha$ peak occurs at $v_r=308$ km s$^{-1}$ in the model, but at $v_r=227$ km s$^{-1}$ in HH 32. Also, while the model shows a single condensation for $200<v_r<340$ km s$^{-1}$, HH 32 shows two or three compact condensations, and it does not converge to a single condensation until $v_r\geq 350$ km s$^{-1}$. Our simple model considers only one bow shock observed at an angle of $\approx 70^\circ$, comparison with the observations shows that a more complicated structure of 2 (or more) bow shocks would better fit the observations (see Raga et al. 2003). We computed line ratio maps from our bow shock model and did not obtain results that qualitatively agree with the observations presented in Figures 5, 6 and 7. This discrepancy is not surprising because of the differences seen in the H$\alpha$ channel maps predicted by the model and observed in HH 32. The significant structure observed in the line ratio maps and the lack of consistency with our calculations is difficult to interpret uniquely, as a superposition of several shocks along the viewing geometry of HH 32 will complicate the models. There are two main criticisms for the 3/2-D bow shock model we have utilized to describe the observations of HH 32. First, it is obvious that the simple, parametrized shape (equation 1) which we have assumed for the emitting shock structure is not appropriate for describing the more complex line of sight superposition of shocks in HH 32. Secondly, a 3/2-D shock model is based on the assumption that the post-bow shock cooling distance is small compared to the size of the bow shock; this is probably not the case in this scenario. As can be seen from the models of Hartigan et al. (1987), the cooling distance behind the head of a 250 km/s shock has a value $d_c\sim 10^{16}$ cm, which is comparable to the sizes of condensations A1 and A2. The model we have used is qualitatively successful at describing the observations of HH 32, but the assumptions on which the 3/2-D bow shock models are based are not ideal for describing this object. We present and discuss a more detailed 3D simulation of multiple working surfaces in a different paper (Raga et al. 2003) Conclusions =========== We have obtained a high resolution spectrum of HH 32A with 2D spatial resolution which provides a wealth of information about the excitation and kinematics of the HH 32 outflow. From these spectra, we have constructed channel maps at velocities ranging from $-20$ to $+400$ km s$^{-1}$ for the H$\alpha$, \[O III\] 5007, \[O I\] 6300, \[N II\] 6583, \[S II\] 6716 and \[S II\] 6731 emission lines; the H$\alpha$ maps are presented in Figure 4 and \[O III\] is shown in the central panels of Figure 7. The channel maps of the emission lines of H$\alpha$, \[O I\], \[N II\], and \[S II\] have a qualitatively similar behavior. In the lower heliocentric radial velocity maps ($v_r<100$ km s$^{-1}$), HH 32A has a morphology of two or three superimposed arc-like features. At higher radial velocities, the emission becomes progressively more concentrated, showing two or three compact condensations. Finally, for $v_r>350$ km s$^{-1}$, a single condensation is seen. Figures 5, 6 and 7 show complex structures in the line ratio maps. The local maxima and minima do not coincide necessarily with the observed knots in the H$\alpha$ images, and the ratios vary for each emission line over the velocity width of the feature. The complex morphological structures suggest that we are likely observing the emission from several compact knots, superimposed on a diffuse emission component, with the knots and the diffuse component having different spectral properties. This interpretation has been previously suggested to explain 2D spatial resolution spectra of HH 2 (Böhm & Solf 1992). The observations of HH 32 can be interpreted in a qualitative way with a simple “3/2-D” bow shock model. We constructed model H$\alpha$ velocity channel maps that show a transition between a single arc-like feature at low radial velocities, to a more concentrated and centrally peaked condensation at high radial velocities (Figure 9). This basic change in morphology is also seen in the velocity channel maps of HH 32A (Figure 4), but the structures in the data are much more complex. From comparison with the simple bow shock model, we conclude that HH 32A is likely a superposition of two or three bow shocks, corresponding to different working surfaces along the HH 32 outflow. HH 32A shows at least two condensations (A1 and A2) with spatial and kinematic properties which resemble the predictions from a single bow shock model. The fact that we see such a superposition is not surprising; the $\phi\approx 70^\circ$ orientation of HH 32 with respect to the plane of the sky will lead naturally to a line of sight overlap of features along the outflow axis. HH 32 might be intrinsically similar to the collimated outflow observed in HH 34, we might be viewing two working surfaces catching up with each other as seen in HH 34S (Reipurth et al. 2002; Raga et al. 2002). The arc-like structures and multiple condensations observed in the data of HH 32A could then correspond to a line of sight superposition of the working surfaces. We present a study of this scenario based on more detailed 3D numerical simulations in Raga et al. (2003). We thank Inger J[ø]{}rgensen, GMOS-N instrument scientist, for approving this project for System Verification observations. A. Riera acknowledges the ICN-UNAM for support during her sabbatical. A. Riera was supported by the MCyT grant AYA2002-00205 (Spain). The work of A. C. Raga was supported by the CONACyT grant 36572-E. Allington-Smith, J. et al. 2002, PASP, 114, 892 Böhm, K. H., Böhm-Vitense, E., 1984, ApJ, 277, 216 Böhm, K. H., Solf, J., 1992, AJ, 104, 1193 Brugel, E. W., Böhm, K. H., Mannery, E., 1981, ApJS, 47, 117 Curiel, S., Raga, A. C., Raymond, J. C., Noriega-Crespo, A., Cantó, 1997, AJ, 114, 2736 Davis, C. J., Eislöffel, J., Smith, M. D., 1996, ApJ, 463, 246 Dopita, M. A., 1978, ApJS, 37, 117 Eislöffel, J., Solf, J., Böhm, K. H., 1990, A&A, 237, 369 Hartigan, P., Mundt, R., Stocke, J., 1986, AJ, 91, 1357 Hartigan, P., Raymond, J. C., Hartmann, L. W., 1987, ApJ, 316, 323 Herbig, G. H., 1974, LicOB, 658 Hook, I. et al. 2002, SPIE, 4841 Herbig, G. H., Jones, B. F., 1983, AJ, 88, 1040 Lee, M. G., Böhm, K. H., Temple, S. D., Raga, A. C., Mateo, M. L., Brugel, E. W., Mundt, R., 1988, AJ, 96, 1690 Moro-Martín, A., Noriega-Crespo, A., Böhm, K. H., Raga, A. C., 1996, RMxAA, 32, 75 Morse, J. A., Hartigan, P., Cecil, G., Raymond, J. C., Heathcote, S., 1992, ApJ, 399, 231 Raga, A. C., Böhm, K. H., 1985, ApJS, 58, 201 Raga, A. C., Böhm, 1986, ApJ, 308, 829 Raga, A. C., Böhm, K. H., Solf, J., 1986, AJ, 92, 119 Raga, A. C., Böhm, K. H., Cantó, J., 1996, RMxAA, 32, 161 Raga, A. C., Cantó. J., Curiel, S., Noriega-Crespo, A., Raymond, J. C., 1997, RMxAA, 33, 157 Raga, A. C., Velázquez, P. F., Cantó, J., Masciadri, E., 2002, A&A, 395, 647 Raga, A. C., et al. 2003, submitted to AJ Reipurth, B., Heathcote, S., Morse, J., Hartigan, P., Bally, J., 2002, AJ, 123, 362 Solf, J., Böhm, K. H., Raga, A. C., 1986, ApJ, 305, 795 Solf, J., Raga, A. C., Böhm, K. H., Noriega-Crespo, A., 1991, AJ, 102, 1147 Zealey, W. J., Williams, P. M., Taylor, K. N. R., Storey, J. W. V., Sandell, G., 1986, A&A, 158, 9 Zinnecker, H., Mundt, R., Geballe, T. R., Zealey, W. J., 1989, ApJ, 342, 337
{ "pile_set_name": "ArXiv" }
--- abstract: | We use a conditional Karhunen-Loève (KL) model to quantify and reduce uncertainty in a stochastic partial differential equation (SPDE) problem with partially-known space-dependent coefficient, $Y(x)$. We assume that a small number of $Y(x)$ measurements are available and model $Y(x)$ with a KL expansion. We achieve reduction in uncertainty by conditioning the KL expansion coefficients on measurements. We consider two approaches for conditioning the KL expansion: In Approach 1, we condition the KL model first and then truncate it. In Approach 2, we first truncate the KL expansion and then condition it. We employ the conditional KL expansion together with Monte Carlo and sparse grid collocation methods to compute the moments of the solution of the SPDE problem. Uncertainty of the problem is further reduced by adaptively selecting additional observation locations using two active learning methods. Method 1 minimizes the variance of the PDE coefficient, while Method 2 minimizes the variance of the solution of the PDE. We demonstrate that conditioning leads to dimension reduction of the KL representation of $Y(x)$. For a linear diffusion SPDE with uncertain log-normal coefficient, we show that Approach 1 provides a more accurate approximation of the conditional log-normal coefficient and solution of the SPDE than Approach 2 for the same number of random dimensions in a conditional KL expansion. Furthermore, Approach 2 provides a good estimate for the number of terms of the truncated KL expansion of the conditional field of Approach 1. Finally, we demonstrate that active learning based on Method 2 is more efficient for uncertainty reduction in the SPDE’s states (i.e., it leads to a larger reduction of the variance) than active learning using Method 2. address: 'Pacific Northwest National Laboratory, P.O. Box 999, MSIN K7-90, Richland, WA 99352' author: - Ramakrishna Tipireddy - 'David A Barajas-Solano' - Alexandre Tartakovsky bibliography: - 'conditional\_jcp.bib' title: 'Conditional Karhunen-Loève expansion for uncertainty quantification and active learning in partial differential equation models' --- Conditioned Karhunen-Loève expansion ,machine learning ,uncertainty reduction ,uncertainty quantification ,polynomial chaos ,Monte Carlo Introduction {#sec:intro} ============ Uncertainty quantification in partial differential equations (PDE) problems with partially known parameters (e.g., coefficients and source terms) is often performed by modeling these partially known quantities as random variables with appropriate probability distribution functions. Spectral methods such as Polynomial chaos (PC)-based stochastic Galerkin [@RK:Ghanem1991; @Babuka2002] and stochastic collocation [@RK:Xiu2002; @Babuka2010] are commonly used for solving PDEs with random parameters. In spectral methods, random fields are represented in terms of their Karhunen-Loève (KL) expansions. While infinite KL expansions are necessary to exactly represent the two-point statistics of a random field, numerical treatment of PDEs requires truncating KL expansions. These KL expansions are truncated based on the decay of their eigenvalues, so that the random fields can be reconstructed with sufficient accuracy using the retained KL terms. The computational cost of spectral methods exponentially increases with the dimensionality of the stochastic problem (i.e., the number of terms in the truncated KL expansion) [@Venturi2013JCP; @Lin2010JSC; @Lin2010JCP; @Lin2009AWR; @barajassolano-2016-stochastic]. Various approaches have been recently proposed to address this issue: by finding the solution in reduced dimensional spaces using basis adaption [@RK:Tipireddy2014], domain decomposition methods [@TIPIREDDY2017203; @tipireddy2017stochastic], sliced inverse regression [@li2016inverse; @yang2017sliced], and sparsity enhancing together with the active subspace method [@yang2016enhancing], among others. In this work, we propose reducing the computational cost of spectral methods by conditioning the KL expansion on available data. We also demonstrate that conditioning on data reduces uncertainty of predictions, i.e., reduces the variance of quantities of interest of the governing stochastic PDEs. Most of existing UQ methods, including all work referenced above, is based on stochastic models of unknown fields with constant variance and stationary covariance functions, that is, covariance functions of the form $C(x,y) = C(|x-y|)$. The unconditional statistics of fields is estimated from data using the so-called ergodicity assumption, where fields are treated as realizations of a random process, and the spatial statistics of the fields is assumed to be the same as the ensemble statistics of the generating random process. However, there is no reason to assume that the variance at locations where measurements are available should be the same as the variance in locations with no measurements. Gaussian process (GP) regression, also known as kriging, has been used in geostatistics and hydrogeological modeling since its introduction in the sixties [@matheron1963principles] to represent partially observed properties of materials (e.g., permeability of geological porous media) as random field conditioned on observations [@zhang-2002-stochastic; @cressie-2015-geostatistics]. A characteristic feature of GP models of random fields is that their *conditional* statistics, i.e., statistics conditioned on data, are not stationary. For example, the conditional variance of a GP model is a function of space (e.g., in the absence of measurement errors, it is zero at the observation locations), and the conditional covariance function $C^c(x,y)$ depends explicitly on both $x$ and $y$. There are very few studies of PDEs with non-stationary random fields conditioned on data, including the conditional Moment Method [@neuman1993prediction; @morales2006non] limited to parameters with small variances, and a few papers on *conditional* PC methods [@LU2004859; @ossiander2014; @li2014conditional]. Two main approaches have been proposed for conditional spectral methods: (i) first truncating the KL expansion of the random parameters and then conditioning the resulting truncated expansion on data [@ossiander2014], and (ii) first conditioning the infinite KL expansion and then truncating it [@li2014conditional]. Here, we demonstrate that the truncating and conditioning operations do not commute as the two approaches produce different results. A detailed analysis of spectral methods for PDEs with non-stationary random inputs is clearly lacking. In this work, we study the application of conditional KL models to quantify and reduce uncertainty in physical systems modeled with stochastic PDEs (SPDEs). We compare the solution of the SPDE in terms of its conditional mean and variance obtained using both the conditioning first and truncating first approaches, and discuss the merits of both constructions. Our results show that the approach of conditioning first and then truncating (with $N$ terms) approach is more accurate than the approach of truncating first (with $N$ terms) and then conditioning. We also show that for the truncating-first approach the final dimension of the conditional model is reduced from $N$ to $N-N_s$, where $N_s$ is the number of measurements. Furthermore, we adopt active learning [@zhao2017active; @raissi2017inferring] to identify additional observation locations in order to efficiently reduce predictive uncertainty in terms of the variance of the solution of the SPDE. We consider two criteria for identifying observation locations. In the first criterion, we choose the location that minimizes a norm of the variance of the conditional KL expansion of the model parameters. In the second criterion, we propose a novel, GP regression-based approximation to the conditional variance of the solution of the SPDE, and choose locations that minimize this approximation. We demonstrate that the second strategy leads to higher reduction of predictive uncertainty for the same number of additional measurements. This manuscript is organized as follows: In Section \[sec:spde-gp\] we formulate a steady-state stochastic diffusion equation with random coefficient and the GP model for the random coefficient. In Section \[sec:cond\_rand\], we describe two approaches for constructing finite-dimensional conditional KL models of the random coefficient. In Section \[sec:active\_learning\], we present active learning criteria for further reducing uncertainty in conditional KL models. Numerical examples are given in Section \[sec:numerical\] and conclusions are presented in Section \[sec:conclusions\]. Governing Equations {#sec:spde-gp} =================== We study a two-dimensional steady-state diffusion equation with a random diffusion coefficient $k(\mathbf{x}, \omega) : D \times \Omega \to \mathbb{R}$, $D \subset \mathbb{R}^2$ that is bounded and strictly positive: $$\label{RK:eq:rf_bound} 0 < k_l \leq k(\mathbf{x},\omega) \leq k_u < \infty \quad \text{a.e.} \quad \text{in} \quad D \times \Omega.$$ We seek the stochastic solution $u(\mathbf{x},\omega) : D \times \Omega \to \mathbb{R}$ to the problem $$\begin{aligned} \label{RK:eq:spde} \nabla \cdot \left [ k(\mathbf{x},\omega) \nabla u(\mathbf{x},\omega) \right ] &= 0 && \text{in}~D\times \Omega,\\ u(\mathbf{x},\omega) &= u_{\Gamma} && \text{on}~\Gamma \subset \partial D,\\ \nabla u(\mathbf{x},\omega) \cdot n &= u_{\Gamma'} && \text{on}~\Gamma' = \partial D \backslash \Gamma, \end{aligned}$$ where the boundary conditions $u_{\Gamma}$ and $u_{\Gamma'}$ are deterministic and known, and $n$ denotes the outward-pointing unit vector normal to $\partial D$. In many practical applications, the full probabilistic characterization of the coefficient $k(\mathbf{x},\omega)$ is not known, but measurements of $k$ are available at a few spatial locations. In this work, we assume that the distribution of $k$ is known and is log-normal [@Ghanem1999; @zhang-2002-stochastic], i.e., $k(\mathbf{x},\omega) \equiv \exp[g(\mathbf{x},\omega)]$, where $g(\mathbf{x},\omega) \equiv \log k(\mathbf{x}, \omega)$ is a Gaussian random field. We construct the GP prior or *unconditional* (i.e., not conditioned on measurements) model for $g$ employing the following two-step approach common in geostatistics [@cressie-2015-geostatistics]: we first select a parameterized GP prior covariance kernel $C_g(\mathbf{x}, \mathbf{x}' \mid \bm{\theta}) : D \times D \to \mathbb{R}$, with hyperparameters $\bm{\theta}$; and next, we compute an estimate $\hat{\bm{\theta}}$ for the hyperparameters from available measurements of $\log k$ by type-II maximum likelihood estimation [@cressie-2015-geostatistics; @rasmussen2006gaussian]. In the first step, we assume that $g$ is wide-sense stationary with zero mean and squared exponential covariance function $$\begin{gathered} \label{gp_prior} g \sim \mathcal{GP}(0, C_g(\mathbf{x}, \mathbf{x}' \mid \bm{\theta})),\\ \label{cov_g} C_g(\mathbf{x}, \mathbf{x}' \mid \bm{\theta}) = \sigma^2 \exp \left [ -\frac{(x_1-x'_1)^2}{l_1^2} -\frac{(x_2-x'_2)^2}{l_2^2} \right ],\end{gathered}$$ where $\bm{\theta} = \{ \sigma, l_1, l_2 \}$ is the set of hyperparameters of the covariance kernel: $\sigma$ is the standard deviation, and $l_1$ and $l_2$ are the correlation lengths along the $x_1$ and $x_2$ spatial coordinates, respectively. To estimate $\bm{\theta}$, we assume that measurements are of the form $y_i = g(\mathbf{x}^{*}_i) + \epsilon_i$, $i \in [1, N_s]$, where $N_s$ is the number of observations, $\mathbf{x}^{*}_i$ is the measurement location for the $i$th measurement, and the $\epsilon_i \sim \mathcal{N}(0, \sigma_{\epsilon})$ are iid measurement errors with standard deviation $\sigma_{\epsilon}$ independent of $g(\mathbf{x}, \omega)$. The observations and observation locations are arranged into the vector $\mathbf{y} = (y_1, \dots, y_{N_s})^{\top}$ and the matrix $X = (\mathbf{x}^{*}_1, \dots, \mathbf{x}^{*}_{N_s})$, respectively. Similarly, the values of $g$ at the observation locations are arranged into the vector $\mathbf{g} = (g(\mathbf{x}^{*}_1), \dots, (\mathbf{x}^{*}_{N_s}))^{\top}$, so that $\mathbf{y}$ and $\mathbf{g}$ are related by $$\label{eq:g-obs} \mathbf{y} = \mathbf{g} + \bm{\epsilon}, \quad \bm{\epsilon} \sim \mathcal{N}(0, \sigma^2_{\epsilon} \mathbf{I}), \quad \mathbb{E}[ \mathbf{g} \bm{\epsilon}^{\top}] = 0.$$ As stated above, in this work we employ type II maximum likelihood estimation [@cressie-2015-geostatistics; @rasmussen2006gaussian] to estimate the hyperparameters of the GP prior and of the likelihood of the observations: $$\label{eq:typeII-ML} (\bm{\theta}, \sigma_{\epsilon}) \equiv \operatorname*{arg\,max}_{\bm{\theta}, \sigma_{\epsilon}} L(\bm{\theta}, \sigma_{\epsilon}, X, \mathbf{y}),$$ where $L(\bm{\theta}, \sigma_{\epsilon}, X, \mathbf{y})$ is the log-marginal likelihood function $$\label{eq:LML} L(\bm{\theta}, \sigma_{\epsilon}, X, \mathbf{y}) = - \frac{1}{2} y^{\top} | C_s(\bm{\theta}) + \sigma^2_{\epsilon} I |^{-1} y - \frac{1}{2} \ln |C_s(\bm{\theta}) + \sigma^2_{\epsilon} I| - \frac{N_s}{2} \ln 2 \pi,$$ and $C_s(\bm{\theta})$ is the observation covariance matrix with $ij$th entry given by $C_g(\mathbf{x}^{*}_i, \mathbf{x}^{*}_j \mid \bm{\theta})$. To compute the solution of the SPDE  conditioned on the observations $\mathbf{y}$, we adopt the stochastic collocation approach, which requires a finite-dimension stochastic representation of the random field $g(\mathbf{x}, \omega)$. For constructing this representation, we will consider two strategies that rely on GP regression and the KL expansion of random fields: 1. In the first approach [@li2014conditional], the conditional random field $g^c(\mathbf{x},\omega)$ is obtained using GP regression and then discretized by calculating its KL expansion in terms of standard Gaussian random variables. Then, the KL expansion is truncated to an appropriate number of random dimensions $d^c$. 2. In the second approach [@ossiander2014], the unconditioned random field $g(\mathbf{x},\omega)$ is first discretized in terms of its KL expansion and unconditioned standard Gaussian random variables $\bm{\xi}$. Then, conditional Gaussian random variables $\tilde{\bm{\xi}}$, conditioned on the observations $\mathbf{y}$, are obtained by projection. We describe these two approaches in details in Section \[sec:cond\_rand\]. Conditional KL models {#sec:cond_rand} ===================== Approach 1: truncated KL expansion of the conditioned GP field {#sec:trunc-cond-kl} -------------------------------------------------------------- In the first approach, the conditional random field $g^c(\mathbf{x},\omega)$ is approximated with a KL expansion written in terms of standard Gaussian random variables. We assume that the hyperparameters of the prior covariance function $C_g(\mathbf{x}, \mathbf{x}')$ have been estimated and are thus dropped from the notation. The mean and covariance of conditioned Gaussian random field $g^c(\mathbf{x},\omega) \sim \mathcal{GP} \left (\mu_g^c, C_g^c(\mathbf{x}, \mathbf{x}') \right )$ are computed using GP regression [@cressie-2015-geostatistics; @rasmussen2006gaussian]: $$\begin{aligned} \label{eq:cond_mean} \mu_g^c(\mathbf{x}) &= C_g (\mathbf{x}, X) \left [ C_s + \sigma^2_{\epsilon} \right ]^{-1} \mathbf{y},\\ \label{eq:cond_cov} C_g^c(\mathbf{x}, \mathbf{x}') &= C_g(\mathbf{x}, \mathbf{x}') - C_g(\mathbf{x}, X) \left [ C_s + \sigma^2_{\epsilon} \right ]^{-1} C_g(\mathbf{x}, \mathbf{x}').\end{aligned}$$ The conditional field is then expanded using a truncated KL expansion [@RK:Ghanem1991] as $$\label{eq:cond_then_kl} g^c(\mathbf{x},\omega) = \mu^c_g(\mathbf{x}) + \sum_{i=1}^{d^c} \sqrt{\lambda_i^c} \phi^c_i(\mathbf{x}) \xi_i,$$ where $\{ \xi_i\sim \mathcal{N}(0,1) \}^{d^c}_{i = 1}$ are standard Gaussian random variables, and $\{ \lambda_i^c,\phi_i^c \}^{d^c}_{i = 1}$ are the first $d^c$ pairs of eigenvalues and eigenfunctions stemming from the eigenvalue problem $$\int_D C_g^c(\mathbf{x}, \mathbf{x}') \phi^c(\mathbf{x}') \, \mathrm{d} \mathbf{x}' = \lambda^c \phi^c(\mathbf{x}),$$ where $C^c_g$ is given by . The conditional solution of the SPDE, $u^c$, can then be computed using MC or sparse grid collocation methods by sampling $g^c(\mathbf{x},\omega)$ using , or by the PC method by constructing a spectral approximation of $u^c$ in terms of the $\xi_i$, $i \in [1, d^c]$. Approach 2. Conditioning truncated KL expansion of the unconditioned field {#sec:cond-trunc-kl} -------------------------------------------------------------------------- In this approach, introduced in [@ossiander2014], the KL expansion of $g(\mathbf{x}, \omega)$ is first truncated, and the resulting set of random variables are conditioned on the observations $\mathbf{y}$. Here, we demonstrate that this approach reduces the number of random dimensions of the representation of $g^c$ by the number of observations and propose a method for rewriting the representation of $g^c$ in terms of the reduced number of random variables. Effectively, this reduces dimensionality of the SPDE solution. To present our approach we first summarize the conditional KL construction presented in [@ossiander2014] and next describe our proposed reduced-dimension conditional KL representation. We start with the KL expansion of the unconditional Gaussian random field $g(\mathbf{x},\omega)$: $$\label{eq:g-KL} g(\mathbf{x},\omega) = \sum_{i=1}^{\infty} \sqrt{\lambda_i} \phi_i(\mathbf{x}) \xi_i(\omega),$$ where the $\{\xi_i \sim \mathcal{N}(0, 1)\}^{\infty}_{i = 1}$ are iid standard Gaussian random variables. In matrix notation, $$\label{eq:g-KL-matrix} g(\mathbf{x},\omega) = \Phi^{\top}(\mathbf{x}) \Lambda^{1 / 2} \bm{\xi}(\omega),$$ where $\Phi(\mathbf{x}) = (\phi_1(\mathbf{x}), \phi_2(\mathbf{x}), \dots )^{\top}$, $\Lambda = \operatorname{diag}(\lambda_1, \lambda_2, \dots)$, and $\bm{\xi}(\omega)$ is the infinite-dimensional random vector $\bm{\xi}(\omega) = (\xi_1(\omega), \xi_2(\omega), \dots)^{\top}$. By Mercer’s theorem, the covariance function of $g$ can be represented as the infinite sum $$\label{eq:cov-g-mercer} C_g(\mathbf{x}, \mathbf{x}') = \sum_{i=1}^{\infty} \lambda_i \phi_i(\mathbf{x}) \phi_i(\mathbf{x}'),$$ where the eigenpairs $\{ \lambda_i, \phi_i \}^{\infty}_{i = 1}$ are the solutions to the eigenvalue problem $$\int_D C_g(\mathbf{x}, \mathbf{x}') \phi(\mathbf{x}') \, \mathrm{d} \mathbf{x}' = \lambda \phi(\mathbf{x}).$$ In matrix notation, the Mercer expansion  can be written as $$\label{eq:cov-g-mercer-matrix} C_g(\mathbf{x}, \mathbf{x}') = \Phi^{\top}(\mathbf{x}) \Lambda \Phi(\mathbf{x}).$$ In [@ossiander2014], the KL expansion  is conditioned on the observations $\mathbf{y}$ by conditioning $\bm{\xi}$ on these observations as described in the following Eqs –. Evaluating  at $X$ and substituting into  yields $$\label{eq:y-KL} \mathbf{y} = \Phi^{\top}(X) \Lambda^{1 / 2} \bm{\xi} + \bm{\epsilon}.$$ Therefore, the joint distribution of $\mathbf{y}$ and $\bm{\xi}$ is given by $$\label{eq:y-xi-joint} \begin{bmatrix} \mathbf{y} \\ \bm{\xi} \end{bmatrix} \sim \mathcal{N} \left ( \begin{bmatrix} 0 \\ 0 \end{bmatrix}, \begin{bmatrix} C_s + \sigma^2_{\epsilon} I & \Phi^{\top}(X) \Lambda^{1 / 2} \\ \Lambda^{1 / 2} \Phi^{\top}(X) & I \end{bmatrix} \right ),$$ where we have employed the relation $\mathbb{E}[ \mathbf{g} \bm{\epsilon}^{\top}] = 0$ and introduced the notation $C_s \equiv \mathbb{E}[ \mathbf{g} \mathbf{g}^{\top}]$. It follows that the distribution of $\bm{\xi}$ conditioned on the measurements is $\bm{\xi} \mid X, \mathbf{y} \sim \mathcal{N}(\bm{\mu}, M)$, where $$\begin{aligned} \bm{\mu} &= \Lambda^{1 / 2} \Phi(X) \mathbf{y},\\ M &= I - \Lambda^{1 / 2} \Phi(X) (C_s + \sigma^2_{\epsilon} I)^{-1} \Phi^{\top}(X) \Lambda^{1 / 2}.\end{aligned}$$ For simplicity, we denote the GP conditional on $(X, \mathbf{y})$ by $\tilde{g}$, and the conditional random vector by $\tilde{\bm{\xi}} = (\tilde{\xi}_1, \tilde{\xi}_2, \dots)^{\top}$. The conditional GP then reads $$\label{sec:g-KL-cond} \tilde{g}(\mathbf{x},\omega) = \Phi^{\top}(\mathbf{x}) \Lambda^{1 / 2} \tilde{\bm{\xi}}(\omega).$$ We now apply the process of conditioning random variables to the truncated KL expansion of the unconditional random field. The KL expansion of the unconditional random field $g(\mathbf{x},\omega)$, truncated to $d$ terms, reads $$\begin{aligned} \label{eq:trunc-KL-g} g^d(\mathbf{x}, \omega) &= \sum_{i = 1}^{d} \sqrt{\lambda_i} \phi_i(\mathbf{x}) \xi_i(\omega), \\ &= (\Phi^d)^{\top}(\mathbf{x}) (\Lambda^d)^{1 / 2} \bm{\xi}^d(\omega),\end{aligned}$$ where $\Phi^d(\mathbf{x}) = (\phi_1(\mathbf{x}), \dots, \phi_d(\mathbf{x}))^{\top}$, $\Lambda^d = \operatorname{diag}(\lambda_1, \dots, \lambda_d)$, and $\bm{\xi}^d(\omega) = (\xi_1(\omega), \dots, \xi_d(\omega))^{\top}$. This expansion corresponds to the truncated covariance $$\label{eq:cov-g-mercer-trunc} C^d_g(\mathbf{x}, \mathbf{x}') = (\Phi^d)^{\top}(\mathbf{x}) \Lambda^d \Phi^d(\mathbf{x}'),$$ obtained by substituting $\Phi^d(\mathbf{x})$ and $\Lambda^d$ for $\Phi$ and $\Lambda$ in Eq . The truncated representation  can be conditioned on the data $(X, \mathbf{y})$ by following the procedure outlined above, resulting in the conditional model $$\label{eq:trunc-KL-g-cond} \tilde{g}^d(\mathbf{x}, \omega) = (\Phi^d)^{\top}(\mathbf{x}) (\Lambda^d)^{1 / 2} \tilde{\bm{\xi}}^d,$$ where $\tilde{\bm{\xi}}^d \equiv \bm{\xi}^d \mid X, \mathbf{y} \sim \mathcal{N}(\bm{\mu}^d, M^d)$, with $\bm{\mu}^d$ and $M^d$ given by $$\begin{aligned} \label{M_matrix} \bm{\mu}^d &= (\Lambda^d)^{1 / 2} \Phi^d(X) \mathbf{y},\\ M^d &= I - (\Lambda^d)^{1 / 2} \Phi^d(X) (C^d_s + \sigma^2_{\epsilon} I)^{-1} (\Phi^d)^{\top}(X) (\Lambda^d)^{1 / 2},\end{aligned}$$ and where $C^d_s \equiv (\Phi^d)^{\top}(X) \Lambda^d \Phi^d(X)$ is the truncated measurement covariance matrix. Note that both and  employ the same set of eigenpairs derived from the unconditioned covariance function $C_g$. Nevertheless, $\tilde{\bm{\xi}}^d$ and $(\tilde{\xi}_1, \dots, \tilde{\xi}_d)$ in  have different joint distribution, which implies that the conditional truncated model is different than the model  after truncating to $d$ terms. Due to the conditioning on $N_s$ measurements, the rank of $M ^d$ is $$\label{eq:r-dim} \operatorname{rk} (M ^d) = r \equiv d - N_s,$$ so that the model  is effectively of dimension $r$. In other words, conditioning the truncated model results in dimension reduction of the GP model for $g$, and reduction of the stochastic dimensionality of the SPDE problem. To leverage this dimension reduction, we propose rewriting the model  in terms of $r$ random variables. We write the eigendecomposition of $M^d$ as $M^d = Q D Q^{-1}$, where $D$ is the diagonal matrix of the form $$\label{eq:D-block-form} D = \begin{bmatrix} D^r& 0\\ 0 & 0 \end{bmatrix},$$ since $\operatorname{rk}(M^d) = r$. Substituting the eigendecomposition of $M^d$ into , we obtain $$\label{eq:trunc-KL-g-cond-r} \tilde{g}^d(\mathbf{x}, \omega) = \hat{g}^d(\mathbf{x}) + (\Phi^d)^{\top}(\mathbf{x}) (\Lambda^d)^{1 / 2} Q D Q^{-1} \bm{\zeta}^d(\omega),$$ where $\hat{g}^d(\mathbf{x}) \equiv (\Phi^d)^{\top}(\mathbf{x}) (\Lambda^d)^{1 / 2} \bm{\mu}^d$ and $\bm{\zeta}^d \sim \mathcal{N}(0, I_d)$. Let $\bm{\eta} = Q^{-1}\bm{\zeta}^d$; then, substituting $\bm{\eta}$ into , we obtain $$\begin{gathered} \tilde{g}^d(\mathbf{x},\omega)- \hat{g}^d(\mathbf{x}) = \\ \begin{bmatrix} (\Phi^d_r)^{\top}(\mathbf{x}) & (\Phi^d)^{\top}_{r'}(\mathbf{x}) \end{bmatrix} \begin{bmatrix} \Lambda^d_r & 0\\ 0 & \Lambda^d_{r'} \end{bmatrix}^{1 / 2} \begin{bmatrix} Q_r & Q_{rr'}\\ Q_{r'r} & Q_{r'} \end{bmatrix} \begin{bmatrix} D_r\bm{\eta}_r \\ 0 \end{bmatrix}.\end{gathered}$$ Similarly, let $\tilde{\bm{\eta}}_r = D_r \bm{\eta}_r$; then, $\tilde{g}^d$ can be expanded in terms of the reduced set of random variables $\tilde{\bm{\eta}}_r$ as $$\begin{aligned} \tilde{g}^d(\mathbf{x},\omega) - \hat{g}(\mathbf{x}) &= \begin{bmatrix} (\Phi^d_r)^{\top}(\mathbf{x}) & (\Phi^d)^{\top}_{r'}(\mathbf{x}) \end{bmatrix} \begin{bmatrix} \Lambda^d_r & 0\\ 0 & \Lambda^d_{r'} \end{bmatrix}^{1 / 2} \begin{bmatrix} Q_r & Q_{rr'}\\ Q_{r'r} & Q_{r'} \end{bmatrix} \begin{bmatrix} \tilde{\bm{\eta}}_r \\ 0 \end{bmatrix}\\ &= [(\Phi^d_r)^{\top}(\mathbf{x}) (\Lambda^d_r)^{1 / 2} Q_r + (\Phi^d_{r'})^{\top}(\mathbf{x}) (\Lambda^d_{r'})^{1 / 2} Q_{r'r}] \tilde{\bm{\eta}}_r. \end{aligned}$$ Here, the components of the random vector $\tilde{\bm{\eta}}_r$ are correlated with one another; nevertheless, $\tilde{\bm{\eta}}_r$ can be converted to a set of uncorrelated random variables using an orthogonalization method such as Gram-Schmidt process, so that $\bm{\zeta}^r = A \tilde{\bm{\eta}}_r$ and $\tilde{\bm{\eta}}_r = A^{-1} \bm{\zeta}^r,$ where $\bm{\zeta}^r \sim \mathcal{N}(0, I_r)$. Now, $\tilde{g}^d$ can be expanded in terms of the new set of random variables $\bm{\zeta}^r$ as $$\label{eq:trunc-KL-g-cond-Psi-r} \tilde{g}^d(\mathbf{x},\omega) = \hat{g}^d(\mathbf{x}) + (\tilde{\Psi}^r)^{\top}(\mathbf{x}) \bm{\zeta}^r,$$ where $$\label{eq:Psi-r} (\tilde{\Psi}^r)^{\top}(\mathbf{x}) = (\Phi^d_r)^{\top}(\mathbf{x}) (\Lambda^d_r)^{1 / 2} Q_r A^{-1} + (\Phi^d_{r'})^{\top}(\mathbf{x}) (\Lambda^d_{r'})^{1 / 2} Q_{r'r} A^{-1}.$$ The $r$ components of $\bm{\zeta}^r$ are uncorrelated Gaussian random variables and hence are also independent. Therefore, the conditional KL expansion  can be used in combination with standard PC-based or stochastic collocation-based methods for solving stochastic PDEs. Active learning for uncertainty reduction {#sec:active_learning} ========================================= The conditioning of $g(\mathbf{x},\omega)$ KL expansion presented in Section \[sec:cond\_rand\] leads to reduced uncertainty in $u(\mathbf{x},\omega)$. This uncertainty can be further reduced by collecting additional measurements of $g$. Giving limited ability to collect additional $g$ measurements, it is important to identify locations for additional measurements that optimally minimize the uncertainty in $u$. This process of optimal data acquisition is referred to as *active learning* [@cohn-1996-active; @zhao2017active; @raissi2017inferring]. In this section, we discuss the standard active learning method (Method 1) based on minimizing the variance of $g^c$  [@zhao2017active; @raissi2017inferring] and propose an alternative approach (Method 2) based on minimizing the variance of $u^c$. In Method 1, the variance of $u^c$ is reduced by minimizing the variance of $g^c$, but, as we show below, it does not lead to the maximum reduction of the $u^c$ variance. It is worth noting that active learning also reduces the cost of computing the conditional solution $u^c$, as solving SPDE problems with coefficients having smaller variance requires smaller number of MC simulations, lower order of polynomial chaos in the stochastic Galerkin method, and lower sparse grid level in the stochastic collocation method. Method 1: minimization of the conditional variance of $g^c$ {#sec:active-learning-g} ----------------------------------------------------------- In the standard active learning method [@zhao2017active; @raissi2017inferring], a new location $\mathbf{x}^{*}$ for sampling $g$ is selected to minimize the variance of the conditional model $g^c$ conditioned on the full set of observations, including the new observation. This can be done in closed form as, per Eq. , the conditional variance $C^c_g$ depends only on the observation locations and not the observed values. Therefore, the new observation location $\mathbf{x}^{*}$ is selected following the acquisition policy $$\label{eq:active-learning-criteria-g} \mathbf{x}^* \equiv \operatorname*{arg\,min}_{\mathbf{x}'} \int_D C^c_g \left ( \mathbf{x} , \mathbf{x} \mid [X, \mathbf{x}'] \right ) \, \mathrm{d} \mathbf{x},$$ where $C^c_g(\cdot, \cdot \mid [X, \mathbf{x}'])$ denotes the covariance of $g$ conditioned on the original set of observation locations, $X$, and a new location $\mathbf{x}'$, and is computed using . In practice, the minimization problem  is approximately solved by choosing $\mathbf{x}^*$ as $\operatorname*{arg\,max}_{\mathbf{x}} C^c_g \left (\mathbf{x}, \mathbf{x} \mid X \right)$ [@zhao2017active; @raissi2017inferring]. Reducing the variance of $g^c$ also reduces the variance of $u^c$. Nevertheless, there is no reason to assume that the observation locations provided by the acquisition strategy  will lead to optimal variance reduction for $u^c$. Method 2: By minimizing the conditional covariance of $u^c$ {#sec:active-learning-u} ----------------------------------------------------------- In Method 2, we chose location for measurements that minimize the conditional variance of $u$, that is, $$\label{eq:active-learning-criteria-u} \mathbf{x}^{*} \equiv \operatorname*{arg\,min}_{\mathbf{x}'} \int_D {C}^c_u(\mathbf{x}, \mathbf{x} \mid [X, \mathbf{x}']) \, \mathrm{d} \mathbf{x},$$ where $C^c_u(\cdot, \cdot \mid [X, \mathbf{x}'])$ denotes the covariance of $u$ conditioned on the original set of observation locations, $X$, and a new location $\mathbf{x}'$. Note that solving this minimization problem is significantly more challenging than solving , as it requires (i) constructing the conditional model $g^c$ conditioned on the new observation, and (ii) forward uncertainty propagation of $g^c$ through the SPDE problem to estimate $C^c_u$. Here, step (i) is the most challenging as it requires knowledge of $g(\mathbf{x}')$ in order to construct the conditioned model $g^c$  [@cohn-1996-active]. Note that $g(\mathbf{x})$ is only partially known; otherwise, this problem would not be uncertain. As an alternative, we propose modeling $g^c$ and $u^c$ as components of a bivariate Gaussian field $h^c$ in order to derive an approximation to $C^c_u(\mathbf{x}, \mathbf{x} \mid [X, \mathbf{x}'])$ that does not require knowledge of $g(\mathbf{x}')$. Specifically, let $g^c$ and $u^c$ denote the fields conditioned on the original data set $(\mathbf{x}, \mathbf{y})$, and let $h^c(\mathbf{x}, \omega) = [g^c(\mathbf{x}, \omega), u^c(\mathbf{x}, \omega)]^{\top}$ be a bivariate Gaussian field. It is possible to compute the conditional covariance of $h^c$ conditioned on an additional measurement location $x'$ by employing . After marginalizing the $u^c$ component, we obtain the approximation $$\begin{gathered} \label{eq:GP_cov_u} C^c_u(\mathbf{x}, \mathbf{x} \mid [X, \mathbf{x}']) \approx C^c_u(\mathbf{x}, \mathbf{x} \mid X )\\ - C^c_{ug}(\mathbf{x}, \mathbf{x}' \mid X ) [ C^c_g(\mathbf{x}', \mathbf{x}' \mid X) ]^{-1} C^c_{gu}(\mathbf{x}', \mathbf{x} \mid X),\end{gathered}$$ where $C^c_{ug}(\mathbf{x}, \mathbf{x}' \mid X)$ denotes the $u$–$g$ cross-covariance conditioned on $X$. Note that this approximation only requires the observation location $x'$ and not the observation value $g(\mathbf{x}')$. In order to apply the approximation , it is necessary to compute the conditional covariances $C^c_{u}(\mathbf{x}, \mathbf{x}' \mid X)$ and $C^c_{ug}(\mathbf{x}, \mathbf{x}' \mid X)$, which, unlike $C^c_{g}(\mathbf{x}, \mathbf{x}' \mid X)$ given by , are not available in closed form; therefore, we compute sample approximations of these covariances. For this purpose, we draw $M$ realizations of $g^c(\mathbf{x}, \omega)$ with conditioned mean and covariance given by and , resulting in the ensemble of synthetic fields $\{ g^c_{(i)}(\mathbf{x}) \equiv g^c(\mathbf{x}, \omega_{(i)}) \}^M_{i = 1}$. For each member of the ensemble we solve the corresponding deterministic PDE problem, resulting in the ensemble of solutions $\{ u^c_{(i)} \}^M_{i = 1}$. Employing both ensembles, we compute the sample covariances $\hat{C}^c_g(\cdot, \cdot)$, $\hat{C}^c_u(\cdot, \cdot)$ and $\hat{C}^c_{ug}(\cdot, \cdot)$. Substituting these sample covariances into , and the result into , leads to the acquisition policy $$\mathbf{x}^{*} \equiv \operatorname*{arg\,min}_{\mathbf{x}'} \int_D \left \{ \hat{C}^c_u(\mathbf{x}, \mathbf{x}) - \hat{C}^c_{ug}(\mathbf{x}, \mathbf{x}') [ \hat{C}^c_g(\mathbf{x}', \mathbf{x}') ]^{-1} \hat{C}^c_{gu}(\mathbf{x}', \mathbf{x}) \right \} \, \mathrm{d} \mathbf{x}. \label{minimization}$$ Numerical Experiments {#sec:numerical} ===================== In this section, we apply the conditional KL modeling approaches presented in Section \[sec:cond\_rand\] and the active learning methods presented in Section \[sec:active\_learning\] to solve the stochastic diffusion equation problem . We consider the following two-dimensional steady state diffusion equation with a random diffusion coefficient over the domain $D = [0,2] \times [0,1]$, subject to Dirichlet and Neumann boundary conditions: $$\label{RK:eq:spde_nr} \begin{aligned} -\nabla \cdot (k(\mathbf{x},\omega) \nabla u(\mathbf{x},\omega)) &= f(\mathbf{x},\omega) && u(\mathbf{x},\omega)\in D\times \Omega,\\ u(\mathbf{x},\omega) &= 1 && x_1 = 0,\\ u(\mathbf{x},\omega) &= 0 && x_1 = 2,\\ n \cdot \nabla u(\mathbf{x},\omega) &= 0 && x_2 = \{ 0, 1 \}.\\ \end{aligned}$$ The reference $g = \log k$ field is constructed synthetically by drawing a realization of the GP process and with the parameters $\sigma_g = 0.65, l_x=0.15$ and $l_y=0.2$, shown in Figure \[obs\_sample\_nobs40\_mrst\]. From this reference field, we draw $40$ observations at random locations to be used for constructing the conditional GP model $g^c$ and computing conditional solution of Eq . [0.48]{} ![Synthetic field $k=\exp(g)$ and locations of the observations.[]{data-label="obs_sample_nobs40_mrst"}](figures/g_obs_sample_nobs40_mrst "fig:") [0.48]{} ![Synthetic field $k=\exp(g)$ and locations of the observations.[]{data-label="obs_sample_nobs40_mrst"}](figures/k_obs_sample_nobs40_mrst "fig:") Conditional GP models {#sec:numerical-conditional-gp} --------------------- We construct finite-dimensional conditional GP models for the field $g$ based on the synthetic dataset using the two approaches introduced in Section \[sec:cond\_rand\] and employ these models to compute the mean and standard deviation of the conditional solution $u^c$ of the SPDE problem . We employ MC simulations to compute reference unconditional and conditional mean and standard deviation of $g$ and $u$. We use Approach 2 of Section \[sec:cond-trunc-kl\], which provides a quantitative measure of dimension reduction due to conditioning, to estimate the dimensionality of the reduced conditional model $g^c$. We construct finite-dimensional models for $g^c$ of the estimated dimension using both Approaches 1 and 2 and propagate uncertainty through the SPDE problem  using the stochastic collocation method [@RK:Xiu2002; @Babuka2010]. ### Reference Monte Carlo unconditional and conditional solutions {#sec:numerical-reference} To sample the reference unconditional $g$ field, we employ the unconditional KL expansion truncated to $d = 60$ terms, where $d$ is chosen such as to retain $99\%$ of the unconditional variance, i.e., $\sum_{i=1}^d \lambda_i \geq 0.99 \operatorname{Tr}(C_g)$, where $\operatorname{Tr}(C_g)$ is the trace of $C_g$. The unconditional mean and covariance of $u$ are computed from MC with samples. The resulting unconditional mean and standard deviation of $g$ and $u$ are presented in Figure \[fig:u\_mc\_001\_uncond\_mrst\_no\_obs\_plot\]. In Figure \[fig:u\_xi\_mc\_uncond\_mean\_sdev\_norm\_15k\] we present a convergence study of the MC estimators of the unconditional mean and standard deviation of $u$, where the $L_2$ norm of the estimators is plotted against the number of MC samples. This figure demonstrates that samples are sufficient to compute these estimators. Similarly, to sample the reference conditional field, we compute the mean and covariance using Eqs and , compute the KL expansion of the conditional covariance, and truncate the resulting conditional KL expansion to $d = 53$ terms to retain $99\%$ of the conditional variance. The conditional mean and covariance of $g$ and $u$ are computed using MC with samples. The results are presented in Figure \[fig:g\_u\_mean\_exact\_cond\_mc\_mrst\]. The convergence study of the MC estimators of the conditional mean and standard deviation of $u$ is shown in Figure \[fig:u\_xi\_mc\_cond\_mean\_sdev\_norm\_15k\] and shows that 15000 samples are sufficient to compute these estimators. [0.48]{} ![Unconditional (a) mean of $g(\mathbf{x},\omega)$, (b) standard deviation of of $g(\mathbf{x},\omega)$, (c) mean of $u(\mathbf{x},\omega)$, and (d) standard deviation of $u(\mathbf{x},\omega)$ computed via MC simulation with realizations.[]{data-label="fig:u_mc_001_uncond_mrst_no_obs_plot"}](figures/g_mean_mc_001_uncond_mrst "fig:") [0.48]{} ![Unconditional (a) mean of $g(\mathbf{x},\omega)$, (b) standard deviation of of $g(\mathbf{x},\omega)$, (c) mean of $u(\mathbf{x},\omega)$, and (d) standard deviation of $u(\mathbf{x},\omega)$ computed via MC simulation with realizations.[]{data-label="fig:u_mc_001_uncond_mrst_no_obs_plot"}](figures/g_sdev_mc_001_uncond_mrst "fig:") [0.48]{} ![Unconditional (a) mean of $g(\mathbf{x},\omega)$, (b) standard deviation of of $g(\mathbf{x},\omega)$, (c) mean of $u(\mathbf{x},\omega)$, and (d) standard deviation of $u(\mathbf{x},\omega)$ computed via MC simulation with realizations.[]{data-label="fig:u_mc_001_uncond_mrst_no_obs_plot"}](figures/u_mean_mc_001_uncond_mrst_no_obs_plot "fig:") [0.48]{} ![Unconditional (a) mean of $g(\mathbf{x},\omega)$, (b) standard deviation of of $g(\mathbf{x},\omega)$, (c) mean of $u(\mathbf{x},\omega)$, and (d) standard deviation of $u(\mathbf{x},\omega)$ computed via MC simulation with realizations.[]{data-label="fig:u_mc_001_uncond_mrst_no_obs_plot"}](figures/u_sdev_mc_001_uncond_mrst_no_obs_plot "fig:") [0.95]{} ![$L_2$ norm of unconditional (a) mean and (b) standard deviation of the unconditional solution $u(\mathbf{x},\omega)$, versus the number of MC realizations.[]{data-label="fig:u_xi_mc_uncond_mean_sdev_norm_15k"}](figures/u_xi_mc_uncond_mean_norm_15k "fig:") [0.95]{} ![$L_2$ norm of unconditional (a) mean and (b) standard deviation of the unconditional solution $u(\mathbf{x},\omega)$, versus the number of MC realizations.[]{data-label="fig:u_xi_mc_uncond_mean_sdev_norm_15k"}](figures/u_xi_mc_uncond_sdev_norm_15k "fig:") [0.48]{} ![(a) Mean and (b) standard deviation of $g^c(\mathbf{x},\omega)$ obtained from KL expansion with $53$ terms, and the corresponding (c) mean and (d) standard deviation of $u^c(\mathbf{x},\omega)$ computed via MC simulation with realizations[]{data-label="fig:g_u_mean_exact_cond_mc_mrst"}](figures/g_mean_exact_cond_mc_mrst "fig:") [0.48]{} ![(a) Mean and (b) standard deviation of $g^c(\mathbf{x},\omega)$ obtained from KL expansion with $53$ terms, and the corresponding (c) mean and (d) standard deviation of $u^c(\mathbf{x},\omega)$ computed via MC simulation with realizations[]{data-label="fig:g_u_mean_exact_cond_mc_mrst"}](figures/g_sdev_exact_cond_mc_mrst "fig:") [0.48]{} ![(a) Mean and (b) standard deviation of $g^c(\mathbf{x},\omega)$ obtained from KL expansion with $53$ terms, and the corresponding (c) mean and (d) standard deviation of $u^c(\mathbf{x},\omega)$ computed via MC simulation with realizations[]{data-label="fig:g_u_mean_exact_cond_mc_mrst"}](figures/u_mean_exact_cond_mc_mrst_no_obs_plot "fig:") [0.48]{} ![(a) Mean and (b) standard deviation of $g^c(\mathbf{x},\omega)$ obtained from KL expansion with $53$ terms, and the corresponding (c) mean and (d) standard deviation of $u^c(\mathbf{x},\omega)$ computed via MC simulation with realizations[]{data-label="fig:g_u_mean_exact_cond_mc_mrst"}](figures/u_sdev_exact_cond_mc_mrst_no_obs_plot "fig:") [0.95]{} ![$L_2$ norm of (a) mean and (b) standard deviation of the conditional solution $u(\mathbf{x},\omega)$, versus the number of realizations in the MC solution (black line) and stochastic collocation points (red circles). The number of collocation points are , and corresponding to the sparse grid levels 2, 3, and 4, respectively.[]{data-label="fig:u_xi_mc_cond_mean_sdev_norm_15k"}](figures/u_xi_mc_quan_cond_mean_norm_15k.pdf "fig:") [0.95]{} ![$L_2$ norm of (a) mean and (b) standard deviation of the conditional solution $u(\mathbf{x},\omega)$, versus the number of realizations in the MC solution (black line) and stochastic collocation points (red circles). The number of collocation points are , and corresponding to the sparse grid levels 2, 3, and 4, respectively.[]{data-label="fig:u_xi_mc_cond_mean_sdev_norm_15k"}](figures/u_xi_mc_quad_cond_sdev_norm_15k.pdf "fig:") [0.48]{} ![Absolute point error in (a) mean of $g^c(\mathbf{x},\omega)$, (b) standard deviation of $g^c(\mathbf{x},\omega)$, (c). mean of $u^c(\mathbf{x},\omega)$, and (d) standard deviation of $u^c(\mathbf{x},\omega)$ obtained with KL expansion truncated first and then conditioned (Approach 2) and the sparse grid collocation method.[]{data-label="g_gauss_xi_d20_red_cond_mrst"}](figures/gpm_est_exact-g_mu_gauss_xi_error "fig:") [0.48]{} ![Absolute point error in (a) mean of $g^c(\mathbf{x},\omega)$, (b) standard deviation of $g^c(\mathbf{x},\omega)$, (c). mean of $u^c(\mathbf{x},\omega)$, and (d) standard deviation of $u^c(\mathbf{x},\omega)$ obtained with KL expansion truncated first and then conditioned (Approach 2) and the sparse grid collocation method.[]{data-label="g_gauss_xi_d20_red_cond_mrst"}](figures/C_pg_est_exact-g_sdev_gauss_xi_error "fig:") [0.48]{} ![Absolute point error in (a) mean of $g^c(\mathbf{x},\omega)$, (b) standard deviation of $g^c(\mathbf{x},\omega)$, (c). mean of $u^c(\mathbf{x},\omega)$, and (d) standard deviation of $u^c(\mathbf{x},\omega)$ obtained with KL expansion truncated first and then conditioned (Approach 2) and the sparse grid collocation method.[]{data-label="g_gauss_xi_d20_red_cond_mrst"}](figures/u_mu_exact_cond_mc-u_mu_gauss_xi_error_no_obs_plot "fig:") [0.48]{} ![Absolute point error in (a) mean of $g^c(\mathbf{x},\omega)$, (b) standard deviation of $g^c(\mathbf{x},\omega)$, (c). mean of $u^c(\mathbf{x},\omega)$, and (d) standard deviation of $u^c(\mathbf{x},\omega)$ obtained with KL expansion truncated first and then conditioned (Approach 2) and the sparse grid collocation method.[]{data-label="g_gauss_xi_d20_red_cond_mrst"}](figures/u_sdev_exact_cond_mc-u_sdev_gauss_xi_error_no_obs_plot "fig:") The comparison of Figures \[RK:fig:g\_sdev\_mc\_001\_uncond\_mrst\] and \[RK:fig:g\_sdev\_exact\_cond\_mc\_mrst\] show that the standard deviation of $g$ is reduced after conditioning on $g$ measurements. Similarly, the standard deviation of $u$, shown in Figures \[RK:fig:u\_sdev\_mc\_001\_uncond\_mrst\_no\_obs\_plot\] and \[RK:fig:u\_sdev\_exact\_cond\_mc\_mrst\_no\_obs\_plot\], also is reduced after conditioning on $g$ measurements. ### Conditional solution computed using Approach 2: Conditioning truncated KL expansion of the unconditioned field {#sec:numerical-approach-2} In this section we employ Approach 2, presented in Section \[sec:cond-trunc-kl\], for constructing a finite-dimensional conditional KL model of the random coefficient of the SPDE problem . In Section \[sec:numerical-reference\], we determined that 60 dimensions are required to represent the unconditional $g$ field. By virtue of , it follows that conditioning these $d = 60$ dimensions on $N_s = 40$ observations reduces the dimensionality of the $g^c$ KL representation to $r = 20$. As this approach for constructing a discretized conditional model disregards information provided by the higher eigenpairs of the unconditional expansion, we expect the resulting conditional moments (mean and standard deviation) of $g^c$ to differ from the reference moments computed in Section \[sec:numerical-reference\]. The absolute point-wise errors in the mean and standard deviation of $g^c$ are shown in Figures \[RK:fig:gpm\_est\_exact-g\_mu\_gauss\_xi\_error\] and \[RK:fig:C\_pg\_est\_exact-g\_sdev\_gauss\_xi\_error\], respectively. Next, we employ the conditional KL expansion to estimate the mean and variance of $u^c$ using the sparse grid collocation method [@RK:Xiu2002; @Babuka2010] with , , and quadrature points. Figure \[fig:u\_xi\_mc\_cond\_mean\_sdev\_norm\_15k\] shows the $L_2$ norm of the estimators. It can be seen that points are sufficient to obtain a convergent solution for both mean and standard deviation of $u^c$. Absolute point-wise errors of the mean and standard deviation of $u^c$ with respect to the reference conditional solution are shown in Figures \[RK:fig:u\_mu\_exact\_cond\_mc-u\_mu\_gauss\_xi\_error\_no\_obs\_plot\] and \[RK:fig:u\_sdev\_exact\_cond\_mc-u\_sdev\_gauss\_xi\_error\_no\_obs\_plot\], respectively. ### Conditioned solution computed using Approach 1: Truncated KL expansion of the conditional field {#sec:numerical-approach-1} Next, we apply Approach 1 for constructing a finite-dimensional conditional KL model for the random coefficient of the SPDE problem . In the previous section, it was determined that the conditional KL expansion obtained with Approach 2 has 20 dimensions. To compare the accuracy of the two conditional KL approaches, here we truncate the conditional KL expansion of $g^c$ to $20$ dimensions. Note that in Section \[sec:numerical-reference\] we demonstrated that $53$ dimensions are necessary to retain $99\%$ of the variance of the exact conditional field $g^c$. Therefore, by truncating the KL expansion to $20$ dimensions, we incur in the absolute point-wise errors in the conditional mean and standard deviation of $g^c$ shown in Figs. \[RK:fig:gpm\_est\_exact-g\_mu\_gauss\_xi\_cond\_trunc\_error\] and \[RK:fig:C\_pg\_est\_exact-g\_sdev\_gauss\_xi\_cond\_trunc\_error\]. As in the previous section, we employ the conditional KL expansion to estimate the mean and variance of $u^c$ using the sparse grid collocation method [@RK:Xiu2002; @Babuka2010]. Figures \[RK:fig:u\_mu\_exact\_cond\_mc-u\_mu\_gauss\_xi\_cond\_trunc\_error\_no\_obs\_plot\] and \[RK:fig:u\_sdev\_exact\_cond\_mc-u\_sdev\_gauss\_xi\_cond\_trunc\_error\_no\_obs\_plot\] show the absolute point-wise error in the mean and standard deviation of $u^c$ with respect to the reference moments computed in Section \[sec:numerical-reference\]. [0.48]{} ![Error in (a) mean of $g^c(\mathbf{x},\omega)$, (b) standard deviation of $g^c(\mathbf{x},\omega)$, (c) mean of $u^c(\mathbf{x},\omega)$, and (d) standard deviation of $u^c(\mathbf{x},\omega)$ obtained using conditioning first and then truncated KL expansion (Approach 1) and the sparse grid collocation method.[]{data-label="g_u_gauss_xi_d20_cond_trunc_mrst."}](figures/gpm_est_exact-g_mu_gauss_xi_cond_trunc_error "fig:") [0.48]{} ![Error in (a) mean of $g^c(\mathbf{x},\omega)$, (b) standard deviation of $g^c(\mathbf{x},\omega)$, (c) mean of $u^c(\mathbf{x},\omega)$, and (d) standard deviation of $u^c(\mathbf{x},\omega)$ obtained using conditioning first and then truncated KL expansion (Approach 1) and the sparse grid collocation method.[]{data-label="g_u_gauss_xi_d20_cond_trunc_mrst."}](figures/C_pg_est_exact-g_sdev_gauss_xi_cond_trunc_error "fig:") [0.48]{} ![Error in (a) mean of $g^c(\mathbf{x},\omega)$, (b) standard deviation of $g^c(\mathbf{x},\omega)$, (c) mean of $u^c(\mathbf{x},\omega)$, and (d) standard deviation of $u^c(\mathbf{x},\omega)$ obtained using conditioning first and then truncated KL expansion (Approach 1) and the sparse grid collocation method.[]{data-label="g_u_gauss_xi_d20_cond_trunc_mrst."}](figures/u_mu_exact_cond_mc-u_mu_gauss_xi_cond_trunc_error_no_obs_plot "fig:") [0.48]{} ![Error in (a) mean of $g^c(\mathbf{x},\omega)$, (b) standard deviation of $g^c(\mathbf{x},\omega)$, (c) mean of $u^c(\mathbf{x},\omega)$, and (d) standard deviation of $u^c(\mathbf{x},\omega)$ obtained using conditioning first and then truncated KL expansion (Approach 1) and the sparse grid collocation method.[]{data-label="g_u_gauss_xi_d20_cond_trunc_mrst."}](figures/u_sdev_exact_cond_mc-u_sdev_gauss_xi_cond_trunc_error_no_obs_plot "fig:") Comparing Figures \[g\_gauss\_xi\_d20\_red\_cond\_mrst\] and \[g\_u\_gauss\_xi\_d20\_cond\_trunc\_mrst.\], it can be seen that, for a set number of stochastic dimensions of the $g^c$ KL expansion, Approach 1 provides a more accurate approximation of the moments of $g^c$ and $u^c$. On the other hand, through Eq. , Approach 2 provides a *priori* estimate of the dimensionality of the conditional KL model sufficient to obtain an accurate solution. In order to study why Approach 1 provides a more accurate approximation of conditional moments, we compare the eigendecompositions of $g^c(x,\omega)$ provided by Approaches 1 and 2. Here, we note that Approach 2 does not provide an explicit eigendecomposition (cf. Eq. ). Therefore, we compute the corresponding eigendecomposition by first computing the covariance matrix induced by Approach 2, given by $(\tilde{\Psi}^r)^{\top} (\tilde{\Psi}^r)$, and then compute its eigendecomposition. Figure \[u\_xi\_mc\_eigen\_uncod\_cond\] shows the eigenvalues of $g^c(x,\omega)$ resulting from Approaches 1 and 2, together with the eigenvalues of $g(x,\omega)$. It can be seen that the magnitude of the eigenvalues of $g^c(x,\omega)$ is smaller than those of $g(x,\omega)$, which follows from the fact that the variance of the conditional KL expansions is smaller than the variance of the unconditional KL expansion. It can also be seen that the conditional eigenvalues decay faster than the unconditional eigenvalues, especially for the first ten eigenvalues. Figure \[fig:u\_xi\_mc\_eigen\_cond\_eig\_func\] shows the first and second eigenfunctions of $g^c(x,\omega)$ resulting from Approaches 1 and 2, together with the eigenvalues of $g(x,\omega)$. Here, we note that, by construction, the eigenfunctions of Approach 2 are calculated from the first 20 eigenfunctions of $g(\mathbf{x}, \omega)$. In contrast, in Approach 1, all the eigenfunctions of $g(\mathbf{x}, \omega)$ contribute to the first 20 eigenfunctions of $g^c(\mathbf{x}, \omega)$; therefore, the eigenfunctions of Approach 1 can resolve finer-scale features than the eigenfunctions of Approach 2. We attribute the superior accuracy of Approach 1 for approximating conditional moments to its superior capacity for resolving fine-scale features of $g^c(\mathbf{x}, \omega)$. [0.48]{} ![Eigenvalue decay of $g(\mathbf{x}, \omega)$ (black) and $g^c(\mathbf{x}, \omega)$ (Approach 1, red, Approach 2, blue) random fields. (a) First 60 eigenvalues and (b) first 20 eigenvalues.[]{data-label="u_xi_mc_eigen_uncod_cond"}](figures/u_xi_mc_eigen_uncod_cond.pdf "fig:") [0.48]{} ![Eigenvalue decay of $g(\mathbf{x}, \omega)$ (black) and $g^c(\mathbf{x}, \omega)$ (Approach 1, red, Approach 2, blue) random fields. (a) First 60 eigenvalues and (b) first 20 eigenvalues.[]{data-label="u_xi_mc_eigen_uncod_cond"}](figures/u_xi_mc_eigen_uncod_cond_red.pdf "fig:") [0.48]{} ![First (left) and second (right) eigenfunctions of $g(\mathbf{x},\omega)$ and $g^c(\mathbf{x},\omega)$[]{data-label="fig:u_xi_mc_eigen_cond_eig_func"}](figures/u_xi_mc_eigen_uncond_eig_func_1.pdf "fig:") [0.48]{} ![First (left) and second (right) eigenfunctions of $g(\mathbf{x},\omega)$ and $g^c(\mathbf{x},\omega)$[]{data-label="fig:u_xi_mc_eigen_cond_eig_func"}](figures/u_xi_mc_eigen_uncond_eig_func_2 "fig:") [0.48]{} ![First (left) and second (right) eigenfunctions of $g(\mathbf{x},\omega)$ and $g^c(\mathbf{x},\omega)$[]{data-label="fig:u_xi_mc_eigen_cond_eig_func"}](figures/u_xi_mc_eigen_cond_eig_func_approach1_1.pdf "fig:") [0.48]{} ![First (left) and second (right) eigenfunctions of $g(\mathbf{x},\omega)$ and $g^c(\mathbf{x},\omega)$[]{data-label="fig:u_xi_mc_eigen_cond_eig_func"}](figures/u_xi_mc_eigen_cond_eig_func_approach1_2.pdf "fig:") [0.48]{} ![First (left) and second (right) eigenfunctions of $g(\mathbf{x},\omega)$ and $g^c(\mathbf{x},\omega)$[]{data-label="fig:u_xi_mc_eigen_cond_eig_func"}](figures/u_xi_mc_eigen_cond_eig_func_1.pdf "fig:") [0.48]{} ![First (left) and second (right) eigenfunctions of $g(\mathbf{x},\omega)$ and $g^c(\mathbf{x},\omega)$[]{data-label="fig:u_xi_mc_eigen_cond_eig_func"}](figures/u_xi_mc_eigen_cond_eig_func_2 "fig:") Active learning {#sec:active-learning} --------------- Here, we apply the two active learning data acquisition methods presented in Section \[sec:active\_learning\] to identify additional measurement locations for the $g$ field. We use these additional observations together with the previously available observations to construct the conditional KL expansion of $g$ using Approach 1, as it was shown in Section \[sec:numerical-approach-1\] to be more accurate than Approach 2 for the considered application. Here, we consider again the SPDE problem , and we aim to explore the behavior of the presented data acquisition methods for two choices of $\sigma_g$, namely $0.65$ and $1.3$. For $\sigma_g = 0.65$, Figure \[norm\_g\_u\_sdv\_k\_and\_k\_u\] shows the $L^2$-norm of the standard deviation of $g^c$ and $u^c$ as a function of the number of additional measurements, $N_{am}$ identified with both active learning methods. In Method 2, the conditional covariances $\hat{C}^c_{u}$ and $\hat{C}^c_{ug}$ are computed using MC realizations. This small number of MC realizations is justified as empirical observation shows that very accurate estimates of the covariances are not necessary for obtaining good estimates of the new observation locations from the minimization problem . Figure \[RK:fig:norm\_g\_sdv\_k\_and\_k\_u\] shows that the decay rate of the norm of the $g^c$ standard deviation is approximately the same for both methods. As expected, the variance reduction in Method 1 is larger (by approximately $5\%$) than in Method 2, because by construction, Method 1 reduces the variance of $g^c$ optimally. On the other hand, Figure \[RK:fig:norm\_u\_sdv\_k\_and\_k\_u\] shows that Method 2 reduces the $L^2$-norm of $\sigma^c_u$ more than Method 1 for most $N_{am}$, and by more than $15\%$ for $N_{am}>11$. For some $N_{am}<11$, Method 2 has larger norm of $\sigma^c_u$ than Method 1, which we attribute to the error in the $\sigma^c_u$ approximation . We obtain qualitatively similar (but more pronounced) results for $\sigma_g = 1.3$, as shown in Figure \[norm\_g\_u\_sdv\_k\_and\_k\_u\_sdev\_1\_03\]. Method 1 leads to a sharper decrease of the $\sigma^c_g$ norm and Method 2 results in a sharper decrease of the $\sigma^c_u$ norm than what we observe for $\sigma_g = 0.65$. From Figures \[norm\_g\_u\_sdv\_k\_and\_k\_u\] and \[norm\_g\_u\_sdv\_k\_and\_k\_u\_sdev\_1\_03\], we conclude that Method 2 is more efficient than Method 1 for reducing uncertainty in $u^c$, and that provides a sufficiently accurate approximation for solving the minimization problem . For the case $\sigma_g=0.65$, Figures \[C\_pg\_exact\_nobs40\_mrst\_x\_star\] and \[u\_sdev\_exact\_nobs40\_mrst\_x\_star\_k\_u\] show the $\sigma^c_g$ and $\sigma^c_u$ fields for $N_{am} = 1$, $5$, and $10$ obtained using both active learning methods. Similarly, $\sigma^c_g$ and $\sigma^c_u$ for $\sigma_g=1.3$ are shown in Figures \[C\_pg\_exact\_nobs40\_mrst\_x\_star\_sdev\_1\_03\] and \[u\_sdev\_exact\_nobs40\_mrst\_x\_star\_k\_u\_sdev\_1\_03\], respectively. As expected, the locations of the additional observations obtained with Methods 1 and 2 are different. Our results show that if the main objective is to predict states, i.e., $u(\mathbf{x})$, rather than coefficients, i.e., $k(\mathbf{x})$, than Method 2 is more efficient than Method 1. [0.48]{} ![$L_2$ norm of standard deviation of (a) $g^c(\mathbf{x},\omega)$ and (b) $u(\mathbf{x},\omega)$ as a function of the number of additional measurements. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="norm_g_u_sdv_k_and_k_u"}](figures/norm_g_sdv_k_and_k_u "fig:") [0.48]{} ![$L_2$ norm of standard deviation of (a) $g^c(\mathbf{x},\omega)$ and (b) $u(\mathbf{x},\omega)$ as a function of the number of additional measurements. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="norm_g_u_sdv_k_and_k_u"}](figures/norm_u_sdv_k_and_k_u "fig:") [0.48]{} ![$L_2$ norm of standard deviation of (a) $g^c(\mathbf{x},\omega)$ and (b) $u(\mathbf{x},\omega)$ as a function of the number of additional measurements. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.3$.[]{data-label="norm_g_u_sdv_k_and_k_u_sdev_1_03"}](figures/norm_g_sdv_k_and_k_u_sdev_1_03 "fig:") [0.48]{} ![$L_2$ norm of standard deviation of (a) $g^c(\mathbf{x},\omega)$ and (b) $u(\mathbf{x},\omega)$ as a function of the number of additional measurements. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.3$.[]{data-label="norm_g_u_sdv_k_and_k_u_sdev_1_03"}](figures/norm_u_sdv_k_and_k_u_sdev_1_03 "fig:") [0.48]{} ![ (a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="C_pg_exact_nobs40_mrst_x_star"}](figures/C_pg_est_exact_nobs40_mrst "fig:") \ [0.48]{} ![ (a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="C_pg_exact_nobs40_mrst_x_star"}](figures/C_pg_exact_nobs40_mrst_x_star_1 "fig:") [0.48]{} ![ (a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="C_pg_exact_nobs40_mrst_x_star"}](figures/C_pg_exact_nobs40_mrst_x_star_k_u_1 "fig:") [0.48]{} ![ (a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="C_pg_exact_nobs40_mrst_x_star"}](figures/C_pg_exact_nobs40_mrst_x_star_5 "fig:") [0.48]{} ![ (a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="C_pg_exact_nobs40_mrst_x_star"}](figures/C_pg_exact_nobs40_mrst_x_star_k_u_5 "fig:") [0.48]{} ![ (a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="C_pg_exact_nobs40_mrst_x_star"}](figures/C_pg_exact_nobs40_mrst_x_star_10 "fig:") [0.48]{} ![ (a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="C_pg_exact_nobs40_mrst_x_star"}](figures/C_pg_exact_nobs40_mrst_x_star_k_u_10 "fig:") [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u"}](figures/u_sdev_exact_cond_mc_mrst_no_obs_plot "fig:") \ [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u"}](figures/u_sdev_exact_nobs40_mrst_x_star_1 "fig:") [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u"}](figures/u_sdev_exact_nobs40_mrst_x_star_k_u_new_1 "fig:") [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u"}](figures/u_sdev_exact_nobs40_mrst_x_star_5 "fig:") [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u"}](figures/u_sdev_exact_nobs40_mrst_x_star_k_u_new_5 "fig:") [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u"}](figures/u_sdev_exact_nobs40_mrst_x_star_10 "fig:") [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 0.65$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u"}](figures/u_sdev_exact_nobs40_mrst_x_star_k_u_new_10 "fig:") [0.48]{} ![(a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="C_pg_exact_nobs40_mrst_x_star_sdev_1_03"}](figures/C_pg_est_exact_nobs40_mrst_sdev_1_03 "fig:") \ [0.48]{} ![(a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="C_pg_exact_nobs40_mrst_x_star_sdev_1_03"}](figures/C_pg_exact_nobs40_mrst_x_star_sdev_1_03_1 "fig:") [0.48]{} ![(a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="C_pg_exact_nobs40_mrst_x_star_sdev_1_03"}](figures/C_pg_exact_nobs40_mrst_x_star_k_u_sdev_1_03_1 "fig:") [0.48]{} ![(a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="C_pg_exact_nobs40_mrst_x_star_sdev_1_03"}](figures/C_pg_exact_nobs40_mrst_x_star_sdev_1_03_5 "fig:") [0.48]{} ![(a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="C_pg_exact_nobs40_mrst_x_star_sdev_1_03"}](figures/C_pg_exact_nobs40_mrst_x_star_k_u_sdev_1_03_5 "fig:") [0.48]{} ![(a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="C_pg_exact_nobs40_mrst_x_star_sdev_1_03"}](figures/C_pg_exact_nobs40_mrst_x_star_sdev_1_03_10 "fig:") [0.48]{} ![(a) Standard deviation of $g^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $g^c(\mathbf{x},\omega)$ with additional measurements: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="C_pg_exact_nobs40_mrst_x_star_sdev_1_03"}](figures/C_pg_exact_nobs40_mrst_x_star_k_u_sdev_1_03_10 "fig:") [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u_sdev_1_03"}](figures/u_sdev_exact_cond_mc_mrst_sdev_1_03 "fig:") \ [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u_sdev_1_03"}](figures/u_sdev_exact_nobs40_mrst_x_star_sdev_1_03_1 "fig:") [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u_sdev_1_03"}](figures/u_sdev_exact_nobs40_mrst_x_star_k_u_sdev_1_03_1 "fig:") [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u_sdev_1_03"}](figures/u_sdev_exact_nobs40_mrst_x_star_sdev_1_03_5 "fig:") [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u_sdev_1_03"}](figures/u_sdev_exact_nobs40_mrst_x_star_k_u_sdev_1_03_5 "fig:") [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u_sdev_1_03"}](figures/u_sdev_exact_nobs40_mrst_x_star_sdev_1_03_10 "fig:") [0.48]{} ![(a) Standard deviation of $u^c(\mathbf{x},\omega)$ without additional measurements. Standard deviation of $u^c(\mathbf{x},\omega)$ with additional measurements of $g$: (b) $N_{am}=1$, Method 1; (c). $N_{am}=1$, Method 2; (d) $N_{am}=1$, Method 5; (e) $N_{am}=5$, Method 2; (f) $N_{am}=10$, Method 1; and (g) $N_{am}=10$, Method 2. Additional observation locations of $g$ are shown in magenta and the original measurements are shown in red. Standard deviation of unconditional $g(\mathbf{x},\omega)$ is $\sigma_g = 1.63$.[]{data-label="u_sdev_exact_nobs40_mrst_x_star_k_u_sdev_1_03"}](figures/u_sdev_exact_nobs40_mrst_x_star_k_u_sdev_1_03_10 "fig:") Conclusions {#sec:conclusions} =========== We presented two methods for constructing finite-dimensional conditional Karhunen-Loève (KL) expansions of partially known parameters in PDE problems. We demonstrated that conditioning on data reduces the dimensionality of KL expansions and, most importantly, reduces uncertainty in the solution of PDE problems. Finally, conditioning on data reduces the computational cost of solving stochastic PDEs. We also present a new active learning strategy for acquiring new observations of the data based on minimizing variance of the conditional PDE solution (referred to as Method 2 in the paper) and compared it with the standard active learning method based on minimizing the conditional variance of the partially known parameters (referred to in the paper as Method 1). In the first approach for constructing finite-dimensional conditional GPs, presented in [@li2014conditional], the parameter field is conditioned first on data and then discretized by computing its KL expansion. In the second approach, presented in [@ossiander2014], the parameter field is discretized first by computing its KL expansion, and then the resulting KL expansion is conditioned on data. For the second approach, we demonstrated that conditioning leads to dimension reduction of the conditional representation, and we proposed a method for constructing a reduced representation in terms of the effective number of random dimensions. For a linear diffusion SPDE with uncertain log-normal coefficient, we show that Approach 1 provides a more accurate approximation of the conditional log-normal coefficient and solution of the SPDE than Approach 2 for the same number of random dimensions in a conditional KL expansion. Furthermore, Approach 2 provides a good estimate for the number of terms of the truncated KL expansion of the conditional field of Approach 1. Finally, we demonstrate that the proposed active learning method (Method 2) is more efficient for reducing uncertainty in the solution of the SPDE under consideration (i.e., it leads to a larger reduction of the variance) than the standard active learning method (Method 1). The difference between two methods increases as the variance of the partially known coefficient increases.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We calculate correlation functions of topological sigma model (A-model) on C.Y.hypersurfaces in $CP^{N-1}$ using torus action method. We also obtain path-integral representation of free energy of the theory coupled to gravity.' author: - | Masao Jinzenji\ \ *Department of Phisics, University of Tokyo\ *Bunkyo-ku, Tokyo 113, Japan** title: | UT-HEP-703-95\ Construction of Free Energy of Calbi-Yau manifold\ embedded in $CP^{N-1}$ via Torus Actions --- Introduction ============ Recently vast development occurs in the field of topological sigma model, especially from the mirror symmetry of the models having Calabi-Yau manifolds as target spaces [@cogp]. In [@nj], Nagura and the author treated the topological sigma model on the Calabi-Yau hypersurfaces $M_{N}$ in $CP^{N-1}$. We derived $N-2$ point correlation function by the analysis of the solution of period integral equation which emerges from the deformation of the complex structure of $\tilde{M}_{N}$,the mirror counterpart of $M_{N}$(B-model). And we showed that translation of the calculation of B-model into the A-model by mirror map actually gives the correlation function of topological sigma model on $M_{N}$, i.e., the number of holomorphic maps satisfying the conditions imposed by external operator insertion. But the explicit evaluation was limited only for the degree $1$ case. So there remains a problem of the calculation of the correlation functions from the A-model point of view for higher degree cases. Of course, search for the deeper understanding of mirror symmetry is very important, but we limit our interest to the above problem in this paper. As we saw in [@nj], correlation functions of the A-model have geometrically very clear meaning. And their direct calculation relies on the explicit construction of moduli spaces for holomorphic maps from $CP^{1}$ to target space (in our case $M_{N}$). For degree 1, this can be done by taking the zero locus of a section of $Sym^{N}(U^{*})$ of $Gr(2,N)$ ($U$ is the universal bundle of Grassmannian). S.Katz constructed degree $2$ moduli space as the zero locus of a section of $Sym^{N}(U^{*})/Sym^{N-2}(U^{*}) \otimes {\cal O}_{P}(-1)$ on $Gr(3,N)$.(${\cal O}_{P}(1)$ is the taughtological sheaf on $P(Sym^{2}(U))$.) These construction seems to become more complicated as the degree rises. But in [@tor], Kontsevich constructed the compact moduli space from $CP^{1}$ to complex manifold $V$ by introducing stable maps. Roughly speaking, these are maps from branched $CP^{1}$ (stable curves) to $V$. In the case where $V$ is $CP^{N}$ or hypersurfaces of $CP^{N}$, he also did some calculations of correlation functions (Gromov-Witten invariants) using this construction and by means of fixed point theorem under the torus action flow of $CP^{N}$. We find that his treatment of quintic hypersurface in $CP^{4}$ is very much like our elementary approach for $M_{N}$. So we thought we can calculate Gromov-Witten invariants for $M_{N}$ by this method. We have to notice one difference. In [@nj],we treated the matter theory, but in Kontsevich’s formulation correlation functions are those for the theory coupled with gravity because of introduction of stable maps. After all, we found this method also works for $M_{N}$. Using the fact that $3$-point functions are identical for both theories and that fusion rule holds in the matter theory (See Greene,Morrison,Plesser [@gmp]), we reconstruct $N-2$ point functions for matter theory and derived some identity. This method has a by-product. It has the structure of the sum of tree graph amplitudes, so by using Feynmann rules, we can construct path-integral representation of the generating function of correlation functions (Free Energy). We will represent it at the end of this paper. In section $2$, we introduce the topological sigma model (A-model) and construct correlation functions as integrals of forms on moduli spaces. In section $3$, we review the torus action method and in section $4$, we do some explicit calculation of amplitudes and see these results are compatible with those of Greene, Morrison, Plesser [@gmp],and S.Katz [@katz]. In section 5 we construct path-integral representation of free energy for $M_{N}$ (coupled with gravity). This paper’s treatment is rather non-rigolous in comparison with the original one of Kontsevich, but we think it is more accessible and may arise some insight for generalization. Correlation Function as an Integral of Forms on Moduli Spaces ============================================================= Topological Sigma Model (A-Model) --------------------------------- ### lagrangian and weak coupling limit Topological sigma model can be obtained by twisting $N=2$ Supersymmetric Sigma Model on M. A-model corresponds to A-twist, which turns $\psi_{+}^i$ and $\psi_{-}^{\bar{i}}$ in $N=2$ Sigma Model into $\chi^{i}$,$\chi^{\bar{i}}$, and $\psi_{+}^{\bar{i}},\psi_{-}^i$ into $\psi_{z}^{\bar{i}},\psi_{\bar{z}}^{i}$. In other words, A-twist means subtraction of half of $U(1)$ charge from world sheet spin quantum number. Then we obtain the following Lagrangian for the A-model, $$L = 2t \int_{\Sigma} d^{2}z(\frac{1}{2}g_{IJ}\partial_{z}\phi^{I} \partial_{\bar{z}}\phi^{J} + i\psi_{z}^{\bar{i}}D_{\bar{z}}\chi^{i} g_{i\bar{i}} + i\psi_{\bar{z}}^{i}D_{z}\chi^{\bar{i}}g_{\bar{i}i} -R_{i\bar{i}j\bar{j}}\psi_{\bar{z}}^{i}\psi_{z}^{\bar{i}}\chi^{j} \chi^{\bar{j}}) \label{a1}$$ (\[a1\]) is invariant under the BRST - transformation, $$\begin{aligned} & \delta\phi^{i} = i\alpha\chi^{i} \nonumber\\ & \delta\phi^{\bar{i}} = i\alpha\chi^{\bar{i}} \nonumber\\ & \delta \chi^{i} = \delta \chi^{\bar{i}} = 0 \nonumber\\ & \delta\psi_{z}^{\bar{i}} = -\alpha\partial_{z}\phi^{\bar{i}} -i\alpha\chi^{\bar{j}}\Gamma^{\bar{i}}_{\bar{j}\bar{m}} \psi_{z}^{\bar{m}} \nonumber\\ & \delta\psi_{\bar{z}}^{i} = -\alpha\partial_{\bar{z}}\phi^{i} -i\alpha\chi^{j}\Gamma^{i}_{jm}\psi_{\bar{z}}^{m} \label{a2}\end{aligned}$$ This invariance allows us to consider only BRST-invariant observables. As in section2, we define BRST operator $Q$ by $\delta V = -i\alpha\{Q,V\}$ for any field V. Of course, $Q^2 = 0$. We can rewrite the lagrangian (\[a1\]) modulo the $\psi$ equation of motion as; $$L = it\int_{\Sigma}d^2z \{Q,V\} + t\int_{\Sigma}\Phi^{*}(e) \label{a3}$$ where $$V = g_{i\bar{j}}(\psi_{z}^{\bar{i}}\partial_{\bar{z}}\phi^{j} + \partial_{z} \phi^{\bar{i}}\psi_{\bar{z}}^{j}) \label{a4}$$ and $$\int_{\Sigma}\Phi^*(e) = \int_{\Sigma}(\partial_{z}\phi^{i}\partial_{\bar{z}} \phi^{\bar{j}}g_{i\bar{j}}- \partial_{\bar{z}}\phi^{i} \partial_{z}\phi^{\bar{j}}g_{i\bar{j}}) \label{a5}$$ (\[a5\]) is the integral of the pullback of the Kähler form $e$ of M, and it depends only on the intersection number between $\Phi_{*}(\Sigma)$ and $PD(e)$($PD(e)$ denotes the Poincare Dual of $e$), which equals the degree of $\Phi$. By an appropriate normalization of $g_{i\bar{j}}$,we have $$\int_{\Sigma}\Phi^{*}(e) = n \label{a60}$$ where n is the degree of $\Phi$. Next, we consider the correlation function of BRST-invariant observables $\{ {\cal O}_{i}\}$, i.e. $$\langle \prod_{i=1}^{k} {\cal O}_{i}\rangle = \int {\cal D}\phi {\cal D}\psi{\cal D}\chi e^{-L} \prod_{i=1}^{k} {\cal O}_{i} \label{a6}$$ We have seen $\int_{\Sigma}\Phi^{*}(e) = n $ and we decompose the space of maps $\phi$ into different topological sectors $\{B_{n}\}$ in each of which $deg(\Phi)$ is a fixed integer. We can rewrite (\[a6\]) as follows. $$\langle \prod_{i=1}^{k} {\cal O}_{i}\rangle = \int {\cal D}\phi {\cal D}\psi{\cal D}\chi e^{-L} \prod_{i=1}^{k} {\cal O}_{i} = \sum_{n=0}^{\infty}e^{-nt}\int_{B_n} {\cal D}\phi {\cal D}\psi{\cal D}\chi e^{-it\int_{\Sigma}d^2 z \ \{Q,V\}} \prod_{i=1}^{k} {\cal O}_{i} \label{a7}$$ And we set $$\langle \prod_{i=1}^{k} {\cal O}_{i}\rangle_{n} = \int_{B_n} {\cal D}\phi {\cal D}\psi{\cal D}\chi e^{-it\int_{\Sigma}d^2 z \ \{Q,V\}} \prod_{i=1}^{k} {\cal O}_{i} \label{a8}$$ We can easily see that $\int_{\Sigma}d^{2}z\{Q,V\} = \{Q,\int_{\Sigma}d^{2} zV\}$,i.e. lagrangian is BRST exact. It follows from this and $\{Q,{\cal O}_i \}=0$ that $\langle \prod_{i=1}^{k}{\cal O}_i\rangle_{n}$ doesn’t depend on the coupling constant t and we can take weak coupling limit $t \rightarrow \infty $ in evaluating the path integral. In this limit,the saddle point approximation of the path integral becomes exact. Saddle points of the lagrangian are given by $$\partial_{\bar{z}}\phi^{i} = 0\;\;\; \partial_{z}\phi^{\bar{i}} = 0 \label{a9}$$ (\[a9\]) shows that the path-integral is reduced to an integral over the moduli space of holomorphic maps from $\Sigma$ to $M$ of degree $n$. We will focus our attention to the case where $\Sigma$ has genus 0. We denote this space as ${\cal M}_{0,n}^{M}$. ### The Ghost Number anomally and BRST observables In the previous subsection, we have seen that the path integral (\[a6\]) is reduced to an integral over ${\cal M}_{0,n}^{M}$ weighted by one loop determinants of the non zero modes. But in general, there are fermion zero modes which are given as the solution of $D_{\bar{z}}\chi^{i}= D_{z}\chi^{\bar{i}}=0$ and $D_{\bar{z}}\psi_{z}^{\bar{i}}= D_{z}\psi_{\bar{z}}^{i} = 0$. Let $a_n$ (resp.$b_n$) be the number of $\chi$ (resp.$\psi$) zero modes. If $M$ is a Calabi-Yau manifold, we can see from Riemann-Roch Theorem, $$w_n=a_n-b_n=\mbox{dim}(M) \label{a10}$$ The existence of fermion zero mode is understood as Ghost number anomally, because lagrangian (\[a1\]) classically conserves the ghost number. In path integration,these zero modes appear only in the integration measure except in $\prod_{i=1}^k{\cal O}_i$,and the correlation function $\langle \prod_{i=1}^k{\cal O}_i \rangle_{n}$ vanishes unless the sum of the ghost number of ${\cal O}_i$ is equal to $w_n$. $D_{\bar{z}}\chi^{i}=0$ (resp.$D_{z}\chi^{\bar{i}}=0$) can be considered as the linearization of the equation $\partial_{\bar{z}}\phi^{i}=0$ (resp.$\partial_{z} \phi^{\bar{i}}=0$) and we can regard $\chi$ zero mode as $T{\cal M}_{0,n}^{M}$. $w_n$ is usually called “virtual dimension “ of ${\cal M}_{0,n}^{M}$. In generic case $b_n=0$ and $\mbox{dim}({\cal M}_{0,n}^{M}) = w_n$ holds. Then we have $\mbox{dim}({\cal M}_{0,n}^{M})=a_n$. BRST cohomology classes of the A-model are constructed from the de Rham cohomology classes $H^{*}(M)$ of the manifold M. Let $W= W_{I_1I_2\cdots I_n}(\phi)d\phi^{I_1}\wedge\cdots\wedge d\phi^{I_n}$ be an $n$ form on $M$. Then we define a corresponding local operator of the A-Model, $${\cal O}_{W}(P) = W_{I_1\cdots I_n}\chi^{I_1}\cdots\chi^{I_n}(P) \label{a11}$$ From (\[a2\]) we have $$\{Q,{\cal O}_{W}\}=-{\cal O}_{dW} \label{a12}$$ which shows that if $W \in H^{*}(M)$,${\cal O}_{W}(P)$ is BRST-closed. ### Evaluation of the Path Integral Now we discuss how we can evaluate $\langle \prod_{i=1}^{k}{\cal O}_i \rangle _n$ . We take ${\cal O}_i$ to be ${\cal O}_{W_i}$ which is induced from $W_i \in H^{*}(M)$. By adding appropriate exact forms we can make $W_i$ into the differential form which has delta function support on $PD(W_i)$. Then ${\cal O}_{W_i}(P_i)$ is non zero only if $$\phi(P_i) \in PD(W_i) \label{a13}$$ Then integration over ${\cal M}_{0,n}^{M}$ is restricted to $\tilde{\cal M}_{0,n}^{M}$ , which consists of $\phi \in {\cal M}_{0,n}^{M}$ satisfying (\[a13\]). In evaluating $\langle \prod_{i=1}^{k}{\cal O}_{W_i}\rangle_n$, (\[a13\]) imposes $\sum_{i=1}^{k}\mbox{dim}(W_i)$ conditions, so $\mbox{dim}(\tilde{\cal M}_{0,n}^{M}) = \mbox{dim}({\cal M}_{0,n}^{M})- \sum_{i=1}^{k}\mbox{dim}(W_i) = w_n + b_n -\sum_{i=1}^{k}\mbox{dim}(W_i)$. But from the fact that ghost number of ${\cal O}_{W_i}$ equals $\mbox{dim}(W_i)$ (contribution from $\chi$) and anomally canselation condition,we have $\mbox{dim}(\tilde{\cal M}_{0,n}^{M}) = b_n$. In generic case where $b_n=0$, $\tilde{\cal M}_{0,n}^{M}$ turns into finite set of points. Then we perform an one loop integral over each of these points. The result is a ratio of boson and fermion determinants,which cancel each other. Then contributions to $\langle \prod_{i=1}^{k} {\cal O}_{W_i} \rangle_{n}$ in the generic case equals the number of instantons which satisfies (\[a13\]),i.e. $$\langle \prod_{i=1}^{k} {\cal O}_{W_i} \rangle_{generic} = {}^{\sharp} \tilde{\cal M}_{0,n}^{M} \label{a14}$$ When $dim(\tilde{\cal M}_{0,n}^{M}) = b_n \geq 1$, there are $b_n$ $\psi$ zero modes which we can regard as the fiber of the vector bundle $\nu$ on $\tilde{\cal M}_{0,n}^{M}$. In this case, contributions to $\langle \prod_{i=1}^{k} {\cal O}_{W_i} \rangle_{n}$are known as the integration of Euler class $\chi(\nu)$ on $\tilde{\cal M}_{0,n}^{M}$. If we consider $\nu$ as a 0-dimensional vector bundle on a point in the generic case, we can apply the same logic there. We denote each component of $\tilde{\cal M}_{0,n}^{M}$ as $\tilde{\cal M}_{0,n,m}^{M}$ and obtain $$\langle \prod_{i=1}^{k} {\cal O}_{W_i} \rangle_{n} = \sum_{m} \int_{\tilde{\cal M}_{0,n,m}^{M}}\chi(\nu) \label{a15}$$ Hence from (\[a7\]) $$\langle \prod_{i=1}^{k} {\cal O}_{W_i} \rangle = \sum_{n=0}^{\infty} \sum_{m=1}^{m_n}e^{-nt}\int_{\tilde{\cal M}_{0,n,m}^{M}}\chi(\nu) \label{a16}$$ In algebraic geometry, generic instantons of degree $n$ corresponds to irreducible maps of degree n, and instantons of degree n with non-zero $\psi$ zero mode to reducible maps which are $j$-th multiple cover of irreducible maps of degree $n/j$ $(j|n)$. Let $\tilde{\cal M}_{0,n,j,m}^{M}$ be the $m$-th connected component of moduli spaces which are $j$-th multiple cover of $n/j$-th irreducible instantons, and $\nu_{j,m}$ be vector bundle of $\psi$ zero modes on $\tilde{\cal M}_{0,n,j,m}^{M}$. Then we have from (\[a16\]), $$\langle \prod_{i=1}^{k} {\cal O}_{W_i} \rangle = \sum_{n=0}^{\infty} \sum_{j|n}\sum_{m=1}^{m_{n,j}}e^{-nt}\int_{\tilde{\cal M}_{0,n,j,m}^{M}} \chi(\nu_{j,m}) \label{a17}$$ Reduction to an integral of Forms on Moduli Spaces -------------------------------------------------- In this paper, we treat the topological sigma model (A-Model) of Calabi-Yau manifold embedded in $CP^{N-1}$. This manifold is realized as the zero-locus of section of $-K$ ($K$ is canonical line bundle of $CP^{N-1}$.).Since $-K$ is equivalent to $ N H $ as a line bundle ($H$ is hyperplane bundle of $CP^{N-1}$), we can take homogeneous polynomial of degree $N$ as the defining equation of $M_{N}$. For example, $$M_{N} := \{(X_{1},X_{2},\cdots ,X_{N}) / \mbox{\bf C}^{\times} \in CP^{N-1}| X_{1}^{N}+X_{2}^{N}+\cdots + X_{N}^{N}=0 \}$$ Observables of this model can be constructed from elements of $w \in H^{*}(M_{N})$ which we denote as ${\cal O}_{w}$, and in the following discussion we consider the observables which are included from the subring of $H^{*}(M_{N},C)$ generated by Kähler form $e$ of $M_{N}$ (we denote it as $H_{CP^{N-1}}(M_{N},C)$). One of the reason why we take this subring is that we can obtain it directly from $H^{*}(CP^{N-1},C)$ and Poincare dual of its elements are analytic submanifold of $M_{N}$. More explicitly, elements of $H_{CP^{N-1}}(M_{N},C)$ are given as $e^{k}$ ($k = 1,2,\cdots ,N-2 $) and Poincare dual of $e^{k}$ is the intersection of the zero locus of the section of $H^{0}(CP^{N-1},{\cal O}(k\cdot H))$ and $M_{N}$. So in the following discussion we treat the observables [^1] $${\cal O}_{1},{\cal O}_{e},{\cal O}_{e^{2}},\cdots ,{\cal O}_{e^{N-2}}$$ Then the fact that Lagrangian of the topological sigma model is BRST-exact allows us to take the strong coupling limit and correlation functions of this model reduces to the integral of closed forms corresponding to the BRST closed observables on moduli spaces of holomorphic maps $f$ from Riemann surface $\Sigma_{g}$ to target space $M_{N}$ (we focus our attention to the case of $g=0$,i.e,$CP^{1}$). When the target space is a hypersurface of simple projective space $CP^{N-1}$,we can classify moduli spaces by the degree $n= {}^{\sharp}(f(CP^{1})\cap PD(e))$ and we denote the moduli space of degree $n$ as ${\cal M}^{M_{N}}_{0,n}$. Dimension of ${\cal M}^{M_{N}}_{0,n}$ which counts the number of $\chi$-zero modes is evaluated by the Riemann-Roch Theorem as follows. $$\begin{aligned} \mbox{dim}({\cal M}^{M_{N}}_{0,n})&:=& \mbox{dim}(H^{0}(CP^{1},f^{*}(T^{\prime}M_{N}))) \nonumber\\ &=& \mbox{dim}(M_{N}) + {\mbox deg}(f) \cdot c_{1}(KM_{N}) + \mbox{dim}(H^{1}(CP^{1},f^{*}(T^{\prime}M_{N}))) \nonumber\\ &=& \mbox{dim}(M_{N}) + \mbox{dim}(H^{1}(CP^{1},f^{*}(T^{\prime}M_{N}))) \nonumber\\ &=& N-2 + \mbox{dim}(H^{1}(CP^{1},f^{*}(T^{\prime}M_{N}))) \label{roch}\end{aligned}$$ where we used the Calabi-Yau condition $c_{1}(K_{M_{N}})=0$.(\[roch\]) tells us that the dimension of moduli space is independent of degree $n$. First, we consider the generic case where $\mbox{dim}(H^{1}(CP^{1},f^{*}(T^{\prime}M_{N})))=0$. From the above argument, we can heuristically represent correlation functions, $$\begin{aligned} \langle {\cal O}_{e^{j_{1}}}(z_{1}){\cal O}_{e^{j_{2}}}(z_{2}) \cdots {\cal O}_{e^{j_{k}}}(z_{k}) \rangle _{n,generic}\nonumber\\ = \int_{{\cal M}^{M_{N}}_{0,n}} \alpha({\cal O}_{e^{j^{1}}})\wedge \alpha({\cal O}_{e^{j^{2}}}) \wedge \cdots \wedge \alpha({\cal O}_{e^{j^{k}}})\end{aligned}$$ where $ \alpha({\cal O}_{e^{j}})$ is the closed form on ${\cal M}^{M_{N}}_{0,n}$ induced from ${\cal O}_{e^{j}}$. Since the form degree of $ \alpha({\cal O}_{e^{j}})$ equals the ghost number of ${\cal O}_{e^{j}}(=\mbox{dim}(e^{j})=j)$,correlation functions are nonzero only if the following conditions are satisfied. $$\begin{aligned} \mbox{dim}({\cal M}^{M_{N}}_{0,n})&=& \sum_{i=1}^{k} j_{i} \nonumber\\ \Longleftrightarrow N-2 &=& \sum_{i=1}^{k} j_{i} \label{sel}\end{aligned}$$ If we take $e^{j}$ as the forms which has the delta function support on $PD(e^{j})$,then from ( ), $\alpha ({\cal O}_{e^{j}})$ can be interpreted as the constraint condition on $f$, $$\alpha ({\cal O}_{e^{j_{i}}}) \leftrightarrow f(z_{i}) \in PD(e^{j_{i}}) \label{con}$$ (\[con\]) imposes $({\mbox dim}(e^{j})-1)+1 = j$ independent conditions on ${\cal M}^{M_{N}}_{0,n}$ ($(\mbox{dim}(e^{j})-1)$ corresponds to the degree of freedom which makes $f(CP^{1}) \cap PD(e^{j})\neq \emptyset$ and $1$ to the one which sends $f(z_{i})$ into $PD(e^{j})$). And from (\[sel\]), what remains is the discrete point set of holomorphic maps $f$ which satisfy (\[con\]) for all $i$. Then we have $$\begin{aligned} \langle {\cal O}_{e^{j_{1}}}(z_{1}){\cal O}_{e^{j_{2}}}(z_{2}) \cdots {\cal O}_{e^{j_{k}}}(z_{k}) \rangle _{n,generic}\nonumber\\ = {}^{\sharp}\{f : CP^{1} \stackrel{hol.}{\rightarrow} M_{N} |f(z_{i}) \in PD(e^{j_{i}})\}\end{aligned}$$ Now, let us consider non-generic case.In this case, $\mbox{dim}(H^{1}(CP^{1},f^{*}(T^{\prime}M_{N}))>0$ and moduli space have additional $\mbox{dim}(H^{1}(CP^{1},f^{*}(T^{\prime}M_{N}))$ degrees of freedom. We can see these degrees of freedom correspond to multiple cover maps by the following argument. A multiple cover map $f$ can be decomposed into the form $f = \tilde{f}\circ \varphi$ where $\tilde{f}$ is irreducible map from $CP^{1}$ to $M_{N}$ and $\varphi$ represents the map from $CP^{1}$ to $CP^{1}$ of degree $d \geq 2$. Then let us count $\mbox{dim}({\cal M}^{M_{N}}_{0,n})$ by taking the (holomorphic) variation of $\tilde{f} \circ \varphi$. $$\begin{aligned} \delta(\tilde{f}\circ\varphi)\nonumber\\ = \delta\tilde{f}\circ\varphi + \tilde{f}\circ\delta\varphi\end{aligned}$$ $\delta\tilde{f}$ corresponds to the generic degrees of freedom,i.e, $$\mbox{dim}(\delta\tilde{f}\circ\varphi)= N-2$$ $\tilde{f}\circ\delta\varphi$ counts the deformation of the multiple cover map which can be realized using the section $\varphi^{z} \in H^{0}(CP^{1},\varphi^{*}(T^{\prime}CP^{1}))$ as $\varphi^{z}\partial_{z}\tilde{f}$. We can count these by using Riemann-Roch, $$\begin{aligned} \mbox{dim}(\tilde{f}\circ\delta\varphi)&=& \mbox{dim}(H^{0}(CP^{1},\varphi^{*}(T^{\prime}CP^{1})))-3\nonumber\\ &=&1 + {\mbox deg}(\varphi)\cdot c_{1}(T^{\prime}CP^{1})-3 \nonumber\\ &=& 2d -2 \label{multi}\end{aligned}$$ In (\[multi\]) we subtract the double counted $SL(2,C)$ which comes from the indetermination of the decomposition of $f$,i.e, $$\begin{aligned} f &=& \tilde{f}\circ\varphi \nonumber\\ &=& \tilde{f}\circ u \circ u^{-1} \circ \varphi \qquad u \in SL(2,C)\end{aligned}$$ After all we find additional $2d-2$ $\chi$ zero modes.But we can also construct $2d-2$ $\psi_{\bar{z}}^{i}$ which comes from $H^{1}(CP^{1},f^{*}(T^{\prime}M_{N}))$. By the Kodaira-Serre duality ,the following equation holds. $$\begin{aligned} \mbox{dim}(H^{1}(CP^{1},f^{*}(T^{\prime}M_{N})) &=& \mbox{dim}(H^{0}(CP^{1},K\otimes f^{*}(T^{\prime *}M_{N})))\nonumber\\ (&=& \mbox{dim}(H^{0}(CP^{1},\bar{K}\otimes f^{*}(\bar{T}^{\prime *}M_{N}))))\end{aligned}$$ and $$g^{i \bar{j}}\tilde{\psi}_{\bar{z},\bar{j}} \qquad (\tilde{\psi}_{\bar{z},\bar{j}} \in H^{0}(CP^{1},\bar{K}\otimes f^{*}(\bar{T}^{*,\prime}M_{N})))$$ Lagrangian (\[a1\]) has the $R_{i \bar{i} j \bar{j}}\psi_{\bar{z}}^{i} \psi_{z}^{\bar{i}}\chi^{j}\chi^{\bar{j}}$ term, so we can “kill” additional $2d-2$ $\chi$ and $\psi$ zero modes by expanding $\exp(R_{i \bar{i} j \bar{j}} \psi_{\bar{z}}^{i}\psi_{z}^{\bar{i}}\chi^{j}\chi^{\bar{j}})$. So we don’t have to add extra operator insertions. Then,by integrating $\psi$ zero-modes first, we have the Euler class $\chi(\nu)$ where $ \nu \simeq H^{1}(CP^{1},\varphi^{*}(T^{\prime}M_{N}))$. This leads to $$\begin{aligned} \langle {\cal O}_{e^{j_{1}}}(z_{1}){\cal O}_{e^{j_{2}}}(z_{2}) \cdots {\cal O}_{e^{j_{k}}}(z_{k}) \rangle _{n} \nonumber\\ = \int_{{\cal M}_{0,n}^{M_{N}}} \chi(\nu) \wedge \alpha({\cal O}_{e^{j_{1}}}(z_{1})) \wedge \cdots \alpha({\cal O}_{e^{j_{k}}}(z_{k})) \label{13}\end{aligned}$$ We can refine (\[13\]) by using the argument which leads to (\[con\]) and define the evaluation map, $$\varphi_{i} : {\cal M}_{0,n}^{M_{N}} \rightarrow M_{N}:f \mapsto f(z_{i})$$ We have $$\begin{aligned} \langle {\cal O}_{e^{j_{1}}}(z_{1}){\cal O}_{e^{j_{2}}}(z_{2}) \cdots {\cal O}_{e^{j_{k}}}(z_{k}) \rangle _{n} \nonumber\\ = \int_{{\cal M}_{0,n}^{M_{N}}} \chi(\nu) \wedge \varphi^{*}_{1}(e^{j_1}) \wedge \cdots \varphi^{*}_{k}(e^{j_{k}}) \label{15}\end{aligned}$$ We can relate the non-generic part the correlation functions to ones of lower degree, because in such case $f$ decomposes into $f= \tilde{f}\circ\varphi$ where $\mbox{deg}(\varphi)=d$ and $\mbox{deg}(\tilde{f})= n/d < N$. But good results are given only in the case of $k=3$, which was derived by Greene, Aspinwall, Morrison and Plesser [@gmp] [@am]. Of course, if we use the fusion rule that holds in the matter theory, we can reduce the correlation functions into the product of three point functions and formally distinguish the non-generic part from the generic ones.But geometrical meaning is still not clear. Then we slightly change our point of view. Since $M_{N}$ is a hypersurface in $CP^{N-1}$,we can see ${\cal M}_{0,n}^{M_{N}}$ as a submanifold of ${\cal M}_{0,n}^{CP^{N-1}}$ which consists of maps satisfying the following condition. $$\begin{aligned} f: CP^{1} \rightarrow CP^{N-1}\nonumber\\ f(z) \in M_{N} \qquad \mbox{for all} \quad z \in CP^{1} \label{16}\end{aligned}$$ If we can realize the condition (\[16\]) as the closed forms $c_{n}(M_{N})$ on ${\cal M}_{0,n}^{CP^{N-1}}$, we have an alternate representation for the correlation functions as follows, $$\begin{aligned} \langle {\cal O}_{e^{j_{1}}}(z_{1}){\cal O}_{e^{j_{2}}}(z_{2}) \cdots {\cal O}_{e^{j_{k}}}(z_{k}) \rangle _{n,alt} \nonumber\\ = \int_{{\cal M}_{0,n}^{CP^{N-1}}} c_{n}(M_{N}) \wedge \tilde{\varphi}^{*}_{1}(e^{j_1}) \wedge \cdots \wedge \tilde{\varphi}^{*}_{k}(e^{j_{k}})\nonumber\\ \tilde{\varphi}_{i} : {\cal M}_{0,n}^{M_{N}} \rightarrow CP^{N-1}:f \mapsto f(z_{i}) \label{17}\end{aligned}$$ In (\[17\]), we can drop off the Euler class $\chi(\nu)$. This is because $$\begin{aligned} \mbox{dim}(H^{1}(CP^{1},\tilde{\varphi}^{*}(T^{\prime}CP^{N-1}))) &=&\mbox{dim}(H^{0}(CP^{1},K\otimes(\tilde{\varphi}^{*} (T^{\prime}M_{N})))) =0 \nonumber \\ (c_{1}(K_{CP^{1}}\otimes \varphi^{*}(K_{CP^{N-1}})) &=& \mbox{deg}(\varphi) \cdot (-2) \cdot N < 0)\end{aligned}$$ Dimension of moduli space does not jump in this case. Then naturally arises the question about the relation between (\[15\]) and (\[17\]). However we want to proceed further with the formula (\[17\]). Then we want to use the torus action method invented by Kontsevich, and we couple gravity to the topological sigma model. Roughly speaking, we add to the moduli space “puncture” degrees of freedom. So for k-point correlation function, dimension of moduli space (we denote it as ${\cal M}_{0,n,k}^{CP^{N-1}}$) increases by $k-3$. $-3$ corresponds to deviding by automorphism of $CP^{1}$, i.e.$SL(2,C)$ which is induced by $c$-ghost zero-modes. And topological selection rule (\[sel\]) is changed into $$\begin{aligned} N-2+k-3= \sum_{i=1}^{k} j_{i}\nonumber\\ \Longleftrightarrow \quad N-5 =\sum_{i=1}^{k}(j_{i}-1) \label{18}\end{aligned}$$ ${\cal M}_{0,n,k}^{CP^{N-1}}$ can be roughly represented as follows, $${\cal M}_{0,n,k}^{CP^{N-1}} \simeq \{(z_{1},z_{2},\cdots,z_{k}),f \}/SL(2,C) \qquad f \in {\cal M}_{0,n}^{CP^{N-1}}$$ where $u \in SL(2,C)$ acts $$u \circ \{(z_{1},z_{2},\cdots,z_{k}),f\} =\{(u(z_{1}),\cdots,u(z_{k})), (u^{-1})^{*}\circ f \}$$ This action of $SL(2,C)$ is compatible with the “evaluation map”, $$\begin{aligned} \phi_{i} : {\cal M}_{0,n,k}^{CP^{N-1}} \mapsto CP^{N-1}\nonumber\\ \{(z_{1},z_{2},\cdots,z_{k}),f\} \mapsto f(z_{i}) \label{19}\end{aligned}$$ because $(u^{-1})^{*}f(u(z_{i})) = f(z_{i})$. Then the integral representation of amplitudes (\[17\]) turns into $$\begin{aligned} \langle {\cal O}_{e^{j_{1}}}(z_{1}){\cal O}_{e^{j_{2}}}(z_{2}) \cdots {\cal O}_{e^{j_{k}}}(z_{k}) \rangle _{n,alt,grav.} \nonumber\\ = \int_{{\cal M}_{0,n,k}^{CP^{N-1}}} c_{n}(M_{N}) \wedge \phi^{*}_{1}(c_{1}^{j_1}(H)) \wedge \cdots \wedge \phi^{*}_{k} (c_{1}^{j_{k}}(H)) \label{20}\end{aligned}$$ where we used the fact that $e$ corresponds to the first chern class of hyperplane bundle $H$. Then we have to find find the realization of $c_{n}(M_{N})$. We can roughly do it as follows. First consider the coordinate representation of ${\cal M}_{0,n}^{CP^{N-1}}$, $$f:(s,t) \mapsto (\sum_{i=0}^{n}a_{1}^{i}s^{n-i}t^{i}, \cdots, \sum_{i=0}^{n}a_{N}^{i}s^{n-i}t^{i} ) \label{21}$$ where $(a_{i}^{j})$’s are the coordinates of ${\cal M}_{0,n}^{CP^{N-1}}$. Then the condition imposed by $c_{n}(M_{N})$ is equal to $$\begin{aligned} &&f(CP^{1}) \in M_{N} \quad \mbox{for all} (s,t)\nonumber\\ &\Longleftrightarrow& (\sum_{i=0}^{n}a_{1}^{i}s^{n-i}t^{i})^{N}+ \cdots +(\sum_{i=0}^{n}a_{N}^{i}s^{n-i}t^{i})^{N}=0 \quad \mbox{for all}(s,t)\nonumber\\ &\Longleftrightarrow& f^{m}(a_{j}^{i}) = 0\qquad (m=0,1,\cdots ,Nn) \label{22}\end{aligned}$$ where $f^{m}(a_{j}^{i})$’s are the coefficient polynomial of $s^{m}t^{Nn-m}$ of the l.h.s of the second line of (\[22\]).This imposes $Nn+1$ condition on ${\cal M}_{0,n}^{CP^{N-1}}$. We can describe this condition mathematically in terms of gravitatinal moduli space ${\cal M}_{0,n,k}^{CP^{N-1}}$. Let $\pi_{j}$ be a forgetfull map $\pi_{j}:{\cal M}_{0,n,j}^{CP^{N-1}} \rightarrow {\cal M}_{0,n,j-1}^{CP^{N-1}}$. Then for $j=1$, the fiber of $\pi_{1}$ is $CP^{1}$. And consider the sheaf $\phi_{1}^{*}(N H)$ on ${\cal M}_{0,n,1}^{CP^{N-1}}$ where $N H$ corresponds to defining polynomial of $M_{N}$ and $H^{0}({\cal M}_{0,n,1}^{CP^{N-1}},\phi_{1}^{*}(N H))$ to the second line of (\[22\]) modulo $SL(2,C)$ equivalence. Then direct limit sheaf $R^{0}_{\pi_{1}}(\phi^{*}_{1}(N H))$(we denote it as ${\cal E}_{Nn+1}$).It locally equals $H^{0}(CP^{1},f^{*}{\cal O}(N H))$ has rank $(Nn+1)$.We can translate the operation in going from the second line of (\[22\]) to the third one into the evaluation of the zero locus of the section of ${\cal E}_{Nn+1}$. Considering the map, $$\tilde{\pi}_{k}:= \pi_{1} \circ \pi_{2} \circ \cdots \circ \pi_{k} : {\cal M}_{0,n,k}^{CP^{N-1}} \mapsto {\cal M}_{0,n,0}^{CP^{N-1}} \label{23}$$ We have $$c(M_{N})= c_{T}(\tilde{\pi}_{k}^{*}({\cal E}_{Nn+1})) \label{24}$$ Finally the representation (\[20\]) turns into $$\begin{aligned} \langle {\cal O}_{e^{j_{1}}}(z_{1}){\cal O}_{e^{j_{2}}}(z_{2}) \cdots {\cal O}_{e^{j_{k}}}(z_{k}) \rangle _{n,alt,grav.} \nonumber\\ = \int_{{\cal M}_{0,n,k}^{CP^{N-1}}} c_{T}(\tilde{\pi}_{k}^{*}({\cal E}_{Nn+1})) \wedge \phi^{*}_{1}(c_{1}^{j_1}(H)) \wedge \cdots \wedge \phi^{*}_{k}(c_{1}^{j_{k}}(H)) \label{25}\end{aligned}$$ Then we can calculate the correlation functions using the torus action method. Review of the Torus Action Method ================================= Introduction of the Torus Action and the Bott Residue Formula ------------------------------------------------------------- Torus action method is the strategy to use the Bott residue formula [@bott] which reduces the integral of Chern classes of vector bundle on $X$ to the one on $X_{f}$ of the fixed point set of the torus action flow on $X$ to the case where $X$ is ${\cal M}_{0,n,k}^{CP^{N-1}}$. First, let us introduce the torus action flow on $CP^{N-1}$, $$\begin{aligned} T_{t}: CP^{N-1}& \mapsto& CP^{N-1}\nonumber\\ (X_{1},X_{2},\cdots,X_{N})& \mapsto & (e^{\lambda_{1}t}X_{1},e^{\lambda_{2}t}X_{2},\cdots,e^{\lambda_{N}t}X_{N}) \nonumber\\ && (t \in C^{\times}) \label{26}\end{aligned}$$ where $\lambda_{i} \in C$ is the character of the flow.Then (\[26\]) induce the flow on ${\cal M}_{0,n,k}^{CP^{N-1}}$ from the compatibility with the evaluation map. $$\begin{aligned} &&\phi_{i}(T_{t}((z_{1},z_{2},\cdots ,z_{k},f)/\sim))\nonumber\\ &:=& T_{t} \circ \phi_{i}((z_{1},z_{2},\cdots ,z_{k},f)/\sim)\nonumber\\ & =& T_{t} \circ f(z_{i}) \label{27}\end{aligned}$$ Next, we introduce the Bott residue formula. For simplicity, we use $X$ for ${\cal M}_{0,n,k}^{CP^{N-1}}$.Let ${\cal E}_{1},\cdots ,{\cal E}_{m}$ be a holomorphic vector bundle on $X$, and $X_{f}$ be the fixed point set of $X$ under the flow $(\ref{27})$. We can decompose $X_{f}$ as the sum of the connected components $X_{\gamma}$. $$X_{f} = \bigcup_{\gamma} X_{\gamma} \label{28}$$ Then consider ${\cal E}_{i}|_{X_{\gamma}}$ and the normal bundle ${\cal N}_{\gamma} \simeq T^{\prime}X|_{X_{\gamma}}/T^{\prime}X_{\gamma}$ and decompose them into the eigen vector bundle under the torus action $T_{t}$,i.e., $$\begin{aligned} {\cal E}_{i}|_{X_{\gamma}} &\cong& \bigoplus_{j=1}^{m_{{\cal E}_{i}}} {\cal E}_{i,j}^{\gamma,f_{i,j}(\lambda_{*})}\nonumber\\ {\cal N}_{\gamma} &\cong& \bigoplus_{j=1}^{m_{{\cal N}}} {\cal N}_{j}^{\gamma,g_{j}(\lambda_{*})} \label{29}\end{aligned}$$ where $$\begin{aligned} T_{t}({\cal E}_{i,j}^{\gamma,f_{i,j}(\lambda_{*})})& = & e^{f_{i,j}(\lambda_{*})t}{\cal E}_{i,j}^{\gamma,f_{i,j}(\lambda_{*})} \nonumber\\ T_{t}({\cal N}_{j}^{\gamma,g_{j}(\lambda_{*})})& = & e^{g_{j}(\lambda_{*})t} {\cal N}_{j}^{\gamma,g_{j}(\lambda_{*})} \label{31}\end{aligned}$$ and we set $$\begin{aligned} rk({\cal E}_{i,j}^{\gamma,f_{i,j}(\lambda_{*})}) = r_{\cal E}(i,j)\nonumber\\ rk({\cal N}_{j}^{\gamma,g_{j}(\lambda_{*})}) = r_{\cal N}(j) \label{32}\end{aligned}$$ We can represent the total Chern class of ${\cal E}_{i,j}^{\gamma,f_{i,j} (\lambda_{*})}$ and ${\cal N}_{j}^{\gamma,g_{j}(\lambda_{*})}$ as the product of first Chern class of formal line bundles as follows. $$\begin{aligned} c({\cal E}_{i,j}^{\gamma,f_{i,j}(\lambda_{*})}|_{X_{\gamma}} ) &=& \prod_{j=1}^{m_{{\cal E}_{i}}}\prod_{k=1}^{r_{\cal E}(i,j)} (1+t \cdot e_{i,j,k}^{\gamma,f_{i,j}(\lambda_{*})}) \nonumber\\ c({\cal N}_{\gamma}) &=& \prod_{j=1}^{m_{{\cal N}}}\prod_{k=1}^{r_{\cal N}(j)} (1+t \cdot n_{j,k}^{\gamma,g_{j}(\lambda_{*})}) \label{33}\end{aligned}$$ Top Chern classes are given as the coefficient form of $t^{k}$ of highest degree. With these preparations, we introduce the Bott residue formula. $$\begin{aligned} \lefteqn{\int_{X} \prod_{i} c_{T}^{\alpha_{i}}({\cal E}_{i}) =} \nonumber\\ &&\sum_{\gamma} \int_{X_{\gamma}} \frac{\prod_{i}\prod_{j=1}^{m_{{\cal E}_{i}}} \prod_{k=1}^{r_{{\cal E}}(i,j)}(e_{i,j,k}^{\gamma,f_{i,j}(\lambda_{*})} +f_{i,j}(\lambda_{*}))^{\alpha_{i}}}{\prod_{j=1}^{m_{\cal N}} \prod_{k=1}^{r_{\cal N}(i)} (n_{i,j}^{\gamma,g_{j}(\lambda_{*})}+g_{j}(\lambda_{*}))} \label{34}\end{aligned}$$ Construction of Fixed Point Set ------------------------------- Fixed points of $CP^{N-1}$ under $T_{t}$ are given by the projective equivalence $$p_{i} := (0,0,\cdots,0,\overbrace{1}^{i},0,\cdots,0) \label{35}$$ Then, we can find the fundamental maps $l_{i,j}^{d}$ from $CP^{1} \mapsto CP^{N-1}$ which remain fixed under $T_{t}$ as the degree d maps which connect $p_{i}$ and $p_{j}$. $$l_{i,j}^{d}: (s,t) \mapsto (0,\cdots,0,\overbrace{s^{d}}^{i}, 0,\cdots,0,\overbrace{t^{d}}^{j},0,\cdots,0) \label{36}$$ Of course $l_{i,j}^{d}$ is kept fixed under $SL(2,C)$ equivalence. But now that we have coupled gravity with the theory, we have to consider the boundary components of moduli space of $CP^{1}$, i.e., stable curves. Stable curve ${\cal C}$ with $k$-punctures is constructed with the set of $CP^{1}$’s $\{C_{\alpha}\}$ with punctures assigned on them and additional punctures of double singularity which connect two components of $C_{\alpha}$’s. Then we can translate the condition into the condition that the genus of stable curve is zero into imposing its arithmetic genus to be zero.In geometrical language, if we represent $C_{\alpha}$ as a line and define a figure with lines which intersect at sigular punctures, this is equivalent to the non-existence of closed loops in it. This addition makes us to introduce stable maps which map stable curves to $CP^{N-1}$. With these considerations, we can label the connected components of the fixed point set ${\cal M}_{0,n,k,f}^{CP^{N-1}}$ with a tree graph $\Gamma$ with the following structure. We denote them by ${\cal M}_{0,n,k}^{CP^{N-1}}(\Gamma)$. The rules of correspondences are, - The vertices $v \in Vert(\Gamma)$ correspond to the connected component $C_{v}$ of\ $f^{-1}(p_{1},\cdots,p_{N})$. $C_{v}$ can be a sum of connected irreducible components of ${\cal C}$ or be a point. - The edges ${\alpha} \in Edge(\Gamma)$ correspond to the irreducible component $C_{\alpha}$mapped to $l_{i,j}^{d}$. Then we have to add the additional structures to $\Gamma$, - We label each $v \in Vert(\Gamma)$ by $f_{v} \in \{1,2,\cdots ,N\}$ which is defined by $p_{f_{v}} = f(C_{v})$. - The $k$-punctures are distributed among the vertices $v \in Vert(\Gamma)$. We represent this distribution by $S_{v} \in \{1,2,\cdots,k\}$. - We attach degree $d_{\alpha}$ to each $\alpha \in Edge(\Gamma)$ defined by the degree of $l_{i,j}^{d}$. We have to set punctures on the vertices $Vert(\Gamma)$ because if we put punctures on $C^{\alpha}$, they move with the flow $T_{t}$, which contradicts with the assumption of fixed point sets. Then we can construct ${\cal M}_{0,n,k}^{CP^{N-1}}(\Gamma)$ under conditions that emerges from the above three structures, - If ${\alpha} \in Edge(\Gamma)$ connects $v,u \in Vert(\Gamma)$, $f_{u} \neq f_{v}$. - $\{1,2,\cdots,k\} = \coprod_{v \in Vert(\Gamma)} S_{v}$. - $\sum_{\alpha \in Edge(\Gamma)} d_{\alpha} = n$ Then we have $${\cal M}_{0,n,k}^{CP^{N-1}}(\Gamma) \cong \prod_{v \in Vert(\Gamma)} ({\cal M}_{0,S_{v}})/(Aut(\Gamma)) \label{38}$$ where ${\cal M}_{0,S_{v}}$ is the moduli space of complex structure of $CP^{1}$ with $S_{v}$ punctures. It represents the gravitational degree of freedom of $C_{v}$.According to Kontsevich, division by $Aut(\Gamma)$ reflects the orbispace structure of ${\cal M}_{0,n,k}^{CP^{N-1}}$. It may reflect the multiplicity of the degeneration of stable maps. Determination of the contribution from Normal and Vector bundles ---------------------------------------------------------------- ### Contributions from ${\cal N}_{{\cal M}(\Gamma)}^{abs}$ With these preparations, we determine the contribution from ${\cal M}_{0,n,k}^{CP^{N-1}}$ (in the following discussion we abbriviate the notation as ${\cal M}(\Gamma))$ to (\[34\]). First, we calculate the contribution from ${\cal N}_{{\cal M}(\Gamma)}$. Following Kontsevich,we will use the expression of vector bundles as the K-group \[ \],which translates sum and quotient operations into addition and subtraction. Then we have $$[{\cal N}_{{\cal M}(\Gamma)}] = [T^{\prime}{\cal M}|_{{\cal M}(\Gamma)}] -[T^{\prime}{\cal M}(\Gamma)] \label{39}$$ If we set $${\cal C}= \tilde{\bigcup}_{\alpha} C_{\alpha}$$ (where $$\tilde{\bigcup}_{\alpha}$$ means a sum with double-singularity gluing operation.), $[T^{\prime}{\cal M}|_{{\cal M}(\Gamma)}]$ consists of the following degrees of freedom, - Moving $f({\cal C})$ in $CP^{N-1}$. - Resolution of singularities of ${\cal C}$, i.e., from $xy =0$ to $xy = \epsilon$. - Moving puncture degrees of freedom. And we have $$\begin{aligned} [T^{\prime}{\cal M}|_{{\cal M}(\Gamma)}] &=& [H^{0}({\cal C},f^{*}(T^{\prime}CP^{N-1}))]\nonumber\\ &+& \sum_{{z \in C_{\alpha} \cap C_{\beta}} \atop {{\alpha}\ne{\beta}}} [T^{\prime}_{z}C_{\alpha}\otimes T^{\prime}_{z}C_{\beta}]\nonumber\\ &+& \sum_{{z \in C_{\alpha} \cap C_{\beta}} \atop {{\alpha}\ne {\beta}}} [T^{\prime}_{z}C_{\alpha}]+ [T^{\prime}_{z}C_{\beta}]\nonumber\\ &+&\sum_{z_{i},i \in \{1,\cdots,k\}}[T^{\prime}_{z_{i}}{\cal C}]- \sum_{\alpha}[H^{0}(C^{\alpha},T^{\prime}C^{\alpha})] \label{40}\end{aligned}$$ The last term of (\[40\]) corresponds to devision by $SL(2,C)$ of each component $C_{\alpha}$. From (\[38\]) ${\cal M}(\Gamma)$ has continuous degrees of freedom which come only from $C_{\alpha}$ mapped to a point, we have $$\begin{aligned} [T^{\prime}{\cal M}(\Gamma)]& =& \sum_{{z \in C_{\alpha} \cap C_{\beta}} \atop{\alpha \ne \beta\;\alpha,\beta \notin Edge(\Gamma)}} [T^{\prime}_{z}C_{\alpha}\otimes T^{\prime}_{z}C_{\beta}] \nonumber\\ &+& \sum_{{z \in C_{\alpha} \cap C_{\beta}} \atop {\alpha \ne \beta, \alpha \notin Edge(\Gamma)}} [T^{\prime}_{z}C_{\alpha}]\nonumber\\ &+& \sum_{z_{i},i \in \{1,2,\cdots,k\}} [T^{\prime}_{z_{i}}{\cal C}]\nonumber\\ &-& \sum_{\alpha\notin Edge(\Gamma)}[H^{0}(C^{\alpha},T^{\prime}C^{\alpha})] \label{41}\end{aligned}$$ where we used the fact that all the punctures lie in the component mapped to a point . From (\[40\]) and (\[41\]), we have $$[{\cal N}_{{\cal M}(\Gamma)}] = [H^{0}({\cal C},f^{*}(T^{\prime}CP^{N-1}))] + [{\cal N}_{{\cal M}(\Gamma)}^{abs}] \label{42}$$ where $$\begin{aligned} [{\cal N}_{{\cal M}(\Gamma)}^{abs}] &:= &\sum_{{z \in C_{\alpha} \cap C_{\beta}}\atop{\alpha \ne \beta\; \alpha,\beta \in Edge(\Gamma)}} [T^{\prime}_{z}C_{\alpha}\otimes T^{\prime}_{z}C_{\beta}] \\ \label{43} &+& \sum_{{z \in C_{\alpha} \cap C_{\beta}} \atop {\alpha \in Edge(\Gamma),\beta\notin Edge(\Gamma)}} [T^{\prime}_{z}C_{\alpha}\otimes T^{\prime}_{z}C_{\beta}]\\ \label{44} &+& \sum_{{z \in C_{\alpha} \cap C_{\beta}}\atop{\alpha \ne \beta \;\alpha\in Edge(\Gamma)}} [T^{\prime}_{z}C_{\alpha}] - \sum_{\alpha\in Edge(\Gamma)} [H^{0}(C^{\alpha},T^{\prime}C^{\alpha})] \label{45}\end{aligned}$$ Then we determine the contribution from the first term of (\[42\]),and (\[43\]),(\[44\]),and (\[45\]). First, consider the contribution from (\[43\]). Since $\alpha,\beta \in Edge(\Gamma)$, $T_{z}C_{\alpha}$ and $T_{z}C_{\beta}$’s are trivial as the line bundle on ${\cal M}(\Gamma)$. Let $C_{\alpha}$ and $C_{\beta}$ be mapped to $l_{i,j}^{d_{\alpha}}$ and $l_{i,l}^{d_{\beta}}$. $$\begin{aligned} C_{\alpha} : (z_{1},z_{2}) \mapsto (0,\cdots,0,\overbrace{z_{1}^{d_{\alpha}}}^{i},0,\cdots,0, \overbrace{z_{2}^{d_{\alpha}}}^{j},0,\cdots,0) \\ \label{46} C_{\beta} : (w_{1},w_{2}) \mapsto (0,\cdots,0, \overbrace{w_{1}^{d_{\beta}}}^{i} ,0,\cdots,0,\overbrace{w_{2}^{d_{\beta}}}^{l},0,\cdots,0) \\ \label{47}\end{aligned}$$ Local coordinate around $z \in C_{\alpha} \cap C_{\beta}$ on $C_{\alpha}$ and $C_{\beta}$ are $\frac{z_{2}}{z_{1}}$ and $\frac{w_{2}}{w_{1}}$ ,and we have $$T^{\prime}_{z}C_{\alpha}\otimes T^{\prime}_{z}C_{\beta} \cong \frac{d}{d(\frac{z_{2}}{z_{1}})}\otimes\frac{d}{d(\frac{w_{2}}{w_{1}})} \label{48}$$ Definition of torus action (\[27\]) leads us to $$\begin{aligned} z_{1} \mapsto z_{1}e^{\frac{\lambda_{i}}{d_{\alpha}}t}&z_{2} \mapsto z_{2}e^{\frac{\lambda_{j}}{d_{\alpha}}t}\nonumber\\ w_{1} \mapsto w_{1}e^{\frac{\lambda_{i}}{d_{\beta}}t}&w_{2} \mapsto w_{2}e^{\frac{\lambda_{l}}{d_{\beta}}t} \label{49}\end{aligned}$$ and $$T^{\prime}_{z}C_{\alpha}\otimes T^{\prime}_{z}C_{\beta} \mapsto e^{(\frac{\lambda_{i}-\lambda_{j}}{d_{\alpha}}+ \frac{\lambda_{i}-\lambda_{l}}{d_{\beta}})t} T^{\prime}_{z}C_{\alpha}\otimes T^{\prime}_{z}C_{\beta} \label{50}$$ The result is, $$(\mbox {Contribution from (\ref{43}) to (\ref{34})}) = \prod_{{C_{\alpha}\cap C_{\beta}\neq\emptyset}\atop {\alpha\neq\beta\;\alpha,\beta\in Edge(\Gamma)}} \frac{1}{\frac{\lambda_{i}-\lambda_{j}}{d_{\alpha}}+ \frac{\lambda_{i}-\lambda_{l}}{d_{\beta}}} \label{51}$$ Again following Kontsevich, we introduce the notation “Flag” $F=(v,\alpha)$ which represents edge $\alpha$ with a direction specified by the source vertex $v$. We define $$w_{F}:= \frac{\lambda_{f_{v}}-\lambda_{f_{u}}}{d_{\alpha}} \label{52}$$ Then the r.h.s of (\[51\]) can be rewritten as follows. $$\prod_{{v\in Vert(\Gamma)}\atop{val(v)=2}\quad{{}^{\sharp}S_{v}}=0} \frac{1}{w_{F_{1}(v)}+w_{F_{2}(v)}} \label{53}$$ where $val(v)$ represents the valency of $v$ and $F_{1}(v)$ and $F_{2}(v)$ are the flags whose sources are $v$. Note that in this case $f^{-1}(v)$ is a point. Next we consider the contributions from (\[44\]). Again from (\[41\]), $T_{z}C_{\alpha}$ is trivial as the line bundle on ${\cal M}(\Gamma)$ but has an eigenvalue $w_{F}$ as in the derivation of (\[50\]).On the other hand, $T^{\prime}_{z}C_{\beta}$ has trivial torus action (because $C_{\beta}$ is mapped to a point) but non trivial line bundle on ${\cal M}(\Gamma)$.And if ${}^{\sharp}(\mbox {punctures on} C_{v})\geq 3$,${\cal M}_{0,S_{v}}$ is well-defined and we have $$\begin{aligned} (\mbox{Contribution from (\ref{44}) to (\ref{34})}) &=& \prod_{v \in Vert(\Gamma)}(\int_{{\cal M}_{0,val(v)+{}^{\sharp}S_{v}}} \prod_{{flags}\atop {F=(v,\alpha)}} \frac{1}{w_{F}+c_{1}(T^{\prime}_{z_{F}}(C_{v}))})\nonumber\\ &&(val(v)+{}^{\sharp}S_{v} \geq 3) \label{54}\end{aligned}$$ where $z_{F}$ represents the gluing point of $C_{v}$ and $F$. We can evaluate the r.h.s of (\[54\]) by expanding in terms of $\frac{1}{w_{F}}$ and using the fact that $c_{1}(T^{\prime}_{z_{F}}C_{v}) = - c_{1}(T^{\prime,*}_{z_{F}}C_{v})$. Expansion coefficients are intersection numbers of Mumford-Morita class on the $CP^{1}$-moduli space ,which is identified as the correlation function of gravitational descendants by Witten [@witten2].Continuing the calculation,we have $$(\mbox{r.h.s of (\ref{54})})= \prod_{v \in Vert(\Gamma)} (\sum_{{d_{1},\cdots, d_{val(v)}\geq 0}\atop{\sum d_{i}= val(v)+{}^{\sharp}S_{v}-3}} \prod_{{flags}\atop{F=(v,\alpha)}}w_{F}^{-d_{i}-1}\langle \sigma_{d_{1}}\cdots\sigma_{d_{val{v}}} \overbrace{P\cdots P}^{{}^{\sharp}S_{v}\mbox{times}}\rangle) \label{55}$$ $\langle \sigma_{d_{1}}\cdots\sigma_{d_{val{v}}}\overbrace{P\cdots P}^ {{}^{\sharp}S_{v}\mbox{times}}\rangle$ is calculated in [@dij], $$\langle \sigma_{d_{1}}\cdots\sigma_{d_{val{v}}}\overbrace{P\cdots P}^ {{}^{\sharp}S_{v}\mbox{times}}\rangle = \frac{(val(v)+{}^{\sharp}S_{v}-3)!}{d_{1}!\cdots d_{val(v)}!} \label{56}$$ Combining (\[54\]),(\[55\]) and (\[56\]), we have $$\begin{aligned} (\mbox{Contribution from (\ref{44}) to (\ref{34})}) &=& \prod_{v \in Vert(\Gamma)}\prod_{{flags}\atop{F=(v,\alpha)}} w_{F}^{-1}(\sum_{{flags}\atop {F=(v,\alpha)}}w_{F}^{-1})^{val(v)+ {}^{\sharp}S_{v}-3}\nonumber\\ &&(val(v)+{}^{\sharp}S_{v} \geq 3) \label{57}\end{aligned}$$ Then we consider (\[45\]). Contributions of the first terms are, as before $$\prod_{{C_{\alpha}\cap C_{\beta}}\atop {\alpha\neq\beta\;\alpha\in Edge(\Gamma)}}\frac{1}{w_{F_{i}(\alpha)}} \label{58}$$ where $F_{i}(\alpha)$’s are two flags having $\alpha$ as their edges. The second terms which represents the automorphism group degrees of freedom of edge components can be expressed by the tangent bundles on the inverse images of two vertices of the edges and scaling transformation degree of freedom fixing the punctures (We denote it as $[0]$). In terms of the K-group, we have $$\begin{aligned} -\sum_{\alpha \in Edge(\Gamma)}[H^{0}(C_{\alpha},T^{\prime}C_{\alpha})] \nonumber\\ = - \sum_{\alpha \in Edge(\Gamma)}([T^{\prime}_{z_{1}(\alpha)}C_{\alpha}] +[0]+[T^{\prime}_{z_{2}(\alpha)}C_{\alpha}]) \label{59}\end{aligned}$$ And contributions to (\[34\]) are $$\prod_{\alpha \in Edge(\Gamma)}w_{F_{1}}(\alpha)\cdot w_{F_{2}}(\alpha)\cdot C([0]) \label{60}$$ where $C([0])$ represents the factor from $[0]$. Multiplying (\[58\]) and (\[60\]), what remains except for $C([0])$ is the products of $w_{F}$’s whose edges have only one double singularity. In other words, the corresponding $F= (v,\alpha)$ has $val(v) = 1$ and $f^{-1}(v)$ is a point. We have $$(\mbox{Contributions from (\ref{45})}) = \prod_{{v\in Vert(\Gamma)}\atop{{val(v)=1}\quad{{}^{\sharp}S_{v}=0}}} \prod_{{flags}\atop {F=(v,\alpha)}}w_{F}\prod_{\alpha\in Edge(\Gamma)}C([0]) \label{61}$$ After all, from (\[53\]),(\[57\]) and (\[61\]), we put all the factors from $[{\cal N}_{{\cal M}(\Gamma)}^{abs}]$ into the form, $$\prod_{v\in Vert(\Gamma)}\prod_{{flags}\atop {F=(v,\alpha)}}w_{F}^{-1} (\sum_{{flags}\atop{F=(v,\alpha)}}w_{F}^{-1})^{val(v) +{}^{\sharp}S_{v}-3}\prod_{\alpha\in Edge(\Gamma)}C([0]) \label{62}$$ ### Determination of the contributions from $[H^{0}({\cal C},f^{*}(T^{\prime}CP^{N-1}))]$ Since $$f({\cal C})= \tilde{\bigcup}_{\alpha\in Edge(\Gamma)}f(C_{\alpha})$$, we can construct $[H^{0}({\cal C},f^{*}(T^{\prime}CP^{N-1}))]$ by gluing $$\bigoplus_{\alpha\in Edge(\Gamma)} [H^{0}(C_{\alpha},f^{*}(T^{\prime}CP^{N-1}))]$$ at $p_{f_{v}}$.This process can be described using exact sequences, $$\begin{aligned} 0 \mapsto H^{0}({\cal C},f^{*}(T^{\prime}CP^{N-1}))\mapsto\nonumber\\ \bigoplus_{\alpha\in Edge(\Gamma)}H^{0}(C_{\alpha},f^{*}(T^{\prime}CP^{N-1})) \mapsto \bigoplus_{v \in Vert(\Gamma)}C^{val(v)-1}\otimes T^{\prime}_{p_{f_{v}}}CP^{N-1} \mapsto 0 \label{63}\end{aligned}$$ This contribution is then given as the contribution from the second term divided by the one from from the third term. As the independent basis of $H^{0}(C_{\alpha},f^{*}(T^{\prime}CP^{N-1}))$ describing the deformation of $f(C_{\alpha})$ in $CP^{N-1}$ where $C_{\alpha}$ is $$C_{\alpha}:\quad (z_{1},z_{2})\mapsto (0,\cdots,0, \overbrace{z_{1}^{d_{\alpha}}}^{i},0,\cdots,0, \overbrace{z_{2}^{d_{\alpha}}}^{j}, 0,\cdots,0) \label{64}$$ , we have $$\begin{aligned} (0,\cdots,0,\overbrace{z_{1}^{d_{\alpha}}+ \epsilon z_{1}^{m}z_{2}^{d_{\alpha}-m}}^{i}, 0,\cdots,0,\overbrace{z_{2}^{d_{\alpha}}}^{j}, 0,\cdots,0)\\ \label{65} (0,\cdots,0,\overbrace{z_{1}^{d_{\alpha}}}^{i},0,\cdots,0, \overbrace{z_{2}^{d_{\alpha}}+ \epsilon z_{1}^{m}z_{2}^{d_{\alpha}-m}}^{j}, 0,\cdots,0)\\ \label{66} (0,\cdots,0,\overbrace{\epsilon z_{1}^{m}z_{2}^{d_{\alpha}-m}}^{k},0,\cdots, 0,\overbrace{z_{1}^{d_{\alpha}}}^{i},0,\cdots,0, \overbrace{z_{2}^{d_{\alpha}}}^{j},0,\cdots,0) \label{67}\end{aligned}$$ We can write these basis in more sophisticated form, $$\begin{aligned} a)\qquad &(\frac{z_{1}}{z_{2}})^{m}X_{i}\frac{\partial}{\partial X_{i}}& \qquad(-d_{\alpha}\leq m \leq d_{\alpha})\\ \label{68} b)\qquad &(\frac{z_{1}}{z_{2}})^{m}X_{j}\frac{\partial}{\partial X_{k}}& \qquad(0\leq m \leq d_{\alpha})\qquad k\neq i,j \label{69}\end{aligned}$$ This expression directly leads us to $$\begin{aligned} (\mbox{Contribution from (\ref{68})})& = &\frac{1}{m\cdot w_{F}}\qquad (m\neq 0)\\ \label{70} &&\frac{1}{C([0])}\qquad (m=0)\\ \label{70b} (\mbox{Contribution from (\ref{69})})& = &\frac{1}{m\cdot w_{F}+\lambda_{j}-\lambda_{k}} \label{71}\end{aligned}$$ And from $T^{\prime}_{p_{f_{v}}}CP^{N-1}\simeq \oplus_{j\neq p_{f_{v}}}\frac{\partial}{\partial(\frac{X_{j}}{X_{f_{v}}})}$, $$(\mbox{Contributions from $C^{val(v)-1}\otimes T^{\prime}_{p_{f_{v}}}CP^{N-1}$}) =\prod_{v\in Vert(\Gamma)} (\lambda_{f_{v}}-\lambda_{j})^{val(v)-1} \label{72}$$ Combining (\[70\]),(\[70b\]),(\[71\]) and (\[72\]), we have $$\begin{aligned} \prod_{flags\quad F=(v,\alpha)}(w_{F})^{-d_{\alpha}} \prod_{{\alpha\in Edge(\Gamma)}\atop{(u,v): vertices\;of\:\alpha}} \prod_{k\neq f_{u},f_{v}}\prod_{m=0}^{d_{\alpha}} (\frac{m\lambda_{f_{u}}+ (d_{\alpha}-m)\lambda_{f_{v}}}{d_{\alpha}}-\lambda_{k})^{-1} \nonumber\\ \prod_{\alpha\in Edge(\Gamma)}(d_{\alpha}!)^{-2} (C([0]))^{-1}\prod_{v\in Vert(\Gamma)}(\prod_{j:j\neq f_{v}} (\lambda_{f_{v}}-\lambda_{j}))^{val(v)-1} \label{73}\end{aligned}$$ ### Factors from Vector Bundles ${\cal E}_{i}$ First, we calculate the factors from ${\cal E}_{Nn+1}$. As we have mentioned in section 2, this fiber locally corresponds to $[H^{0}({\cal C},f^{*}({\cal O}(N\cdot H)))]$. We can construct it as in (\[62\]), by the exact sequences, $$\begin{aligned} 0 \mapsto H^{0}({\cal C},f^{*}({\cal O}(N H)))\mapsto \nonumber\\ \bigoplus_{\alpha\in Edge(\Gamma)}H^{0}(C_{\alpha},f^{*}({\cal O} (N H))) \mapsto \bigoplus_{v \in Vert(\Gamma)}C^{val(v)-1}\otimes {\cal O}_{p_{f_{v}}} (N H) \mapsto 0 \label{74}\end{aligned}$$ Then since the basis of $H^{0}(C_{\alpha},f^{*}({\cal O}(N\cdot H)))$ are given as $$\{z_{1}^{N d_{\alpha}},z_{1}^{N d_{\alpha}-1}z_{2},\cdots,z_{2}^{Nd_{\alpha}}\} \label{75}$$ and the section of ${\cal O}_{p_{f_{v}}}(N H)$ is $X_{f_{v}}^{N}$, we have $$\begin{aligned} (\mbox{Contributions from $c_{T}({\cal E}_{Nn+1})$}) &=& \prod_{{\alpha\in Edge(\Gamma)}\atop{(v_{1},v_{2}): vertices\; of\; \alpha}} \prod_{a=0}^{Nd_{\alpha}}(\frac{a\lambda_{f_{v_{1}}}+(N-a) \lambda_{f_{v_{2}}}}{d_{\alpha}})\nonumber\\ &&\prod_{v\in Vert(\Gamma)} (N\lambda_{f_{v}})^{1-val(v)} \label{76}\end{aligned}$$ Next, we determine the factor from $\phi^{*}_{i}(c_{1}^{j_{i}}(H))$. From the argument of $\S 3.2$,puncture $i$ lies on the vertex $v(i)$ of $\Gamma$,and $\phi^{*}_{i}(c_{1}^{j_{i}}(H))$ reduces to ${\cal O}_{p_{f_{v(i)}}}(j_{i}H)$. This leads to $$(\mbox{Contributions from $\phi^{*}_{i}(c_{1}^{j_{i}}(H))$}) = \lambda^{j_{i}}_{f_{v(i)}} \label{77}$$ ### Local Appendix We have to divide the above factors by ${}^{\sharp}Aut(\Gamma)$ coming from (\[38\]) and in practice, we have to multiply a factor $\frac{1}{d_{\alpha}}$ for each edge $\alpha$. We cannot justify the reason for this factor at this stage. Some Explict Calculation of Amplitudes ====================================== From the argument of section 2, we constructed representation of amplitudes as an integral of forms on ${\cal M}_{0,n,k}^{CP^{N-1}}$, and the strategy of section 3 enables us to compute them as a sum of amplitudes corresponding to tree graphs. For some examples, we calculate $\langle {\cal O}_{e^{N-4}} \rangle_{1}$,$\langle {\cal O}_{e^{N-4}}\rangle_{2}$ and $\langle {\cal O}_{e^{N-4}}\rangle_{3}$. First, we write out tree graphs that contribute to the amplitudes up to degree 3. (See [**Fig.1**]{}.) In [**Fig.1**]{}, we abbriviate the external insertion of “punctures”. So in calculation,we have to add all the cases of external operator insertions of $\langle {\cal O}_{e^{N-4}}\rangle $ to vertices. Note that the two character numbers (for example “i”) of neighboring vertices never coincide with each other. Then direct application of the argument of section 3 leads to the following formula. $$\begin{aligned} \langle {\cal O}_{e^{N-4}}\rangle_{1} &=& \frac{1}{2}\sum_{i\ne j}(\prod_{k\ne i,j}(\lambda_{i}-\lambda_{k})^{-1} (\lambda_{j}-\lambda_{k})^{-1}\prod_{a=0}^{N}(a\lambda_{i}+(N-a)\lambda_{j}) (\frac{\lambda_{i}^{N-4}-\lambda_{j}^{N-4}}{w_{F_{1}}}))\nonumber\\ &&(\mbox{from (a)})\nonumber\\ \langle {\cal O}_{e^{N-4}}\rangle_{2} &=& \frac{1}{2}\sum_{{i\ne j}\atop{j\ne k}}(\frac{1}{N\lambda_{j}} (\frac{\lambda^{N-4}_{i}w_{F_{4}} + \lambda^{N-4}_{k}w_{F_{1}}} {w_{F_{2}}+w_{F_{3}}} + \lambda_{j}^{N-4})\nonumber\\ &&\frac{1}{w_{F_{1}}w_{F_{2}}w_{F_{3}}w_{F_{4}}} \prod_{n\ne j}(\lambda_{j}-\lambda_{n})\nonumber\\ &&\prod_{m_{1}\ne i,j}(\lambda_{i}-\lambda_{m_{1}})^{-1} (\lambda_{j}-\lambda_{m_{1}})^{-1} \prod_{m_{2}\ne j,k}(\lambda_{j}-\lambda_{m_{2}})^{-1} (\lambda_{k}-\lambda_{m_{2}})^{-1}\nonumber\\ &&\prod_{a_{1}=0}^{N}(a_{1}\lambda_{i}+(N-a_{1})\lambda_{j}) \prod_{a_{2}=0}^{N}(a_{2}\lambda_{j}+(N-a_{2})\lambda_{k})\nonumber\\ &&(\mbox{from (b)})\nonumber\\ &+&\frac{1}{4}\sum_{i\ne j}((\lambda_{i}^{N-4}-\lambda_{j}^{N-4})w_{F_{2}} \frac{1}{w_{F_{1}}^{2}w_{F_{2}}^{2}}\nonumber\\ &&\prod_{k\ne i,j}(\lambda_{i}-\lambda_{k})^{-1} (\lambda_{j}-\lambda_{k})^{-1}(\frac{\lambda_{i}+\lambda_{j}}{2} -\lambda_{k})^{-1}\nonumber\\ &&\prod_{a=0}^{2N}(\frac{a\lambda_{j}+(2N-a)\lambda_{k}}{2}))\nonumber\\ &&(\mbox{from (c)})\nonumber\\ \langle {\cal O}_{e^{N-4}} \rangle_{3}&=& \frac{1}{2}\sum_{{i\ne j}\atop{j\ne k,k\ne l}}(( \lambda_{i}^{N-4}\frac{1}{w_{F_{2}}+w_{F_{3}}}\frac{1}{w_{F_{4}}+w_{F_{5}}} w_{F_{6}}\nonumber\\ &+&w_{F_{1}}\frac{\lambda_{j}^{N-4}}{w_{F_{2}}w_{F_{3}}} \frac{1}{w_{F_{4}}+w_{F_{5}}} w_{F_{6}}\nonumber\\ &+&w_{F_{1}}\frac{1}{w_{F_{2}}+w_{F_{3}}} \frac{\lambda_{k}^{N-4}}{w_{F_{4}}w_{F_{5}}} w_{F_{6}}\nonumber\\ &+&w_{F_{1}}\frac{1}{w_{F_{2}}+w_{F_{3}}} \frac{1}{w_{F_{4}}+w_{F_{5}}} \lambda_{l}^{N-4})\nonumber\\ &&\frac{1}{w_{F_{1}}w_{F_{2}}w_{F_{3}}w_{F_{4}}w_{F_{5}}w_{F_{6}}} \frac{1}{(N\lambda_{j}N\lambda_{k})}\nonumber\\ &&\prod_{n_{1}\ne j}(\lambda_{j}-\lambda_{n_{1}}) \prod_{n_{2}\ne k}(\lambda_{k}-\lambda_{n_{2}})\nonumber\\ &&\prod_{m_{1}\ne i,j}(\lambda_{i}-\lambda_{m_{1}})^{-1} (\lambda_{j}-\lambda_{m_{1}})^{-1}\nonumber\\ &&\prod_{m_{2}\ne j,k}(\lambda_{j}-\lambda_{m_{2}})^{-1} (\lambda_{k}-\lambda_{m_{2}})^{-1}\nonumber\\ &&\prod_{m_{3}\ne k,l}(\lambda_{k}-\lambda_{m_{3}})^{-1} (\lambda_{l}-\lambda_{m_{3}})^{-1}\nonumber\\ &&\prod_{a_{1}=0}^{N}(a_{1}\lambda_{i}+(N-a_{1})\lambda_{j}) \prod_{a_{2}=0}^{N}(a_{2}\lambda_{j}+(N-a_{2})\lambda_{k})\nonumber\\ &&\prod_{a_{3}=0}^{N}(a_{3}\lambda_{k}+(N-a_{3})\lambda_{l}))\nonumber\\ &&(\mbox{from (d)})\nonumber\\ &+&\frac{1}{2}\sum_{{i\ne j}\atop{j\ne k}}(\frac{1}{4} (\lambda_{i}^{N-4}\frac{1}{w_{F_{2}}+w_{F_{3}}}w_{F_{4}}+ w_{F_{1}}\frac{\lambda_{j}^{N-4}}{w_{F_{2}}w_{F_{3}}}w_{F_{4}}+ w_{F_{1}}\frac{1}{w_{F_{2}}+w_{F_{3}}}\lambda_{k}^{N-4})\nonumber\\ &&\frac{1}{N\lambda_{j}}\frac{1}{w_{F_{1}}^{2}w_{F_{2}}^{2}} \frac{1}{w_{F_{3}}w_{F_{4}}} \prod_{n\ne j}(\lambda_{j}-\lambda_{n})\nonumber\\ &&\prod_{m_{1}\ne i,j}(\lambda_{i}-\lambda_{m_{1}})^{-1} (\lambda_{j}-\lambda_{m_{1}})^{-1}(\frac{\lambda_{i}+\lambda_{j}}{2} -\lambda_{m_{1}})^{-1}\nonumber\\ &&\prod_{m_{2}\ne j,k}(\lambda_{j}-\lambda_{m_{2}})^{-1} (\lambda_{k}-\lambda_{m_{2}})^{-1}\nonumber\\&&\prod_{a_{1}=0}^{2N}(\frac{a_{1}\lambda_{j}+(2N-a_{1})\lambda_{k}}{2}) \prod_{a_{2}=0}^{N}(a_{2}\lambda_{j}+(N-a_{2})\lambda_{k}))\nonumber\\ &&(\mbox{from (e)})\nonumber\\ &+&\frac{1}{6}\sum_{i\ne j}(\frac{1}{36}(w_{F_{2}}(\lambda_{i}^{N-4}- \lambda_{j}^{N-4})\frac{1}{w_{F_{1}}^{3}w_{F_{2}}^{3}}\nonumber\\ &&\prod_{m\ne i,j}(\lambda_{i}-\lambda_{m})^{-1} (\frac{2\lambda_{i}+\lambda_{j}}{3}-\lambda_{m})^{-1} (\frac{\lambda_{i}+2\lambda_{j}}{3}-\lambda_{m})^{-1} (\lambda_{j}-\lambda_{m})^{-1}\nonumber\\ &&\prod_{a=0}^{2N}(\frac{a\lambda_{i}+(3N-a)\lambda_{j}}{3}))\nonumber\\ &&(\mbox{from (f)})\nonumber\\ &+&\frac{1}{6}\sum_{{i\ne j}\atop{i\ne k,i\ne l}}( \lambda_{j}^{N-4}\frac{1}{w_{F_{1}}w_{F_{3}}w_{F_{5}}}w_{F_{4}}w_{F_{6}}+ \lambda_{l}^{N-4}\frac{1}{w_{F_{1}}w_{F_{3}}w_{F_{5}}}w_{F_{2}}w_{F_{6}} \nonumber\\ &+&\lambda_{k}^{N-4}\frac{1}{w_{F_{1}}w_{F_{3}}w_{F_{5}}}w_{F_{2}}w_{F_{4}} \nonumber\\ &+&w_{F_{2}}w_{F_{4}}w_{F_{6}}\lambda_{i}^{N-4} \frac{1}{w_{F_{1}}w_{F_{3}}w_{F_{5}}} (\frac{1}{w_{F_{1}}}+\frac{1}{w_{F_{3}}}+\frac{1}{w_{F_{5}}}))\nonumber\\ &&\frac{1}{(N\lambda_{i})^{2}} \frac{1}{w_{F_{1}}w_{F_{2}}w_{F_{3}}w_{F_{4}}w_{F_{5}}w_{F_{6}}}\nonumber\\ &&\prod_{m_{1}\ne i,j}(\lambda_{i}-\lambda_{m_{1}})^{-1} (\lambda_{j}-\lambda_{m_{1}})^{-1}\nonumber\\ &&\prod_{m_{2}\ne i,l}(\lambda_{i}-\lambda_{m_{2}})^{-1} (\lambda_{l}-\lambda_{m_{2}})^{-1}\nonumber\\ &&\prod_{m_{3}\ne i,k}(\lambda_{i}-\lambda_{m_{3}})^{-1} (\lambda_{k}-\lambda_{m_{3}})^{-1}\nonumber\\ &&\prod_{a_{1}=0}^{N}(a_{1}\lambda_{i}+(N-a_{1})\lambda_{j})\nonumber\\ &&\prod_{a_{2}=0}^{N}(a_{2}\lambda_{i}+(N-a_{2})\lambda_{k})\nonumber\\ &&\prod_{a_{3}=0}^{N}(a_{3}\lambda_{i}+(N-a_{3})\lambda_{l})\nonumber\\ &&\prod_{n\ne i}(\lambda_{i}-\lambda_{n})^{2}\nonumber\\ &&(\mbox{from (g)}) \label{78}\end{aligned}$$ These results are generically indpendent of the values $\lambda_{i}$, so we set $\lambda_{i}$ equals to $3^{i}$. Similarly we calculate the amplitudes $\langle {\cal O}_{e^{\alpha}}{\cal O}_{e^{\beta}} \rangle_{1}$,$\langle {\cal O}_{e^{\alpha}}{\cal O}_{e^{\beta}} \rangle_{2}$,$(\alpha + \beta =N-3)$. The results are collected in [**Table 1**]{},$\sim$ [**Table 3**]{}. Note that $\langle {\cal O}_{e^{N-4}}\rangle_{n} = \langle {\cal O}_{e}{\cal O}_{e^{N-4}}\rangle_{n}\cdot n$. This implies that the Käler equation of Gromov-Witten invariants holds for the amplitudes defined by (\[25\]). Assuming this relation for all amplitudes, the results of [**Table 1**]{}$\sim$ [**Table 3**]{} coincides with the ones calculated from mirror symmetry [@gmp]. [^2] We can calculate the amplipudes for matter theory for example $\langle\overbrace{{\cal O}_{e} \cdots {\cal O}_{e} }^{N-2\mbox{times}} \rangle$ by a rather cunning way. Fusion rules hold in the matter theory, so we can reduce the amplitudes into the products of three-point functions. Consider the “matter” expansion $$\langle {\cal O}_{e}{\cal O}_{e^{\alpha}}{\cal O}_{e^{\beta}}\rangle = N + \sum_{k=1}^{\infty} \langle {\cal O}_{e}{\cal O}_{e^{\alpha}} {\cal O}_{e^{\beta}}\rangle_{k}e^{-kt} \label{79}$$ where $t$ is the deformation parameter coupled to the Kähler form. By using fusion rules, and flat metric $\eta_{\alpha\beta}= N\cdot \delta_{\alpha+\beta,N-2}$, $$\begin{aligned} \langle\overbrace{{\cal O}_{e} \cdots {\cal O}_{e} }^{N-2\mbox{times}}\rangle&=& N^{5-N}\prod_{i=1}^{N-4}\langle{\cal O}_{e} {\cal O}_{e^{i}}{\cal O}_{e^{N-3-i}}\rangle\nonumber\\ &=& N + \sum_{k=1}^{\infty}\langle \overbrace{{\cal O}_{e} \cdots {\cal O}_{e} }^{N-2\mbox{times}}\rangle_k e^{-kt} \label{80}\end{aligned}$$ Then for example,$\langle{\overbrace{{\cal O}_{e} \cdots {\cal O}_{e} }^{N-2\mbox{times}}}\rangle_{1}$ can be calculated as $$\begin{aligned} \langle{\overbrace{{\cal O}_{e} \cdots {\cal O}_{e} }^{N-2\mbox{times}}}\rangle_{1}&=& -\frac{1}{2}\sum_{i\ne j}\prod_{k\ne i,j}(\lambda_{i}-\lambda_{k})^{-1} (\lambda_{j}-\lambda_{k})^{-1} \prod_{a=0}^{N}(a\lambda_{i}+(N-a)\lambda_{j})\nonumber\\ &&((N-4)(\lambda_{j}^{N-2}-\lambda_{i}^{N-2})-(N-2) (\lambda_{j}^{N-3}\lambda_{i}-\lambda_{i}^{N-3}\lambda_{j}))\nonumber\\ &&\frac{1}{(\lambda_{i}-\lambda_{j})^3}\\ \label{81} &=& N^{N+1}-(N-2)\cdot N\cdot N!(\frac{N-1}{1}+\cdots +\frac{1}{N-1}) \nonumber\\ &&-2N\cdot N! \label{82}\end{aligned}$$ If we set $\lambda_{i}=i$, we can derive (\[81\]) from (\[82\]) by a rather clumsy but elementary calculation. So theoretically we can see the coincidence of the calculation from A-model and B-model to the arbitrary degree $n$. Construction of Free Energy =========================== In section $4$, we see that we can calculate the amplitudes $\langle *\rangle_{n,grav.alt.}$ for tpological sigma model on $M_{N}$ coupled to gravity by torus action method. As we have seen in section $3$ and section $4$ , this method has a structure of summing over tree graphs, so we can construct a representation of Path-Integral form of the generating function of all amplitudes, i.e., free energy. First, let us write out explicitly the contribution from ${\cal M}(\Gamma)$ to the amplitude $\langle {\cal O}_{e^{j_{1}}}\cdots {\cal O}_{e^{j_{k}}} \rangle_{n,alt.grav.}$ $$\begin{aligned} \lefteqn{(\mbox{Contribution from ${\cal M}(\Gamma)$ to $\langle {\cal O}_{e^{j_{1}}}\cdots {\cal O}_{e^{j_{k}}} \rangle_{n,alt.grav.}$})}\nonumber\\ &&=\frac{1}{{}^{\sharp}Aut(\Gamma)}(\prod_{v\in Vert(\Gamma)} \lambda_{f_{v}(i)}^{j_{i}}\nonumber\\ &&\prod_{v\in Vert(\Gamma)}\prod_{{flags}\atop {F=(v,\alpha)}}w_{F}^{-1} (\sum_{{flags}\atop{F=(v,\alpha)}}w_{F}^{-1})^{val(v) +{}^{\sharp}S_{v}-3}\nonumber\\ &&\prod_{v\in Vert(\Gamma)}(\prod_{j\ne f_{v}} (\lambda_{f_{v}}-\lambda_{j}))^{val(v)-1}\nonumber\\ &&\prod_{\alpha\in Edge(\Gamma)}\frac{1}{d_{\alpha}}\nonumber\\ &&\prod_{{flags}\atop {F=(v,\alpha)}}(w_{F})^{-d_{\alpha}} \prod_{{\alpha\in Edge(\Gamma)}\atop{(u,v): vertices\;of\:\alpha}} \prod_{k\neq f_{u},f_{v}}\prod_{m=0}^{d_{\alpha}} (\frac{m\lambda_{f_{u}}+ (d_{\alpha}-m)\lambda_{f_{v}}}{d_{\alpha}}-\lambda_{k})^{-1} \nonumber\\ &&\prod_{\alpha\in Edge(\Gamma)}(d_{\alpha}!)^{-2} \prod_{v\in Vert(\Gamma)}(\prod_{j:j\neq f_{v}} (\lambda_{f_{v}}-\lambda_{j}))^{val(v)-1}\nonumber\\ &&\prod_{{\alpha\in Edge(\Gamma)}\atop{(v_{1},v_{2}): vertices\; of\;\alpha}} \prod_{a=0}^{Nd_{\alpha}}(\frac{a\lambda_{f_{v_{1}}}+(N-a) \lambda_{f_{v_{2}}}}{d_{\alpha}})\prod_{v\in Vert(\Gamma)} (N\lambda_{f_{v}})^{1-val(v)}) \label{83}\end{aligned}$$ Then we classify the factors into two groups. One is the factors from edges, and the other from vertices. The factors from the edges are $$\begin{aligned} \lefteqn{\mbox{(i)} \prod_{{flags}\atop{F=(v,\alpha)}}w_{F}^{-1}} \nonumber\\ && \mbox{(ii)}\prod_{v\in Vert(\Gamma)}(\prod_{j\ne f_{v}} (\lambda_{f_{v}}-\lambda_{j}))^{val(v)}\nonumber\\ && \mbox{(iii)}\prod_{{flags}\atop {F=(v,\alpha)}}(w_{F})^{-d_{\alpha}} \prod_{{\alpha\in Edge(\Gamma)}\atop{(u,v): vertices\;of\:\alpha}} \prod_{k\neq f_{u},f_{v}}\prod_{m=0}^{d_{\alpha}} (\frac{m\lambda_{f_{u}}+ (d_{\alpha}-m)\lambda_{f_{v}}}{d_{\alpha}}-\lambda_{k})^{-1} \prod_{\alpha\in Edge(\Gamma)}(d_{\alpha}!)^{-2}\nonumber\\ &&\mbox{(iv)}\prod_{{\alpha\in Edge(\Gamma)}\atop{(v_{1},v_{2}): vertices\; of\; \alpha}} \prod_{a=0}^{Nd_{\alpha}}(\frac{a\lambda_{f_{v_{1}}}+(N-a) \lambda_{f_{v_{2}}}}{d_{\alpha}})\prod_{v\in Vert(\Gamma)} (N\lambda_{f_{v}})^{-val(v)}\nonumber\\ &&\mbox{(v)}\prod_{Edge(\Gamma)}\frac{1}{d_{\alpha}} \label{84}\end{aligned}$$ And the factors we can push into the contribution from vertices are, $$\begin{aligned} \lefteqn{\mbox{(i)}\prod_{v\in Vert(\Gamma)} \lambda_{f_{v}(i)}^{j_{i}}}\nonumber\\ &&\mbox{(ii)}\prod_{v\in Vert(\Gamma)} (\sum_{{flags}\atop{F=(v,\alpha)}}w_{F}^{-1})^{val(v) +{}^{\sharp}S_{v}-3}\nonumber\\ &&\mbox{(iii)}\prod_{v\in Vert(\Gamma)}(\prod_{j\ne f_{v}} (\lambda_{f_{v}}-\lambda_{j}))^{-1}\nonumber\\ &&\mbox{(iv)}\prod_{v\in Vert(\Gamma)}(N\lambda_{f_{v}}) \label{85}\end{aligned}$$ Then we introduce the field variables $\phi_{ij,d}$,propagator $g_{ij,d;i^{\prime}j^{\prime},d^{\prime}}$, vertex\ $C_{i_{1}j_{1},d_{1};\cdots i_{k}j_{k},d_{k}}\phi_{i_{1}j_{1},d_{1}}\cdots\phi_{i_{k}j_{k},d_{k}}$ and external field source parameters $t_{1},\cdots,t_{N-4}$. In this formulation, field variables correspond to the edges with characters i and j and degree d, $g_{ij,d;i^{\prime}j^{\prime},d^{\prime}}$ remains nonzero only if $i=j^{\prime},j=i^{\prime},d=d^{\prime}$, and the nonzero value of propagator is given as the reciprocal of the product of (\[84\]) (i)$\sim$(v). Then we have $$g_{ij,d}:=g_{ij,d;ji,d}= \frac{-d^{3}(\lambda_{i}-\lambda_{j})^{2} \prod_{k=1}^{N}\prod_{a=1}^{d-1}(a\lambda_{i}+(d-a)\lambda_{j}- d\lambda_{k})}{\prod_{a=1}^{Nd-1}(a\lambda_{i}+(Nd-a)\lambda_{j})} \label{86}$$ Vertex $C_{i_{1}j_{1},d_{1};\cdots i_{k}j_{k},d_{k}}\phi_{i_{1}j_{1},d_{1}}\cdots\phi_{i_{k}j_{k},d_{k}}$ are constructed with pairing the factor $\lambda_{i}^{k}$ to $t_{k}$ as follows. $$\begin{aligned} \lefteqn{\sum_{k=1}^{\infty}\sum_{l=1}^{\infty} \frac{1}{l!} \sum_{i=1}^{N} \frac{N\lambda_{i}}{\prod_{j\ne i}(\lambda_{i}-\lambda_{j})} \sum_{{d_{1},\cdots,d_{k},d_{*}\geq 1}\atop{j_{1},\cdots,j_{k},j_{*}\ne i}} \sum_{\tilde{j_{1}},\cdots,\tilde{j_{l}} \in \{ 1,2,\cdots,N-4\}}}\nonumber\\ &&(v_{ij_{1},d_{1}}+\cdots+v_{ij_{k},d_{k}})^{k+l-3} \phi_{ij_{1},d_{1}}\cdots\phi_{ij_{k},d_{k}} t_{\tilde{j_{1}}}\cdots t_{\tilde{j_{k}}} (\lambda_{i}^{\tilde{j_{1}}+\cdots +\tilde{j_{k}}})\nonumber\\ &&=\sum_{i=1}^{N} \frac{N\lambda_{i}}{\prod_{j\ne i}(\lambda_{i}-\lambda_{j})} \sum_{k=1}^{\infty}\frac{1}{k!} \sum_{{d_{1},\cdots,d_{k},d_{*}\geq 1}\atop{j_{1},\cdots,j_{k},j_{*}\ne i}} (v_{ij_{1},d_{1}}+\cdots+v_{ij_{k},d_{k}})^{k-3} \phi_{ij_{1},d_{1}}\cdots\phi_{ij_{k},d_{k}}\nonumber\\ &&\exp((t_{1}\lambda_{i}+\cdots+t_{N-4}\lambda_{i}^{N-4}) (v_{ij_{1},d_{1}}+\cdots+v_{ij_{k},d_{k}}))\nonumber\\ &&v_{ij,d}:=\frac{d}{\lambda_{i}-\lambda_{j}} \label{87}\end{aligned}$$ where $1/k!$ is the factor that produces $1/{}^{\sharp}Aut(\Gamma)$ and $1/l!$ is the combinatorial factor in the insertion of the external operator. With these preparation, we have the path-integral representation of the free energy. $$\begin{aligned} \lefteqn{F_{M_{N}}(t_{1},\cdots,t_{N-4})} \nonumber\\ &&:= \sum_{n_{1},\cdots,n_{N-4}\geq 0}\langle {\cal O}^{n_{1}}_{e^{1}} \cdots {\cal O}^{n_{N-4}}_{e^{N-4}}\rangle\frac{t_{1}^{n_{1}}\cdots t_{N-4}^{n_{N-4}}}{n_{1}!\cdots n_{N-4}!}\nonumber\\ &&=Res_{z}Res_{{h}}(\frac{1}{z}\log(\mbox{det}(g^{-1})\frac{1}{{h}} \int d\phi_{ij,d}\nonumber\\ &&(-\frac{1}{2}\sum_{i,j,d}\frac{-d^{3}(z\lambda_{i}-z\lambda_{j})^{2} \prod_{k=1}^{N}\prod_{a=1}^{d-1}(az\lambda_{i}+(d-a)z\lambda_{j}- dz\lambda_{k})}{\prod_{a=1}^{Nd-1}(az\lambda_{i}+(Nd-a)z\lambda_{j})} \phi_{ij,d}\phi_{ji,d}\nonumber\\ &&+\sum_{i=1}^{N} \frac{Nz\lambda_{i}}{\prod_{j\ne i}(z\lambda_{i}-z\lambda_{j})} \sum_{k=1}^{\infty}\frac{1}{k!} \sum_{{d_{1},\cdots,d_{k},d_{*}\geq 1}\atop{j_{1},\cdots,j_{k}, j_{*}\ne i}} (\frac{v_{ij_{1},d_{1}}}{z}+\cdots+\frac{v_{ij_{k},d_{k}}}{z})^{k-3} \phi_{ij_{1},d_{1}}\cdots\phi_{ij_{k},d_{k}}\nonumber\\ &&\exp((t_{1}\lambda_{i}z+\cdots+t_{N-4}\lambda_{i}^{N-4}z^{N-4}) (\frac{v_{ij_{1},d_{1}}}{z}+\cdots+\frac{v_{ij_{k},d_{k}}}{z}))))) \label{88}\end{aligned}$$ where we introduce $h$ and dummy variable $z$ to pick up the portion that comes from tree graphs and satisfies the topological selection rule (\[18\]). Conclusion ========== I believe that results of this paper are reasonably clear, so I just point out what remains to show, or consider in the future. First, the relation between the representations of $\langle *\rangle_{n,grav.}$ and $\langle* \rangle_{n,grav.alt.}$. We think that the situation corresponds to the case described by Witten [@witten3], i.e., the zero locus of the section of ${\cal E}_{Nn+1}$ and external forms are not always points but submanifolds of ${\cal M}_{0,n,k}^{CP^{N-1}}$. In such cases, Euler classes on these submanifolds arise and $\langle * \rangle_{n,grav.alt.}$ reduces to $\langle *\rangle_{n,grav.}$. Second, the difference between the moduli space of matter theory and the one coupled to gravity. For the matter theory we asserted in [@nj] that moduli spaces can be constructed by subtracting boundary points from simple projective space. This statement is indirectly supported by the work of Morrison,Plesser [@mp] by use of gauged linear sigma model. But stable map approach seems to go in the opposite way, i.e., it adds boundary points to compactify the moduli space. So the problem of construction of moduli spaces of matter theory still remains.\ For generalization to the worldsheet of higher genera, we can think of two naive additional approaches. One is the addition of loop amplitudes. The other is the introduction of gravitaional correlation functions for higher genus in the calculation of (\[55\]). We tried the calculation of genus 1 amplitudes with these factors but cannot get good results. So, further consideration is needed.\ I’d like to thank Dr.K.Hori for many useful discussions. I also thank Dr.Y.Sun,Dr.M.Nagura and Prof.T.Eguchi for kind encouragement. (130,180) (11,130)[(1,0)[8]{}]{} (12,137)[(1,0)[6]{}]{} (15,140) (15,116) (15,110) (18,123)[(-1,0)[6]{}]{} (7,130) (23,130) (15,133) (11,90)[(1,0)[8]{}]{} (15,100) (15,76) (15,70) (12,97)[(1,0)[6]{}]{} (18,83)[(-1,0)[6]{}]{} (15,93) (7,90) (23,90) (41,90)[(1,0)[8]{}]{} (45,100) (45,76) (42,97)[(1,0)[6]{}]{} (48,83)[(-1,0)[6]{}]{} (45,93) (51,90)[(1,0)[8]{}]{} (50,70) (55,100) (55,76) (52,97)[(1,0)[6]{}]{} (58,83)[(-1,0)[6]{}]{} (55,93) (37,90) (50,93) (63,90) (11,50)[(1,0)[8]{}]{} (15,60) (15,36) (15,30) (12,57)[(1,0)[6]{}]{} (18,43)[(-1,0)[6]{}]{} (7,50) (23,50) (15,53) (41,50)[(1,0)[8]{}]{} (45,60) (45,36) (42,57)[(1,0)[6]{}]{} (48,43)[(-1,0)[6]{}]{} (51,50)[(1,0)[8]{}]{} (60,30) (55,60) (55,36) (52,57)[(1,0)[6]{}]{} (58,43)[(-1,0)[6]{}]{} (37,50) (50,53) (63,50) (45,53) (55,53) (81,50)[(1,0)[8]{}]{} (85,60) (85,36) (82,57)[(1,0)[6]{}]{} (88,43)[(-1,0)[6]{}]{} (91,50)[(1,0)[8]{}]{} (95,60) (95,36) (92,57)[(1,0)[6]{}]{} (98,43)[(-1,0)[6]{}]{} (101,50)[(1,0)[8]{}]{} (105,60) (105,36) (102,57)[(1,0)[6]{}]{} (108,43)[(-1,0)[6]{}]{} (95,30) (77,50) (90,53) (100,53) (85,53) (95,53) (105,53) (113,50) (50,21)[(0,1)[8]{}]{} (43,22)[(0,1)[6]{}]{} (57,28)[(0,-1)[6]{}]{} (37,25) (60,25) (53,30) (53,25) (50,19)[(3,-2)[8]{}]{} (63,20) (52,3) (46,3) (33,20) (37,18)[(3,2)[6]{}]{} (48,12)[(-3,-2)[6]{}]{} (58,7)[(-3,2)[6]{}]{} (57,21)[(3,-2)[6]{}]{} (53,20) (55,17) (37,11) (50,19)[(-3,-2)[8]{}]{} (61,11) (45,17) (60,1) (50,20) (50,30) (59,13) (41,13) (10,130) (20,130) (10,90) (20,90) (10,50) (20,50) (40,90) (50,90) (60,90) (40,50) (50,50) (60,50) (80,50) (90,50) (100,50) (110,50) $\langle {\cal O}_{e^{N-4}}\rangle_{1}$ $\langle {\cal O}_{e^{N-4}}\rangle_{2}$ $\langle {\cal O}_{e^{N-4}}\rangle_{3}$ ------ ----------------------------------------- ----------------------------------------- ----------------------------------------- N=5 2875 $\frac{4876875}{4}$ $\frac{8564575000}{9}$ N=6 60480 440899200 6255156284160 N=7 1009792 122240038536 $\frac{274758045710330728}{9}$ N=8 15984640 33397163702784 $\frac{1386812286427888746496}{9}$ N=9 253490796 9757818404032059 897560654227562367535680 N=10 4120776000 3151991359959750000 6298886011657402651840000000 : $\langle {\cal O}_{e^{N-4}}\rangle_{alt,grav}$ --------------------------------------------------------------- N=5 $\langle {\cal O}_{e}{\cal O}_{e}\rangle_{1} = 2875$ ------ -------------------------------------------------------- N=6 $\langle {\cal O}_{e}{\cal O}_{e^{2}}\rangle_{1}= 60480$ N=7 $\langle {\cal O}_{e}{\cal O}_{e^{3}} \rangle_{1}= 1009792$ $\langle {\cal O}_{e^{2}}{\cal O}_{e^{2}} \rangle_{1}= 1707797$ N=8 $\langle {\cal O}_{e}{\cal O}_{e^{4}}\rangle_{1}= 15984640$ $\langle {\cal O}_{e^{2}}{\cal O}_{e^{3}}\rangle_{1}= 37502976$ N=9 $\langle {\cal O}_{e}{\cal O}_{e^{5}}\rangle_{1}= 253490796$ $\langle {\cal O}_{e^{2}}{\cal O}_{e^{4}}\rangle_{1}= 763954092$ $\langle {\cal O}_{e^{3}}{\cal O}_{e~{3}}\rangle_{1}= 1069047153$ N=10 $\langle {\cal O}_{e}{\cal O}_{e^{6}}\rangle_{1}= 4120776000$ $\langle {\cal O}_{e^{2}}{\cal O}_{e^{5}}\rangle_{1}= 15274952000$ $\langle {\cal O}_{e^{3}}{\cal O}_{e^{4}}\rangle_{1}= 27768048000$ --------------------------------------------------------------- : $\langle {\cal O}_{e^{\alpha}}{\cal O}_{e^{\beta}} \rangle_{1,alt,grav}$ --------------------------------------------------------------- N=5 $\langle {\cal O}_{e}{\cal O}_{e}\rangle_{2} = \frac{4876875}{2}$ ------ -------------------------------------------------------- N=6 $\langle {\cal O}_{e}{\cal O}_{e^{2}}\rangle_{2}= 881798400$ N=7 $\langle {\cal O}_{e}{\cal O}_{e^{3}} \rangle_{2}= 244480077072$ $\langle {\cal O}_{e^{2}}{\cal O}_{e^{2}} \rangle_{2}= \frac{1021577199083}{2}$ N=8 $\langle {\cal O}_{e}{\cal O}_{e^{4}}\rangle_{2}= 66794327405568$ $\langle {\cal O}_{e^{2}}{\cal O}_{e^{3}}\rangle_{2}= 224340722909184$ N=9 $\langle {\cal O}_{e}{\cal O}_{e^{5}}\rangle_{2}= 19515636808064118$ $\langle {\cal O}_{e^{2}}{\cal O}_{e^{4}}\rangle_{2}= 93777295510651590$ $\langle {\cal O}_{e^{3}}{\cal O}_{e~{3}}\rangle_{2}= \frac{312074853388012521}{2}$ N=10 $\langle {\cal O}_{e}{\cal O}_{e^{6}}\rangle_{2}= 6303982719919500000$ $\langle {\cal O}_{e^{2}}{\cal O}_{e^{5}}\rangle_{2}= 40342298393756700000$ $\langle {\cal O}_{e^{3}}{\cal O}_{e^{4}}\rangle_{2}= 100290980414189400000$ --------------------------------------------------------------- : $\langle {\cal O}_{e^{\alpha}}{\cal O}_{e^{\beta}} \rangle_{2,alt,grav}$ [99]{} M.Kontsevich. hepth 9405035 R.Bott Jour. Diff. Geom. 1(1967)311-330 M.Jinzenji and M.Nagura UT-680 B.R.Greene, D.R.Morrison and M.R.Plesser CLNS-93/1253,IASSNS-HEP-HEP-94/2,YCTP-P31-92 E.Witten in Essays on Mirror Manifolds, ed. S.-T.Yau (Int.Press. Co.,Hong Kong, 1992) 120-180 T.Eguchi and S.K.Yang Mod. Phys. Lett. A, Vol.5, No.21 (1990) 1693-1701 P.Candelas, X.de la Ossa, P.Green and L.Parkes Phys.Lett.[**258B**]{} (1991) 118; Nucl.Phys. [**B359**]{} (1991) 21 P.Griffiths and J.Harrith Wiley 1978 S.Katz OSU-M-92-3, 1992 R.Dijkgraaf IASSNS-HEP-91/91 E.Witten Nucl.Phys.[**B340**]{} (1990) 281 M.Kontsevich Commun.Math.Phys.[**147**]{} (1992),1-23 D.R.Morrison and M.R.Plesser DUKE-TH-94-78 IASSNS-HEP-94/82 hep-th/9412236 K.Hori UT-694 P.S.Aspinwall and D.R.Morrison Commun.Math.Phys. [**151**]{} (1993),245-262 C.Vafa and E.Witten hep-th/940874,HUTP-94/A017,IASSNS-HEP-94-54 [^1]: When coupled to gravity,${\cal O}_{1}$ correspond to puncture operator $P$ but in the small phase, space $P$ insertion is suppressed except for constant map sector because of puncture equation and as we know from the later discussion of topological selection rule, ghost number of inserted operator must be less than $N-3$. So it suffices to consider only $N-4$ elements ${\cal O}_{e},{\cal O}_{e^2},\cdots,{\cal O}_{e^{N-4}}$. [^2]: Note that for three-point function,amplitudes of the matter theory and the ones of theory coupled with gravity coincide.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Using the Keldysh Green function technique, we calculate the finite-frequency correlator between the electrical current and the heat current flowing through a quantum dot connected to reservoirs. At equilibrium, we find that this quantity, called mixed noise, is linked to the thermoelectric ac-conductance by the fluctuation-dissipation theorem. Out-of-equilibrium, we discuss its spectrum and find evidence of the close relationship between the mixed noise and the thermopower. We study the spectral coherence and identify the conditions to have a strong correlation between the electrical and heat currents. The change in the spectral coherence due to the presence of a temperature gradient between the reservoirs is also highlighted.' author: - 'Paul Eyméoud$^{1,2}$' - 'Adeline Crépieux$^{1}$' title: 'Mixed electrical-heat noise spectrum in a quantum dot' --- Introduction ============ The electrical noise spectrum in quantum systems is accessible experimentally through very sensitive techniques such as spectrum analyzer [@Picciotto1997; @Saminadayar1997], use of a superconductor-insulator-superconductor tunnel junction as an on-chip spectrum analyzers [@Deblock2003] and measurement of photon emission spectrum [@Schull2009]. It exists also a proposal to measure the heat statistics in quantum devices [@Sanchez2012]. Given the fast progress of detection techniques, it is not forbidden to imagine that in the next few years, the measurement of the noise correlator between the electrical current and the heat current (mixed noise) would be possible. In parallel, calculations of finite-frequency electrical-heat mixed noise are needed for quantum systems. There exist very few works on the zero-frequency mixed noise [@Giazotto2006; @Sanchez2013; @Battista2014_proc; @Crepieux2015; @Crepieux2016] and not even one concerning the finite-frequency mixed noise. Theoretically, the studies are limited to the electrical noise spectrum (see [@Blanter2000; @Martin2005; @Zamoum2016] and references therein), to the energy noise spectrum [@Averin2010; @Sergi2011; @Agarwalla2015], to the statistics of the energy current in the presence of time-dependent excitation [@Battista2014; @Yu2016], and to the heat noise spectrum [@Zhan2011]. This is regrettable since it has been shown recently that the zero-frequency mixed noise contains information on the thermoelectric response of the system [@Crepieux2015; @Crepieux2016]: it gives the figure of merit in the linear response regime and it is related to the thermoelectric efficiency in the weak transmission regime (Schottky regime). At finite-frequency, the mixed noise should bring information on the dynamics of the thermoelectric conversion, in particular on the thermoelectric response of time-modulated systems, which is the study of an increasing number of works [@Crepieux2011; @Goker2012; @Goker2013; @Boehnke2013; @Bagheri2014; @Chirla2014; @Zhou2015; @Dare2016; @Ludovico2016; @Okada2016]. In this paper, we fill this lacuna by calculating the mixed noise spectrum of a quantum dot (QD) using the Keldysh out-of-equilibrium Green function technique. We focus on the non-symmetrized noise spectrum since this is the quantity which is relevant for quantum systems, due to the fact that the current operators do not commute with each other [@Lesovik1997; @Gavish2000; @Deblock2003]. The paper is organized as follows: We present the model and give the definition of electrical and heat currents in Sec. II. The results for the noise spectra are presented in Sec. III, and discussed in Sec. IV. We conclude in Sec. V. Model ===== To model the QD connected to left ($L$) and right ($R$) reservoirs, we use the Hamiltonian $H=H_L+H_R+H_{D}+H_T$, where $H_{\alpha=L,R}=\sum_{k\in \alpha} \varepsilon_{k \alpha} c_{k \alpha}^{\dag}c_{k \alpha}$ describes the energy of electrons in the reservoir $\alpha$, with $c_{k \alpha}^{\dag}$ ($c_{k \alpha}$), the creation (annihilation) operator, $H_{D}= \varepsilon_d d^{\dag}d$ describes the QD with a single energy level $\varepsilon_d$, with $d^{\dag}$ ($d$) the creation (annihilation) operator, and $H_T=\sum_{\alpha=L,R}\sum_{k\in \alpha}(V_{k \alpha} c_{k \alpha}^{\dag} d+h.c.)$ describes the transfer of electrons from the reservoirs to the QD and vice versa. The left and right reservoirs are assumed to be at equilibrium with temperature $T_{L,R}$ and chemical potential $\mu_{L,R}$ (see Fig. \[figure1\]). ![Picture of the QD connected to left and right reservoirs with distinct temperatures and chemical potentials (we take $eV=\mu_L-\mu_R$ and $\Delta T=T_L-T_R$). The black arrows indicate the convention chosen for the currents’ direction.[]{data-label="figure1"}](figure1.eps){width="7cm"} The charge current, $\hat{I}^0_{\alpha}$, and heat current, $\hat{I}^1_{\alpha}$, flowing from the reservoir $\alpha$ to the QD, are given by the time derivatives [@Jauho1994; @Haug2007; @Crepieux2012] of the operators *number* of electrons in the reservoir $\alpha$, $N_{\alpha}$, and *energy* of electrons in the reservoir $\alpha$, $H_{\alpha}$: $\hat{I}^0_{\alpha}=-e\dot{N}_{\alpha}$, and $\hat{I}^1_{\alpha}=-\dot{H}_{\alpha}+\mu_{\alpha}\dot{N}_{\alpha}$, with $N_{\alpha}=\sum_{k\in\alpha}c^{\dag}_{k \alpha}c_{k \alpha}$. The time derivatives of these two quantities are equal to $\dot{N}_{\alpha}=i\hbar^{-1}\big[H,N_{\alpha}\big]$, and $\dot{H}_{\alpha}=i\hbar^{-1}\big[H,H_{\alpha}\big]$, which lead after calculation to $$\begin{aligned} \label{N} \dot{N}_{\alpha}&=&\frac{i}{\hbar}\sum_{k\in\alpha}\big(-V_{k \alpha}c^{\dag}_{k \alpha}d+V^{*}_{k \alpha}d^{\dag}c_{k \alpha}\big)~,\\ \label{H} \dot{H}_{\alpha}&=&\frac{i}{\hbar}\sum_{k\in\alpha}\varepsilon_{k \alpha}\big(-V_{k \alpha}c^{\dag}_{k \alpha}d+V^{*}_{k \alpha}d^{\dag}c_{k \alpha}\big)~.\end{aligned}$$ Injecting Eqs. (\[N\]) and (\[H\]) in the definitions of $\hat{I}^0_{\alpha}$ and $\hat{I}^1_{\alpha}$, we obtain: $$\begin{aligned} \hat{I}^0_{\alpha}&=&\frac{ie}{\hbar}\sum_{k\in\alpha}\big(V_{k \alpha}c^{\dag}_{k \alpha}d-V^{*}_{k \alpha}d^{\dag}c_{k \alpha}\big)~,\\ \hat{I}^1_{\alpha}&=&\frac{i}{\hbar}\sum_{k\in\alpha}(\varepsilon_{k \alpha}-\mu_{\alpha})\big(V_{k \alpha}c^{\dag}_{k \alpha}d-V^{*}_{k \alpha}d^{\dag}c_{k \alpha}\big)~,\end{aligned}$$ which give in a compact form $$\begin{aligned} \label{c1} \hat I^p_\alpha&=&\frac{ie^{1-p}}{\hbar}\sum_{k\in\alpha} \big(\varepsilon_{k \alpha}-\mu_\alpha\big)^p\big( V_{k \alpha} c_{k \alpha}^{\dag} d-V_{k \alpha}^{\ast} d^{\dag} c_{k \alpha} \big)~.\nonumber\\\end{aligned}$$ Noise spectrum ============== The non-symmetrized noise spectrum is defined as $$\begin{aligned} \label{def_S} \mathcal{S}^{pq}_{\alpha\beta}(\omega)=\int_{-\infty}^{\infty} \mathcal{S}^{pq}_{\alpha\beta}(t,0)e^{-i\omega t}dt~,\end{aligned}$$ where $\mathcal{S}^{pq}_{\alpha\beta}(t,0)=\langle \Delta \hat{I}^p_\alpha(t) \Delta \hat{I}^q_\beta(0) \rangle$ is the current-current time-correlator, and $\Delta \hat{I}^p_\alpha(t)=\hat{I}^p_\alpha(t)-\langle \hat I^p_\alpha\rangle$, with $\hat{I}^p_\alpha$, the electrical ($p=0$) or heat ($p=1$) current operator from the reservoir $\alpha$ to the central region through the barrier $\alpha$. The finite-frequency non-symmetrized noise $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ quantifies the correlation between the currents $\hat{I}^{p}_{\alpha}$ and $\hat{I}^{q}_{\beta}$ at finite frequency $\omega$. Such a general definition embeds three types of noise: (i) the charge noise, $\mathcal{S}^{00}_{\alpha\beta}(\omega)$, corresponding to the correlator between the electrical current and itself; (ii) the mixed noises, $\mathcal{S}^{01}_{\alpha\beta}(\omega)$ and $\mathcal{S}^{10}_{\alpha\beta}(\omega)$, corresponding to the correlators between the electrical current and the heat current, and (iii) the heat noise, $\mathcal{S}^{11}_{\alpha\beta}(\omega)$, corresponding to the correlator between the heat current and itself. We first compute the time-correlator $\mathcal{S}^{pq}_{\alpha\beta}(t,t')$ using the Keldysh out-of-equilibrium formalism [@Keldysh1964], and next calculate its Fourier transform in order to get $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$. To achieve this task, we insert the current operator, given by Eq. (\[c1\]), in the definition of the noise, given by Eq. (\[def\_S\]), and perform the calculation of the average of the product of four creation/annihilation operators, making the following assumptions: non-interacting electrons, wide-band approximation, and symmetrical coupling strength between the reservoirs and the QD (i.e., symmetrical left and right barriers). The details of the calculation are given in the Appendix \[appendixA\]. The final expression of finite-frequency non-symmetrized noise we obtain is $$\begin{aligned} \label{exp_noise} &&\mathcal{S}^{pq}_{\alpha\beta}(\omega)=\frac{e^{2-p-q}}{h}\int_{-\infty}^{\infty} d\varepsilon \nonumber\\ && \times\Big[(\varepsilon-\mu_\alpha)^{p}(\varepsilon-\mu_\beta)^{q}\mathcal{A}_{\alpha\beta}(\varepsilon,\omega)\nonumber\\ &&+(\varepsilon-\mu_\alpha)^{p}(\varepsilon-\hbar\omega-\mu_\beta)^{q}\mathcal{B}_{\alpha\beta}(\varepsilon,\omega)\nonumber\\ &&+(\varepsilon-\hbar\omega-\mu_\alpha)^{p}(\varepsilon-\mu_\beta)^{q}\mathcal{B}^*_{\beta\alpha}(\varepsilon,\omega)\nonumber\\ && +(\varepsilon-\hbar\omega-\mu_\alpha)^{p}(\varepsilon-\hbar\omega-\mu_\beta)^{q}\mathcal{C}_{\alpha\beta}(\varepsilon,\omega)\Big]~,\end{aligned}$$ with $$\begin{aligned} \label{A} &&\mathcal{A}_{\alpha\beta}(\varepsilon,\omega) =\mathcal{T}(\varepsilon-\hbar\omega)f_M^h(\varepsilon-\hbar\omega)\Big[\mathcal{T}(\varepsilon)f_M^e(\varepsilon)\nonumber\\ &&+[\delta_{\alpha\beta}-t(\varepsilon)]f_\alpha^e(\varepsilon)+[\delta_{\alpha\beta}-t^*(\varepsilon)]f_\beta^e(\varepsilon)\Big]~,\end{aligned}$$ $$\begin{aligned} \label{B} &&\mathcal{B}_{\alpha\beta}(\varepsilon,\omega) =t(\varepsilon)t(\varepsilon-\hbar\omega)\left[f_\alpha^e(\varepsilon)-t^*(\varepsilon)f_M^e(\varepsilon)\right]\nonumber\\ &&\times \left[f_\beta^h(\varepsilon-\hbar\omega)-t^*(\varepsilon-\hbar\omega)f_M^h(\varepsilon-\hbar\omega)\right]~,\end{aligned}$$ and $$\begin{aligned} \label{C} \mathcal{C}_{\alpha\beta}(\varepsilon,\omega) &=&\mathcal{T}(\varepsilon)f_M^e(\varepsilon)\Big[\mathcal{T}(\varepsilon-\hbar\omega)f_M^h(\varepsilon-\hbar\omega)\nonumber\\ &&+[\delta_{\alpha\beta}-t^*(\varepsilon-\hbar\omega)]f_\alpha^h(\varepsilon-\hbar\omega)\nonumber\\ &&+[\delta_{\alpha\beta}-t(\varepsilon-\hbar\omega)]f_\beta^h(\varepsilon-\hbar\omega)\Big]~,\end{aligned}$$ where $f_\alpha^e(\varepsilon)=[1+\exp((\varepsilon-\mu_\alpha)/k_BT_\alpha)]^{-1}$ is the Fermi-Dirac distribution function for electrons, $f_\alpha^h(\varepsilon)=1- f_\alpha^e(\varepsilon)$ is the distribution function for holes, $f_M^{e,h}(\varepsilon)=[f_L^{e,h}(\varepsilon)+f_R^{e,h}(\varepsilon)]/2$ is the average left and right distribution, $t(\varepsilon)$ is the transmission amplitude, and $\mathcal{T}(\varepsilon)=|t(\varepsilon)|^2$ is the transmission coefficient. The transmission amplitude is related to the retarded Green function of the QD, $G^r(\varepsilon)$, through the relation $t(\varepsilon)=i\Gamma G^r(\varepsilon)$, where $\Gamma$ is the coupling strength between the QD and the reservoirs [@Zamoum2016]. Equation (\[exp\_noise\]) gives the electrical noise when $p=q=0$, it gives the mixed noise when either $p=0$ and $q=1$, or vice versa, and it gives the heat noise when $p=q=1$. $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ is a real quantity when $p=q$ and $\alpha=\beta$ (auto-correlator), but can be complex otherwise (cross-correlator). We have checked that the electrical noise $\mathcal{S}^{00}_{\alpha\beta}(\omega)$ extracted from Eq. (\[exp\_noise\]) coincides with the results of the literature [@Hammer2011; @Zamoum2016], and that the heat noise $\mathcal{S}^{11}_{\alpha\beta}(\omega)$ extracted from Eq. (\[exp\_noise\]) coincides with the existing results of the literature in the limit of energy-independent transmission amplitude [@Sergi2011]. The expressions for the mixed noises $\mathcal{S}^{01}_{\alpha\beta}(\omega)$ and $\mathcal{S}^{10}_{\alpha\beta}(\omega)$ are novels. This is the central result of this paper. It is valid at any frequency $\omega$, coupling strength $\Gamma$ between the QD and the reservoirs, QD energy level $\varepsilon_d$, and for any temperature and voltage gradients between the left and right reservoirs. In the following, we choose first to restrict our study to the case where temperatures for the left and right reservoirs are equal, $T_L=T_R=T$, and for $\varepsilon_d=0$ (electron-hole symmetry point), and we discuss the mixed noise spectrum in three situations: (i) at equilibrium, (ii) for energy-independent transmission amplitude, and (iii) for an Anderson-type transmission amplitude. In the latter case, we also discuss the spectral coherence in the presence of a temperature gradient between the two reservoirs. Discussion ========== At equilibrium -------------- At equilibrium, i.e., $\mu_L=\mu_R=\varepsilon_F$, where $\varepsilon_F$ is the Fermi energy for electrons in the reservoirs, and for equal left and right reservoir temperatures, i.e., $T_L=T_R=T$, we have from Eqs. (\[exp\_noise\])-(\[C\]) $$\begin{aligned} \label{S_eq} &&\mathcal{S}^{pq}_{\alpha\beta}(\omega)=\frac{e^{2-p-q}}{h}\int_{-\infty}^{\infty} d\varepsilon f_M^e(\varepsilon)f_M^h(\varepsilon-\hbar\omega)\nonumber\\ && \times\Big[\varepsilon^{p+q}\mathcal{\widetilde A}_{\alpha\beta}(\varepsilon,\omega)+\varepsilon^{p}(\varepsilon-\hbar\omega)^{q}\mathcal{\widetilde B}_{\alpha\beta}(\varepsilon,\omega)\nonumber\\ &&+(\varepsilon-\hbar\omega)^{p}\varepsilon^{q}\mathcal{\widetilde B}^*_{\beta\alpha}(\varepsilon,\omega) +(\varepsilon-\hbar\omega)^{p+q}\mathcal{\widetilde C}_{\alpha\beta}(\varepsilon,\omega)\Big]~,\nonumber\\\end{aligned}$$ with $$\begin{aligned} \mathcal{\widetilde A}_{\alpha\beta}(\varepsilon,\omega) &=&\mathcal{T}(\varepsilon-\hbar\omega) \left[2\delta_{\alpha\beta}-\mathcal{T}(\varepsilon)\right]~,\\ \mathcal{\widetilde B}_{\alpha\beta}(\varepsilon,\omega) &=&t(\varepsilon)t(\varepsilon-\hbar\omega)\nonumber\\ &&\times\left[1-t^*(\varepsilon)\right]\left[1-t^*(\varepsilon-\hbar\omega)\right]~,\\ \mathcal{\widetilde C}_{\alpha\beta}(\varepsilon,\omega) &=&\mathcal{T}(\varepsilon) \left[2\delta_{\alpha\beta}-\mathcal{T}(\varepsilon-\hbar\omega)\right]~,\end{aligned}$$ since for isothermal reservoirs at equilibrium, we have $f^{e,h}_L(\varepsilon)=f^{e,h}_R(\varepsilon)=f^{e,h}_M(\varepsilon)$. From Eq. (\[S\_eq\]), it can be shown using $\mathcal{S}^{pq}_{\alpha\beta}(-\omega)=e^{\hbar\omega/k_BT}\mathcal{S}^{qp}_{\alpha\beta}(\omega)$ that the noise spectrum obeys the relation $$\begin{aligned} \mathcal{S}^{pq}_{\alpha\beta}(\omega)=N(\hbar\omega)[\mathcal{S}^{qp}_{\alpha\beta}(-\omega)-\mathcal{S}^{pq}_{\alpha\beta}(\omega)]~,\end{aligned}$$ where $N(\hbar\omega)=[\exp(\hbar\omega/k_BT)-1]^{-1}$ is the Bose-Einstein distribution function. Removing the reservoirs’ index and using the definitions [@Andrade2015] of the electrical ac-conductance, $G(\omega)=[\mathcal{S}^{00}(-\omega)-\mathcal{S}^{00}(\omega)]/2\hbar\omega$, the thermal ac-conductance $K(\omega)=[\mathcal{S}^{11}(-\omega)-\mathcal{S}^{11}(\omega)]/2\hbar\omega T$, and the thermoelectric ac-conductance, $X(\omega)=[\mathcal{S}^{10}(-\omega)-\mathcal{S}^{01}(\omega)]/2\hbar\omega T$ which is the product of the ac-thermopower (i.e., Seebeck coefficient) by the electrical ac-conductance, we establish that the noises are related at equilibrium to the ac-conductances through the following fluctuation-dissipation relations $$\begin{aligned} \label{G} \mathcal{S}^{00}(\omega)&=&2\hbar\omega N(\hbar\omega)G(\omega)~,\\\label{X} \mathcal{S}^{01}(\omega)&=&2\hbar\omega T N(\hbar\omega)X(\omega)~,\\\label{K} \mathcal{S}^{11}(\omega)&=&2\hbar\omega T N(\hbar\omega)K(\omega)~.\end{aligned}$$ Through these relations, we can state that in a similar way that the finite-frequency electrical noise contains information on the dynamics of the charge transfer, the finite-frequency heat noise contains information on the dynamic of the heat transfer (since $G(\omega)$ and $K(\omega)$ are the response to an excitation modulated in time). Moreover, Eq. (\[X\]) confirms the key role played by the mixed noise $\mathcal{S}^{01}(\omega)$ to quantify the thermoelectric conversion. Note that in the limit of zero-frequency, Eqs. (\[G\])-(\[K\]) reduce to the relations given in Ref. , since we have in that limit $N(\hbar\omega)\rightarrow k_BT/\hbar\omega$. Energy-independent transmission ------------------------------- For an energy-independent transmission amplitude, $t$, the real parts of the electrical, mixed, and heat noise spectra are given by Fig. \[figure2\] in the low-temperature limit. We do not plot their imaginary parts whose magnitudes are smaller with a factor 100 comparing to the ones of the real parts. Let us now discuss the features appearing on Fig. \[figure2\]. First, we notice that similarly to the electrical noise, which cancels when the frequency is larger than the voltage, $\hbar\omega>eV$, the mixed and heat noises cancel as well. The reason is the following: knowing that the noise is called emission noise at positive frequency and absorption noise at negative frequency [@Basset2010], we understand that the system can not emit an energy larger than the energy provided to it, here the voltage since temperature is taken small. Second, we observe that the electrical noise varies linearly or by plateaus with both voltage and frequency, due to the fact that when transmission is energy independent, the system works in the linear regime. Third, the mixed noise can change its sign whereas the electrical and heat noises keep a single sign. Fourth, the electrical and mixed correlators between distinct reservoirs, $\mathcal{S}^{pq}_{LR}$, are equal in absolute values to the correlators in the same reservoir, $\mathcal{S}^{pq}_{LL}$, and nearly equal for the heat correlator [@note0]. ![Noises spectra as a function of frequency $\hbar\omega/\Gamma\in[-10,10]$ and voltage $eV/\Gamma\in[-10,10]$ for $\mathcal{T}=0.01$ and $t=\mathcal{T}+i[\mathcal{T}(1-\mathcal{T})]^{1/2}$, at low temperature $k_BT/\Gamma=0.01$. $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ is plotted in units of $e^{2-p-q}\Gamma^{1+p+q}/h$. The right reservoir is grounded ($\mu_R=0$).[]{data-label="figure2"}](figure2.eps){width="8cm"} In the limit of weak or perfect transmission, i.e., $\mathcal{T}\ll 1$ or $\mathcal{T}=1$ respectively, the integration over energy in Eq. (\[exp\_noise\]) can be performed analytically (see Appendix \[appendixB\] for the details of the calculation). The expressions of the noises, which are all real in these limits, are given in Table \[table1\]. These expressions constitute a generalization of the fluctuation-dissipation theorem to an out-of-equilibrium situation since the Bose-Einstein distribution function is estimated at frequency shifted by $\pm eV/\hbar$. Concerning the electrical noise, its expression at $\mathcal{T}\ll 1$ is in full agreement with the result of perturbative calculations [@Roussel2016]. Concerning the mixed and heat noises expressions, there is no previous work to compare in the literature. Note that at zero-voltage, the mixed noise cancels in both limits ($\mathcal{T}\ll 1$ and $\mathcal{T}=1$) but not in the intermediate regime: the mixed noise is then given by Eq. (\[X\]) which is a priori non-zero. It is also worth to notice that $\mathcal{S}^{11}_{\alpha\beta}(\omega)$ contains a contribution which is proportional to $\mathcal{S}^{00}_{\alpha\beta}(\omega)$ with a proportionality factor equal to $\mathcal{L}T^2$, where $\mathcal{L}=\pi^2k_B^2/3e^2$ is the Lorenz number. Since in certain limits, the heat noise is related to the thermal conductance and the electrical noise to the electrical conductance, as through Eqs. (\[K\]) and (\[G\]) at equilibrium for example, it is not surprising to find a relation which involves the Lorenz number between the heat noise and the electrical noise thanks to the Wiedemann-Franz law, or between the thermal conductance and the electrical noise as obtained in Ref. . Table \[table1\] gives also the sum over reservoirs of the electrical, mixed, and heat noises, $\sum_{\alpha\beta}\mathcal{S}^{pq}_{\alpha\beta}(\omega)$. Contrary to the total electrical and mixed noises, which are equal to zero in the limits we consider (no charging effect on the QD), the total heat noise takes a finite value which indeed corresponds to the heat power fluctuations. At zero-frequency, the power fluctuations are conserved, i.e., the heat power fluctuations are equal to the electrical power fluctuations [@Crepieux2015]. It is also true at finite-frequency provided that $\mathcal{T}=1$, since in that limit we have from Table \[table1\], $\sum_{\alpha\beta}\mathcal{S}^{11}_{\alpha\beta}(\omega)=V^2\mathcal{S}^{00}_{\alpha\alpha}(\omega)$. At zero-temperature, we get for $\mathcal{T}=1$: $\sum_{\alpha\beta}\mathcal{S}^{11}_{\alpha\beta}(\omega)=2|\hbar\omega|(eV)^2\Theta(-\omega)/h$, and for $\mathcal{T}\ll 1$, $\sum_{\alpha\beta}\mathcal{S}^{11}_{\alpha\beta}(\omega)=4\mathcal{T}|\hbar\omega|^3\Theta(-\omega)/h$ at zero-voltage, and $\sum_{\alpha\beta}\mathcal{S}^{11}_{\alpha\beta}(\omega)=\mathcal{T}|eV|^3/h$ at zero-frequency in good agreement with Ref. . ---------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- Type of noise Notation $\mathcal{T}\ll 1$ $\mathcal{T}=1$ --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- ------------------------------------------------------------ ---------------------------------------------------------------------------------------------------------------- ------------------------------------------------------------------------------------------------------------------ Electrical noise $\mathcal{S}^{00}_{\alpha\beta}(\omega)$ $\frac{e^2\mathcal{T}}{h}(2\delta_{\alpha\beta}-1)\sum_{\pm}(\hbar\omega\pm eV)N(\hbar\omega\pm eV)$ $\frac{e^2}{h}(2\delta_{\alpha\beta}-1)2\hbar\omega N(\hbar\omega)$ Total electrical noise $\sum_{\alpha\beta}\mathcal{S}^{00}_{\alpha\beta}(\omega)$ $0$ $0$ Mixed noise $\mathcal{S}^{01}_{\alpha\beta}(\omega)$ $\frac{e\mathcal{T}}{h}(2\delta_{\alpha L}-1)\sum_{\pm}\mp\frac{(\hbar\omega\pm eV)^2}{2}N(\hbar\omega\pm eV)$ $\frac{e}{h}(1-2\delta_{\alpha L})eV\hbar\omega N(\hbar\omega)$ Total mixed noise $\sum_{\alpha\beta}\mathcal{S}^{01}_{\alpha\beta}(\omega)$ $0$ $0$ Heat noise $\mathcal{S}^{11}_{\alpha\alpha}(\omega)$ $\mathcal{L}T^2\mathcal{S}^{00}_{\alpha\alpha}(\omega)$ $\left(\mathcal{L}T^2+\frac{V^2}{2}+\frac{\hbar^2\omega^2}{12e^2}\right)\mathcal{S}^{00}_{\alpha\alpha}(\omega)$ *[(auto-correlator)]{}&&$+\frac{\mathcal{T}}{h}\left[\hbar^3\omega^3N(\hbar\omega)+\sum_{\pm}\frac{(\hbar\omega\pm eV)^3}{3}N(\hbar\omega\pm eV)\right]$&$+\frac{\hbar^2\omega^2}{4h}\sum_{\pm}(\hbar\omega\pm eV)N(\hbar\omega\pm eV)$\ Heat noise&$\mathcal{S}^{11}_{\alpha\bar\alpha}(\omega)$& $\mathcal{L}T^2\mathcal{S}^{00}_{\alpha\bar\alpha}(\omega)$&$\mathcal{L}T^2\mathcal{S}^{00}_{\alpha\bar\alpha}(\omega)-\frac{\hbar^2\omega^2}{12e^2}\mathcal{S}^{00}_{\alpha\alpha}(\omega)$\ *[(cross-correlator)]{}&&$+\frac{\mathcal{T}}{h}\sum_{\pm}\frac{(\hbar\omega\pm eV)^3}{6}N(\hbar\omega\pm eV)$&$-\frac{\hbar^2\omega^2}{4h}\sum_{\pm}(\hbar\omega\pm eV)N(\hbar\omega\pm eV)$\ Total heat noise& $\sum_{\alpha\beta}\mathcal{S}^{11}_{\alpha\beta}(\omega)$&$\frac{\mathcal{T}}{h}\left[2\hbar^3\omega^3N(\hbar\omega)+\sum_{\pm}(\hbar\omega\pm eV)^3N(\hbar\omega\pm eV)\right]$&$V^2\mathcal{S}^{00}_{\alpha\alpha}(\omega)$\ ** ---------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- : Expressions of the electrical, mixed, and heat noises in the energy-independent weak/perfect transmission limits [@note1]. We have $\bar\alpha=R$ when $\alpha=L$, and vice versa. The total electrical, mixed and heat noises summed over both reservoirs are also given. The details of the calculations are performed in Appendix B.[]{data-label="table1"} Anderson-type energy transmission --------------------------------- ![Noises spectra as a function of frequency $\hbar\omega/\Gamma\in[-10,10]$ and voltage $eV/\Gamma\in[-10,10]$ for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, at low temperature $k_BT/\Gamma=0.01$. $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ is plotted in units of $e^{2-p-q}\Gamma^{1+p+q}/h$. The right reservoir is grounded ($\mu_R=0$).[]{data-label="figure3"}](figure3.eps){width="8cm"} For an Anderson-type transmission amplitude of the form $t(\varepsilon)=i\Gamma/[(\varepsilon-\varepsilon_d)+i\Gamma]$, both the real and imaginary parts of the electrical, mixed, and heat noise spectra are given by Fig. \[figure3\] in the low temperature limit. Note that the imaginary parts of $\mathcal{S}^{00}_{LL}(\omega)$ and $\mathcal{S}^{11}_{LL}(\omega)$ are both zero since the auto-correlators are real quantities, and that the real and imaginary parts of the cross-correlators are of the same order of magnitude, contrary to the energy independent transmission case. The main observation is the dramatically distinct spectra that we have for the auto-correlators, $\mathcal{S}^{00}_{LL}(\omega)$ and $\mathcal{S}^{11}_{LL}(\omega)$, in comparison to the cross-correlators, $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ with $p\ne q$ or/and $\alpha\ne\beta$. Whereas the auto-correlator spectra are quite similar to the ones obtained in the case of an energy-independent transmission amplitude (compare to Fig. \[figure2\]), excepted an additional structure in the region of small positive frequency, the cross-correlator spectra exhibit the following features: (i) their sign can change, (ii) it appears a new region with specific behavior close to small frequency, but (iii) we still have a cancellation of the noises for $\hbar\omega>eV$, again due to the fact that the system can not emit energy larger than the one provided to it. We remark that in any situations, those depicted in Figs. \[figure2\] and \[figure3\] and those summarized in Table I, the mixed noise cancels at zero-voltage, meaning that the cancellation of the ratio $-V/\Delta T$, which is equal to the Seebeck coefficient $S_T$ for open circuit, causes the cancellation of the mixed noise. This is one evidence that thermopower and mixed noise are closely connected. ![Electrical noises $\mathcal{S}^{00}_{LL}(\omega)$ and $\mathcal{S}^{00}_{LR}(\omega)$ (left column) and heat noises $\mathcal{S}^{11}_{LL}(\omega)$ and $\mathcal{S}^{11}_{LR}(\omega)$ (right column), as a function of frequency $\hbar\omega/eV$, for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, with $\Gamma/eV=0.01$, and for varying values of the temperature $k_BT/eV$: 0.01 (red line), 0.5 (orange line), 1 (blue line), 2 (brown line), and 10 (black line). $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ is plotted in units of $e^{2-p-q}(eV)^{1+p+q}/h$. The right reservoir is grounded ($\mu_R=0$).[]{data-label="figure4"}](figure4a.eps "fig:"){width="4cm"} ![Electrical noises $\mathcal{S}^{00}_{LL}(\omega)$ and $\mathcal{S}^{00}_{LR}(\omega)$ (left column) and heat noises $\mathcal{S}^{11}_{LL}(\omega)$ and $\mathcal{S}^{11}_{LR}(\omega)$ (right column), as a function of frequency $\hbar\omega/eV$, for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, with $\Gamma/eV=0.01$, and for varying values of the temperature $k_BT/eV$: 0.01 (red line), 0.5 (orange line), 1 (blue line), 2 (brown line), and 10 (black line). $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ is plotted in units of $e^{2-p-q}(eV)^{1+p+q}/h$. The right reservoir is grounded ($\mu_R=0$).[]{data-label="figure4"}](figure4b.eps "fig:"){width="4cm"} ![Electrical noises $\mathcal{S}^{00}_{LL}(\omega)$ and $\mathcal{S}^{00}_{LR}(\omega)$ (left column) and heat noises $\mathcal{S}^{11}_{LL}(\omega)$ and $\mathcal{S}^{11}_{LR}(\omega)$ (right column), as a function of frequency $\hbar\omega/eV$, for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, with $\Gamma/eV=0.01$, and for varying values of the temperature $k_BT/eV$: 0.01 (red line), 0.5 (orange line), 1 (blue line), 2 (brown line), and 10 (black line). $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ is plotted in units of $e^{2-p-q}(eV)^{1+p+q}/h$. The right reservoir is grounded ($\mu_R=0$).[]{data-label="figure4"}](figure4c.eps "fig:"){width="4cm"} ![Electrical noises $\mathcal{S}^{00}_{LL}(\omega)$ and $\mathcal{S}^{00}_{LR}(\omega)$ (left column) and heat noises $\mathcal{S}^{11}_{LL}(\omega)$ and $\mathcal{S}^{11}_{LR}(\omega)$ (right column), as a function of frequency $\hbar\omega/eV$, for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, with $\Gamma/eV=0.01$, and for varying values of the temperature $k_BT/eV$: 0.01 (red line), 0.5 (orange line), 1 (blue line), 2 (brown line), and 10 (black line). $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ is plotted in units of $e^{2-p-q}(eV)^{1+p+q}/h$. The right reservoir is grounded ($\mu_R=0$).[]{data-label="figure4"}](figure4d.eps "fig:"){width="4cm"} ![Electrical noises $\mathcal{S}^{00}_{LL}(\omega)$ and $\mathcal{S}^{00}_{LR}(\omega)$ (left column) and heat noises $\mathcal{S}^{11}_{LL}(\omega)$ and $\mathcal{S}^{11}_{LR}(\omega)$ (right column), as a function of frequency $\hbar\omega/eV$, for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, with $\Gamma/eV=0.01$, and for varying values of the temperature $k_BT/eV$: 0.01 (red line), 0.5 (orange line), 1 (blue line), 2 (brown line), and 10 (black line). $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ is plotted in units of $e^{2-p-q}(eV)^{1+p+q}/h$. The right reservoir is grounded ($\mu_R=0$).[]{data-label="figure4"}](figure4e.eps "fig:"){width="4cm"} ![Electrical noises $\mathcal{S}^{00}_{LL}(\omega)$ and $\mathcal{S}^{00}_{LR}(\omega)$ (left column) and heat noises $\mathcal{S}^{11}_{LL}(\omega)$ and $\mathcal{S}^{11}_{LR}(\omega)$ (right column), as a function of frequency $\hbar\omega/eV$, for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, with $\Gamma/eV=0.01$, and for varying values of the temperature $k_BT/eV$: 0.01 (red line), 0.5 (orange line), 1 (blue line), 2 (brown line), and 10 (black line). $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ is plotted in units of $e^{2-p-q}(eV)^{1+p+q}/h$. The right reservoir is grounded ($\mu_R=0$).[]{data-label="figure4"}](figure4f.eps "fig:"){width="4cm"} To have a deeper insight in the electrical, mixed, and heat noises, we plot their real and imaginary parts as a function of frequency at weak coupling strength $\Gamma$, for increasing temperatures in Figs. \[figure4\] and \[figure5\]. All the types of noise exhibit an asymmetric spectrum at low temperature (red curves) and a nearly symmetrical spectrum at large temperature with a vanishing imaginary part (black curves), due to the fact that when the temperature increases we are leaving the quantum regime. Thus, at large temperature, it is no longer necessary to make the distinction between non-symmetrized and symmetrized noises since the currents are no longer operators but just scalars (classical regime). The electrical and heat auto-correlators (see Figs. \[figure4\](a) and \[figure4\](b)) are real and positive quantities. The electrical auto-correlator, $\mathcal{S}^{00}_{LL}(\omega)$, is strongly frequency dependent at low temperature with a down staircase-like behavior starting from the value $2\pi\Gamma e^2/h$ and going to the value $0$ (see red curve in Fig. \[figure4\](a)), but resembles to a white noise at large temperature [@noteHT] with a constant value equals to $\pi\Gamma e^2/h$, except in a narrow low frequency region (see black curve in Fig. \[figure4\](a)). At large temperature, the heat auto-correlator, $\mathcal{S}^{11}_{LL}(\omega)$, presents a power-law variation with frequency, given by $\hbar^2\omega^2\pi\Gamma/h$ (see black curve in Fig. \[figure4\](b)) whereas the real part of $\mathcal{S}^{11}_{LR}(\omega)$ decreases linearly with temperature [@noteHT]. The electrical and heat cross-correlators, depicted in Figs. \[figure4\](c)-\[figure4\](f), are complex quantities whose imaginary parts cancel at large temperature (black curves), making the cross-correlators real quantities in that limit. The electrical auto-correlator and the real part of the electrical cross-correlator have distinct profiles but coincide at zero-frequency in absolute value since due to charge conservation we have $\mathcal{S}^{00}_{LL}(\omega=0)=-\mathcal{S}^{00}_{LR}(\omega=0)=\pi\Gamma e^2/2h$ in that limit. ![Mixed noises $\mathcal{S}^{01}_{LL}(\omega)$ (left column) and $\mathcal{S}^{01}_{LR}(\omega)$ (right column), as a function of frequency $\hbar\omega/eV$. Same parameters as in Fig. \[figure4\].[]{data-label="figure5"}](figure5a.eps "fig:"){width="4cm"} ![Mixed noises $\mathcal{S}^{01}_{LL}(\omega)$ (left column) and $\mathcal{S}^{01}_{LR}(\omega)$ (right column), as a function of frequency $\hbar\omega/eV$. Same parameters as in Fig. \[figure4\].[]{data-label="figure5"}](figure5b.eps "fig:"){width="4cm"} ![Mixed noises $\mathcal{S}^{01}_{LL}(\omega)$ (left column) and $\mathcal{S}^{01}_{LR}(\omega)$ (right column), as a function of frequency $\hbar\omega/eV$. Same parameters as in Fig. \[figure4\].[]{data-label="figure5"}](figure5c.eps "fig:"){width="4cm"} ![Mixed noises $\mathcal{S}^{01}_{LL}(\omega)$ (left column) and $\mathcal{S}^{01}_{LR}(\omega)$ (right column), as a function of frequency $\hbar\omega/eV$. Same parameters as in Fig. \[figure4\].[]{data-label="figure5"}](figure5d.eps "fig:"){width="4cm"} We turn now our interest to the mixed noise depicted in Fig. \[figure5\]. Similarly than for electrical and heat noises, increasing the temperature changes the mixed spectrum from an asymmetric profile to a symmetric profile with frequency, and cancels its imaginary part, again due to the fact that we are leaving the quantum regime. At low temperature, we also see that the imaginary parts of the mixed noises, $\mathcal{S}^{01}_{LL}(\omega)$ and $\mathcal{S}^{01}_{LR}(\omega)$, have a staircase-like profile which is a reminiscent of the electrical noise auto-correlator (compare Figs. \[figure5\](c) and \[figure5\](d) to Fig. \[figure4\](a)). Besides, the real parts of the mixed noises present quite particular profiles at low temperature: a linear profile in frequency for $\mathcal{S}^{01}_{LL}(\omega)$ (see the red curve in Fig. \[figure5\](a)) and vanishing value when $|\hbar\omega|>|eV|$ for $\mathcal{S}^{01}_{LR}(\omega)$ (see the red curve in Fig. \[figure5\](b)). At large temperature, $\mathcal{S}^{01}_{LL}(\omega)$ becomes frequency independent with an asymptotic value equal to $-\pi\Gamma e^2V/h$, and $\mathcal{S}^{01}_{LR}(\omega)$ cancels [@noteHT]. Here again, we find that the mixed noise is related to the Seebeck coefficient $S_T$, since both quantities vary linearly with voltage. ![Spectral coherence for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, for $\Gamma/eV=0.02$ (left column) and $\Gamma/eV=0.2$ (right column), at $k_BT_R/eV=0.1$ and temperature gradient equals to: $\Delta T/eV=0$ (blue curve), $\Delta T/eV=0.5$ (green curve), and $\Delta T/eV=1$ (red curve). The dashed black line shows the maximal possible value for the spectral coherence, i.e., 1.[]{data-label="figure6"}](figure6a.eps "fig:"){width="4cm"} ![Spectral coherence for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, for $\Gamma/eV=0.02$ (left column) and $\Gamma/eV=0.2$ (right column), at $k_BT_R/eV=0.1$ and temperature gradient equals to: $\Delta T/eV=0$ (blue curve), $\Delta T/eV=0.5$ (green curve), and $\Delta T/eV=1$ (red curve). The dashed black line shows the maximal possible value for the spectral coherence, i.e., 1.[]{data-label="figure6"}](figure6b.eps "fig:"){width="4cm"} ![Spectral coherence for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, for $\Gamma/eV=0.02$ (left column) and $\Gamma/eV=0.2$ (right column), at $k_BT_R/eV=0.1$ and temperature gradient equals to: $\Delta T/eV=0$ (blue curve), $\Delta T/eV=0.5$ (green curve), and $\Delta T/eV=1$ (red curve). The dashed black line shows the maximal possible value for the spectral coherence, i.e., 1.[]{data-label="figure6"}](figure6c.eps "fig:"){width="4cm"} ![Spectral coherence for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, for $\Gamma/eV=0.02$ (left column) and $\Gamma/eV=0.2$ (right column), at $k_BT_R/eV=0.1$ and temperature gradient equals to: $\Delta T/eV=0$ (blue curve), $\Delta T/eV=0.5$ (green curve), and $\Delta T/eV=1$ (red curve). The dashed black line shows the maximal possible value for the spectral coherence, i.e., 1.[]{data-label="figure6"}](figure6d.eps "fig:"){width="4cm"} ![Spectral coherence for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, for $\Gamma/eV=0.02$ (left column) and $\Gamma/eV=0.2$ (right column), at $k_BT_R/eV=0.1$ and temperature gradient equals to: $\Delta T/eV=0$ (blue curve), $\Delta T/eV=0.5$ (green curve), and $\Delta T/eV=1$ (red curve). The dashed black line shows the maximal possible value for the spectral coherence, i.e., 1.[]{data-label="figure6"}](figure6e.eps "fig:"){width="4cm"} ![Spectral coherence for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, for $\Gamma/eV=0.02$ (left column) and $\Gamma/eV=0.2$ (right column), at $k_BT_R/eV=0.1$ and temperature gradient equals to: $\Delta T/eV=0$ (blue curve), $\Delta T/eV=0.5$ (green curve), and $\Delta T/eV=1$ (red curve). The dashed black line shows the maximal possible value for the spectral coherence, i.e., 1.[]{data-label="figure6"}](figure6f.eps "fig:"){width="4cm"} ![Spectral coherence for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, for $\Gamma/eV=0.02$ (left column) and $\Gamma/eV=0.2$ (right column), at $k_BT_R/eV=0.1$ and temperature gradient equals to: $\Delta T/eV=0$ (blue curve), $\Delta T/eV=0.5$ (green curve), and $\Delta T/eV=1$ (red curve). The dashed black line shows the maximal possible value for the spectral coherence, i.e., 1.[]{data-label="figure6"}](figure6g.eps "fig:"){width="4cm"} ![Spectral coherence for $\mathcal{T}(\varepsilon)=\Gamma^2/(\varepsilon^2+\Gamma^2)$, for $\Gamma/eV=0.02$ (left column) and $\Gamma/eV=0.2$ (right column), at $k_BT_R/eV=0.1$ and temperature gradient equals to: $\Delta T/eV=0$ (blue curve), $\Delta T/eV=0.5$ (green curve), and $\Delta T/eV=1$ (red curve). The dashed black line shows the maximal possible value for the spectral coherence, i.e., 1.[]{data-label="figure6"}](figure6h.eps "fig:"){width="4cm"} For completeness, we discuss the spectral coherence of the cross-correlators, defined as $C^{pq}_{\alpha\beta}(\omega)=|\mathcal{S}^{pq}_{\alpha\beta}(\omega)|^2/[\mathcal{S}^{pp}_{\alpha\alpha}(\omega)\mathcal{S}^{qq}_{\beta\beta}(\omega)]$, and plot their profiles on Fig. \[figure6\]. Thanks to Cauchy-Schwarz inequality, we have $0\le C^{pq}_{\alpha\beta}(\omega)\le 1$, where the value zero for the spectral coherence means that the currents $I^p_\alpha$ and $I^q_\beta$ are uncorrelated, whereas the value one means that the currents $I^p_\alpha$ and $I^q_\beta$ are fully correlated. The plots on the left column of Fig. \[figure6\] are obtained for a weak coupling strength ($\Gamma/eV=0.02$), whereas the plots on the right column correspond to an intermediate coupling strength ($\Gamma/eV=0.2$). In the weak coupling strength limit, we remark that $C^{pq}_{\alpha\ne \beta}(\omega)$ is equal to zero at negative frequency, meaning that the absorbed signals in distinct reservoirs are uncorrelated. Moreover, we see in Fig. \[figure6\](a) that the left and right electrical currents are well correlated only at zero-frequency, i.e., in the large-time limit, due to charge conservation which imposes $C^{00}_{LR}(\omega=0)=C^{00}_{LL}(\omega=0)=1$. When the coupling strength increases, the spectral coherence $C^{00}_{LR}(\omega)$ is broadened to non-zero frequencies (see Fig. \[figure6\](b)) and can even reach $40\%$ at positive frequencies, an effect which is amplified around $\hbar\omega=eV$ when a temperature gradient is applied (see red curve in Fig. \[figure6\](b)). The electrical and heat currents inside a single reservoir, here $L$, are well correlated at $\Delta T=0$ (see blue curves in Figs. \[figure6\](c) and \[figure6\](d)) provided that $\hbar\omega >-eV$, excepted in a narrow region around $\hbar\omega=eV$ where a minimum of $C^{01}_{LL}(\omega)$ is observed. At the same frequency, $C^{01}_{LR}(\omega)$ exhibits a maximum (see blue curves in Figs. \[figure6\](e) and \[figure6\](f)) meaning that the left electrical current and the right heat current are maximally correlated in that region of frequency. The increasing of the coupling strength reinforces this effect with values of $C^{01}_{LR}(\omega)$ up to $80\%$ (see blue curve in Fig. \[figure6\](f)). This result allows us to make the prediction that the thermoelectric conversion could be optimal when the voltage applied to the QD is time-modulated with a frequency equals to the dc-voltage, i.e., $eV/\hbar$. This effect is, however, suppressed in the presence of a temperature gradient (see green and red curves in Fig. \[figure6\](f)). On the contrary, $C^{11}_{LR}(\omega)$ can be increased at high frequency by the application of a temperature gradient (see Fig. \[figure6\](h)). Conclusion ========== We have used the Keldysh out-of-equilibrium Green function technique to calculate the finite-frequency mixed noise, and have shown that its spectrum presents a rich and specific profile, which differs from the ones of the electrical and heat noises. At equilibrium, it is related to the thermoelectric ac-conductance, meaning that the finite-frequency mixed noise gives access to the dynamics of the thermoelectric conversion. Out-of-equilibrium, by a careful study of the spectral coherence, we find that the electrical current in one reservoir is strongly correlated to the heat current in the other reservoir when the frequency is of the order of the applied voltage. The method developed here constitutes an adequate framework which can be used in future works on this quantity in more complicated quantum systems, including multi-terminals, multi-channels, and interactions in some extent. We would like to thank M. Guigou, M. Lavagna, T. Martin, F. Michelini, and R. Zamoum for useful discussions. We acknowledge financial support from the CNRS Cellule Energie funding project ICARE, and from A\*Midex. Correlators of charge and heat currents in a QD {#appendixA} =============================================== Computation of the time-correlator $\mathcal{S}^{pq}_{\alpha\beta}(t,t')$ ------------------------------------------------------------------------- We use an approach analog to the one developed by Haug and Jauho for the calculation of electrical noise [@Haug2007], but instead of calculating the symmetrized noise, we calculate the non-symmetrized one since it is this latter quantity which is accessible in the experiments measuring the electrical current noise. We perform this calculation for each type of noise, the electrical, mixed and heat ones, using the general framework exposed in this Appendix. ### Expression of $\mathcal{S}^{pq}_{\alpha\beta}(t,t')$ in terms of the two-particle Green function of the QD, $G^{dd}_i$ We report Eq. (\[c1\]) in Eq. (\[def\_S\]) and get $$\begin{aligned} &&\mathcal{S}^{pq}_{\alpha\beta}(t,t')=-\frac{e^{2-p-q}}{\hbar^{2}}\sum_{\substack{k\in\alpha,k'\in\beta}}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p}\left(\varepsilon_{k'\beta}-\mu_{\beta}\right)^{q}\nonumber\\ &&\times\Big[V_{k\alpha}V_{k'\beta}\langle c^{\dag}_{k\alpha}(t)d(t)c^{\dag}_{k'\beta}(t')d(t')\rangle\nonumber\\ && -V_{k\alpha}V^{*}_{k'\beta}\langle c^{\dag}_{k\alpha}(t)d(t)d^{\dag}(t')c_{k'\beta}(t')\rangle\nonumber\\ &&-V^{*}_{k\alpha}V_{k'\beta}\langle d^{\dag}(t)c_{k\alpha}(t)c^{\dag}_{k'\beta}(t')d(t')\rangle\nonumber\\ && +V^{*}_{k\alpha}V^{*}_{k'\beta}\langle d^{\dag}(t)c_{k\alpha}(t)d^{\dag}(t')c_{k'\beta}(t')\rangle\Big]-\langle\hat{I}^{p}_{\alpha}\rangle\langle\hat{I}^{q}_{\beta}\rangle~.\end{aligned}$$ Defining the following [*greater*]{} two-particle Green functions [@Haug2007] $$\begin{aligned} G^{cd,>}_{1}(t,t')&=&i^{2}\langle Tc^{\dag}_{k\alpha}(t)d(t)c^{\dag}_{k'\beta}(t')d(t')\rangle~,\\ G^{cd,>}_{2}(t,t')&=&i^{2}\langle Tc^{\dag}_{k\alpha}(t)d(t)d^{\dag}(t')c_{k'\beta}(t')\rangle~,\\ G^{cd,>}_{3}(t,t')&=&i^{2}\langle Td^{\dag}(t)c_{k\alpha}(t)c^{\dag}_{k'\beta}(t')d(t')\rangle~,\\ G^{cd,>}_{4}(t,t')&=&i^{2}\langle Td^{\dag}(t)c_{k\alpha}(t)d^{\dag}(t')c_{k'\beta}(t')\rangle~,\end{aligned}$$ and using the Keldysh formalism [@Keldysh1964], the non-equilibrium contour-ordered counterparts of the correlation function can be expressed in terms of $G^{cd}_{i}(\tau,\tau')$, the contour-ordered counterparts of the two-particle Green functions, $G^{cd,>}_{i}(t,t')$, through $$\begin{aligned} \label{g1} &&\mathcal{S}^{pq}_{\alpha\beta}(\tau,\tau')=\frac{e^{2-p-q}}{\hbar^{2}}\sum_{\substack{k\in\alpha,k'\in\beta}}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p}\left(\varepsilon_{k'\beta}-\mu_{\beta}\right)^{q}\nonumber\\ &&\times\Big[V_{k\alpha}V_{k'\beta}G^{cd}_{1}(\tau,\tau')-V_{k\alpha}V^{*}_{k'\beta}G^{cd}_{2}(\tau,\tau')\nonumber\\ & &-V^{*}_{k\alpha}V_{k'\beta}G^{cd}_{3}(\tau,\tau')+V^{*}_{k\alpha}V^{*}_{k'\beta}G^{cd}_{4}(\tau,\tau')\Big]-\langle\hat{I}^{p}_{\alpha}\rangle\langle\hat{I}^{q}_{\beta}\rangle~.\nonumber\\\end{aligned}$$ The next step is to express the two-particle Green functions, $G^{cd}_i$, mixing $c$ and $d$ operators in terms of the two-particle Green functions of the QD, $G^{dd}_i$, and of the bare Green function of the reservoirs, $g_{k\alpha}$. We have [@Haug2007] $$\begin{aligned} &&G^{cd}_{1}(\tau,\tau')=-\frac{V^{*}_{k\alpha}V^{*}_{k'\beta}}{\hbar^2}\iint d\tau_1 d\tau_2 \nonumber\\ &&\times g_{k\alpha}(\tau_{1},\tau) g_{k'\beta}(\tau_{2},\tau')G^{dd}_{1}(\tau,\tau',\tau_{1},\tau_{2})~,\\ &&G^{cd}_{2}(\tau,\tau')=-\delta_{kk'}\delta_{\alpha\beta}g_{k\alpha}(\tau',\tau)G(\tau,\tau') -\frac{V^{*}_{k\alpha}V_{k'\beta}}{\hbar^2}\nonumber\\ &&\times\iint d\tau_1 d\tau_2g_{k\alpha}(\tau_{2},\tau)g_{k'\beta}(\tau',\tau_{1})G^{dd}_{2}(\tau,\tau',\tau_{1},\tau_{2})~,\\ &&G^{cd}_{3}(\tau,\tau')=-\delta_{kk'}\delta_{\alpha\beta}g_{k\alpha}(\tau,\tau')G(\tau',\tau)+\frac{V_{k\alpha}V^{*}_{k'\beta}}{\hbar^2}\nonumber\\ &&\times\iint d\tau_1 d\tau_2g_{k\alpha}(\tau,\tau_{1})g_{k'\beta}(\tau_{2},\tau')G^{dd}_{3}(\tau,\tau',\tau_{1},\tau_{2})~,\\ &&G^{cd}_{4}(\tau,\tau')=-\frac{V_{k\alpha}V_{k'\beta}}{\hbar^2}\iint d\tau_1 d\tau_2\nonumber\\ &&\times g_{k\alpha}(\tau,\tau_{1})g_{k'\beta}(\tau',\tau_{2})G^{dd}_{4}(\tau,\tau',\tau_{1},\tau_{2})~,\end{aligned}$$ with $$\begin{aligned} G^{dd}_{1}(\tau,\tau',\tau_{1},\tau_{2})&=&i^{2}\langle T_{C}d(\tau)d(\tau')d^{\dag}(\tau_{1})d^{\dag}(\tau_{2})\rangle~,\\ G^{dd}_{2}(\tau,\tau',\tau_{1},\tau_{2})&=&i^{2}\langle T_{C}d(\tau)d^{\dag}(\tau')d(\tau_{1})d^{\dag}(\tau_{2})\rangle~,\\ G^{dd}_{3}(\tau,\tau',\tau_{1},\tau_{2})&=&i^{2}\langle T_{C}d^{\dag}(\tau)d(\tau')d(\tau_{1})d^{\dag}(\tau_{2})\rangle~,\\ G^{dd}_{4}(\tau,\tau',\tau_{1},\tau_{2})&=&i^{2}\langle T_{C}d^{\dag}(\tau)d^{\dag}(\tau')d(\tau_{1})d(\tau_{2})\rangle~.\end{aligned}$$ Injecting the above expressions of $G^{cd}_{i}$ in Eq. (\[g1\]), we get $$\begin{aligned} \label{g2} &&\mathcal{S}^{pq}_{\alpha\beta}(\tau,\tau')=\frac{e^{2-p-q}}{\hbar^{2}}\Bigg(\delta_{\alpha\beta}\sum_{k\in\alpha}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p+q}\nonumber\\ &&\times|V_{k\alpha}|^{2}\big[g_{k\alpha}(\tau',\tau)G(\tau,\tau')+g_{k\alpha}(\tau,\tau')G(\tau',\tau)\big]\nonumber\\ & &+\sum_{\substack{k\in\alpha,k'\in\beta}}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p}\left(\varepsilon_{k'\beta}-\mu_{\beta}\right)^{q}\frac{|V_{k\alpha}V_{k'\beta}|^{2}}{\hbar^2}\nonumber\\ & &\times\iint d\tau_1 d\tau_2\Big[-g_{k\alpha}(\tau_{1},\tau)g_{k'\beta}(\tau_{2},\tau')G^{dd}_{1}(\tau,\tau',\tau_{1},\tau_{2})\nonumber\\ & &+g_{k\alpha}(\tau_{2},\tau)g_{k'\beta}(\tau',\tau_{1})G^{dd}_{2}(\tau,\tau',\tau_{1},\tau_{2})\nonumber\\ & & -g_{k\alpha}(\tau,\tau_{1})g_{k'\beta}(\tau_{2},\tau')G^{dd}_{3}(\tau,\tau',\tau_{1},\tau_{2})\nonumber\\ & &-g_{k\alpha}(\tau,\tau_{1})g_{k'\beta}(\tau',\tau_{2})G^{dd}_{4}(\tau,\tau',\tau_{1},\tau_{2})\Big]\Bigg) -\langle\hat{I}^{p}_{\alpha}\rangle\langle\hat{I}^{q}_{\beta}\rangle~,\nonumber\\\end{aligned}$$ where $G(\tau,\tau')=-i\langle T_{C}d(\tau)d^{\dag}(\tau')\rangle$ is the one-particle Green function of the QD. $ \quad \mathcal{P}_0(t,t')\quad $ $g^{<}_{k\alpha}(t',t)G^{>}(t,t')+g^{>}_{k\alpha}(t,t')G^{<}(t',t)$ -------------------------------------- ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- $ \mathcal{P}_1(t,t') $ $-\int dt_{1}\big[G^{r}(t',t_{1})g^{<}_{k\alpha}(t_{1},t)+G^{<}(t',t_{1})g^{a}_{k\alpha}(t_{1},t)\big]\int dt_{2}\big[G^{r}(t,t_{2})g^{>}_{k'\beta}(t_{2},t')+G^{>}(t,t_{2})g^{a}_{k'\beta}(t_{2},t')\big]$ $ \mathcal{P}_2(t,t') $ $G^{>}(t,t')\iint dt_{1}dt_{2}\Big[g^{r}_{k'\beta}(t',t_{1})G^{r}(t_{1},t_{2})g^{<}_{k\alpha}(t_{2},t) $ $ +g^{r}_{k'\beta}(t',t_{1})G^{<}(t_{1},t_{2})g^{a}_{k\alpha}(t_{2},t)+g^{<}_{k'\beta}(t',t_{1})G^{a}(t_{1},t_{2})g^{a}_{k\alpha}(t_{2},t)\Big]$ $ \mathcal{P}_3(t,t') $ $G^{<}(t',t)\iint dt_{1}dt_{2}\Big[g^{r}_{k\alpha}(t,t_{1})G^{r}(t_{1},t_{2})g^{>}_{k'\beta}(t_{2},t') $ $ +g^{r}_{k\alpha}(t,t_{1})G^{>}(t_{1},t_{2})g^{a}_{k'\beta}(t_{2},t')+g^{>}_{k\alpha}(t,t_{1})G^{a}(t_{1},t_{2})g^{a}_{k'\beta}(t_{2},t')\Big]$ $ \mathcal{P}_4(t,t') $ $-\int dt_{1}\big[g^{r}_{k\alpha}(t,t_{1})G^{>}(t_{1},t')+g^{>}_{k\alpha}(t,t_{1})G^{a}(t_{1},t')\big]$ $\times\int dt_{2}\big[g^{r}_{k'\beta}(t',t_{2})G^{<}(t_{2},t)+g^{<}_{k'\beta}(t',t_{2})G^{a}(t_{2},t)\big]$ : Expressions of the $\mathcal{P}_{i}(t,t')$ coefficients appearing in Eq. (\[s2\]).[]{data-label="tablePtime"} ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- $ \quad\mathcal{P}_0(\omega)\quad $ $\int d\varepsilon\Big(g^{<}_{k\alpha}(\varepsilon)G^{>}(\varepsilon-\hbar\omega)+g^{>}_{k\alpha}(\varepsilon-\hbar\omega)G^{<}(\varepsilon)\Big)$ --------------------------------------- ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- $ \mathcal{P}_1(\omega) $ $-\int d\varepsilon \Big[G^{r}(\varepsilon)g^{<}_{k\alpha}(\varepsilon)G^{r}(\varepsilon-\hbar\omega)g^{>}_{k'\beta}(\varepsilon-\hbar\omega)+G^{r}(\varepsilon)g^{<}_{k\alpha}(\varepsilon)G^{>}(\varepsilon-\hbar\omega)g^{a}_{k'\beta}(\varepsilon-\hbar\omega)$ $ +G^{<}(\varepsilon)g^{a}_{k\alpha}(\varepsilon)G^{r}(\varepsilon-\hbar\omega)g^{>}_{k'\beta}(\varepsilon-\hbar\omega)+G^{<}(\varepsilon)g^{a}_{k\alpha}(\varepsilon)G^{>}(\varepsilon-\hbar\omega)g^{a}_{k'\beta}(\varepsilon-\hbar\omega)\Big]$ $ \mathcal{P}_2(\omega) $ $\int d\varepsilon G^{>}(\varepsilon-\hbar\omega)\Big[g^{r}_{k'\beta}(\varepsilon)G^{r}(\varepsilon)g^{<}_{k\alpha}(\varepsilon)+g^{r}_{k'\beta}(\varepsilon)G^{<}(\varepsilon)g^{a}_{k\alpha}(\varepsilon)+g^{<}_{k'\beta}(\varepsilon)G^{a}(\varepsilon)g^{a}_{k\alpha}(\varepsilon)\Big]$ $ \mathcal{P}_3(\omega) $ $\int d\varepsilon G^{<}(\varepsilon) \Big[g^{r}_{k\alpha}(\varepsilon-\hbar\omega)G^{r}(\varepsilon-\hbar\omega)g^{>}_{k'\beta}(\varepsilon-\hbar\omega)$ $ +g^{r}_{k\alpha}(\varepsilon-\hbar\omega)G^{>}(\varepsilon-\hbar\omega)g^{a}_{k'\beta}(\varepsilon-\hbar\omega) +g^{>}_{k\alpha}(\varepsilon-\hbar\omega)G^{a}(\varepsilon-\hbar\omega)g^{a}_{k'\beta}(\varepsilon-\hbar\omega)\Big]$ $ \mathcal{P}_4(\omega) $ $-\int d\varepsilon \Big[g^{r}_{k\alpha}(\varepsilon-\hbar\omega)G^{>}(\varepsilon-\hbar\omega)+g^{>}_{k\alpha}(\varepsilon-\hbar\omega)G^{a}(\varepsilon-\hbar\omega)\Big] \Big[g^{r}_{k'\beta}(\varepsilon)G^{<}(\varepsilon)+g^{<}_{k'\beta}(\varepsilon)G^{a}(\varepsilon)\Big]$ ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- : Expressions of the $\mathcal{P}_{i}(\omega)$ coefficients appearing in Eq. (\[s3\]).[]{data-label="tablePomega"} $A_{\alpha\beta}(\varepsilon,\omega)$ $-\frac{1}{2}\big[f^h_{\alpha}(\varepsilon-\hbar\omega)+f^h_{\bar{\alpha}}(\varepsilon-\hbar\omega)\big]\big([f^e_{\alpha}(\varepsilon)-f^e_{\beta}(\varepsilon)]t(\varepsilon)\mathcal{T}(\varepsilon-\hbar\omega)-\frac{1}{2}[f^e_{\bar{\beta}}(\varepsilon)-3f^e_{\beta}(\varepsilon)]\mathcal{T}(\varepsilon)\mathcal{T}(\varepsilon-\hbar\omega)\big)$ --------------------------------------- --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- $B_{\alpha\beta}(\varepsilon,\omega)$ $f^e_{\alpha}(\varepsilon)f^h_{\beta}(\varepsilon-\hbar\omega)t(\varepsilon)t(\varepsilon-\hbar\omega)-\frac{1}{2}f^e_{\alpha}(\varepsilon)\big[f^h_{\alpha}(\varepsilon-\hbar\omega)+f^h_{\bar{\alpha}}(\varepsilon-\hbar\omega)\big]t(\varepsilon)\mathcal{T}(\varepsilon-\hbar\omega)$ $-\frac{1}{2}f^h_{\beta}(\varepsilon-\hbar\omega)\big[f^e_{\alpha}(\varepsilon)+f^e_{\bar{\alpha}}(\varepsilon)\big]\mathcal{T}(\varepsilon)t(\varepsilon-\hbar\omega)+\frac{1}{4}\big[f^e_{\alpha}(\varepsilon)+f^e_{\bar{\alpha}}(\varepsilon)\big]\big[f^h_{\alpha}(\varepsilon-\hbar\omega)+f^h_{\bar{\alpha}}(\varepsilon-\hbar\omega)\big]\mathcal{T}(\varepsilon)\mathcal{T}(\varepsilon-\hbar\omega)$ $C_{\alpha\beta}(\varepsilon,\omega)$ $-\frac{1}{2}f^h_{\beta}(\varepsilon-\hbar\omega)\big[f^e_{\beta}(\varepsilon)+f^e_{\bar{\beta}}(\varepsilon)\big]\mathcal{T}(\varepsilon)t(\varepsilon-\hbar\omega)-\frac{1}{2}f^h_{\alpha}(\varepsilon-\hbar\omega)\big[f^e_{\beta}(\varepsilon)+f^e_{\bar{\beta}}(\varepsilon)\big]\mathcal{T}(\varepsilon)t^{*}(\varepsilon-\hbar\omega) $ $ +\frac{1}{4}\big[f^e_{\beta}(\varepsilon)+f^e_{\bar{\beta}}(\varepsilon)\big]\big[f^h_{\beta}(\varepsilon-\hbar\omega)+f^h_{\bar{\beta}}(\varepsilon-\hbar\omega)\big]\mathcal{T}(\varepsilon)\mathcal{T}(\varepsilon-\hbar\omega)$ $D_{\alpha}(\varepsilon,\omega)$ $f^e_{\alpha}(\varepsilon)\big[f^h_{\alpha}(\varepsilon-\hbar\omega)+f^h_{\bar{\alpha}}(\varepsilon-\hbar\omega)\big]\mathcal{T}(\varepsilon-\hbar\omega)$ $E_{\alpha}(\varepsilon,\omega)$ $\big[f^e_{\alpha}(\varepsilon)+f^e_{\bar{\alpha}}(\varepsilon)\big]f^h_{\alpha}(\varepsilon-\hbar\omega)\mathcal{T}(\varepsilon)$ : Expressions of the coefficients appearing in Eq. (\[s5\]).[]{data-label="tablecoeff"} ### Evaluation of the two-particle Green functions $G^{dd}_i$ using decoupling procedure In the absence of Coulomb interactions, we can express fully the two-particle Green functions of the QD, $G^{dd}_i(\tau,\tau',\tau_{1},\tau_{2})$, in terms of the one-particle Green function of the QD, $G(\tau,\tau')$, through $$\begin{aligned} G^{dd}_{1}(\tau,\tau',\tau_{1},\tau_{2})&=&G(\tau,\tau_{2})G(\tau',\tau_{1})-G(\tau,\tau_{1})G(\tau',\tau_{2})~,\\ G^{dd}_{2}(\tau,\tau',\tau_{1},\tau_{2})&=&G(\tau,\tau')G(\tau_{1},\tau_{2})-G(\tau,\tau_{2})G(\tau_{1},\tau')~,\\ G^{dd}_{3}(\tau,\tau',\tau_{1},\tau_{2})&=&G(\tau_{1},\tau)G(\tau',\tau_{2})-G(\tau',\tau)G(\tau_{1},\tau_{2})~,\\ G^{dd}_{4}(\tau,\tau',\tau_{1},\tau_{2})&=&G(\tau_{2},\tau)G(\tau_{1},\tau')-G(\tau_{1},\tau)G(\tau_{2},\tau')~.\end{aligned}$$ Injecting these four expressions in Eq. (\[g2\]), we obtain a result which can be separated into two parts, a *connected* part and a *disconnected* part [@Haug2007] $$\begin{aligned} \mathcal{S}^{pq}_{\alpha\beta}(\tau,\tau')&=&\mathcal{S}^{pq}_{\substack{\alpha\beta,disc}}(\tau,\tau')+\mathcal{S}^{pq}_{\substack{\alpha\beta,conn}}(\tau,\tau')-\langle\hat{I}^{p}_{\alpha}\rangle\langle\hat{I}^{q}_{\beta}\rangle~,\nonumber\\\end{aligned}$$ with $$\begin{aligned} &&\mathcal{S}^{pq}_{\substack{\alpha\beta,disc}}(\tau,\tau')=\frac{e^{2-p-q}}{\hbar^{2}}\nonumber\\ &&\times\sum_{\substack{k\in\alpha,k'\in\beta}}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p}\left(\varepsilon_{k'\beta}-\mu_{\beta}\right)^{q}\frac{|V_{k\alpha}V_{k'\beta}|^{2}}{\hbar^{2}}\nonumber\\ && \times\bigg[\int d\tau_{1}g_{k\alpha}(\tau_{1},\tau)G(\tau,\tau_{1})\int d\tau_{2}g_{k'\beta}(\tau_{2},\tau')G(\tau',\tau_{2})\nonumber\\ && -\int d\tau_{2}g_{k\alpha}(\tau_{2},\tau)G(\tau,\tau_{2})\int d\tau_{1}g_{k'\beta}(\tau',\tau_{1})G(\tau_{1},\tau')\nonumber\\ & &-\int d\tau_{1}g_{k\alpha}(\tau,\tau_{1})G(\tau_{1},\tau)\int d\tau_{2}g_{k'\beta}(\tau_{2},\tau')G(\tau',\tau_{2})\nonumber\\ && +\int d\tau_{1}g_{k\alpha}(\tau,\tau_{1})G(\tau_{1},\tau)\int d\tau_{2}g_{k'\beta}(\tau',\tau_{2})G(\tau_{2},\tau')\bigg]~,\nonumber\\\end{aligned}$$ and $$\begin{aligned} \label{s4} &&\mathcal{S}^{pq}_{\substack{\alpha\beta,conn}}(\tau,\tau')=\frac{e^{2-p-q}}{\hbar^{2}}\bigg[\delta_{\alpha\beta}\sum_{k\in\alpha}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p+q}|V_{k\alpha}|^{2}\nonumber\\ &&\times\big[g_{k\alpha}(\tau',\tau)G(\tau,\tau')+g_{k\alpha}(\tau,\tau')G(\tau',\tau)\big]\nonumber\\ & &+\sum_{\substack{k\in\alpha,k'\in\beta}}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p}\left(\varepsilon_{k'\beta}-\mu_{\beta}\right)^{q}\frac{|V_{k\alpha}V_{k'\beta}|^{2}}{\hbar^{2}}\nonumber\\ & &\times\iint d\tau_1 d\tau_2\Big[-g_{k\alpha}(\tau_{1},\tau)g_{k'\beta}(\tau_{2},\tau')G(\tau,\tau_{2})G(\tau',\tau_{1})\nonumber\\ &&+g_{k\alpha}(\tau_{2},\tau)g_{k'\beta}(\tau',\tau_{1})G(\tau,\tau')G(\tau_{1},\tau_{2})\nonumber\\ & & +g_{k\alpha}(\tau,\tau_{1})g_{k'\beta}(\tau_{2},\tau')G(\tau',\tau)G(\tau_{1},\tau_{2})\nonumber\\ &&-g_{k\alpha}(\tau,\tau_{1})g_{k'\beta}(\tau',\tau_{2})G(\tau_{2},\tau)G(\tau_{1},\tau')\Big]\bigg]~.\end{aligned}$$ The disconnected part can be calculated directly. We obtain $\mathcal{S}^{pq}_{\substack{\alpha\beta,disc}}(\tau,\tau')=\langle\hat{I}^{p}_{\alpha}\rangle\langle\hat{I}^{q}_{\beta}\rangle$, thus finally $\mathcal{S}^{pq}_{\alpha\beta}(\tau,\tau')=\mathcal{S}^{pq}_{\substack{\alpha\beta,conn}}(\tau,\tau')$. ### Analytic continuation of the connected part We perform now the analytic continuation of Eq. (\[s4\]) to get its $\tau>\tau'$ component $$\begin{aligned} &&\mathcal{S}^{pq}_{\alpha\beta}(t,t')= \frac{e^{2-p-q}}{\hbar^{2}}\Bigg[\delta_{\alpha\beta}\sum_{k\in\alpha}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p+q}|V_{k\alpha}|^{2}\nonumber\\ & &\times\underbrace{\big[g_{k\alpha}(\tau',\tau)G(\tau,\tau')+g_{k\alpha}(\tau,\tau')G(\tau',\tau)\big]_{\tau>\tau'}}_{\mathcal{P}_{0}(t,t')}\\ & &+\sum_{\substack{k\in\alpha,k'\in\beta}}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p}\left(\varepsilon_{k'\beta}-\mu_{\beta}\right)^{q}\frac{|V_{k\alpha}V_{k'\beta}|^{2}}{\hbar^{2}}\\ & &\times\bigg(-\underbrace{\big[\iint d\tau_{1}d\tau_{2}G(\tau',\tau_{1})g_{k\alpha}(\tau_{1},\tau)G(\tau,\tau_{2})g_{k'\beta}(\tau_{2},\tau')\big]_{\tau>\tau'}}_{\mathcal{P}_{1}(t,t')}\nonumber\\ & & +\underbrace{\big[G(\tau,\tau')\iint d\tau_{1}d\tau_{2}g_{k'\beta}(\tau',\tau_{1})G(\tau_{1},\tau_{2})g_{k\alpha}(\tau_{2},\tau)\big]_{\tau>\tau'}}_{\mathcal{P}_{2}(t,t')}\\ & & +\underbrace{\big[G(\tau',\tau)\iint d\tau_{1}d\tau_{2}g_{k\alpha}(\tau,\tau_{1})G(\tau_{1},\tau_{2})g_{k'\beta}(\tau_{2},\tau')\big]_{\tau>\tau'}}_{\mathcal{P}_{3}(t,t')}\nonumber\\ & & -\underbrace{\big[\iint d\tau_{1}d\tau_{2}g_{k\alpha}(\tau,\tau_{1})G(\tau_{1},\tau')g_{k'\beta}(\tau',\tau_{2})G(\tau_{2},\tau)\big]_{\tau>\tau'}}_{\mathcal{P}_{4}(t,t')}\bigg)\Bigg]~.\end{aligned}$$ The five $\mathcal{P}_{i}(t,t')$ contributions are computed using analytic continuation rules [@Haug2007]. Their expressions are given in Table \[tablePtime\]. Using these notations, the noise reads as $$\begin{aligned} \label{s2} &&\mathcal{S}^{pq}_{\alpha\beta}(t,t')=\frac{e^{2-p-q}}{\hbar^2}\Bigg[\delta_{\alpha\beta}\sum_{k\in\alpha}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p+q}|V_{k\alpha}|^{2}\mathcal{P}_{0}(t,t')\nonumber\\ &&+\sum_{\substack{k\in\alpha\\k'\in\beta}}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p}\left(\varepsilon_{k'\beta}-\mu_{\beta}\right)^{q}\frac{|V_{k\alpha}V_{k'\beta}|^{2}}{\hbar^{2}}\sum^{4}_{i=1}\mathcal{P}_{i}(t,t')\Bigg]~.\nonumber\\\end{aligned}$$ Finite-frequency non-symmetrized noise $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ ------------------------------------------------------------------------------- ### Fourier transform of $\mathcal{S}^{pq}_{\alpha\beta}(t,t')$ and exact result for $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ Performing the Fourier transform of Eq. (\[s2\]) and using the fact that in the stationary case the Green functions depend on the difference of their time arguments only, we obtain: $$\begin{aligned} \label{s3} &&\mathcal{S}^{pq}_{\alpha\beta}(\omega)=\frac{e^{2-p-q}}{\hbar}\Bigg[\delta_{\alpha\beta}\sum_{k\in\alpha}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p+q}|V_{k\alpha}|^{2}\mathcal{P}_{0}(\omega)\nonumber\\ &&+\sum_{\substack{k\in\alpha\\k'\in\beta}}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p}\left(\varepsilon_{k'\beta}-\mu_{\beta}\right)^{q}|V_{k\alpha}V_{k'\beta}|^{2}\sum^{4}_{i=1}\mathcal{P}_{i}(\omega)\Bigg]~,\nonumber\\\end{aligned}$$ where the expressions of $\mathcal{P}_{i}(\omega)$ are given in Table \[tablePomega\]. Using the expressions of the bare Green functions of the reservoirs, $g^{>}_{k\alpha}(\varepsilon)$ and $g^{<}_{k\alpha}(\varepsilon)$, in terms of the Fermi-Dirac distribution function for electrons, $f^e_{\alpha}(\varepsilon)$, and Fermi-Dirac distribution function for holes, $f^h_{\alpha}(\varepsilon)=1-f^e_{\alpha}(\varepsilon)$, $$\begin{aligned} &&g^{<}_{k\alpha}(\varepsilon)=2\pi if^e_{\alpha}(\varepsilon)\delta(\varepsilon-\varepsilon_{k\alpha})~,\\ &&g^{>}_{k\alpha}(\varepsilon)=-2\pi if^h_{\alpha}(\varepsilon)\delta(\varepsilon-\varepsilon_{k\alpha})~,\end{aligned}$$ we can rewrite Eq. (\[s3\]) under the form $$\begin{aligned} \label{s1} &&\mathcal{S}^{pq}_{\alpha\beta}(\omega)=\frac{e^{2-p-q}}{h}\int d\varepsilon \Bigg[\delta_{\alpha\beta}i F^{p+q}_{\alpha}(\varepsilon)f^e_{\alpha}(\varepsilon)\nonumber\\ & &\times\Big[G^{<}(\varepsilon-\hbar\omega)+G^{r}(\varepsilon-\hbar\omega)-G^{a}(\varepsilon-\hbar\omega)\Big]\nonumber\\ & &-\delta_{\alpha\beta}i F^{p+q}_{\alpha}(\varepsilon-\hbar\omega)f^h_{\alpha}(\varepsilon-\hbar\omega)G^{<}(\varepsilon)\nonumber\\ & & -G^{r}(\varepsilon)G^{r}(\varepsilon-\hbar\omega)f^e_{\alpha}(\varepsilon)f^h_{\beta}(\varepsilon-\hbar\omega)F^{p}_{\alpha}(\varepsilon)F^{q}_{\beta}(\varepsilon-\hbar\omega)\nonumber\\ & &-G^{a}(\varepsilon)G^{a}(\varepsilon-\hbar\omega)f^e_{\beta}(\varepsilon)f^h_{\alpha}(\varepsilon-\hbar\omega)F^{p}_{\alpha}(\varepsilon-\hbar\omega)F^{q}_{\beta}(\varepsilon)\nonumber\\ & & +iG^{r}(\varepsilon)G^{>}(\varepsilon-\hbar\omega) f^e_{\alpha}(\varepsilon)F^{p}_{\alpha}(\varepsilon)H^{q*}_{\beta}(\varepsilon)\nonumber\\ & &+iG^{<}(\varepsilon)G^{r}(\varepsilon-\hbar\omega)f^h_{\beta}(\varepsilon-\hbar\omega)F^{q}_{\beta}(\varepsilon-\hbar\omega)H^{p}_{\alpha}(\varepsilon)\nonumber\\ & & +iG^{a}(\varepsilon)G^{>}(\varepsilon-\hbar\omega) f^e_{\beta}(\varepsilon)F^{q}_{\beta}(\varepsilon)H^{p}_{\alpha}(\varepsilon)\nonumber\\ & &+iG^{<}(\varepsilon)G^{a}(\varepsilon-\hbar\omega)f^h_{\alpha}(\varepsilon-\hbar\omega)F^{p}_{\alpha}(\varepsilon-\hbar\omega)H^{q*}_{\beta}(\varepsilon)\nonumber\\ & & +G^{<}(\varepsilon)G^{>}(\varepsilon-\hbar\omega)H^{p}_{\alpha}(\varepsilon,p)H^{q*}_{\beta}(\varepsilon)\Bigg]~,\end{aligned}$$ where we have introduced the two following functions $$\begin{aligned} F^p_{\alpha}(\varepsilon)&=&2\pi\sum_{k\in\alpha}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p}|V_{k\alpha}|^{2}\delta(\varepsilon-\varepsilon_{k\alpha})\nonumber\\ &=& \left(\varepsilon-\mu_{\alpha}\right)^{p}\underbrace{2\pi |V_\alpha(\varepsilon)|^{2}\rho_\alpha(\varepsilon)}_ {=\Gamma_{\alpha}(\varepsilon)}~,\end{aligned}$$ and $$\begin{aligned} H^p_{\alpha}(\varepsilon)&=&\sum_{k\in\alpha}\left(\varepsilon_{k\alpha}-\mu_{\alpha}\right)^{p}|V_{k\alpha}|^{2}\big[g^{a}_{k\alpha}(\varepsilon)-g^{r}_{k\alpha}(\varepsilon-\hbar\omega)\big]~,\nonumber\\\end{aligned}$$ with $\rho_\alpha(\varepsilon)$ the density of states associated to the reservoirs $\alpha$, and $\Gamma_\alpha=2\pi |V_\alpha(\varepsilon)|^{2}\rho_\alpha(\varepsilon)$, the coupling strength between the QD and the reservoir $\alpha$. Note that Eq. (\[s1\]), given $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$, has been obtained without making any approximation at this stage: it is the exact result for a non-interacting QD. ### $\mathcal{S}^{pq}_{\alpha\beta}(\omega)$ for symmetrical barriers in the wide-band limit To continue further the calculation, we make two simplifying assumptions: (i) wide-band limit, i.e., we are working on an interval of energy in which the density of states is constant, $\rho_{\alpha}(\varepsilon)=cst$, and we assume as well that $V_{\alpha}(\varepsilon)$ is energy independent, consequently, we have $\Gamma_\alpha(\varepsilon)=2\pi |V_\alpha(\varepsilon)|^2 \rho_{\alpha}(\varepsilon)\equiv\Gamma_\alpha$, and (ii) symmetrical barriers, i.e., we assume that the left and right barriers are symmetrical ($\Gamma_{L}=\Gamma_{R}\equiv\Gamma$). In that case, we have the remarkable relation [@Zamoum2016]: $t(\varepsilon)+t^{*}(\varepsilon)=2\mathcal{T}(\varepsilon)$, with $t(\varepsilon)=i\Gamma G^{r}(\varepsilon)$, the transmission amplitude and $\mathcal{T}(\varepsilon)$, the transmission coefficient. Within these two simplifying assumptions, we have $$\begin{aligned} F^p_{\alpha}(\varepsilon)&=&\Gamma(\varepsilon-\mu_{\alpha})^{p}~,\end{aligned}$$ and $$\begin{aligned} H^p_{\alpha}(\varepsilon)&=&\frac{i\Gamma}{2}\left[(\varepsilon-\mu_{\alpha})^{p}+(\varepsilon-\hbar\omega-\mu_{\alpha})^{p}\right]~.\end{aligned}$$ Injecting these two last expressions in Eq. (\[s1\]), rearranging the terms, and using the relations [@Mahan2000; @Haug2007] $$\begin{aligned} &&G^{>}(\varepsilon)-G^{<}(\varepsilon)=G^{r}(\varepsilon)-G^{a}(\varepsilon)~,\\ &&G^{<}(\varepsilon)=i\Gamma G^{r}(\varepsilon)G^{a}(\varepsilon)\big[f^e_{\alpha}(\varepsilon)+f^e_{\bar{\alpha}}(\varepsilon)\big]~,\\ &&G^{r}(\varepsilon)-G^{a}(\varepsilon)=-2i\Gamma G^{r}(\varepsilon)G^{a}(\varepsilon)~,\\ &&\mathcal{T}(\varepsilon)=\Gamma^2G^{r}(\varepsilon)G^{a}(\varepsilon)~.\end{aligned}$$ We finally get $$\begin{aligned} \label{s5} &&\mathcal{S}^{pq}_{\alpha\beta}(\omega)=\frac{e^{2-p-q}}{h}\int d\varepsilon \bigg[\delta_{\alpha\beta}\left(\varepsilon-\mu_{\alpha}\right)^{p+q}D_{\alpha}(\varepsilon,\omega)\nonumber\\ & & +\delta_{\alpha\beta}\left(\varepsilon-\hbar\omega-\mu_{\alpha}\right)^{p+q}E_{\alpha}(\varepsilon,\omega)\nonumber\\ & & +\left(\varepsilon-\mu_{\alpha}\right)^{p}\left(\varepsilon-\mu_{\beta}\right)^{q}A_{\alpha\beta}(\varepsilon,\omega) \nonumber\\ & & +\left(\varepsilon-\mu_{\alpha}\right)^{p}\left(\varepsilon-\hbar\omega-\mu_{\beta}\right)^{q}B_{\alpha\beta}(\varepsilon,\omega)\nonumber\\ & & +\left(\varepsilon-\hbar\omega-\mu_{\alpha}\right)^{p}\left(\varepsilon-\mu_{\beta}\right)^{q}B^{*}_{\beta\alpha}(\varepsilon,\omega)\nonumber\\ & & +\left(\varepsilon-\hbar\omega-\mu_{\alpha}\right)^{p}\left(\varepsilon-\hbar\omega-\mu_{\beta}\right)^{q}C_{\alpha\beta}(\varepsilon,\omega)\bigg]~,\end{aligned}$$ where the coefficients $A_{\alpha}(\varepsilon,\omega)$, $B_{\alpha}(\varepsilon,\omega)$, $C_{\alpha}(\varepsilon,\omega)$, $D_{\alpha}(\varepsilon,\omega)$, and $E_{\alpha}(\varepsilon,\omega)$ are given in Table \[tablecoeff\]. Equation (\[s5\]) leads to the Eqs. (\[exp\_noise\])-(\[C\]) once we define $$\begin{aligned} \mathcal{A}_{\alpha\beta}(\varepsilon,\omega)&=&A_{\alpha\beta}(\varepsilon,\omega)+\delta_{\alpha\beta}D_{\alpha}(\varepsilon,\omega)~,\\ \mathcal{B}_{\alpha\beta}(\varepsilon,\omega)&=&B_{\alpha\beta}(\varepsilon,\omega)~,\\ \mathcal{C}_{\alpha\beta}(\varepsilon,\omega)&=&C_{\alpha\beta}(\varepsilon,\omega)+\delta_{\alpha\beta}E_{\alpha}(\varepsilon,\omega)~,\\ f^{e,h}_M(\varepsilon)&=&\frac{1}{2}\big[f^{e,h}_\alpha(\varepsilon)+f^{e,h}_{\bar\alpha}(\varepsilon)\big]~,\end{aligned}$$ where $\bar{\alpha}=R$ when $\alpha=L$, and $\bar{\alpha}=L$ when $\alpha=R$. Noise spectrum for $t$ and $\mathcal{T}$ independent of energy {#appendixB} ============================================================== In case of independent energy transmission amplitude and coefficient, Eqs. (\[exp\_noise\])-(\[C\]) reduce to $$\begin{aligned} \label{S00} \mathcal{S}^{00}_{\alpha\beta}(\omega)&=&\frac{e^{2}}{h}\sum_{\gamma\delta}\int_{-\infty}^{\infty} d\varepsilon \mathcal{M}_{\alpha\beta}^{\gamma\delta}f^e_\gamma(\varepsilon)f^h_\delta(\varepsilon-\hbar\omega)~,\nonumber\\\end{aligned}$$ $$\begin{aligned} \label{S01} \mathcal{S}^{01}_{\alpha\beta}(\omega)&=&\frac{e}{h}\sum_{\gamma\delta}\int_{-\infty}^{\infty} d\varepsilon \big[(\varepsilon-\mu_\beta)\mathcal{M}_{\alpha\beta}^{\gamma\delta}-\hbar\omega\mathcal{N}_{\alpha\beta}^{\gamma\delta}\big]\nonumber\\ &&\times f^e_\gamma(\varepsilon)f^h_\delta(\varepsilon-\hbar\omega)~,\end{aligned}$$ $$\begin{aligned} \label{S10} \mathcal{S}^{10}_{\alpha\beta}(\omega)&=&\frac{e}{h}\sum_{\gamma\delta}\int_{-\infty}^{\infty} d\varepsilon \big[(\varepsilon-\mu_\alpha)\mathcal{M}_{\alpha\beta}^{\gamma\delta} -\hbar\omega(\mathcal{N}_{\beta\alpha}^{\gamma\delta})^*\big]\nonumber\\ &&\times f^e_\gamma(\varepsilon)f^h_\delta(\varepsilon-\hbar\omega)~,\end{aligned}$$ and $$\begin{aligned} \label{S11} \mathcal{S}^{11}_{\alpha\beta}(\omega)&=&\frac{1}{h}\sum_{\gamma\delta}\int_{-\infty}^{\infty} d\varepsilon \big[(\varepsilon-\mu_\alpha)(\varepsilon-\mu_\beta)\mathcal{M}_{\alpha\beta}^{\gamma\delta}\nonumber\\ &&-\hbar\omega(\varepsilon-\mu_\alpha)\mathcal{N}_{\alpha\beta}^{\gamma\delta}-\hbar\omega(\varepsilon-\mu_\beta)(\mathcal{N}_{\beta\alpha}^{\gamma\delta})^*\nonumber\\ &&+\hbar^2\omega^2\mathcal{O}_{\alpha\beta}^{\gamma\delta}\big]f^e_\gamma(\varepsilon)f^h_\delta(\varepsilon-\hbar\omega)~,\end{aligned}$$ with the coefficients $\mathcal{M}_{\alpha\beta}^{\gamma\delta}$, $\mathcal{N}_{\alpha\beta}^{\gamma\delta}$, and $\mathcal{O}_{\alpha\beta}^{\gamma\delta}$ given in Tables \[tableM\]-\[tableO\], where $\mathcal{Z}=[\mathcal{T}(1-\mathcal{T})]^{1/2}$ is the imaginary part of $t$, $\mathcal{T}$ being the real part of $t$. These real and imaginary parts are extracted from the two relations: $tt^*=\mathcal{T}$ and $t+t^*=2\mathcal{T}$. $\mathcal{M}_{\alpha\beta}^{\gamma\delta}$ $\gamma=\delta=L$ $\gamma=\delta=R$ $\gamma=L$, $\delta=R$ $\gamma=R$, $\delta=L$ -------------------------------------------- ------------------- ------------------- ------------------------------- ------------------------------- $\alpha=\beta=L$ $\mathcal{T}^2$ $\mathcal{T}^2$ $\mathcal{T}(1-\mathcal{T})$ $\mathcal{T}(1-\mathcal{T})$ $\alpha=\beta=R$ $\mathcal{T}^2$ $\mathcal{T}^2$ $\mathcal{T}(1-\mathcal{T})$ $\mathcal{T}(1-\mathcal{T})$ $\alpha=L$, $\beta=R$ $-\mathcal{T}^2$ $-\mathcal{T}^2$ $-\mathcal{T}(1-\mathcal{T})$ $-\mathcal{T}(1-\mathcal{T})$ $\alpha=R$,$\beta=L$ $-\mathcal{T}^2$ $-\mathcal{T}^2$ $-\mathcal{T}(1-\mathcal{T})$ $-\mathcal{T}(1-\mathcal{T})$ : Expressions of the coefficients $\mathcal{M}_{\alpha\beta}^{\gamma\delta}$ appearing in Eqs. (\[S00\])-(\[S11\]).[]{data-label="tableM"} $\mathcal{N}_{\alpha\beta}^{\gamma\delta}$ $\gamma=\delta=L$ $\gamma=\delta=R$ $\gamma=L$, $\delta=R$ $\gamma=R$, $\delta=L$ -------------------------------------------- ------------------------------------------ ------------------------------------------ ------------------------------------------------------- ------------------------------------------------------- $\alpha=\beta=L$ $\frac{\mathcal{T}^2}{2}+i\mathcal{ZT}$ $\frac{\mathcal{T}^2}{2}$ $-\frac{i\mathcal{ZT}}{2}$ $\mathcal{T}(1-\mathcal{T})-\frac{i\mathcal{ZT}}{2}$ $\alpha=\beta=R$ $\frac{\mathcal{T}^2}{2}$ $\frac{\mathcal{T}^2}{2}+i\mathcal{ZT}$ $\mathcal{T}(1-\mathcal{T})-\frac{i\mathcal{ZT}}{2}$ $-\frac{i\mathcal{ZT}}{2}$ $\alpha=L$, $\beta=R$ $-\frac{\mathcal{T}^2}{2}$ $-\frac{\mathcal{T}^2}{2}-i\mathcal{ZT}$ $-\mathcal{T}(1-\mathcal{T})+\frac{i\mathcal{ZT}}{2}$ $\frac{i\mathcal{ZT}}{2}$ $\alpha=R$,$\beta=L$ $-\frac{\mathcal{T}^2}{2}-i\mathcal{ZT}$ $-\frac{\mathcal{T}^2}{2}$ $\frac{i\mathcal{ZT}}{2}$ $-\mathcal{T}(1-\mathcal{T})+\frac{i\mathcal{ZT}}{2}$ : Expressions of the coefficients $\mathcal{N}_{\alpha\beta}^{\gamma\delta}$ appearing in Eqs. (\[S01\])-(\[S11\]).[]{data-label="tableN"} $\mathcal{O}_{\alpha\beta}^{\gamma\delta}$ $\gamma=\delta=L$ $\gamma=\delta=R$ $\gamma=L$, $\delta=R$ $\gamma=R$, $\delta=L$ -------------------------------------------- ------------------------------------------------------ ------------------------------------------------------ ------------------------------------------------------ ------------------------------------------------------ $\alpha=\beta=L$ $\frac{\mathcal{T}^2}{4}+\mathcal{T}(1-\mathcal{T})$ $\frac{\mathcal{T}^2}{4}$ $\frac{\mathcal{T}^2}{4}$ $\frac{\mathcal{T}^2}{4}+\mathcal{T}(1-\mathcal{T})$ $\alpha=\beta=R$ $\frac{\mathcal{T}^2}{4}$ $\frac{\mathcal{T}^2}{4}+\mathcal{T}(1-\mathcal{T})$ $\frac{\mathcal{T}^2}{4}+\mathcal{T}(1-\mathcal{T})$ $\frac{\mathcal{T}^2}{4}$ $\alpha=L$, $\beta=R$ $-\frac{\mathcal{T}^2}{4}+\frac{i\mathcal{ZT}}{2}$ $-\frac{\mathcal{T}^2}{4}-\frac{i\mathcal{ZT}}{2}$ $-\frac{\mathcal{T}^2}{4}-\frac{i\mathcal{ZT}}{2}$ $-\frac{\mathcal{T}^2}{4}+\frac{i\mathcal{ZT}}{2}$ $\alpha=R$,$\beta=L$ $-\frac{\mathcal{T}^2}{4}-\frac{i\mathcal{ZT}}{2}$ $-\frac{\mathcal{T}^2}{4}+\frac{i\mathcal{ZT}}{2}$ $-\frac{\mathcal{T}^2}{4}+\frac{i\mathcal{ZT}}{2}$ $-\frac{\mathcal{T}^2}{4}-\frac{i\mathcal{ZT}}{2}$ : Expressions of the coefficients $\mathcal{O}_{\alpha\beta}^{\gamma\delta}$ appearing in Eq. (\[S11\]).[]{data-label="tableO"} Preliminary calculations ------------------------ In the following sections, we will meet the integral $$\begin{aligned} I^{(n)}_{\gamma\delta}=\int_{-\infty}^\infty \varepsilon^n d\varepsilon f_\gamma^e(\varepsilon)f_\delta^h(\varepsilon-\hbar\omega)~,\end{aligned}$$ with $n$ = 0, 1 or 2. Here, we calculate this integral considering the isothermal case, $T_L=T_R\equiv T$, thus $$\begin{aligned} && I^{(n)}_{\gamma\delta} =N(\hbar\omega+\mu_\delta-\mu_\gamma)\nonumber\\ &&\times\int_{-\infty}^\infty \varepsilon^n d\varepsilon \frac{\sinh\left(\frac{\hbar\omega+\mu_\delta-\mu_\gamma}{2k_BT}\right)}{2\cosh\left(\frac{\varepsilon-\mu_\gamma}{2k_BT}\right)\cosh\left(\frac{\varepsilon-\mu_\delta-\hbar\omega}{2k_BT}\right)}~,\end{aligned}$$ where we have introduced the Bose-Einstein distribution function: $N(\hbar\omega)=[\exp(\hbar\omega/k_BT)-1]^{-1}$. Using the identity: $\sinh(a-b)/[\cosh(a)\cosh(b)]=\tanh(a)-\tanh(b)$, we end up with $$\begin{aligned} &&I^{(n)}_{\gamma\delta} =\frac{N(\hbar\omega+\mu_\delta-\mu_\gamma)}{2}\nonumber\\ &&\times\int_{-\infty}^\infty \varepsilon^n d\varepsilon \left[\tanh\left(\frac{\varepsilon-\mu_\gamma}{2k_BT}\right)-\tanh\left(\frac{\varepsilon-\mu_\delta-\hbar\omega}{2k_BT}\right)\right]~.\nonumber\\\end{aligned}$$ To go further, we perform a Taylor expansion up to the third order with $x=\omega$, $\mu_\gamma$ or $\mu_\delta$. It leads to $$\begin{aligned} I^{(n)}_{\gamma\delta} &=&\frac{N(\hbar\omega+\mu_\delta-\mu_\gamma)}{2}\Bigg[\frac{\mu_\gamma-\mu_\delta-\hbar\omega}{2k_BT}L^{(n)}_1 \nonumber\\ &&+\frac{\mu_\gamma^2-(\mu_\delta+\hbar\omega)^2}{4k_B^2T^2}L^{(n)}_2 +\frac{\mu_\gamma^3-(\mu_\delta+\hbar\omega)^3}{24k_B^3T^3}L^{(n)}_3\Bigg]~,\nonumber\\\end{aligned}$$ with $L^{(n)}_1=\int_{-\infty}^\infty \varepsilon^n d\varepsilon[\tanh^2(\varepsilon/2k_BT)-1]$, $L^{(n)}_2=\int_{-\infty}^\infty \varepsilon^n d\varepsilon[\tanh^3(\varepsilon/2k_BT)-\tanh(\varepsilon/2k_BT)]$, and $L^{(n)}_3=\int_{-\infty}^\infty \varepsilon^n d\varepsilon[1+3\tanh^4(\varepsilon/2k_BT)-4\tanh^2(\varepsilon/2k_BT)]$. The calculation of these integrals gives $$\begin{aligned} L^{(n)}_1= \left\{\begin{array}{ll} -4k_BT& (n=0)\\ 0 &(n=1)\\ -4\pi^2k_B^3T^3/3 &(n=2) \end{array}\right.~,\end{aligned}$$ $$\begin{aligned} L^{(n)}_2 = \left\{\begin{array}{ll} 0& (n=0)\\ -4k_B^2T^2 &(n=1)\\ 0&(n=2) \end{array}\right.~,\end{aligned}$$ and $$\begin{aligned} L^{(n)}_3= \left\{\begin{array}{ll} 0& (n=0)\\ 0 &(n=1)\\ -16k_B^3T^3 &(n=2) \end{array}\right.~.\end{aligned}$$ Finally, we get $$\begin{aligned} \label{I0} &&I^{(0)}_{\gamma\delta}=(\hbar\omega+\mu_\delta-\mu_\gamma)N(\hbar\omega+\mu_\delta-\mu_\gamma)~,\\\label{I1} &&I^{(1)}_{\gamma\delta}=\frac{(\hbar\omega+\mu_\delta)^2-\mu_\gamma^2}{2}N(\hbar\omega+\mu_\delta-\mu_\gamma)~,\\\label{I2} && I^{(2)}_{\gamma\delta}=\bigg[\frac{(\hbar\omega+\mu_\delta-\mu_\gamma)\pi^2k_B^2T^2}{3}\nonumber\\ &&+\frac{(\hbar\omega+\mu_\delta)^3-\mu_\gamma^3}{3}\bigg]N(\hbar\omega+\mu_\delta-\mu_\gamma)~.\end{aligned}$$ Limit of weak transmission $\mathcal{T}\ll 1$ --------------------------------------------- In this subsection, we give the calculation of the expressions appearing in the first column of Table I. ### Electrical noise spectrum We calculate only $ \mathcal{S}^{00}_{LL}(\omega)$ since when $t$ and $\mathcal{T}$ are independent of energy, we have the relations $$\begin{aligned} \mathcal{S}^{00}_{LL}(\omega)&=&\mathcal{S}^{00}_{RR}(\omega)=-\mathcal{S}^{00}_{LR}(\omega)=-\mathcal{S}^{00}_{RL}(\omega)~.\end{aligned}$$ At weak $\mathcal{T}$, we have $$\begin{aligned} &&\mathcal{S}^{00}_{LL}(\omega)=\frac{e^{2}}{h}\sum_{\gamma\delta}\int_{-\infty}^\infty d\varepsilon \mathcal{M}_{LL}^{\gamma\delta}f^e_\gamma(\varepsilon)f^h_\delta(\varepsilon-\hbar\omega)\nonumber\\ &=&\frac{e^{2}\mathcal{T}}{h}\int_{-\infty}^\infty [f^e_L(\varepsilon)f^h_R(\varepsilon-\hbar\omega) +f^e_R(\varepsilon)f^h_L(\varepsilon-\hbar\omega)]d\varepsilon\nonumber\\ &=&\frac{e^{2}\mathcal{T}}{h}\Big[ I^{(0)}_{LR}+ I^{(0)}_{RL}\Big]~,\end{aligned}$$ which gives $$\begin{aligned} \mathcal{S}^{00}_{LL}(\omega)&=&\frac{e^{2}\mathcal{T}}{h}\big[(\hbar\omega-eV)N(\hbar\omega-eV)\nonumber\\ &&+(\hbar\omega+eV)N(\hbar\omega+eV)\big]~.\end{aligned}$$ It reduces at equilibrium (zero-voltage) to $$\begin{aligned} \mathcal{S}^{00}_{LL}(\omega)&=&\frac{2e^{2}\mathcal{T}}{h}\hbar\omega N(\hbar\omega)~,\end{aligned}$$ and at zero-temperature to $$\begin{aligned} \mathcal{S}^{00}_{LL}(\omega) &=&\frac{e^{2}\mathcal{T}}{h}\big[(eV-\hbar\omega)\Theta(eV-\hbar\omega)\nonumber\\ &&-(\hbar\omega+eV)\Theta(-eV-\hbar\omega)\big]~,\end{aligned}$$ since we have $N(x)=-\Theta(-x)$ when $T\rightarrow 0$. ### Mixed noise spectrum We start from $$\begin{aligned} \mathcal{S}^{01}_{\alpha\beta}(\omega)&=&\frac{e}{h}\sum_{\gamma\delta}\int_{-\infty}^{\infty} d\varepsilon \big[(\varepsilon-\mu_\beta)\mathcal{M}_{\alpha\beta}^{\gamma\delta}\nonumber\\ &&-\hbar\omega\mathcal{N}_{\alpha\beta}^{\gamma\delta}\big]f^e_\gamma(\varepsilon)f^h_\delta(\varepsilon-\hbar\omega)~.\end{aligned}$$ We calculate only $ \mathcal{S}^{01}_{LL}(\omega)$ since at weak $\mathcal{T}$, we have: $$\begin{aligned} \mathcal{S}^{01}_{LL}(\omega)&=&-\mathcal{S}^{01}_{RR}(\omega)=\mathcal{S}^{01}_{LR}(\omega)=-\mathcal{S}^{01}_{RL}(\omega)~.\end{aligned}$$ We have $$\begin{aligned} \mathcal{S}^{01}_{LL}(\omega)&=&\frac{e\mathcal{T}}{h}\int_{-\infty}^{\infty} d\varepsilon \big[(\varepsilon-\mu_L)[f^e_L(\varepsilon)f^h_R(\varepsilon-\hbar\omega)\nonumber\\ &&+f^e_R(\varepsilon)f^h_L(\varepsilon-\hbar\omega)]-\hbar\omega f^e_R(\varepsilon)f^h_L(\varepsilon-\hbar\omega)\big]\nonumber\\ &=&\frac{e\mathcal{T}}{h}\Big[I^{(1)}_{LR}-\mu_L I^{(0)}_{LR}+I^{(1)}_{RL}-(\mu_L+\hbar\omega)I^{(0)}_{RL}\Big]~,\nonumber\\\end{aligned}$$ which gives $$\begin{aligned} \mathcal{S}^{01}_{LL}(\omega)&=&\frac{e\mathcal{T}}{2h}\big[(\hbar\omega-eV)^2N(\hbar\omega-eV)\nonumber\\ &&-(\hbar\omega+eV)^2N(\hbar\omega+eV)\big]~.\end{aligned}$$ It reduces to $\mathcal{S}^{01}_{LL}(\omega)=0$ at equilibrium (zero-voltage) and to $$\begin{aligned} \mathcal{S}^{01}_{LL}(\omega)&=&\frac{e\mathcal{T}}{2h}\big[-(\hbar\omega-eV)^2\Theta(eV-\hbar\omega)\nonumber\\ &&+(\hbar\omega+eV)^2\Theta(-eV-\hbar\omega)\big]~,\end{aligned}$$ at zero-temperature. ### Heat noise spectrum We start from $$\begin{aligned} \mathcal{S}^{11}_{\alpha\beta}(\omega)&=&\frac{1}{h}\sum_{\gamma\delta}\int_{-\infty}^{\infty} d\varepsilon \big[(\varepsilon-\mu_\alpha)(\varepsilon-\mu_\beta)\mathcal{M}_{\alpha\beta}^{\gamma\delta}\nonumber\\ &&-\hbar\omega(\varepsilon-\mu_\alpha)\mathcal{N}_{\alpha\beta}^{\gamma\delta}-\hbar\omega(\varepsilon-\mu_\beta)(\mathcal{N}_{\beta\alpha}^{\gamma\delta})^*\nonumber\\ &&+\hbar^2\omega^2\mathcal{O}_{\alpha\beta}^{\gamma\delta}\big]f^e_\gamma(\varepsilon)f^h_\delta(\varepsilon-\hbar\omega)~,\end{aligned}$$ which gives for $\mathcal{S}^{11}_{LL}(\omega)$ $$\begin{aligned} &&\mathcal{S}^{11}_{LL}(\omega)=\frac{\mathcal{T}}{h}\int_{-\infty}^{\infty} d\varepsilon \big[(\varepsilon-\mu_L)^2[f^e_L(\varepsilon)f^h_R(\varepsilon-\hbar\omega)\nonumber\\ &&+f^e_R(\varepsilon)f^h_L(\varepsilon-\hbar\omega)]-2\hbar\omega(\varepsilon-\mu_L)f^e_R(\varepsilon)f^h_L(\varepsilon-\hbar\omega)\nonumber\\ &&+\hbar^2\omega^2[f^e_L(\varepsilon)f^h_L(\varepsilon-\hbar\omega)+f^e_R(\varepsilon)f^h_L(\varepsilon-\hbar\omega)]\big]\nonumber\\ &&=\frac{\mathcal{T}}{h} \bigg[I^{(2)}_{LR}+I^{(2)}_{RL}-2\mu_L(I^{(1)}_{LR}+I^{(1)}_{RL})+\mu_L^2(I^{(0)}_{LR}\nonumber\\ &&+I^{(0)}_{RL})-2\hbar\omega(I^{(1)}_{RL}-\mu_LI^{(0)}_{RL}) +\hbar^2\omega^2(I^{(0)}_{LL}+I^{(0)}_{RL})\bigg]~.\nonumber\\\end{aligned}$$ We report the expressions of the integrals given by Eqs. (\[I0\])-(\[I2\]) and factorize the various contributions. It gives $$\begin{aligned} \mathcal{S}^{11}_{LL}(\omega)&=&\frac{\mathcal{T}}{h}\bigg[(\hbar\omega)^3N(\hbar\omega)\nonumber\\ &&+\frac{\pi^2k_B^2T^2}{3}\Big[(\hbar\omega-eV)N(\hbar\omega-eV)\nonumber\\ &&+(\hbar\omega+eV)N(\hbar\omega+eV)\Big]\nonumber\\ &&+\frac{(\hbar\omega-eV)^3}{3}N(\hbar\omega-eV)\nonumber\\ &&+\frac{(\hbar\omega+eV)^3}{3}N(\hbar\omega+eV)\bigg]~,\end{aligned}$$ which reduces at equilibrium (zero-voltage) to $$\begin{aligned} \mathcal{S}^{11}_{LL}(\omega)&=&\frac{\mathcal{T}}{h}\bigg[\frac{5}{3}(\hbar\omega)^3+\frac{2\pi^2k_B^2T^2}{3}\hbar\omega \bigg]N(\hbar\omega)~,\nonumber\\\end{aligned}$$ and at zero-temperature to $$\begin{aligned} \mathcal{S}^{11}_{LL}(\omega)&=&\frac{\mathcal{T}}{h}\bigg[-(\hbar\omega)^3\Theta(-\hbar\omega)+\nonumber\\ &&\frac{(eV-\hbar\omega)^3}{3}\Theta(eV-\hbar\omega)\nonumber\\ &&-\frac{(\hbar\omega+eV)^3}{3}\Theta(-eV-\hbar\omega)\bigg]~,\end{aligned}$$ since we have $N(x)=-\Theta(-x)$ when $T\rightarrow 0$. Note that $\mathcal{S}^{11}_{RR}(\omega)$ is obtained from the expression of $\mathcal{S}^{11}_{LL}(\omega)$ by inverting the voltage $V\rightarrow -V$, as a consequence we have $\mathcal{S}^{11}_{RR}(\omega)=\mathcal{S}^{11}_{LL}(\omega)$. We now calculate $$\begin{aligned} \mathcal{S}^{11}_{LR}(\omega)&=&\frac{\mathcal{T}}{h}\int_{-\infty}^{\infty} d\varepsilon \Big[-(\varepsilon-\mu_L)(\varepsilon-\mu_R)\nonumber\\ &&\times[f^e_L(\varepsilon)f^h_R(\varepsilon-\hbar\omega)+f^e_R(\varepsilon)f^h_L(\varepsilon-\hbar\omega)]\nonumber\\ &&+\hbar\omega(\varepsilon-\mu_L)f^e_L(\varepsilon)f^h_R(\varepsilon-\hbar\omega)\nonumber\\ && +\hbar\omega(\varepsilon-\mu_R)f^e_R(\varepsilon)f^h_L(\varepsilon-\hbar\omega)\Big]\nonumber\\ &=&\frac{\mathcal{T}}{h}\Big[-I^{(2)}_{LR}-I^{(2)}_{RL}+(\mu_L+\mu_R)\big[I^{(1)}_{LR}+I^{(1)}_{RL}\big]\nonumber\\ &&-\mu_L\mu_R\big[I^{(0)}_{LR}+I^{(0)}_{RL}\big]+\hbar\omega\big[I^{(1)}_{LR}+I^{(1)}_{RL}\big]\nonumber\\ &&-\hbar\omega\mu_LI^{(0)}_{LR} -\hbar\omega\mu_RI^{(0)}_{RL} \Big]~.\end{aligned}$$ We report the expressions of the integrals given by Eqs. (\[I0\])-(\[I2\]) and factorize the various contributions. It gives $$\begin{aligned} \mathcal{S}^{11}_{LR}(\omega)&=&\frac{\mathcal{T}}{h}\bigg[-\frac{\pi^2k_B^2T^2}{3}\Big[(\hbar\omega-eV)N(\hbar\omega-eV)\nonumber\\ &&+(\hbar\omega+eV)N(\hbar\omega+eV)\Big]\nonumber\\ &&+\frac{(\hbar\omega-eV)^3}{6}N(\hbar\omega-eV)\nonumber\\ &&+\frac{(\hbar\omega+eV)^3}{6}N(\hbar\omega+eV)\bigg]~,\end{aligned}$$ which reduces at equilibrium (zero-voltage) to $$\begin{aligned} \mathcal{S}^{11}_{LR}(\omega)&=&\frac{\mathcal{T}}{h}\bigg[\frac{1}{3}(\hbar\omega)^3-\frac{2\pi^2k_B^2T^2}{3}\hbar\omega \bigg]N(\hbar\omega)~,\nonumber\\\end{aligned}$$ and at zero-temperature to $$\begin{aligned} \mathcal{S}^{11}_{LR}(\omega)&=&\frac{\mathcal{T}}{6h}\bigg[(eV-\hbar\omega)^3\Theta(eV-\hbar\omega)\nonumber\\ &&-(\hbar\omega+eV)^3\Theta(-eV-\hbar\omega)\bigg]~,\end{aligned}$$ since we have $N(x)=-\Theta(-x)$ when $T\rightarrow 0$. Note that $\mathcal{S}^{11}_{RL}(\omega)$ is obtained from the expression of $\mathcal{S}^{11}_{LR}(\omega)$ by inverting the voltage $V\rightarrow -V$, as a consequence we have $\mathcal{S}^{11}_{RL}(\omega)=\mathcal{S}^{11}_{LR}(\omega)$.\ Limit of perfect transmission $\mathcal{T}=1$ --------------------------------------------- In this subsection, we give the calculation of the expressions appearing in the second column of Table I. ### Electrical noise spectrum For $\mathcal{T}=1$, we have $$\begin{aligned} \mathcal{S}^{00}_{\alpha\beta}(\omega)&=&\frac{e^{2}}{h}(2\delta_{\alpha\beta}-1)\int_{-\infty}^\infty d\varepsilon [f^e_L(\varepsilon)f^h_L(\varepsilon-\hbar\omega)\nonumber\\ &&+f^e_R(\varepsilon)f^h_R(\varepsilon-\hbar\omega)]\nonumber\\ &=&\frac{e^{2}}{h}(2\delta_{\alpha\beta}-1)[I^{(0)}_{LL}+I^{(0)}_{RR}]\nonumber\\ &=&\frac{e^{2}}{h}(2\delta_{\alpha\beta}-1)2\hbar\omega N(\hbar\omega)~.\end{aligned}$$ ### Mixed noise spectrum For $\mathcal{T}=1$, we have $$\begin{aligned} &&\mathcal{S}^{01}_{LL}(\omega)=\frac{e}{h}\int_{-\infty}^\infty d\varepsilon \left(\varepsilon-\mu_L-\frac{\hbar\omega}{2}\right)\nonumber\\ &&\times\Big[f^e_L(\varepsilon)f^h_L(\varepsilon-\hbar\omega)+f^e_R(\varepsilon)f^h_R(\varepsilon-\hbar\omega)\Big]\nonumber\\ &&=\frac{e}{h}\left[I^{(1)}_{LL}+I^{(1)}_{RR}-\left(\mu_L+\frac{\hbar\omega}{2}\right)[I^{(0)}_{LL}+I^{(0)}_{RR}]\right]~.\nonumber\\\end{aligned}$$ After simplification, it leads to $$\begin{aligned} \mathcal{S}^{01}_{LL}(\omega)=-\frac{e}{h}\hbar\omega eV N(\hbar\omega)~.\end{aligned}$$ A similar calculation leads to $\mathcal{S}^{01}_{RR}(\omega)=\frac{e}{h}\hbar\omega eV N(\hbar\omega)$. Moreover, we have $ \mathcal{S}^{01}_{LR}(\omega)=-\mathcal{S}^{01}_{RR}(\omega)$, and $\mathcal{S}^{01}_{RL}(\omega)=-\mathcal{S}^{01}_{LL}(\omega)$. ### Heat noise spectrum For $\mathcal{T}=1$, we have $$\begin{aligned} &&\mathcal{S}^{11}_{LL}(\omega)=\frac{1}{h}\int_{-\infty}^\infty d\varepsilon\bigg([(\varepsilon-\mu_L)^2-\hbar\omega(\varepsilon-\mu_L)]\nonumber\\ &&\times[f^e_L(\varepsilon)f^h_L(\varepsilon-\hbar\omega)+f^e_R(\varepsilon)f^h_R(\varepsilon-\hbar\omega)]\nonumber\\ && +\frac{\hbar^2\omega^2}{4}\sum_{\gamma\delta}f^e_\gamma(\varepsilon)f^h_\delta(\varepsilon-\hbar\omega)\bigg)\nonumber\\ &&=\frac{1}{h}\bigg[I^{(2)}_{LL}+I^{(2)}_{RR}-(2\mu_L+\hbar\omega)[I^{(1)}_{LL}+I^{(1)}_{RR}]\nonumber\\ &&+\mu_L(\mu_L+\hbar\omega)[I^{(0)}_{LL}+I^{(0)}_{RR}]+\frac{\hbar^2\omega^2}{4}\sum_{\gamma\delta}I^{(0)}_{\alpha\beta}\bigg]~.\nonumber\\\end{aligned}$$ It gives $$\begin{aligned} &&\mathcal{S}^{11}_{LL}(\omega)=\frac{1}{h}\bigg[\left(\frac{2\hbar\omega\pi^2k_B^2T^2}{3}+\frac{\hbar^3\omega^3}{6}+\hbar\omega e^2V^2\right)N(\hbar\omega)\nonumber\\ &&+\frac{\hbar^2\omega^2}{4}\sum_{\pm}(\hbar\omega\pm eV)N(\hbar\omega\pm eV)\bigg]~.\end{aligned}$$ A similar calculation leads to $$\begin{aligned} &&\mathcal{S}^{11}_{LR}(\omega) =-\frac{1}{h}\bigg[\left(\frac{2\hbar\omega\pi^2k_B^2T^2}{3}+\frac{\hbar^3\omega^3}{6}\right)N(\hbar\omega)\nonumber\\ &&+\frac{\hbar^2\omega^2}{4}\sum_{\pm}(\hbar\omega\pm eV)N(\hbar\omega\pm eV)\bigg]~.\end{aligned}$$ In addition, we have $\mathcal{S}^{11}_{RR}(\omega)=\mathcal{S}^{11}_{LL}(\omega)$ and $\mathcal{S}^{11}_{RL}(\omega)=\mathcal{S}^{11}_{LR}(\omega)$. [99]{} L. Saminadayar, D.C. Glattli, Y. Jin, and B. Etienne, Phys. Rev. Lett. [**79**]{}, 2526 (1997). R. de Picciotto, M. Reznikov, M. Heiblum, V. Umansky, G. Bunin, and D. Mahalu, Nature London [**389**]{}, 162 (1997). R. Deblock, E. Onac, L. Gurevich, and L. Kouwenhoven, Science [**301**]{}, 203 (2003). G. Schull, N. Néel, P. Johansson, and R. Berndt, Phys. Rev. Lett. [**102**]{}, 057401 (2009). R. Sánchez and M. Büttiker, Eur. Phys. Lett. [**100**]{}, 47008 (2012) ; [**104**]{}, 49901 (2013). F. Giazotto, T.T. Heikkilä, A. Luukanen, A.M. Savin, and J.P. Pekola, Rev. Mod. Phys. [**78**]{}, 217 (2006). R. Sánchez, B. Sothmann, A.N. Jordan, and M. Büttiker, New J. Phys. [**15**]{}, 125001 (2013). F. Battista,F. Haupt, and J. Splettstoesser, J. Phys.: Conf. Ser. [**568**]{} 052008 (2014). A. Crépieux and F. Michelini, J. Phys.: Condens. Matter [**27**]{}, 015302 (2015). A. Crépieux and F. Michelini, J. Stat. Mech. 054015 (2016). Y.M. Blanter and M. Büttiker, Phys. Rep. [**336**]{}, 1 (2000). T. Martin, in [*Les Houches Session LXXXI*]{}, edited by H. Bouchiat [*et al.*]{} (Elsevier, Amsterdam, 2005). R. Zamoum, M. Lavagna, and A. Crépieux, Phys. Rev. B [**93**]{}, 235449 (2016). D. Sergi, Phys. Rev. B [**83**]{}, 033401 (2011). D.V. Averin and J.P. Pekola, Phys. Rev. Lett. [**104**]{}, 220601 (2010); Europhys. Lett. [**96**]{}, 67004 (2011). B.K. Agarwalla, J.-H. Jiang, and D. Segal, Phys. Rev. B [**92**]{}, 245418 (2015). F. Battista, F. Haupt, and J. Splettstoesser, Phys. Rev. B [**90**]{}, 085418 (2014). Z. Yu, G.-M. Tang, and J. Wang, Phy. Rev. B [**93**]{}, 195419 (2016). F. Zhan, S. Denisov, P. Hänggi, Phys. Rev. B [**84**]{}, 195117 (2011); Physica Status Solidi (b) [**250**]{}, 2355 (2013). A. Crépieux, F. Simkovic, B. Cambon, and F. Michelini, Phys. Rev. B [**83**]{}, 153417 (2011); Phys. Rev. B [**89**]{}, 239907(E) (2014). A. Goker and B. Uyanik, Physics Letters A [**376**]{}, 2735 (2012). A. Goker and E. Gedik, J. Phys.: Condens. Matter [**25**]{}, 125301 (2013). A. Boehnke, M. Walter, N. Roschewsky, T. Eggebrecht, V. Drewello, K. Rott, M. Münzenberg, A. Thomas, and G. Reiss, Rev. Sci. Instrum. [**84**]{}, 063905 (2013). M. Bagheri Tagani and H. Rahimpour Soleimani, Int. J. Thermophys. [**35**]{}, 136 (2014). R. Chirla and C.P. Moca, Phys. Rev. B [**89**]{}, 045132 (2014). H. Zhou, J. Thingna, P. Hänggi, J.-S. Wang, and B. Li, Sci. Rep. [**5**]{}, 14870 (2015). A.-M. Daré and P. Lombardo, Phys. Rev. B [**93**]{}, 035303 (2016). M.F. Ludovico, F. Battista, F. von Oppen, and L. Arrachea, Phys. Rev. B [**93**]{}, 075136 (2016). H. Okada and Y. Utsumi, arXiv:1605.03809 (2016). G.B. Lesovik and R.  Loosen, Pisma Zh. Eksp. Teor. Fiz. [**65**]{}, 280 (1997) \[JETP Lett. 65, 295 (1997)\]. U. Gavish, Y. Levinson, and Y. Imry, Phys. Rev. B [**62**]{}, R10637(R) (2000). H.J.W. Haug and A.P. Jauho, *Quantum Kinetics in Transport and Optics of Semiconductors*, Springer Series in Solid-State Sciences 123, Second Substantially Revised Edition (2010). A.-P. Jauho, N.-S. Wingreen, and Y. Meir, Phys. Rev. B **50**, 5528 (1994). A. Crépieux and F. Michelini, Int. J. Nanotechnol. **9**, 355 (2012). J. Schwinger, J. Math. Phys. **2**, 407 (1961); L.V. Keldysh, Zh. Eksp. Teor. Fiz. **47**, 1515 (1964) \[Sov. Phys. JETP 20, 1018 (1965)\]. J. Hammer and W. Belzig, Phys. Rev. B [**84**]{}, 085419 (2011); Phys. Rev. B [**87**]{}, 125422 (2013). T. Andrade, S.A. Gentle, and B. Withers, J. High Energ. Phys. [**06**]{}, 134 (2016). J. Basset, H. Bouchiat, and R. Deblock, Phys. Rev. Lett. [**105**]{}, 166801 (2010). For an energy independent transmission amplitude, we have $\mathcal{S}^{00}_{\alpha\alpha}(\omega)=-\mathcal{S}^{00}_{\alpha\bar\alpha}(\omega)$, and $\mathcal{S}^{01}_{L\alpha}(\omega)=-\mathcal{S}^{01}_{R\alpha}(\omega)$. B. Roussel, P. Degiovanni, and I. Safi, Phys. Rev. B [**93**]{}, 045102 (2016). N.V. Gnezdilov, M. Diez, M.J. Pacholski, and C.W.J. Beenakker, Phys. Rev. B [**94**]{}, 115415 (2016). Note that at zero-temperatures, for positive voltage and in weak transmission limit, the expressions for the noises read as $\mathcal{S}^{pq}_{LR}(\omega)=e^{2-p-q}\mathcal{T}h^{-1}[(-1)^{1+p}|eV-\hbar\omega|^{1+p+q}\Theta(eV-\hbar\omega) +(-1)^{1+q}|eV+\hbar\omega|^{1+p+q}\Theta(-eV-\hbar\omega)]/(1+p+q)!$, with $\mathcal{S}^{00}_{LL}(\omega)=-\mathcal{S}^{00}_{LR}(\omega)$, $\mathcal{S}^{01}_{LL}(\omega)=\mathcal{S}^{01}_{LR}(\omega)$, and $\mathcal{S}^{11}_{LL}(\omega)=2\mathcal{S}^{11}_{LR}(\omega)+\mathcal{T}h^{-1}|\hbar\omega|^3\Theta(-\omega)$, where $\Theta$ is the Heaviside function. All the asymptotic values of noises given in the large temperature limit are calculated from Eqs. (\[exp\_noise\])-(\[C\]) taking $f_{\alpha,M}^{e,h}(\varepsilon)=1/2$. We have obtained: $\mathcal{S}^{00}_{\alpha\beta}(\omega)=e^2\pi\Gamma\delta_{\alpha\beta}/h$, $\mathcal{S}^{01}_{\alpha\beta}(\omega)=-e^2V\pi\Gamma\delta_{\alpha\beta}/h$, $\mathcal{S}^{11}_{\alpha\alpha}(\omega)=\hbar^2\omega^2\pi\Gamma/h$, and have checked numerically that $\mathcal{S}^{11}_{\alpha\bar\alpha}(\omega)$ varies linearly with temperature. G.D. Mahan, in *Many-Particle Physics*, Kluwer Academic/Plenum Publishers (2000).
{ "pile_set_name": "ArXiv" }
--- abstract: 'In this work, a shell model for metal clusters up to 220 valence electrons is used to obtain the fractional occupation probabilities of the electronic orbitals. Then, the calculation of a statistical measure of complexity and the Fisher-Shannon information is carried out. An increase of both magnitudes with the number of valence electrons is observed. The shell structure is reflected by the behavior of the statistical complexity. The magic numbers are indicated by the Fisher-Shannon information. So, as in the case of atomic nuclei, the study of statistical indicators also unveil the existence of magic numbers in metal clusters.' address: - | Departamento de Física, Facultad de Ciencias,\ Universidad de Extremadura, E-06071 Badajoz, Spain,\ and BIFI, Universidad de Zaragoza, E-50009 Zaragoza, Spain - | DIIS and BIFI, Facultad de Ciencias,\ Universidad de Zaragoza, E-50009 Zaragoza, Spain author: - Jaime Sañudo - 'Ricardo López-Ruiz' title: | Statistical measures applied to metal clusters:\ evidence of magic numbers --- and Statistical Indicators; Metal Clusters; Shell Structure; Magic Numbers Nowadays the calculation of information-theoretic measures on quantum systems has received a special attention [@gadre1987; @panos2005; @sanudo2008-]. The application of these indicators to atoms or nuclei reveals some properties of the hierarchical organization of these many-body systems [@panos2009; @sanudo2009]. In particular, entropic products such as Fisher-Shannon information and statistical complexity present two main characteristics when applied to the former systems. On one hand, they display an increasing trend with the number of particles, electrons or nucleons. On the other hand, they take extremal values on the closure of shells. Moreover, in the case of nuclei, the trace of magic numbers is displayed by these statistical magnitudes [@lopezruiz2009]. Metal clusters are useful quantum systems to understand how the physical properties evolve in the transition from atom to molecule to small particle to bulk solid [@meiwes2000; @broglia2004]. They also present a shell structure where it can applied the statistical indicators before mentioned. As in the case of atoms and nuclei, the fractional occupation probabilities of valence electrons in the different orbitals can capture the shell structure. This set of probabilities can be used to evaluate the statistical quantifiers for metallic clusters as a function of the number of valence electrons. Similar calculations have been reported for the electronic atomic structure [@panos2009; @sanudo2009] and for nuclei [@lopezruiz2009]. In this work, by following this method, we undertake the calculation of statistical complexity and Fisher-Shannon information for metal clusters. The jellium model provides an accurate description of some simple metal clusters. In this model, a valence electron is assumed to interact with the average potential generated by the other electrons and the ions [@meiwes2000; @broglia2004]. The confinement potential in the Schrödinger equation leading to shell structure is taken as a potential intermediate between the three-dimensional harmonic oscillator and the three-dimensional square well. This yields a filling of shells with a number $N$ of valence electrons given by the series: 2, 8, 18, 20, 34, 40, 58, 68, 70, 92, 106, 112, 138, 156, 166, 168, 198, 220 and so on. Each shell is given by $(nl)^w$, where $l$ denotes the orbital angular momentum ($l=0,1,2,\ldots$), $n$ counts the number of levels with that $l$ value, and $w$ is the number of valence electrons in the shell, $0\leq w\leq 2(2l+1)$. As an example, we explicitly give the shell configuration of a metal cluster formed by $N=58$ valence electrons. It is obtained: $$(N=58): (1s)^2(1p)^6(1d)^{10}(2s)^2(1f)^{14}(2p)^6(1g)^{18}\,.$$ The fractional occupation probability distribution of electron orbitals $\{p_k\}$, $k = 1,2,\ldots,\Pi$, being $\Pi$ the number of shells, can be defined in the same way as it has been done for calculations in atoms and nuclei [@panos2009; @sanudo2009; @lopezruiz2009]. This normalized probability distribution $\{p_k\}$ $(\sum p_k=1)$ is easily found by dividing the superscripts $w$ by the total number $N$ of electrons. Then, from this probability distribution, the different statistical magnitudes (Shannon entropy, disequilibrium, Fisher information, statistical complexity and Fisher-Shannon entropy) can be obtained. Here, we undertake the calculation of entropic products, a statistical measure of complexity $C$ and the Fisher-Shannon entropy $P$, that result from the product of two statistical quantities, one of them representing the information content of the system, and the other one giving an idea of how far the system is from the equilibrium. The classical indicator of information is the Shannon entropy, that in the discrete version is expressed as $$S = -\sum_{k=1}^{\Pi}p_k\log p_k \,.$$ Any monotonous function of $S$ can also be used for this purpose, such as the exponential Shannon entropy [@dembo1991], $H = e^{S}$ or $J = {1\over 2\pi e}\; e^{2S/3}$. The indicator of how much concentrated is the probability distribution of the system can be related with some kind of distance to the equilibrium distribution, that in our case is the equiprobability. The disequilibrium $D$ given by $$D = \sum_{k=1}^{\Pi}(p_k-1/\Pi)^2 \,,$$ and the Fisher information $I$ by $$I = \sum_{k=1}^{\Pi} {(p_{k+1}-p_k)^2\over p_k}\;,$$ where $p_{\Pi+1}=0$, are two useful parameters in this direction. Then, the statistical measure of complexity, $C$, the so-called LMC complexity [@lopez1995; @lopez2002], is defined as $$C = H\cdot D\;,$$ and the Fisher-Shannon information [@vignat2003; @romera2004; @szabo2008], $P$, is given by $$P = J\cdot I \;. \label{eq-p}$$ The statistical complexity, $C$, of metal clusters as a function of the number of valence electrons, $N$, is given in Fig. \[fig1\]. We can observe in this figure that this magnitude fluctuates around an slightly increasing average value $N$. This trend is also found for the electronic structure of atoms [@sanudo2009] and for the shell structure of nuclei [@lopezruiz2009], reinforcing the idea that in general complexity increases with the number of units forming a system. However, the shell model supposes that the system encounters certain ordered rearrangements for some specific number of units (electrons or nucleons) that coincide with closed shells. In the present case, this fact is reflected by the notable increase of $C$ in the metal clusters with one valence electron more than those with closed shells, which are indicated in Fig. \[fig1\], just as happens for atoms when one electron is added to noble gases or when one nucleon is added to a closed shell in nuclei. Observe that some major shells do not show local minima at their closing. This effect is due to the number of valence electrons belonging to each shell: a shell with a few valence electrons displays a local minimum of $C$ when is closed, but this is not the case when the number of valence electrons in a shell increases. The Fisher-Shannon entropy, $P$, of metal clusters as a function of $N$ is given in Fig. \[fig2\]. It presents an increasing trend with $N$. The spiky behavior of $C$ provoked by the shell structure is still present for $P$ but becomes smoother in this case. $P$ displays notable peaks only at a few $N$ related with the filling of some major shells, concretely at the numbers $2,8,18,34,58,92,138,198$. It must be remarked that, similarly as happens with $C$, the maximum values of $P$ are taken on the nuclei with one unit more than the former series, although now the difference is slightly appreciable. Only peaks at $20$ and $40$ disagree with the sequence of magic numbers {2, 8, 18, 20, 34, 40, 58, 92, 138, 198} obtained from experimental data, for instance, for $Na$ clusters [@martin1990] and for $Cs$ clusters [@pedersen1990; @lange1990]. Let us observe that the magic numbers are basically marked by the Fisher information such as it can be seen in Fig. \[fig3\]. In summary, the behavior of the statistical complexity $C$ and the Fisher-Shannon information $P$ with the number of valence electrons in metal clusters has been reported. The increasing trend of these magnitudes with the number of valence electrons, $N$, has been found. The method that uses the fractional occupation probabilities has been applied to calculate these statistical indicators. As in the case of atoms and nuclei, the shell structure is well displayed by the spiky behavior of $C$. On the other hand, $P$ shows an smoother behavior but with relevant peaks just on the major shells that coincide with the series of magic numbers in metal clusters. Therefore, the qualitative study of metal clusters by means of statistical indicators unveil certain physical properties of them. In fact, we can conclude that this type of statistical measures is able to enlighten some conformational aspects of quantum many-body systems. Acknowledgements {#acknowledgements .unnumbered} ================ The authors acknowledge some financial support from the Spanish project DGICYT-FIS2009-13364-C02-01. J.S. also thanks to the Consejería de Economía, Comercio e Innovación of the Junta de Extremadura (Spain) for financial support, Project Ref. GRU09011. [99]{} S.R. Gadre and R.D. Bendale, Phys. Rev. A [**36**]{} (1987) 1932. K.Ch. Chatzisavvas, Ch.C. Moustakidis, and C.P. Panos, J. Chem. Phys. [**123**]{} (2005) 174111. J. Sañudo and R. Lopez-Ruiz, Phys. Lett. A [**372**]{} (2008) 5283. C.P. Panos, N.S. Nikolaidis, K.Ch. Chatzisavvas, and C.C. Tsouros, Phys. Lett. A [**373**]{} (2009) 2343. J. Sañudo and R. Lopez-Ruiz, Phys. Lett. A, [**373**]{} (2009) 2549. R. Lopez-Ruiz and J. Sañudo, arXiv:0906.3148v1 (2009). K.H. Meiwes-Broer (Ed.), [*Metal Clusters at Surfaces: Structure, Quantum Properties, Physical Chemistry*]{}, Springer Series in Cluster Physics, Springer, Berlin, 2000. R.A. Broglia, G. Colo, G. Onida and H.E. Roman, [*Solid State Physics of Finite Systems: Metal Clusters, Fullerenes, Atomic Wires*]{}, Advances Texts in Physics, Springer, Berlin, 2004. A. Dembo, T.A. Cover, and J.A. Thomas, IEEE Trans. Inf. Theory [**37**]{} (1991) 1501. R. Lopez-Ruiz, H.L. Mancini, and X. Calbet, Phys. Lett. A [**209**]{} (1995) 321. R.G. Catalan, J. Garay, and R. Lopez-Ruiz, Phys. Rev. E [**66**]{} (2002) 011102. C. Vignat and J.F. Bercher, Phys. Lett. A [**312**]{} (2003) 27. E. Romera and J.S. Dehesa, J. Chem. Phys. [**120**]{} (2004) 8906. J.B. Szabo, K.D. Sen, and A. Nagy, Phys. Lett. A [**372**]{} (2008) 2428. T.P. Martin, T. Bergmann, H. Göhlich and T. Lange, Chem. Phys. Lett. [**172**]{} (1990) 209; Z. Phys. D[**19**]{} (1991) 25. S. Bjornholm, J. Borggreen, O. Echt, K. Hansen, J. Pedersen and H.D. Rasmussen, Phys. Rev. Lett. [**65**]{} (1990) 1627; Z. Phys. D[**19**]{} (1991) 47. H. Göhlich, T. Lange, T. Bergmann and T.P. Martin, Phys. Rev. Lett. [**65**]{} (1990) 748. ![Statistical complexity, $C$, vs. number of valence electrons, $N$. The arrows indicate the positions of closed shells.[]{data-label="fig1"}](C_expSD){width="12cm"} ![Fisher-Shannon entropy, $P$, vs. the number of valence electrons, $N$. The arrows indicate prominent closed shells that are magic numbers. For details, see the text.[]{data-label="fig2"}](P_IJ){width="12cm"} ![Fisher information, $I$, vs. the number of valence electrons, $N$. The arrows indicate prominent closed shells that are magic numbers.[]{data-label="fig3"}](I_Fisher){width="12cm"}
{ "pile_set_name": "ArXiv" }
--- abstract: 'We consider work performed over the last decade on single-$j$-shell studies. We will discuss four topics.' author: - 'A. Escuderos' - 'L. Zamick' title: Topics in nuclear structure --- A study of the E$(J_{\text{max}})$ interaction ============================================== This is an extension of work done with Apolodor Raduta and Elvira Moya de Guerra when they visited Rutgers on a NATO grant in 2001 [@mrzs03]. In contrast to work on $J=0$ pairing, we here study, in a single $j$ shell, an interaction $E(J_{\text{max}})$ for which all two-body matrix elements vanish except for $J=J_{\text{max}}=2j$. In this work we will focus on the model itself. Comparisons with realistic interactions are given in Ref. [@ze13] For a two-particle (or two-hole) system in a single $j$ shell, the even-$J$ states have isospin $T=1$ (triplets) and the odd-$J$ states have isospin $T=0$ (singlets). Thus, an interaction acting in only $J_{\text{odd}}$ states, including $J=J_{\text{max}}$, cannot occur for two neutrons or two protons in a single $j$ shell—only in the neutron–proton system. This leads to some simplifications. We will consider a system of two protons and two neutrons (or holes), e.g. $^{96}$Cd in the $g_{9/2}$ shell. The basis states are $| [pp(J_p) nn(J_n)]^{I} \rangle$, where $I$ is the total angular momentum. To satisfy the Pauli principle, $J_{p}$ and $J_{n}$ must be even. For $I=0$, the secular Hamiltonian is separable and $J_n=J_p$: $$H_{J_{p},J_{p'}}=V(J_{\text{odd}})f(J_{p})f(J_{p'})\,,$$ where $f(J_{p})$ is twice the unitary 9-$j$ symbol ($U9$-$j$): $$\begin{aligned} f(J_{p}) & = 2\langle(jj)^{J_{p}}(jj)^{J_{p}}|(jj)^{J_{\text{odd}}} (jj)^{J_{\text{odd}}}\rangle^{I=0}= \\ & = 2(2J_{p}+1)(2J_{\text{odd}}+1) \begin{Bmatrix} j & j & J_{p}\\ j & j & J_{p}\\ J_{\text{odd}} & J_{\text{odd}} & 0 \end{Bmatrix} \end{aligned}$$ If we write the wave function as $\sum X_{J_{p}J_{p}} |[pp(J_{p}) nn(J_{p})]^{I=0}\rangle$, then it was shown in Ref. [@mrzs03] that $X_{J_{p}J_{p}}$ is proportional to $f(J_{p})$. The other $I=0^{+}$ eigenstates are degenerate and, if $V(J_{\text{odd}})$ is negative, they are at higher energies. In other words, what we have shown in Ref. [@mrzs03] is that the wave-function components $X_{J_{p}J_{p}}$ of the lowest $I=0^{+}$ state are equal within a normalization to the $U9$-$j$ coefficients: $\sqrt{2} \langle (jj)^{J_{p}} (jj)^{J_{n}} | (jj)^{J_{\text{max}}} (jj)^{J_{\text{max}}} \rangle^{I=0}$. The eigenvalue is given by $$E(I=0^{+})=V(J_{\text{odd}})|\sum f(J_{p})X_{J_{p}J_{p}}|^{2}$$ Note that our very simple interactions are charge independent. This means that the lowest (non-degenerate) $I=0^{+}$ state has good isospin, presumably $T=0$. It is amusing that we can assign the isospin quantum number to a wave function with $U9$-$j$ coefficients. Similarly, we can show that there is also a very simple expression for the $I=1^{+}$ lowest eigenfunction: $2 \langle (jj)^{J_{p}} (jj)^{J_{n}} | (jj)^{J_{\text{max}}} (jj)^{J_{\text{max}}-1} \rangle^{I=1}$. However, for states of angular momentum 2 or higher, the secular Hamiltonian is no longer separable. The eigenvalue equation is $$\begin{gathered} \label{eq:i2ev} 4\sum_{J_{x}}\langle(jj)^{J_{p}}(jj)^{J_{n}}|(jj)^{9}(jj)^{J_{x}}\rangle^{I} \times \\ \times \sum_{J_{p'}J_{n'}}\langle(jj)^{J_{p'}}(jj)^{J_{n'}}| (jj)^{9}(jj)^{J_{x}}\rangle^{I} X_{J_{p'}J_{n'}} = \\ = \lambda X_{J_{p} J_{n}} \,,\end{gathered}$$ where $\lambda$ is the eigenvalue and $X_{J_{p} J_{n}}$ stands for the eigenfunction components. For $I=2^{+}$ there are two terms corresponding to $J_{x}=7$ and 9; for $I=3^{+}$ the values are $J_{x}=6$ and 8, etc. Despite the complexity of the above equation, there are some surprising results. The eigenfunction components of the lowest $2^{+}$ state are numerically extraordinarily close to the single $U9$-$j$ symbols $\sqrt{2}\langle(jj)^{J_{p}}(jj)^{J_{n}}|(jj)^{9}(jj)^{9}\rangle^{I=2}$. Furthermore, the next $2^{+}$ state has also components exceedingly close to $2\langle(jj)^{J_{p}}(jj)^{J_{n}}|(jj)^{9}(jj)^{7}\rangle^{I=2}$. This is by no means obvious because, as mentioned above, the interaction involves a sum of two separable terms corresponding to $J_{x}=7$ and 9. We can explain this result by performing a calculation of the overlap of the two $U9$-$j$’s of the last paragraph. We restrict the sum to even $J_{p}$ and even $J_{n}$. We first note schematically $$\begin{gathered} \label{eq:sch} 4\sum_{\text{even }J_{p},J_{n}} = \sum(1+(-1)^{J_{p}})(1+(-1)^{J_{n}})= \\ = \sum+\sum(-1)^{J_{p}}+\sum(-1)^{J_{n}}+\sum(-1)^{J_{p}+J_{n}}\end{gathered}$$ Using orthogonality relations for 9-$j$ symbols, we can see that the first term vanishes. In the last term, one of the $U9$-$j$’s has two rows that are the same, which means that the only non-vanishing terms in the sum have ($J_{p}+J_{n}$) even. Thus, the last term is the same as the first—zero. The two middle terms are the same, so we get the overlap of the two $U9$-$j$’s to be $$\begin{aligned} \mathop{\sum_{\text{even }}}_{J_{p},J_{n}} = & \frac{1}{2} \mathop{\sum_{\text{even }}}_{J_{p},J_{n}} (-1)^{J_{p}}\langle(jj)^{J_{p}} (jj)^{J_{n}}|(jj)^{9}|(jj)^{9}\rangle^{I=2}\times \\ & \times\langle(jj)^{J_{p}}(jj)^{J_{n}}|(jj)^{9}(jj)^{7}\rangle^{I=2}= \\ = & -\frac{1}{2}\langle(jj)^{9}(jj)^{9}|(jj)^{9}(jj)^{7}\rangle^{I=2}\,. \label{eq:i2}\end{aligned}$$ We obtain the above by using again orthogonality relations for $9j$-symbols. Using similar arguments, one can show that the normalization for the $|[pp(9)nn(9)]^{I=2}\rangle$ state is such that its normalization factor is $$\label{eq:n2-99} \begin{aligned} N(9)^{-2} & = \frac{1}{2}-\frac{1}{2}\langle(jj)^{9}(jj)^{9}| (jj)^{9}(jj)^{9}\rangle^{I=2}=\\ & = \frac{1}{2}-\frac{1}{2}0.00001209813=0.499993950935 \end{aligned}$$ For the $|[pp(9)nn(7)]^{I=2}\rangle$ state, we obtain $$\label{eq:n2-97} \begin{aligned} N(7)^{-2} & = \frac{1}{4}-\frac{1}{2}\langle(jj)^{9}(jj)^{7}| (jj)^{9}(jj)^{7}\rangle^{I=2}= \\ & = \frac{1}{4}+\frac{1}{2}0.00075253477=0.250376267385 \end{aligned}$$ To get this latter result, we use the following relationship $$\sum_{J_p,J_n}(-1)^{(J_{p}+J_{n})}\left|\langle(jj)^{9}(jj)^{7}| (jj)^{J_{p}}(jj)^{J_{n}}\rangle^{I=2}\right|^{2}=0$$ From Eqs.  and , we find that the normalizations are 1.414222 and 1.998497, only slightly different from $\sqrt{2}$ and 2 respectively. Therefore, we obtain that the term in Eq.  is exceedingly small for the $g_{9/2}$ shell, namely 0.00009113 and, if we include the exact normalization factors, we get 0.00025756. We can see in Table \[tab:j2\] that the results for matrix diagonalization for both $I=2^{+}$ states yield wave function components which are very close to the normalized $U9$-$j$ coefficients. In fact, they are so close that one could wonder if they are exactly the same. But they are not. As seen in Eq. , the two $U9$-$j$ sets corresponding to $J_x=9$ and $J_x=7$ are very nearly orthogonal, but not quite. [|c|c|c|c|c|]{} $[J_{p}\,,J_{n}]$ & E(9) & $U9$-$j$ & E(9) & $U9$-$j$ & 1.069 & & 3.0558 & $[0\,,2]$ & 0.5334 & 0.5338 & 0.1349 & 0.1351 $[2\,,2]$ & $-0.4707$ & $-0.4708$ & 0.5569 & 0.5567 $[2\,,4]$ & 0.3035 & 0.3035 & 0.3188 & 0.3189 $[4\,,4]$ & $-0.1388$ & $-0.1390$ & 0.6300 & 0.6299 $[4\,,6]$ & 0.0531 & 0.0531 & 0.1320 & 0.1320 $[6\,,6]$ & $-0.0137$ & $-0.0138$ & 0.1350 & 0.1350 $[6\,,8]$ & 0.0025 & 0.0025 & 0.0114 & 0.0114 $[8\,,8]$ & $-0.0003$ & $-0.0003$ & 0.0052 & 0.0052 \[tab:j2\] It turns out that all the other lowest even-$I$ states have eigenfunctions close although not exactly equal to $\sqrt{2}\langle(jj)^{J_{p}}(jj)^{J_{n}}| (jj)^{9}(jj)^{9}\rangle^{I}$. In Table \[tab:9j\] we compare, as an example, the wave function of the $I=8^{+}$ state. In the second column, we give the single $U9$-$j$ symbols (normalized) and in the third column we give results of diagonalizing the E(9) interaction. Since the coefficient $[J_{p},J_{n}]$ is the same as $[J_{n},J_{p}]$, we list only one of them. The overlap of the two wave funtions is 0.9998. [|c|c|c|]{} $[J_{p}\,,J_{n}]$ & $U9$-$j$ & E(9) $[0\,,8]$ & 0.0630 & 0.0644 $[2\,,6]$ & 0.4299 & 0.4271 $[2\,,8]$ & $-0.0522$ & $-0.0513$ $[4\,,4]$ & 0.7444 & 0.7456 $[4\,,6]$ & $-0.1803$ & $-0.1729$ $[4\,,8]$ & 0.0256 & 0.0280 $[6\,,6]$ & 0.0521 & 0.0657 $[6\,,8]$ & $-0.0076$ & $-0.0012$ $[8\,,8]$ & 0.0011 & 0.0047 \[tab:9j\] We can plot the coupling of Eq.  for various shells, as we can see in Table \[tab:u9j-sh\]. We see that the coupling $U9$-$j$ decreases at least exponentially as we go up in shell. Experts say that this type of 9-$j$ lies in the non-classical region. [|c|c|c|]{} $j$ & $U9$-$j$ & overlap of Eq.  $p_{3/2}$ & $-0.1800$ & 0.2546 $d_{5/2}$ & $-0.021328$ & 0.03016 $f_{7/2}$ & $-0.002074$ & 0.002933 $g_{9/2}$ & $-0.0001822$ & 0.0002577 $h_{11/2}$ & $-0.00001502$ & 0.00002174 $i_{13/2}$ & $-0.000001185$ & 0.000001676 \[tab:u9j-sh\] Degeneracies ------------ With the E$(J_{\text{max}})$ interaction for $g_{9/2}$ \[that is, E(9)\], we get several degenerate states with an absolute energy zero. In some detail, for $I=0^{+}$ there are five states, three with isospin $T=0$ and two with $T=2$. There is one non-degenerate state at an energy $2V(9)$ ($V(9)$ is negative). The other four $I=0^{+}$ states have zero energy. For $I=1^{+}$ all states have isospin $T=1$. There is a single non-degenerate state at $V(9)$, the other three have zero energy. For $I=2^{+}$ there are twelve states—six have $T=0$, four have $T=1$, and two have $T=2$. There are two non-degenerate $T=0$ states with approximate energies $2V(9)$ and $V(9)$ respectively, and one non-degenerate $T=1$ state with energy $V(9)$. The other nine states have zero energy. To understand this, take a wave function $$|\Psi^{\alpha}\rangle = \sum_{J_p, J_n}{C^{\alpha}(J_{p},J_{n})| [pp(J_{p}) nn(J_{n})]^{I}\rangle}$$ and the corresponding energies $E^{\alpha}=\langle\Psi^{\alpha} |H|\Psi^{\alpha}\rangle$. Consider the sum $\sum_{\alpha}{E^{\alpha}}$. We have $$\sum_{\alpha}{C^{\alpha}(J_{p},J_{n})C^{\alpha}(J_{p'},J_{n'})}= \delta_{J_{p},J_{p'}}\delta_{J_{n},J_{n'}}$$ Thus $$\label{eq:sume} \begin{aligned} \sum_{\alpha}{E^{\alpha}} & = \sum_{J_{p}J_{n}}{\langle[pp(J_{p}) nn(J_{n})]^{I}| H|[pp(J_{p}) nn(J_{n})]^{I}\rangle} = \\ & = 4V(9)\mathop{\sum_{J_{p}J_{n}}}_{\text{even}}{\sum_{J_{x}}{\left| \langle(jj)^{J_{p}}(jj)^{J_{n}}|(jj)^{9}(jj)^{J_{x}}\rangle^{I} \right|^{2}}} \end{aligned}$$ This expression does not depend on the detailed wave functions. Referring to Eqs.  and and neglecting the very small correction terms, we see that $N^{-2}$ is equal to $1/2$ for $J_{x}=9$ and to $1/4$ for all other $J_{x}$. Basically then Eq.  becomes $4 V(9) \sum N(J_{x})^{-2}$. Hence we obtain $\sum_{\alpha}{E^{\alpha}}=2V(9)$ for $I=0$, $V(9)$ for $I=1$, and $4V(9)$ for $I=2$. But we can alternately show, using the explicit wave functions, that for $I=0$ the energy of the lowest state is $2V(9)$. Hence, all the other states must have zero energy. A similar story for $I=1$. The $I=2$ state is a bit more complicated because of the coupling between two states, however small it is. Still one can work it through and see that the $4V(9)$ energy is exhausted by the two $T=0$ and the one $T=1$ non-degenerate states. For $I=0$ we have two $T=0$ and two $T=2$ states, all degenerate. One can remove the degeneracies of $T=0$ and $T=2$ by adding to the Hamiltonian an interaction $b\; t(i)\cdot t(j)$. This will not affect the wave functions of the non-degenerate states but will shift the $T=2$ states away from the formerly degenerate $T=0$ states. The number of neutron-proton pairs with even $J$ for a system of two neutrons and two protons is given by [@z07b] $$\sum_{J_{a}} | D^{I}(J_{a} J)| ^{2} (\delta_{T,0} + 4 \delta_{T,2}) ~,$$ where $D^{I}(J_{P} J_{N})$ is the probability amplitude that in a state of total angular momentum $I$ , the protons couple to $J_{P}$ and the neutrons to $J_{N}$. How to handle degeneracies—pairing and $Q\cdot Q$ ================================================= We show how degeneracies, accidental or otherwise, can obscure some interesting physics. But we further show how one can get around this problem. In a 2006 publication, Escuderos and Zamick [@ez06] found some interesting behaviour in the $g_{9/2}$ shell. Unlike the lower shells, e.g. $f_{7/2}$, seniority is not a good quantum number in the $g_{9/2}$ shell. Despite this, it was found that in a matrix diagonalization with four identical particles in the $g_{9/2}$ shell with total angular momentum $I=4$ or 6, one unique state emerged no matter what interaction was used. This problem was also addressed by others. Before the mixing, one has two states with seniority $v=4$ and one with $v=2$. The surprise was that, after the diagonalization, one gets a unique state that is always the same independently of the interaction used. This unique state has seniority $v=4$. The components of the wave function are given in the fourth column of Table \[tab:res\] (labelled “$T=2,$ $v=4$ unique"). The problem to be dealt with was not only why this state did not mix with the $v=2$ state, but also why it does not mix with the other $v=4$ state. But this will not concern us here. Rather we will use this as an example of how degeneracies can obscure interesting physics. We first consider how the unique $T=2$ wave function looks like for a system of three protons and one neutron. This is shown in Table \[tab:uni\]. No matter what interaction is used, this appears as a unique state. [|l|c|]{} $J_{0}$ & $(j^{3} J_{0} j |\} I=4, v=4)$ 3/2 & 0.1222 5/2 & 0.0548 7/2 & 0.6170 9/2 ($v=1$) & 0.0000 9/2 ($v=3$) & 0.0000 11/2 & $-0.4043$ 13/2 & $-0.6148$ 15/2 & $-0.1597$ 17/2 & 0.1853 \[tab:uni\] It was already commented on in a later paper by Zamick [@z07] that cfp’s for identical particles are usually calculated using a pairing interaction. With such an interaction, the two $v=4$ states are degenerate, i.e. they have the same energy. This means that any linear combination of the two states can emerge in a matrix diagonalization. Thus, the emergence of a unique state gets completely lost. The problem was also addressed in Refs. [@zv08; @q11; @ih08]. In this work we consider a less obvious example: a matrix diagonalization of two proton holes and two neutron holes in the $g_{9/2}$ shell, i.e. we consider $^{96}$Cd rather than $^{96}$Pd, the latter consisting of four proton holes (whether we consider holes or particles does not matter). We use a quadrupole–quadrupole interaction $Q\cdot Q$ for the matrix diagonalization. The two-body matrix elements in units of MeV from $J=0$ to $J=9$ are: $-1.0000$, $-0.8788$, $-0.6516$, $-0.3465$, $-0.0152$, 0.2879, 0.4849, 0.4849, 0.1818, and $-0.5454$. We show some relevant results in Table \[tab:res\]. For $I=4$, we get 14 eigenfunctions, but we list only two of them in the first two columns. The reason we single these out is that they are degenerate—both are at an excitation energy of 3.5284 MeV. In the third wave function column, we have the unique state, one that emerges, as we said above, with any interaction, however complicated, e.g. CCGI [@ccgi12]. But now we have to modify the phrase “any interaction”. We do not see this unique state when we use the $Q\cdot Q$ interaction—none of the 14 states looks like the one in column 3. Learning from our experience with the pairing interaction, we suspect that the problem lies with the two degenerate states at 3.5284 MeV. We assumed that the two states were mixtures of one $T=0$ and one $T=2$ state. [|c|c|c|c|c|c|]{} & Mix $T=0,2$ & Mix $T=0,2$ & $T=2$, $v=4$ unique & $T=0$ untangled & other $T=2$, $v=4$ & 3.5284 & 3.5284 & 6.5285 & 3.5284 & $[J_{p}\,,J_{n}]$ & & & & & $[0\,,4]$ & 0.0000 & 0.0000 & 0.0000 & 0.0000 & 0.0000 $[2\,,2]$ & $-0.3250$ & $-0.4170$ & $-0.4270$ & 0.3123 & $-0.0255$ $[2\,,4]$ & $-0.2364$ & $-0.2472$ & $-0.2542$ & 0.2289 & $-0.1986$ $[2\,,6]$ & 0.2168 & 0.3043 & 0.3107 & $-0.2076$ & $-0.1976$ $[4\,,4]$ & 0.0207 & 0.2390 & 0.2395 & $-0.0135$ & $-0.3313$ $[4\,,6]$ & $-0.1826$ & $-0.1364$ & $-0.1418$ & 0.1784 & 0.2245 $[4\,,8]$ & 0.0934 & 0.1540 & 0.1567 & $-0.0888$ & 0.3874 $[6\,,6]$ & $-0.1312$ & 0.1678 & 0.1638 & 0.1362 & 0.5645 $[6\,,8]$ & $-0.1343$ & 0.0357 & 0.0316 & 0.1353 & 0.0247 $[8\,,8]$ & $-0.7421$ & 0.5881 & 0.5625 & 0.7594 & $-0.1087$ \[tab:res\] We can remove the degeneracy without altering the wave functions of the non-degenerate states by adding a $t(1)\cdot t(2)$ interaction to the Hamiltonian. This will shift energies of states of different isospin. What we actually did was equivalent to this. We added $-1.000$ MeV to the two-body $T=0$ matrix elements. These had odd spin $J=1,3,5,7,9$. What emerged is shown in columns 3 and 4. The degeneracy is removed. We have a $T=2$ state in the third column shifted up by 3 MeV and in the fourth column a $T=0$ state unshifted. The wave function components are different from what they are in the first two columns. The $T=2$ state is the unique state we were talking about—one that emerges with any interaction, e.g. CCGI or delta. It is the double analog of a state of four identical proton holes ($^{96}$Pd). The $T=0$ state in the fourth column has an interesting structure with vanishing components for $[0\,,4]$ . However, when interactions other than $Q\cdot Q$ are used, it is no longer an eigenfunction. In the last column, we list the other $T=2$, $v=4$ state. One sees this on the list when one uses a seniority-conserving interaction such as a delta interaction. However, for a general interaction, it does not appear. This is because it gets mixed with the $T=2$, $v=2$ state. Only the state in the third column remains unscathed when we turn on some arbitrary interaction—and only that state does not end up being degenerate with some other state. There is also a unique $J=6^{+}$, $v=4$ state. With the pairing interaction, this is degenerate with another $J=6^{+}$, $v=4$ state and so the uniqueness gets obscured. However, with the $Q\cdot Q$ interaction, unlike the case for $J=4^{+}$, the unique $J=6^{+}$, $v=4$ state is not degenerate with another state. Hence even with $Q\cdot Q$ this unique state appears in the calculation. There are other examples of confusions. The electric dipole moment of the neutron would vanish if parity conservation holds. But at a more important level, it vanishes if time reversal invariance holds. The first $J=1^+$, $T=0$ states in a single-$j$-shell configuration in even–even nuclei. ======================================================================================== The first even–even nucleus for which there are $J=1^+$, $T=0$ states in a single-$j$-shell configuration is $^{48}$Cr. If we limit ourselves to single $j$, there are no $M1$ transitions from these states to any $J=0^+$, $T=0$ states. Absence of $J=1^{+}$, $T=2$ states in $j^{4}$ configurations ------------------------------------------------------------ In early shell model calculations by McCullen et al. [@mbz64] and Ginocchio and French [@gf63], it was noted that, in the $f_{7/2}$ shell, certain combinations of spin and isopin did not exist. For example, there were no $J=0^+$, $T=1$ states in $^{44}$Ti and no $J=1^{+}$ states with $T=T_{\text{min}}+1$, where $T_{\text{min}}= |N-Z|/2$. There were also no $J=1^{+}$ states with $T=T_{\text{max}}$. However, those states are analogous to states of a system of identical particles, i.e. calcium isotopes. Explanations for some of the missing states can be shown by simple techniques as will be discussed later. In the mid eighties, papers were published which counted the states in a more systematic way. They include the works of I. Talmi on recursion relations for counting the states of identical fermions [@t05] and by Zhao and Arima [@za05], who obtained expressions for the number of $T=0$, 1 and 2 states for protons and neutrons in a single $j$ shell. Indeed the latter authors give all the answers to the counting questions addressed in this paper. The first occurrence of $J=1^{+}$, $T=0$ states in the single $j$ shell—$^{48}$Cr --------------------------------------------------------------------------------- There are no $J=1^{+}$, $T=0$ states for four nucleons in the single $j$ shell. To make things more concrete, consider $^{44}$Ti. The two $f_{7/2}$ protons can have angular momenta 0, 2, 4, and 6, all occurring once; likewise the two neutrons. The \[$J_{p}$, $J_{n}$\] configurations that can add up to a total $J=1$ are \[2, 2\], \[4, 4\] and \[6, 6\]. Thus, there are three $J=1^{+}$ states. The possible isospins are 0, 1, and 2. Let us next consider $^{44}$Sc. The three neutrons can have angular momenta 3/2, 5/2, 7/2, 9/2, 11/2, and 15/2, all occurring only once. The states that add up to one are \[7/2, 5/2\], \[7/2, 7/2\], and \[7/2, 9/2\]. Again we have three states. However, since $^{44}$Sc has $|T_{z}|=1$, the isospins can only be one or two. Hence, there are no $J=1^{+}$, $T=0$ states in $^{44}$Ti which are of the $(f_{7/2})^{4}$ configuration. We next consider $^{48}$Cr. The possible states of four protons, including seniority labels are: $v=0$, $J=0$; $v=2$, $J=2,4,6$; $v=4$, $J=2^{*},4^{*},5,8$. The possible $J=1^{+}$ states are \[2, 2\], \[4, 4\], \[6, 6\], \[2$^{*}$, 2$^{*}$\], \[4$^{*}$, 4$^{*}$\], \[5, 5\], \[8, 8\], \[2, 2$^{*}$\], \[2$^{*}$, 2\], \[4, 4$^{*}$\], \[4$^{*}$, 4\], \[4, 5\], \[5, 4\], \[4$^{*}$, 5\], \[5, 4$^{*}$\], \[5, 6\], \[6, 5\]. There are 17 such states with a priori possible isospins $T=0,1,2,3,4$. We next consider $^{48}$V, which consists of three protons and five neutrons. The latter can also be regarded as three neutron holes, so the possible states are the same for neutrons and protons. The three proton states are 3/2, 5/2, 7/2, 9/2, 11/2, and 15/2, all occurring only once. The possible $J=1^{+}$ states are \[3/2, 3/2\], \[5/2, 5/2\], \[7/2, 7/2\], \[9/2, 9/2\], \[11/2, 11/2\], \[15/2, 15/2\], \[3/2, 5/2\], \[5/2, 3/2\], \[5/2, 7/2\], \[7/2, 5/2\], \[7/2, 9/2\], \[9/2, 7/2\], \[9/2, 11/2\], \[11/2, 9/2\]. There are 14 such states and they all must have isospins greater than zero. Hence, the number of $J=1^{+}$, $T=0$ states of the $(f_{7/2})^{8}$ configuration is $(17-14)= 3$. The wave functions of these states are included in a larger compilation by A. Escuderos, L. Zamick, and B.F. Bayman  [@ezb05]. It is there noted that because both protons and neutrons are at mid shell, the quantitiy $s=(-1)^{v}$ is a good quantum number, where $v=(v_p+v_n)/2$. Referring to Ref. [@mbz64] for $J=0^{+}$, $T=0$, there are four states with $s=+1$ and two with $s=-1$. All $J=0^{+}$, $T=1$ states have $s=-1$, while all $T=2$ and $T=4$ states have $s=+1$. There are two $J=1^{+}$, $T=0$ states with $s=-1$ at energies of 7.775 and 9.258 MeV. There is one $s=+1$ state at 9.037 MeV with a rather simple wave function: $1/\sqrt{2} [(4^{*},5) + (5,4^{*})]$. The lowest $J=0^{+}$, $T=0$ state has $s= +1$. $M1$ selection rules -------------------- There is a modern twist to what we are here doing. There has been an extensive review of $M1$ excitations, including spin-flip modes, scissors modes etc., by K. Heyde, P. Von Neumann-Cosel, and A. Richter  [@hnr10]. The mode we are here considering has, to the best of our knowledge, not yet been studied experimentally. There have been studies of $M1$ $T=0 \to T=0$ transitions, e.g. the electro-excitation of $J=1^{+}$, $T=0$ excited states of $^{12}$C, but these involve more than one shell. Isospin impurities are very important for these transitions because the isovector $M1$ coupling constants are much larger than the isoscalar ones. One simple selection rule for $M1$ transitions in this limited model space is that $M1$ $(T=0\rightarrow T=0)$ equals zero. To see this, we note that in the single-$j$-shell space we can replace the $M1$ operator by $g_jJ$. The $M1$ matrix element for a $T=0 \to T=0$ transition is thus proportional to $(g_{j\pi} + g_{j\nu})$, i.e. the isoscalar sum. But if such a term is non-zero, it would imply that the total angular momentum operator $J$ (obtained by setting the two $g$’s above each equal to 1/2) could induce an $M1$ transition, which, of course, it cannot. Another “midshell” selection rule is that the quantum number $s$ has to be the same for the initial $J=1^{+}$, $T=0$ state and for any final state, e.g $J=1^{+}$, $T=1$ or $J=2^{+}$, $T=1$. Although not necessary, it is nevertheless instructive to show in more detail why the $T=0 \to T=0$ matrix element vanishes. Consider a transition from $s=-1$ to $s=-1$. In the wave functions, there will be no amplitude of the configuration $(J_p,J_n) =(2,2)$, but there will be of $(2,2^{*})$ and $(2^{*},2)$. The transition matrix element will have the form $\langle (2,2^{*})^{2} +(2^{*},2)^{2}|| M1 || (2,2^{*})^{1}- (2^{*},2)^{1}\rangle$. This is equal to $\langle (2,2^{*})^{2}||M1|| (2,2^{*})^{1}\rangle - \langle (2^{*},2)^{2}||M1|| (2^{*},2)^{1}\rangle$. Since in the single $j$ shell one can replace $M1$ by $g_j J$, the matrix element $\langle 2||M1||2\rangle$ is equal to $\langle 2^{*}|| M1 ||2^{*}\rangle$. We thus see that the complete matrix element vanishes. Selected Systematics of Odd–Odd Nuclear Spectra =============================================== We consider $T=1$ states of four nucleons with three partices (holes) of one kind and one of the other kind, e.g $^{44}$Sc (three neutrons and one proton) or $^{96}$Ag (three proton holes and one neutron hole). We formulated a $(2j-1)$ rule which will here be presented in a somewhat different way than in Ref. [@ze13]. The rule is that, for these systems, yrast states with angular momenta $I=(2j-1)$ lie lower in energy than neighbouring states with angular momenta $(2j-1)-1$ or $(2j-1)+1$. We give some examples in Table \[tab:2j-1\], with energies in MeV. $I$ Exp. Theory ----------------------- ----- ------- -------- $^{44}$Sc 5 1.513 1.276 6 0.271 0.381 7 0.968 1.272 $^{52}$Mn 5 1.254 1.404 6 0.000 0.000 7 0.870 1.819 $^{96}$Ag 7 ????? 0.861 8 0.000 0.000 9 0.470 0.492 $h_{11/2}$ $Q\cdot Q$ 9 1.30 10 0.21 11 0.85 : Examples of the $(2j-1)$ mentioned in the text. The energies are given in MeV. \[tab:2j-1\] A possible explanation of this rule for nuclei at the end of a closed shell is that for such nuclei the value of the rotational quantum number $K$ is equal to $(2j-1)$. In more detail, the neutron hole has $k_{1}=j$ and the three proton holes have $k_{2}=j-1$, so that $K=k_{1}+k_{2}=(2j-1)$. In Ref. [@ze13], it was noted that it is hard to get two-body matrix elements from experiment. One can get $T=1$ matrix elements from the spectrum of $^{98}$Cd, but the spectrum of the two-hole nucleus $^{98}$In is not known, so we cannot get the $T=0$ two-body matrix elements in a simple direct way. Sorlin and Porquet [@sp08] discussed using a Pandya transformation to get the particle-particle spectrum from the particle-hole spectrum of $^{90}$Nb. This is a priori a reasonable thing to try. They used as input the yrast spectrum of $^{90}$Nb, except for $J=1^{+}$, where they used the second excited $1^{+}$ state. We reproduced the results in [@ze13]. We find that for $^{96}$Cd this method gives a significantly lower excitaton energy for $J=16^{+}$ in $^{96}$Cd than does a realistic CCGI interaction [@ccgi12]: 3.898 MeV vs 5.245 MeV. This may be due to the increasing collectivity as one moves away from the $N=50$ $Z=50$ closed shell. E. Moya de Guerra, A.A. Raduta, L. Zamick, and P. Sarriguren, Nucl. Phys. A **727**, 3 (2003). L. Zamick and A. Escuderos, Phys. Rev. C **87**, 044302 (2013). L. Zamick, Phys. Rev. C **75**, 024307 (2007). A. Escuderos and L. Zamick, Phys. Rev. C **73**, 044302 (2006). L. Zamick, Phys. Rev. C **75**, 064305 (2007). L. Zamick and P. Van Isacker, Phys. Rev. C **78**, 044327 (2008). Chong Qi, Phys. Rev. C **83**, 014307 (2011). P. Van Isacker and S. Heinze, Phys. Rev. Lett. **100**, 052501 (2008). L. Coraggio, A. Covello, A. Gargano, and N. Itaco, Phys. Rev. C **85**, 034335 (2012). J.D. McCullen, B.F. Bayman, and L. Zamick, Phys. Rev. **134**, B515 1964). J.N. Ginocchio and J.B. French, Phys. Lett. **7**, 137 (1963). I. Talmi, Phys. Rev. C **72**, 037302 (2005). Y.M. Zhao and A. Arima, Phys. Rev. C **71**, 047304 (2005); Phys. Rev. C **72**, 064333 (2005). A. Escuderos, L. Zamick, and B.F. Bayman, arXiv:nucl-th/0506050 (2005). K. Heyde, P. von Neumann-Cosel, and A. Richter, Rev. Mod. Phys. **82**, 2365 (2010). O. Sorlin and M.-G. Porquet, Prog. Part. Nucl. Phys. **61**, 602 (2008).
{ "pile_set_name": "ArXiv" }
--- abstract: 'To understand the foundations of quantum mechanics, we have to think carefully about how theoretical concepts are rooted in — and limited by — the nature of experience, as Bohr attempted to show. Geometrical pictures of physical phenomena are favored because of their clarity. Quantum phenomena, however, do not permit them. Instead, the historical and dynamical aspects of description diverge and must be expressed in different but complementary languages. Objective historical facts are recorded in terms of objects, which necessarily have an imprecise, empirical quality. Dynamics is based on quantitative abstraction from recurring patterns. The “quantum of action” is the discontinuity between these two ways of looking at the physical world.' address: - Indiana University Cyclotron Facility - '2401 Milo B. Sampson Lane' - 'Bloomington, IN 47408 USA' author: - 'Doug Bilodeau[^1]' title: Why Quantum Mechanics is Hard to Understand --- ${\left(} \def$[)]{} $${\left[} \def$$[\]]{} §[[S]{}]{} .3in INTRODUCTION ============ There was a time when the Copenhagen Interpretation was commonly thought to have solved the conceptual problems of quantum mechanics. That view is much less fashionable now, but we do not appear to be approaching a consensus on any alternative. Most papers written on this subject start with the mathematical formalism (its predictions so thoroughly validated by experiment) and try to recast it into some new form with a natural and unambiguous physical meaning. These efforts seem to me to take us in the wrong direction — deeper into abstraction (where it is easy to mislead ourselves) and farther from the immediate empirical basis of physical concepts, which is where the difficulty lies. I think Bohr was right, or nearly so. Unfortunately, his writings are so obscure that there is no more unanimity on “what Bohr really thought” than there is on what quantum mechanics is all about. I have tried here to follow what I take to be the spirit of Bohr’s analysis, but in language which will be clearer to physicists today. But not very clear — quantum mechanics is intrinsically hard, because it pushes us to the limits of the very idea of the physical. Some linguists claim that the capacity for language is part of our genetic makeup. Even the grammatical structure of language is to some extent biologically determined, e.g., the distinction between nouns and verbs [@pinker]. Perhaps our brains also favor certain kinds of mathematical and physical concepts. The machines we design and the theorems we prove reflect the hard-wired structure of our thinking. Whether this is true or not, it is true that quantum theory startles us by distinguishing two aspects of the way we think about the physical world — two aspects we are used to describing in a single unified account: what a thing is and what a thing does. The result is that we cannot form a single consistent picture of physical reality at the quantum level based on geometrical structures in physical space and time. When Einstein wrote his first paper on special relativity, he used mathematics already well known in contemporary research on electrodynamics. His profound and original contribution was to show how physical concepts are based on the nature and limitations of observations, and how intuitively appealing concepts like simultaneity are abstractions which may or may not apply to the real world. To understand quantum mechanics, we also have to wrestle with physical concepts. In this brief essay, I cannot treat any particular examples in detail, but I have tried to point out the crucial features of physical analysis which take on new meanings as we go from classical to quantum physics. Needless to say, my discussion here is neither as original nor as revolutionary as Einstein’s was. I have for the most part restated very old ideas in slightly new language, but hopefully new and fresh enough to encourage others to see Copenhagen in a new light. In this paper I present the conceptual system which I think gives the clearest understanding of what it is we are doing when we use quantum mechanics. I do not specifically argue against other views, such as Bohmian mechanics or various versions of decoherence, etc. My criticisms of these ideas I reserve for possible future publications. I assume that the reader is familiar with elementary quantum mechanics, and also with some of the classic literature on conceptual issues, e.g., as collected in [@wz]. OBJECTS VS. THEORY ================== Physics (like all science) begins with physical objects, a category which is indispensable and familiar but still somehow strangely vague. Objects are things which are generally in constant interaction with their environments. We can continue to identify them through many sorts of motions and changes. We can observe them repeatedly and act on them and observe the results. Other people can observe all this along with us. Factual information about objects is in principle verifiable by anyone. Objects are persistently identifiable in this way (at least for a while), but they also change, sometimes unpredictably, sometimes in ways hard to detect or describe. Heraclitus said we cannot step into the same river twice — i.e., it is never quite the same. An object is like a river of events and appearances which we see as a persisting unity. Objects are defined by indicating them — e.g., “That big rock over there next to the tree.” I don’t have to know everything there is to know about the rock to specify it uniquely. There may not even be a “set of all facts” about an object. Is mud caked on the side of the rock part of it, or a separate object? Objects are full of ambiguities like that which we can never completely define away. The important thing is that it can be located within the network of familiar objects and events — the history of our world. This is in contrast to the abstract entities of geometry. All the properties of an equilateral triangle are implied by its definition (given the axioms of the geometrical context in which it is considered). A physical object can always surprise us. The big rock by the tree might turn out to be a hollow prop or a heavy iron meteorite. This contingent quality of genuinely physical things can be annoying to a theorist, because it reminds us that all theoretical descriptions are necessarily approximations and simplifications. But we cannot describe the world without at some point referring to objects — it is only through them that we have a handle on the physical world. They pretty much constitute what we mean by the physical world. The continuity and verifiability of objects help us distinguish “objective” reality from fantasy. Physical thinking begins with physical objects, but physical theory quickly builds up a very different class of concepts for dealing with phenomena (e.g., space-time, point particles, trajectories, momentum, energy, entropy). The shift is obvious in a comparison of the thinking of Aristotle and Galileo. Aristotle still focuses on objects as such and their essences, transformations, and causes. This makes him very obscure for a modern scientifically trained reader. Galileo and his successors look instead for descriptions in terms of quantifiable properties of objects, and especially quantifiable relationships between them (such as relative positions and velocities). Our beautiful and precisely tested theories of matter refer to an abstract description of phenomena which is made possible by restricting our attention to quantifiable properties. This realm of abstraction is one step removed from the concrete reality of things. GEOMETRY VS. DYNAMICS {#sec:geodyn} ===================== Geometrical objects are static. The mind’s eye can scan over them, making comparisons or measurements back and forth repeatedly as much as we like. The fundamental geometrical ideas are congruence and measure. Physical processes and objects are dynamical (a word which implies both change and a chain of causal links among objects). A physical event happens once and can never recur — only an event of the same [*type*]{}. In physics we observe the consequences of events by means of our own transient subjective experiences and by repeatable and objective reference to objects whose properties have been affected. Physical understanding means knowing how to categorize types of events and the types of causal connections which relate “typical situations” (based on observing or setting up relevant conditions and ignoring irrelevant ones) to their possible outcomes. We as observers and experimenters are not external to the physical world. All our actions are part of the web of events and causal relationships. Physical concepts are based on the dynamical properties of objects (including us). When we are accustomed to using a particular physical theory, the process of connecting concrete objects with abstract theory seems very straightforward. The difference between those two things (concrete and abstract) is really very subtle and full of pitfalls. It is convenient to talk as if our theoretical structures were “isomorphic” to the objects out there in the world which they describe, but true isomorphism relates one mathematical entity to another. Only when a physical object can be treated like a geometrical object — e.g., a machined piece of metal which we can inspect and measure repeatedly, examining it with the physical eye in the same way the mind contemplates a geometrical figure — only then can we find something like a quantitative isomorphism. What we have then is a relation between geometrical measure and physical acts of measurement. The latter are in a sense outside of time in this case. The results we get do not depend on when we do the measurements or in what order. Usually, physical analysis is not so simple. The static geometrical view of the physical world treats time as essentially equivalent to space, as one coordinate of a space-time in which the history of the universe exists as a single geometrical object. The units of analysis are geometrical objects — lines, areas, and volumes which indicate the trajectories of objects or the space they occupy. This simple view works at the classical level. Ultimately, however, we are forced to form our concepts on the basis of the practical units of physical analysis in a dynamical universe — interactions between objects. DYNAMICS VS. HISTORY ==================== Suppose we could have a “complete” history of the universe — we know exactly what has happened at every point in space-time. But to say that “q” happened at point $({\bf x},t)$ is meaningless unless q is a member of a set of the possible things which can happen, i.e., it is an abstract property distilled from a knowledge of the range of patterns which occur in the universe as a whole. The way we sort out the [*whens*]{} and [*wheres*]{} in physics is different from the way we sort out the [*whats*]{}. The distinction is almost imperceptible in classical physics. A vector from my own position to the location of a planet determines the gravitational force which the planet’s mass exerts on me, but the vector also tells me where to look to find it in the sky. In quantum theory, however, it becomes clear that locations in space and time take their meaning from the concrete world of objects, but the way we categorize properties comes from dynamics. These are two irreducible aspects of physical description which evolve together — historical and dynamical. A thing is historical insofar as it is objective (can be observed and treated as an object). It then enters into the realm of recordable objective occurrences which can be ordered in historical space and time. It is dynamical insofar as it is defined as an abstract element of a dynamical theory which explains causal relationships between objects. A quantum entity can be both (a common source of confusion). A proton is an object if we are talking about its track through a system of detectors. More often the proton concept appears as an abstract constituent of a dynamical system. Imagine that we could see the universe as omniscient external observers, all of space and time at once, and that what we “see” is a tangle of intersecting particle world-lines (cf. Ch. 1 of Misner, Thorne, and Wheeler [@mtw]). We might detect some patterns which would constitute physical rules or laws in some sense, but it would be quite difficult or impossible to know whether we had found all the important patterns, or to distinguish significant relationships from accidental ones. Even more difficult would be to translate this omniscient description into the kinds of relationships and laws which would be observed by the huge clumsy bunches of world-lines which constitute ourselves. When we set out to investigate Nature, we are not like that external omniscient observer at all. We look for relationships and patterns in the behavior of objects we know. We want to find out – does this kind of object always behave this way under these circumstances? The phrases “this kind,” “this way,” and “these circumstances” imply the ability to abstract relevant or significant features from what are really unique events. They also imply that we can find or (even better) set up many instances of these typical situations. The result is that the concepts we develop to describe physical phenomena depend not only on what we can observe, but also on what we can do. To say that A affects or causes or influences or interacts with B implies a counterfactual: If A had been different, B would have been different, too. The most convincing way to establish a connection is to “wiggle” some parameter in A more or less randomly and then observe the same odd pattern showing up in some property of B [@newton]. If I want to know whether a wall switch controls a certain light, I can flip the switch on and off and observe whether the light follows my actions. There is always the possibility that the light is being controlled by someone else or goes on and off spontaneously; but if I put the switch through a very irregular and spontaneous sequence of changes and the light still follows along, then the probability of a causal connection is very high (barring a conspiracy to deceive the experimenter). Physical theory is possible because we [*are*]{} immersed and included in the whole process — because we can act on objects around us. Our ability to intervene in nature clarifies even the motion of the planets around the sun — masses so great and distances so vast that our role as participants seems insignificant. Newton was able to transform Kepler’s kinematical description of the solar system into a far more powerful dynamical theory because he added concepts from Galileo’s experimental methods — force, mass, momentum, and gravitation. The truly external observer will only get as far as Kepler. Dynamical concepts are formulated on the basis of what we can set up, control, and measure. REDUCTION IS DYNAMICAL, NOT GEOMETRICAL. ======================================== Human beings are adept at contriving machines which we assemble part by part. We try to understand nature by the inverse process of dissection into component parts. Twentieth century physics has probed phenomena at ever smaller scales and we find that matter can be carved into more or less fundamental pieces — atoms and particles. These pieces are nothing like the hard bits of matter Newton or Descartes might have expected, nothing like the machined metal part I mentioned in Section \[sec:geodyn\] above. This should not be surprising, because physical dissection of objects is not like the intellectual dissection of geometrical entities. Machines are an exception because they are designed — they are intellectual patterns imposed on matter. We probe matter by contriving interactions and studying dynamical patterns. The end result is not building blocks of matter, but fundamental units of interaction — quanta. An exchange of one quantum is the minimal dynamical relationship. It is the simplest way one object can interact with another. Quantum particles are not mechanical components. In phenomena where classical approximations are appropriate, it is useful to think of atoms and particles in that way. At a finer scale, it soon becomes apparent that our mechanical concepts are not fundamental. In a sense Plank was correct to say that the electromagnetic radiation in a cavity is not composed of photons, but that they are merely the discrete units of energy which the field exchanges with its environment. The field itself is not discrete — composed of particles. It is simply “quantized.” Quantum fields are not constitutive but dynamical descriptions. They tell us not what the field is but what it can do. CONFIGURATION SPACE VS. PHYSICAL SPACE ====================================== Newtonian mechanics combines the historical and dynamical aspects of physics into one unified description. Any object is made of particles with trajectories ${\bf x}_i (t)$ (where the $i$ indexes the set of particles). These can be thought of as sequences of historical events (a certain particle being in a certain place at at certain time) in the concrete physical world we live in. The same numbers ${\bf x}_i$, along with inherent particle properties such as mass and charge, also determine all the forces at work in the system — i.e., the dynamics. This way of thinking about the physical world is so neat and appealing, it is natural to regard it as an ideal to which physics should conform as much as possible. It may even be difficult to imagine that a physical theory could be anything else. But the quantum method of describing phenomena is radically different. In the familiar Schrödinger approach, we have not trajectories but a wave function $\psi({\bf x}_i ,t)$. Now the ${\bf x}_i$ are independent variables along with the time. This looks at first like a field theory, but $\psi$ cannot be given a realistic historical interpretation in the way that the electromagnetic field can. The ${\bf x}_i$ represent not historical space (the realm of objective events) but the configuration space of the dynamical situation we are analyzing (the possible degrees of freedom of the interaction). The space of states is the space of well-behaved complex functions on the configuration space. The wave function description presupposes a prior objective context. We partition the objective world into chunks with carefully prepared properties (e.g., source, target, and detector). The wave function represents the dynamical relationship among these objects. That is a crucial point and the most common stumbling block. The quantum “system” is usually something small — one or a few particles. The system gives us the degrees of freedom of the dynamics. But the dynamics is a causal relationship among big things — the parts of the apparatus. The quantum experimenter goes to a lot of trouble to make big things interact in small ways, so that the discreteness of the quantum of action shows up in the behavior of objects we can experience directly. From $\psi$ we calculate probabilities of various responses of the detector in our experimental setup (scattering cross-sections, branching ratios, etc.). The wave function or state vector encompasses many possible outcomes, so if we try to take it out of its dynamical context and give it a historical interpretation, then we may imagine as a consequence “many worlds” or “collapse” events. This comes of confusing the configuration space with historical space. How [*are*]{} these two spaces related, if they are not identical? I suggest that it is useful to think of the configuration space as something like a tangent space (a crude metaphorical use of the mathematical term). It is not tangent to a point of historical space-time, but rather is associated with a particular “dynamical partition” of the physical world, i.e., a certain way of dividing the world into objects. A quantum state is a little like the velocity vector of a flow on a manifold. We require different tangent spaces to describe the velocity at different locations. Similarly, when we alter the conditions of an experiment or perform a measurement or otherwise gain objective information, we shift to a new dynamical partition, so that there is a discontinuity between the old state description and the new one appropriate to the new context. Perhaps it is no coincidence that the Schrödinger equation is linear even though quantum transition phenomena have a nonlinear flavor — it may be (again speaking metaphorically) because the Schrödinger equation is defined on a linear tangent space rather than on the full historical manifold. I doubt that there is a well-defined set of all possible dynamical partitions. If there were, then perhaps we could arrive at a realistic, abstract, historical description of quantum dynamics by integrating (another metaphorical abuse of a mathematical term) the states in the configuration spaces over the set of partitions. But even if some such procedure actually produced a meaningful result, it still might not be useful or important. The mathematical structure might be too remote from anything we could observe or measure to be intelligible. COMMON MISCONCEPTIONS ===================== I want to comment briefly on three errors sometimes found in discussions of quantum mechanics, often as a misrepresentation of Bohr’s explanation of the theory by those attacking or even those defending him. Error 1: : “Two theories are needed in physics: classical physics for macroscopic and quantum mechanics for microscopic phenomena.” Bohr’s terminology was a little unfortunate when he talked about the “classical” nature of the measuring apparatus. I think what he meant was that physics develops as a refinement of our ordinary experience with objects. The mathematical concepts of quantum theory are dynamical, not ontological, and we still need the objective aspect of historical description to give them meaning — the same objective aspect which was indistinguishable from the dynamical aspect of classical physics. Now we still use the same historical objective concepts to describe an experiment or observation even though the dynamical language is entirely different. For practical purposes, we use classical physics where it works, but there is no reason to think that a fully quantum analysis of any dynamical system would not give the correct answers if we could do the calculations. Sometimes we use the geometrical structure of physical description to sharpen and criticize our ideas about what is really out there and what really happens; e.g., a pencil in a glass of water is not really broken at the water’s surface – the appearance is due to the refraction of light by water; similarly a rainbow is not really a multicolored ribbon spanning the sky – it is a trick of sunlight refracting through water droplets suspended in the air. This is the kind of refinement of description physics has always pursued. Newtonian mechanics made it appear reasonable to many that all phenomena could ultimately be reduced to an underlying geometrical structure of matter which would be the one true description of what was “really there.” Quantum physics has given up looking for that kind of geometrical ontology. We refine our description of objects as precisely as we can, and then quantum mechanics gives us the dynamical relations. There is no ground floor of quantum description which replaces the pragmatic description of ordinary objects. We have atoms and elementary particles which are useful for describing dynamical systems, their conserved quantities and degrees of freedom, but these are not fundamental in a historical/ontological sense. They are not mechanical components with individual well-defined locations and causal roles at every moment of time. This becomes obvious when we look at field theoretical descriptions of the quantum vacuum. That we still need to base our description of quantum phenomena on observation of macroscopic objects is not a weakness or dilution of quantum mechanical principles. That was the whole idea of quantum dynamics from the very beginning when Heisenberg decided to represent dynamical variables by operators. When Bohr insisted that the quantum revolution was irreversible, that we could never return to classical physics by means of hidden variables or any other trick, he was affirming among other things that our experience with atomic physics had revealed what a sufficiently perspicacious natural philosopher might have realized anyway, that the historical and dynamical modes of description are independent (but complementary) and may take different forms. The quantum of action is the natural fault line between these two ways of looking at things. Error 2: : “There are two kinds of quantum processes: the continuous evolution of quantum states according to the Schrödinger equation, punctuated by discontinuous collapse or projection events which occur when a measurement is made.” From classical physics, we get the habit of thinking of a state as a state of being, independent of context. A quantum state is a dynamical relation defined within an objective context. The continuous evolution of the state is the evolution of interaction amplitudes based on prior objective information. A new measurement changes the context. If we are still considering after the measurement a dynamical relation mediated by the same or same type of particle (or set of particles), then it might be convenient to think of the new state as a projection of the prior state onto a subspace determined by eigenvalues of the observable measured. Actually, as I suggested in the previous section, the two states are defined on entirely different “tangent spaces.” Once this is understood, the Measurement Problem and associated paradoxes disappear. [^2] Many of these paradoxes involve an infinite regression of contexts. Wigner’s friend looks at his apparatus and then Wigner asks his friend the result. If “measurement” causes the state to collapse, is Wigner’s friend in a superposition of states until he gives his answer? The superposition principle applies to dynamical states, not objects. The quantum state of the apparatus is generally irrelevant to the analysis of an experiment. We could in principle determine such a state, but first we would have to specify an objective context in which the apparatus functions as a carrier of dynamical influence. There might be many ways to specify the context, and each one result in a different state, but without any difference in the objective physical reality. It is not really meaningful to talk about the state of an object unless it is treated in our analysis as a dynamical system in an objective context. Meanwhile, an object is what it is — an object, not an abstraction. If the answer is that simple (and my version is not very different from what many others have said or even from Bohr’s explanation in 1927) why is there still controversy after so many years? The emotional and intellectual appeal of the static geometrical picture of space-time and a mechanical view of matter should not be underestimated. If the world is a machine, then perhaps we can control it, especially if the watchmaker who made it is blind and cannot interfere (to use the peculiar metaphor in the title of a book by Richard Dawkins [@dawkins] — can the world be a mechanism without a design engineer, albeit an aimless one?) If the world is simply a collection of atoms in motion, then superstition is without basis; there can be no transcendent evil to fear, as Lucretius taught (but perhaps not much to value either). If the world is nothing but mathematical form, then our minds can penetrate to the ultimate foundation of things and we can escape the dreariness of mundane personal life by contemplating “the thoughts of God when he made the universe.” Of course, we can develop technology, rationally manage our environments, reject superstition, and find joy in Platonic contemplation no matter what physical theory happens to be true. But the spell of geometry and abstraction is powerful, and we can come to depend on it. Error 3: : “Bohr’s view of quantum mechanics implies a shadowy, mystical picture of the world in which the mind of the observer influences events in occult ways and there is no objective independent physical reality.” Classical physics encouraged the belief that whatever cannot be described mathematically does not exist. To Locke and others, the primary attributes of objects were geometrical. The physical world is simply a geometrical form realized in matter. Attributes or qualities which could not be described in this way were secondary – not existing in themselves but only in the subjectivity of the beholder. The more militant version of this view is that physical entities must not only be describable mathematically, they must [*be*]{} mathematical constructs. The universe must be founded on a mathematical ontology. To some, this became the meaning of “objective independent physical reality.” I think only students who have invested a lot of time and effort studying classical mechanics are likely to be shocked to learn that quantum theory does not have such an ontology. It has only been since the 1970s that the general public (or even many physicists) had any idea there was a Problem of Reality in quantum physics. The indeterminacy of quantum dynamics made a little stir earlier in the century because of a possible connection with the classical philosophical problem of free will. Otherwise, relativity got all the press as [*the*]{} weird theory of 20th century physics. In the 1950s it was received teaching that Bohr and von Neumann had solved all the conceptual and mathematical problems of non-relativistic quantum mechanics (cf. for example [@bernstein pp. 44–45]. I doubt that many physicists really understood Bohr’s ideas or realized that Bohr’s treatment of measurement was radically different from von Neumann’s. (Bohr would never have assigned a quantum state to the measuring apparatus.) Perhaps the shift in the center of gravity of physics research from the Europe of Kant and Husserl to pragmatic North America did not help. Our physics departments are good at teaching formal theoretical methods and experimental techniques, but how to understand the connections between one set of concepts and the other is often left as an exercise for the student. Small wonder that the conceptual foundations of quantum mechanics as given in textbooks in confused and perfunctory accounts of the Copenhagen Interpretation looked to students like a closet full of skeletons. The collision of rigid (because rarely examined) orthodoxy with playful revisionism generated a cadre of popularizers, who preached to the multitudes that quantum mechanics was strange in a very cool way, and that in fact perhaps all strange and cool things were linked to it. A pop language was created which popular writers from very different backgrounds were not slow to exploit. This is not necessarily a bad thing. After Newton and the industrial revolution, classical physical terminology often appeared in common speech well out of its proper context, sometimes to legitimize outright frauds or fallacies (animal magnetism), sometimes as useful metaphor (social entropy, a team’s momentum in sports). Why not quantum leaps of achievement and economic superfluidity? Anything radically new seems magical, and so it becomes a metaphor for the magical bursts of insight which constitute the milestones and landmarks of human experience. It remains true nonetheless that objectivity, solidity, and regularity are not compromised by quantum theory. The objectivity of physics comes from its grounding in experience. No formalism can dissolve or erase the basic truths of everyday experience. Schrödinger’s cat is definitely alive or dead, even if the quantum formalism can describe only its dynamical state, not its state of being. Physics gives a complete [*dynamical*]{} description of all phenomena — i.e., all causal links and transformations as measured physically. That doesn’t mean physics explains all higher level emergent phenomena (consciousness, culture, morality) or that it should be expected to. Nor does it provide an underlying ontology for reality in all its aspects [@bilodeau]. There are still some hard cases for which our experience with microscopic processes gives us little help. Quantum dynamics presumes an objective context, but if we are interested in the dynamics of the universe as a whole in an early epoch in which the quantum properties of space-time itself dominate, then there is no conventional recipe for understanding what is going on. Perhaps the present is the objective context for the past. In any case, we should learn from past experience to expect answers only to questions we can put to the universe with action and observation, not with reason and imagination alone. [9]{} Steven Pinker, [*The Language Instinct*]{} (HarperPerennial, New York, 1995). John Archibald Wheeler and Woyciech Hubert Zurek, [ *Quantum Theory and Measurement*]{} (Princeton University Press, Princeton, 1983). Charles W. Misner, Kip S. Thorne, and John Archibald Wheeler, [*Gravitation*]{} (W. H. Freeman, San Francisco, 1973). Roger Newton, “Particles that travel faster than light?”, [*Science*]{} [**167**]{} (3925), 1970, pp. 1569–1574. Richard Dawkins, [*The Blind Watchmaker*]{} (Norton, New York, 1986). Jeremy Bernstein, [*Cranks, Quarks, and the Cosmos*]{} (BasicBooks, New York, 1993). Douglas J. Bilodeau, “Physics, Machines, and the Hard Problem”, [*Journal of Consciousness Studies*]{} [**3**]{}, No. 5/6 (1996), pp. 386–401. Reprinted in Jonathan Shear, ed., [*Explaining Consciousness — The Hard Problem*]{} (MIT Press, Cambridge, 1997), pp. 217–234. [^1]: email:[email protected] [^2]: When people talk about the measurement problem, they focus on the last step in quantum dynamical analysis — the translation from the final quantum state (which may encompass many possible outcomes) to a particular objective result of the measurement. They forget about the first step — the translation from a unique objective situation to an abstract initial quantum state based on properties we consider relevant and can control or measure. There is a discontinuity in mode of description at both ends. The initial state is no more an objective ontological state of being than the final state.
{ "pile_set_name": "ArXiv" }
--- abstract: 'A mostly right-handed sneutrino as the lightest supersymmetric particle (LSP) is an interesting dark matter candidate, leading to LHC signatures which can be quite distinct from those of the conventional neutralino LSP. Using  for testing the model against the limits published by ATLAS and CMS in the context of so-called Simplified Model Spectra (SMS), we investigate to what extent the supersymmetry searches at Run 1 of the LHC constrain the sneutrino-LSP scenario. Moreover, we discuss the most relevant topologies for which no SMS results are provided by the experimental collaborations but which would allow to put more stringent constraints on sneutrino LSPs. These include, for instance, the mono-lepton signature which should be particularly interesting to consider at Run 2 of the LHC.' bibliography: - 'biblio.bib' --- LPSC15065\ HEPHY-PUB 948/15 [**Constraints on sneutrino dark matter from LHC Run 1**]{} [ Chiara Arina$^{1,2}$, Maria Eugenia Cabrera Catalan$^{2,3,4}$, Sabine Kraml$^{5}$, Suchita Kulkarni$^{5,6}$, Ursula Laa$^{5,6}$\ ]{} Introduction {#sec:intro} ============ Before the start of data taking at the LHC, the common perception was that supersymmetry (SUSY), if it has anything to do with stabilizing the electroweak scale, would be discovered quickly, while Higgs physics would need to wait for rather high statistics. In reality, quite the opposite has happened: a Higgs boson has been found [@atlas:2012gk; @cms:2012gu], but there is still no sign of SUSY—or of any new physics beyond the Standard Model (BSM) whatsoever. Indeed, the searches at Run 1 of the LHC at centre-of-mass energies of 7–8 TeV have pushed the mass limits of SUSY particles quite high already, well above 1 TeV for 1st/2nd generation squarks and gluinos [@atlas:susy:twiki; @cms:susy:twiki]. Scenarios with high-scale [@Djouadi:2013vqa], split [@Giudice:2004tc; @Bhattacharyya:2012ct; @Benakli:2013msa] or at least spread [@Hall:2011jd] SUSY are thus becoming increasingly popular in the literature. It should be kept in mind, however, that the current LHC limits sensitively depend on the presence of particular decay modes, and are considerably weakened in case of compressed [@Dreiner:2012gx] or stealth [@Fan:2011yu] spectra. Besides, the squark/gluino mass limits vanish completely in case the neutralino LSP is heavier than about 600 GeV. It should also be kept in mind that the SUSY mass limits depend sensitively on the nature of the LSP. Most experimental analyses indeed assume that the LSP is the lightest neutralino of the Minimal Supersymmetric Standard Model (MSSM). A particularly interesting alternative, and the subject of this paper, is a mainly right-handed (RH) mixed sneutrino in the MSSM augmented by a RH neutrino superfield [@Borzumati:2000mc; @Arkani-Hamed:2000bq]. This case is well motivated by two basic problems: the origin of neutrino masses and the nature of dark matter (DM). Its LHC signatures can be quite distinct from those of the conventional neutralino LSP. The left-handed (LH) sneutrino of the MSSM is excluded as the LSP and as a DM candidate because it has a non-zero hypercharge: its couplings to the $Z$ boson makes it annihilate too efficiently in the early Universe, and hence its final relic abundance is lower than the value $\Omega_{\rm DM} h^2$ measured by the WMAP and Planck satellites [@Hinshaw:2012aka; @Ade:2013zuv]. Very stringent limits come moreover from direct DM detection experiments: the $\tilde{\nu}_L$ scattering off nuclei is mediated by $t$-channel $Z$ boson exchange, giving a spin-independent (SI) cross section of order $10^{-39} {\rm cm^2}$ — a value excluded already a decade ago for DM particles heavier than 10 GeV. A light $\tilde{\nu}_L$ with mass below $m_Z/2$ is also excluded by the $Z$ invisible width. The picture changes dramatically if we include in the MSSM a RH neutrino superfield (MSSM+RN from here on), which gives rise to Dirac neutrino masses. Besides the RH neutrino, the superfield also contains a scalar field, the RH sneutrino $\tilde{N}$ (strictly speaking this is a right-chiral field, but we use the RH notation for simplicity). This field, if at TeV scale, can mix with the LH partner $\tilde{\nu}_L$ and yield a mostly RH sneutrino LSP as a viable thermal DM candidate [@Borzumati:2000mc; @Arkani-Hamed:2000bq].[^1] The phenomenology of this model was investigated in detail in [@Arkani-Hamed:2000bq; @Belanger:2010cd; @Dumont:2012ee]. Indirect detection and cosmology were discussed in [@Hooper:2004dc; @Arina:2007tm; @Choi:2012ap], and LHC signatures in [@Thomas:2007bu; @Belanger:2011ny; @Arina:2013zca] (see also [@BhupalDev:2012ru; @Guo:2013asa; @Harland-Lang:2013wxa] for related LHC studies). Reference [@Arina:2013zca] in fact gave an update of the status of the sneutrino as DM after the Higgs mass measurements, by exploring the SUSY parameter space with the soft breaking terms fixed at the grand unification (GUT) scale, and assessing also the impact of the most recent exclusion bound for DM direct searches from LUX [@Akerib:2013tjd]. In this paper, we extend the work of [@Arina:2013zca] by investigating to what extent the results from SUSY searches at Run 1 of the LHC, published in terms of so-called Simplified Model Spectra (SMS) limits,[^2] constrain the sneutrino-LSP scenario. Moreover, we discuss the most promising topologies for which no SMS results exist but would enhance the LHC sensitivity to sneutrino DM. To this aim, we make use of the  package [@Kraml:2013mwa; @smodels:v1; @smodels:wiki] to compare the predictions of the MSSM+RN model against the SMS limits published by ATLAS and CMS. The strengths of   are that it 1.) automatically decomposes the signal of an arbitrary SUSY spectrum into all its SMS-equivalent topologies, and 2.) includes a large database of more than 60 SMS results from ATLAS and CMS SUSY searches. This allows us to test the limits from a large variety of searches and at the same time draw conclusions about which additional topologies should be considered. The paper is organised as follows. After briefly defining the MSSM+RN in Section \[sec:model\] we describe the numerical procedure in Section \[sec:num\]. In particular in \[sec:sample\] we explain the sampling method and the constraints implemented in the model likelihood function, while in \[sec:sms\] we describe the application of  to the MSSM+RN. Our numerical results are presented in Section \[sec:result\], and the conclusions in Section \[sec:concl\]. Two appendices contain some interesting supplementary material. Appendix \[sec:efficiency\] discusses the validity of applying SMS results from slepton searches (dilepton signature) to chargino-pair production followed by decays into leptons and sneutrinos. Appendix \[sec:lifetime\] gives some details on scenarios with long-lived heavy charged particles, in particular gluinos or stops, which so far cannot be constrained by SMS results. The MSSM+RN model {#sec:model} ================= We use the MSSM+RN model as defined in [@Arkani-Hamed:2000bq; @Borzumati:2000mc; @Arina:2007tm]. (The model used in [@Belanger:2010cd; @Belanger:2011ny; @Dumont:2012ee] differs only slightly in notation.) The superpotential for Dirac RH neutrino superfield is given by $$\label{lrsuppot} W = \epsilon_{ij} (\mu \hat H^{u}_{i} \hat H^{d}_{j} - Y_{l}^{IJ} \hat H^{d}_{i} \hat L^{I}_{j} \hat R^{J} + Y_{\nu}^{IJ} \hat H^{u}_{i} \hat L^{I}_{j} \hat N^{J} )\,,$$ where $Y_{\nu}^{IJ}$ is a matrix in flavor space (which we choose to be real and diagonal), from which the mass of neutrinos are obtained as $m_D^{I} = v_{u}Y_{\nu}^{II}$. Note that lepton-number violating terms are absent in this scheme. The additional scalar fields contribute with new terms in the soft-breaking potential $$\label{lrsoftpot} V_{\rm soft} = (M_{L}^{2})^{IJ} \, \tilde L_{i}^{I \ast} \tilde L_{i}^{J} + (M_{N}^{2})^{IJ} \, \tilde N^{I \ast} \tilde N^{J} - [\epsilon_{ij}(\Lambda_{l}^{IJ} H^{d}_{i} \tilde L^{I}_{j} \tilde R^{J} + \Lambda_{\nu}^{IJ} H^{u}_{i} \tilde L^{I}_{j} \tilde N^{J}) + \mbox{h.c.}]\,,$$ where both matrices $M_{N}^{2}$ and $\Lambda_{\nu}^{IJ}$ are real and diagonal, $M_{N}^{2}={\rm diag}(m^2_{N^k})$ and $\Lambda_{\nu}^{IJ}={\rm diag}(A_{{\tilde{\nu}}}^k)$, with $k=e,\mu,\tau$ being the flavor index. In the sneutrino interaction basis, defined by the vector $\Phi^\dag=(\tilde{\nu}_L^\ast,\, \tilde N^\ast)$, the sneutrino mass potential is $$\begin{aligned} V_{\rm mass}^k =\frac{1}{2}\, \Phi^{\dag}_{LR}\, \mathcal{M}^2_{LR}\, \Phi_{LR}\,,\end{aligned}$$ with the squared–mass matrix $\mathcal{M}^2_{LR}$ $$\mathcal{M}^2_{LR} = \left( \begin{array}{cc} m^2_{L^k} + \frac{1}{2} m_{Z}^{2} \cos(2\beta) + m_D^2 & \; \; \; \frac{1}{\sqrt{2}} A_{{\tilde{\nu}}}^k v\sin\beta - \mu m_D/\tan\beta \\ \frac{1}{\sqrt{2}} A_{{\tilde{\nu}}}^k v\sin\beta - \mu m_D/\tan\beta & m^2_{N^k} + m_D^2 \end{array}\right) \,. \label{eq:masslr}$$ Here, $m^2_{L^k}$ are the soft mass terms for the three SU(2) leptonic doublets, $\tan\beta = v_u/v_d$ and $v^2=v_u^2+v_d^2=(246\, {\rm GeV})^2$, with $v_{u,d}$ the usual Higgs vacuum expectation values (vevs). The Dirac neutrino mass $m_D$ is small and can be safely neglected. The off-diagonal term determines the mixing of the LH and RH fields. If $A_{{\tilde{\nu}}}^k = \eta Y_\nu$, that is if the trilinear term is aligned to the neutrino Yukawa, this term is certainly very small as compared to the diagonal entries and is therefore negligible. However, $A_{{\tilde{\nu}}}^k$ can in general be a free parameter and may naturally be of the order of the other entries of the matrix [@Borzumati:2000mc; @Arkani-Hamed:2000bq], thus inducing a sizable mixing among the interaction eigenstates. The sneutrino mass eigenstates are then given by $$\begin{aligned} \label{lr_eigenstates} \left(\begin{array}{c} {\tilde{\nu}}_{k_1} \\ {\tilde{\nu}}_{k_2} \end{array}\right) = \left(\begin{array}{cc} -\sin\theta^k_{{\tilde{\nu}}} & \cos\theta^k_{{\tilde{\nu}}} \\ \cos\theta^k_{{\tilde{\nu}}} & \sin\theta^k_{{\tilde{\nu}}} \end{array}\right) \left(\begin{array}{c} {\tilde{\nu}}_L^k \\ \tilde{N}^k \end{array}\right) \,.\end{aligned}$$ The relevant parameters at the electroweak (EW) scale for the sneutrino sector are the two mass eigenvalues $m_{{\tilde{\nu}}_{k_1}}$ and $m_{{\tilde{\nu}}_{k_2}}$ and the mixing angle $\theta_{{\tilde{\nu}}}^k$, related to the $A_{{\tilde{\nu}}}^k$ term via $$\sin 2\theta_{{\tilde{\nu}}}^k = \sqrt{2} \frac{A_{{\tilde{\nu}}}^k\, v \sin \beta}{(m^2_{{\tilde{\nu}}_{k 2}}-m^2_{{\tilde{\nu}}_{k 1}})}\, .$$ The sneutrino coupling to the $Z$ boson, which does not couple to SU(2)$_L$ singlets, is largely reduced by a sizeable mixing. This has a relevant impact on the sneutrino phenomenology, as discussed in, [*e.g.*]{}, Refs. [@Arkani-Hamed:2000bq; @Smith:2001hy; @Grossman:1997is; @Arina:2007tm; @Belanger:2010cd]. The renormalization group equations (RGEs) are modified by the new singlet superfields $\hat{N}$ as $$\begin{aligned} \frac{{{\rm d}}m^2_{N^k}}{{{\rm d}}\ln\mu} & = & \frac{4}{16 \pi^2} \left(A^{k}_{{\tilde{\nu}}}\right)^2\,,\\ \nonumber \frac{{{\rm d}}m^2_{L^k}}{{{\rm d}}\ln\mu} & = & \left( \rm MSSM\, terms \right) + \frac{2}{16 \pi^2} \left(A^{k}_{{\tilde{\nu}}}\right)^2\,,\\ \nonumber \frac{{{\rm d}}A_{{\tilde{\nu}}}^k}{{{\rm d}}\ln\mu} & = & \frac{2}{16 \pi^2} \left(- \frac{3}{2} g^2_2 - \frac{3}{10} g_1^2 + \frac{3}{2} Y_t^2 + \frac{1}{2} Y_\tau^2 \right) A_{{\tilde{\nu}}}^k\,,\\ \nonumber \frac{{{\rm d}}m^2_{H_u}}{{{\rm d}}\ln\mu} & = & \left( \rm MSSM\, terms \right) +\sum_{k=e,\mu,\tau} \frac{2}{16 \pi^2} \left(A^{k}_{{\tilde{\nu}}}\right)^2\,, \nonumber\end{aligned}$$ with $\mu$ being the renormalization scale, $g_2$ and $g_1$ the SU(2) and U(1) gauge couplings, $Y_{t,\tau}$ the top and $\tau$ Yukawa respectively. Notice that the RH soft mass receives corrections only from the trilinear term, which affects as well the running of the LH part, as recognized in [@Belanger:2010cd; @Belanger:2011ny]. By neglecting all lepton Yukawas but $Y_\tau$ in the RGEs and by assuming common scalar masses and trilinear couplings for all flavors, the sneutrino tau, ${\tilde{\nu}}_{\tau_1}$, ends up to be the lightest one among the three sneutrino flavors and hence the LSP, while ${\tilde{\nu}}_{e_1}={\tilde{\nu}}_{\mu_1}$. Note that it frequently happens that the mass splitting between ${\tilde{\nu}}_{\tau_1}$ and ${\tilde{\nu}}_{e_1,\mu_1}$ is smaller than 5 GeV, which means that regarding collider phenomenology they are practically degenerate. This will be discussed in more detail in Section \[sec:sms\]. Numerical procedure {#sec:num} =================== Sampling method over the model parameters {#sec:sample} ----------------------------------------- For definiteness, we study the MSSM+RN with soft terms defined at a high scale $M \sim M_{\rm GUT}$ as in [@Arina:2013zca]. Allowing for non-universalities in the gaugino and scalar sectors, our set of free parameters is $$M_1, M_2, M_3, m_L, m_R, m_N, m_Q, m_H, A_l,A_{{\tilde{\nu}}}, A_q, \tan\beta, {\rm sgn}\mu \,. \label{eq:param}$$ Here the $M_i$ are the gaugino masses, $m_L, m_R, m_N$ are the charged slepton and sneutrino masses (equal for all flavors), $m_Q$ is a common squark mass parameter, $m_H\equiv m_{H_u} = m_{H_d}$ denotes the common entry for the two Higgs doublet masses, and $A_l$ and $A_q$ are the scalar trilinear couplings for the sleptons and squarks respectively, same for all flavors. The absolute value of $\mu$ is obtained from the minimization of the Higgs potential, leaving only the sign of $\mu$ as a free parameter. The computation of the mass spectrum follows that explained in [@Arina:2013zca], where all details are provided. Observable Value/Constraint Ref. -------- ------------------------------------------------------- ----------------------------------------------------- ------------------------------- [**]{} $m_h$ $ 125.85 \pm 0.4$ GeV (exp) $ \pm\, 4 $ GeV (theo)  [@atlas:2012gk; @cms:2012gu] BR$(B \to X_s \gamma) \times 10^{4}$ $ 3.55 \pm 0.24 \pm 0.09 $ (exp)  [@Aaij:2012nna] BR$(B_s \to \mu^+ \mu^-) \times 10^{9}$ $ 3.2^{+1.4}_{-1.2} $ (stat) $^{+0.5}_{-0.3}$ (sys)  [@Amhis:2012bh] $\Omega_{\rm DM} h^2$ $ 0.1186 \pm 0.0031 $ (exp) $ \pm\, 20\%$ (theo)  [@Ade:2013zuv] [**]{} $\Delta\Gamma_Z^{\rm invisible}$ $ <2 $ MeV ( 95% CL)  [@PDG] ${\rm BR}(h \to {\rm invisible})$ $ <20\% $ (95% CL)  [@Belanger:2013xza] $m_{{\tilde{\tau}_1^-}}$ $ > 85$ GeV (95% CL)  [@sleptons2004] $m_{\widetilde{\chi}_1^+}, m_{\tilde{e},\tilde{\mu}}$ $> 101$ GeV (95% CL)  [@PDG] $m_{\tilde{g}}$ $ > 308$ GeV (95% CL)  [@Abazov:2007aa] $ \sigma_{n}^{\rm SI} $ $ < \sigma^{\rm SI}_{\rm LUX}$ (90% CL)  [@Akerib:2013tjd] : Summary of the observables and constraints used in this analysis. \[tab:co\] The list of constraints implemented in the model likelihood function is given in Table \[tab:co\]. In particular, besides consistency with $B$-physics constraints, we require the Higgs mass $m_h$ to be compatible with the ATLAS and CMS measurements [@atlas:2012gk; @cms:2012gu], which we combine by a statistical mean, as obtained in [@Cabrera:2012vu]. Its uncertainty is dominated by the theoretical error, estimated to be around 4GeV [@Allanach:2004rh]. We also require that chargino and charged slepton masses fulfill the LEP bounds at 95% confidence level (CL) —notice that the tau slepton has a slightly less stringent lower bound of 85 GeV [@sleptons2004] as compared to selectrons and smuons— and we include the gluino mass bound from the D0 collaboration [@Abazov:2007aa]. If ${\tilde{\nu}}_{\tau_1}$ is light enough to be produced in $Z$ decay, we require its contribution to the $Z$ invisible decay width to be smaller than 2MeV [@ALEPH:2005ab]. Similarly, when the sneutrino mass is lighter than $m_h/2$, the Higgs can decay invisibly into sneutrino pairs. We require that such decays do not contribute more than 20% to the Higgs invisible branching ratio [@Belanger:2013xza]. Regarding DM constraints, we require consistency with the measured relic abundance and with the bounds from direct detection experiments (constraints from indirect DM detection are also fulfilled). The experimental error on $\Omega_{\rm DM} h^2$ has become incredibly small due to the Planck measurement [@Ade:2013zuv], while the theoretical one is still large. We use a conservative estimate of the order 20% [@Boudjema:2011ig] for the latter. Furthermore, we enforce the sneutrino SI scattering cross section off nuclei, $\sigma_n^{\rm SI}$, to be compatible with the recent 90% CL bound from LUX [@Akerib:2013tjd]. To evaluate the experimental observables we first computed the supersymmetric particle spectrum with a modified version of  [@Allanach:2001kg]. For the computation of the sneutrino relic density and elastic scattering cross-section the model has been implemented in  [@Duhr:2011se; @Degrande:2011ua], by adding the appropriate term in the superpotential and in the soft SUSY breaking potential. We generate output files compatible with in order to use the public code  [@Belanger:2013oya]. The $B$-physics observables are computed by interfacing the program with  [@Mahmoudi:2008tp]. The likelihood is constructed in a simple way. For measured quantities, we assume a Gaussian likelihood function with a variance given by combining in quadrature the theoretical and experimental variances. For observables for which only lower or upper limits are available, we use a likelihood modelled as a step function on the $x\%$ CL of the exclusion limit. The total likelihood function is then the product of the individual likelihoods associated to each experimental result. In order to save time in the sampling procedure, the slepton, chargino and gluino mass limits are, however, absorbed into the prior probability density functions: each parameter point generating a mass spectrum that violates one of these bounds is immediately discarded. Given the likelihood function, we sample the posterior probability density function with the algorithm [@Feroz:2007kg; @Feroz:2008xx; @2013arXiv1306.2144F]. In order to cover all phenomenological interesting classes, we run separate chains that look either for light(ish) EW-inos ($m_{\tilde{\chi}_1^{\pm}} < 900$ GeV), light sleptons ($m_{\tilde{l}} < 600$ GeV), or for light squark or gluinos ($m_{\tilde{q}} < 1.5$ TeV or $m_{\tilde{g}} < 1.5$ TeV). As for the choice of priors, we always take logarithmic priors on $M_3, m_Q, A_Q, m_H$, while we use both logarithmic and flat priors for $M_1, M_2, m_L, m_R,m_N,A_L,A_{{\tilde{\nu}}}, \tan\beta$, the sign of $\mu$ is fixed to +1 (details on the prior ranges are provided in [@Arina:2013zca]). In particular we perform two chains, one with log and one with flat priors, for each relevant data set: two chains for light EW-inos (these two data sets coincide with the ones used in [@Arina:2013zca]), two chains for light sleptons and two chains for light squarks or gluinos. In each case, the other masses are left to vary freely from high to low values. The motivation for this is, as mentioned, to cover all potentially interesting cases; the results we will present in Section \[sec:result\] are for all chains combined together. The sampled points correspond to a 95% CL in volume of the posterior. (Since in this study we are not interested in statistical statements on the parameter space, we will however not exploit this feature.) The limits imposed by a step function are of course strictly obeyed by all scan points. Moreover, we have checked that none of the individual constraints implemented by a Gaussian gets a large pull in the final sample. In particular, BR$(B \to X_s \gamma)$ and BR$(B_s \to \mu^+ \mu^-)$ are in full agreement with the 95% CL experimental results [@Aaij:2012nna; @Amhis:2012bh] for all points in the samples. Once the sampling of the parameter space according to the constraints in Table \[tab:co\] is completed, all the points in the chains are confronted against the LHC Run 1 results by means of  as explained in the next subsection. Deriving LHC constraints with SModelS {#sec:sms} -------------------------------------  [@Kraml:2013mwa; @smodels:v1; @smodels:wiki] is designed to decompose the signal of any arbitrary BSM spectrum with a $\mathbb{Z}_2$ symmetry into simplified model topologies and test it against the existing LHC bounds in the SMS context.  uses Pythia6.4 [@Sjostrand:2006za], NLL-fast [@Beenakker:1996ch; @Beenakker:1997ut; @Kulesza:2008jb; @Kulesza:2009kq; @Beenakker:2009ha; @Beenakker:2010nq; @Beenakker:2011fu] and PySLHA [@Buckley:2013jua], and includes a database of more than 60 SMS results from ATLAS and CMS. The decomposition procedure works “out of the box” for the MSSM+RN model with a sneutrino LSP. Nonetheless some subtleties must be taken care of when processing the MSSM+RN scan points with . First, the input to  can be simulated events or an SLHA [@Skands:2003cj] file containing the full mass spectrum and decay tables as well as the SUSY production cross sections, $\sigma$, in the format specified at [@slha-xsext]. We choose the latter option. We use the MSSM+RN model implemented in [micrOMEGAs]{}3.2 [@Belanger:2013oya] (see [@Arina:2013zca]) to compute the decay branching ratios, ${\mbox{\ensuremath{\mathcal{B}}}}$. The production cross sections for sleptons and sneutrinos ( the sector modified with respect to the MSSM) are also computed with [micrOMEGAs]{}3.2. For all other production processes, we use the default   cross section calculator based on Pythia6.4 [@Sjostrand:2006za] and NLL-fast [@Beenakker:1996ch; @Beenakker:1997ut; @Kulesza:2008jb; @Kulesza:2009kq; @Beenakker:2009ha; @Beenakker:2010nq; @Beenakker:2011fu]. Electroweak cross sections are thus computed at leading order while strong productions are computed at NLO+NLL order. Given the information on $\sigma$ and ${\mbox{\ensuremath{\mathcal{B}}}}$ in the SLHA files,  computes ${\mbox{\ensuremath{\sigma\times\mathcal{B}}}\xspace}$ for each topology that occurs. Here, a topology is characterised by the SM particles originating from each vertex, and the mass vector of the SUSY particles in the decays. In order to avoid dealing with a large number of irrelevant processes, which is expensive in terms of computing time, topologies for which ${\mbox{\ensuremath{\sigma\times\mathcal{B}}}\xspace}< \sigma_{\rm cut}$, with $ \sigma_{\rm cut}=0.05$ fb, are discarded. When dealing with an arbitrary spectrum of SUSY particles, it is possible that a part of the decay chain leads to completely invisible decays,  a decay of a heavy sneutrino to a lighter one plus neutrinos in the current scenario. In such cases,  compresses the invisible part of the decay chain as illustrated in Fig. \[fig:invisible-illustration\]. All decays to neutrinos appearing after the last visible decay are disregarded, yielding an “effective LSP” for the particular event, which can be different from the true LSP. This procedure is called “invisible compression”. Likewise, a neutralino may decay invisibly to a sneutrino and a neutrino; in this case the compressed topology resembles an MSSM topology. ![Illustration of “invisible compression” in . The decays of the heavier sneutrino to the lighter one plus neutrinos are discarded in the final topology, leaving the $\tilde\nu_{\tau_2}$ as an effective LSP.[]{data-label="fig:invisible-illustration"}](plots/compression_illustration_snu2snu.png){width="50.00000%"} In addition, if the mass gap between mother and daughter particles is small, the decay products will be too soft to be detected at the LHC. This is taken care of by the so-called “mass compression” in , discarding any SM particle that come from a vertex for which the mass splitting of the R-odd particles is less than a certain threshold. We use $5$ GeV as the minimum required mass difference for the decay products to be visible. Another comment is in order. The experimental constraints currently implemented in the  database require final states containing missing transverse energy (MET). This means that scenarios with long-lived particles ($c\tau>10$ mm) leading to signatures with displaced vertices or heavy charged particle tracks cannot be tested with . In the MSSM, this problem occurs, , in wino-LSP scenarios where the ${\widetilde{\chi}_1^\pm}$ is highly mass-degenerate with the ${\widetilde{\chi}_{1}^0}$ and thus becomes long-lived. In the sneutrino LSP case, not only charginos can be long lived if the mass splitting with the sneutrino is small enough; other possibilities are, , long-lived gluinos or stops, if they are the next-to-lightest SUSY particle (NLSP). We perform a detailed check of all input points to avoid the erroneous application of SMS limits to such cases. Points that have visible decays from long-lived particles or heavy charged particle tracks with cross sections larger than $\sigma_{\rm cut}$ are discarded. (A brief discussion of such scenarios can be found in Appendix \[sec:lifetime\].) Once the decomposition into SMS topologies, including mass and invisible compression, is completed and the checks that the SMS results actually apply are passed, a given point is confronted against the SMS results in the  database. For each experimental constraint that exists,  reports among other things the analysis name, the [**Tx**]{} name identifying the topology,[^3] the predicted signal cross section for the point under consideration and the 95% CL experimental upper limit on it. Finally, the ratio $r$ of the signal cross-section and the upper limit, $r=\sigma(\textrm{predicted})/\sigma(\textrm{excluded})$, is given, where $\sigma$ effectively means ${\mbox{\ensuremath{\sigma\times\mathcal{B}}}\xspace}$ or the weight of the topology. A value of $r\ge 1$ means that the input model is likely excluded by the corresponding analysis. Results {#sec:result} ======= We now turn to analysing the impact of the LHC searches on the MSSM+RN parameter space. As explained in the previous Section, we here consider only points for which the SMS results apply,  we discard points with non-prompt visible decays as well as points with long-lived charged particles. Scanning over the parameter space, we can then distinguish several cases: - the SMS results in principle apply but no SMS constraints actually exist for the specific topologies of the point — these points will be labelled as [*not tested*]{};[^4] - there exist (one or more) SMS results that test the specific topologies of the point but for each topology the total $\sigma \times {\mbox{\ensuremath{\mathcal{B}}}}$ is below the corresponding 95% CL upper limit — these points will be considered as [*allowed*]{}; and - at least one topology has a $\sigma \times {\mbox{\ensuremath{\mathcal{B}}}}$ equal or above its 95% CL upper limit ($r\ge1$) — these points will be considered as [*excluded*]{}. ![For scan points that are excluded by the SMS limits, we show (in color) the breakdown of most constraining analyses in the ${\tilde{\nu}}_{\tau_1}$ vs. $\tilde g$ mass plane. To illustrate the coverage of the parameter space, we also show (in grey) the not excluded or not tested points.[]{data-label="fig:breakdown-snu-gluino"}](plots/excluded_grouped_mgluino_snutau1.png){height="7cm"} ![As Fig. \[fig:breakdown-snu-gluino\] but in the ${\tilde{\nu}}_{\tau_1}$ vs. ${\widetilde{\chi}^\pm}_1$ mass plane.[]{data-label="fig:breakdown-snu-c1"}](plots/excluded_grouped_mC1_snutau1.png){height="7cm"} Let us start with the question which analyses are the most important ones for constraining the model. To this end, Figs. \[fig:breakdown-snu-gluino\] and \[fig:breakdown-snu-c1\] show a breakdown of most constraining analyses in the ${\tilde{\nu}}_{\tau_1}$ versus $\tilde g$ and ${\tilde{\nu}}_{\tau_1}$ versus ${\widetilde{\chi}^\pm}_1$ mass planes, respectively. Looking first at Fig. \[fig:breakdown-snu-gluino\], we see that (the SMS interpretations of) the hadronic SUSY searches [@Aad:2013wta; @Aad:2014wea; @ATLAS-CONF-2013-061; @Chatrchyan:2013wxa; @Chatrchyan:2013lya; @Chatrchyan:2013iqa; @Chatrchyan:2014lfa] are constraining gluino masses up to about $m_{\tilde g}\approx 1200$ GeV and LSP masses up to about $m_{{\tilde{\nu}}_{\tau_1}}\approx 500$ GeV. These searches mostly exclude points where either $\tilde g\to b\bar b{\widetilde{\chi}^0}_i$, $\tilde g\to t\bar t{\widetilde{\chi}^0}_i$ or $\tilde g\to q\bar q{\widetilde{\chi}^0}_i$ decays are dominant, followed by an invisible decay of the neutralino, ${\widetilde{\chi}^0}_i\to \nu{\tilde{\nu}}$. Moreover, dilepton + MET searches [@Aad:2014vma; @Khachatryan:2014qwa] exclude sneutrino LSP masses up to about $m_{{\tilde{\nu}}_{\tau_1}}\approx 210$ GeV, independent of the gluino mass. The process that is constrained here is Drell-Yang production of ${\widetilde{\chi}^+}_1{\widetilde{\chi}^-}_1$ followed by ${\widetilde{\chi}^\pm}_1\to l^\pm {\tilde{\nu}}_{l1}$ ($l=e$ or $\mu$), with the ${\tilde{\nu}}_{l1}\to{\tilde{\nu}}_{\tau_1}+X$ decay being invisible (because of $X$ being genuinely invisible or very soft). Consequently, in Fig. \[fig:breakdown-snu-c1\] we see that chargino masses can be excluded up to about $m_{{\widetilde{\chi}^\pm}_1}\approx 440$ GeV by the dilepton + MET limits. (There is also a small region of parameter space at low masses where $\tau^+\tau^- + {\rm MET}$ [@Aad:2014yka] gives the strongest limit.) It is important to note here that the constraints on ${\widetilde{\chi}^+}_1{\widetilde{\chi}^-}_1\to l^+l^- + {\rm MET}$ actually stem from the $\tilde l^+\tilde l^- \to l^+l^-{\widetilde{\chi}^0}_1{\widetilde{\chi}^0}_1$ simplified model (and analogously for $\tau^+\tau^- + {\rm MET}$), which has the opposite spin configuration than chargino-pair production followed by chargino decays into sneutrinos. The validity of applying the limits from the slepton searches to the case of chargino-pair production is discussed in Appendix \[sec:efficiency\]. ![Scatter plots of points for which SMS results apply. The top row shows the ${\tilde{\nu}}_{\tau_1}$ vs. $\tilde g$, the bottom row the ${\tilde{\nu}}_{\tau_1}$ vs. ${\widetilde{\chi}^\pm}_1$ mass plane. In the panels on the left, the points excluded by the SMS constraints (red) are plotted on top of those which are not excluded (blue); in panels on the right this plotting order is inverted. Also shown (in grey) are the “not tested” points, for which no SMS constraints exist.[]{data-label="fig:summary"}](plots/summary_mgluino_snutau1.png "fig:"){width="44.00000%"} ![Scatter plots of points for which SMS results apply. The top row shows the ${\tilde{\nu}}_{\tau_1}$ vs. $\tilde g$, the bottom row the ${\tilde{\nu}}_{\tau_1}$ vs. ${\widetilde{\chi}^\pm}_1$ mass plane. In the panels on the left, the points excluded by the SMS constraints (red) are plotted on top of those which are not excluded (blue); in panels on the right this plotting order is inverted. Also shown (in grey) are the “not tested” points, for which no SMS constraints exist.[]{data-label="fig:summary"}](plots/summary_mgluino_snutau1_inverted.png "fig:"){width="44.00000%"} ![Scatter plots of points for which SMS results apply. The top row shows the ${\tilde{\nu}}_{\tau_1}$ vs. $\tilde g$, the bottom row the ${\tilde{\nu}}_{\tau_1}$ vs. ${\widetilde{\chi}^\pm}_1$ mass plane. In the panels on the left, the points excluded by the SMS constraints (red) are plotted on top of those which are not excluded (blue); in panels on the right this plotting order is inverted. Also shown (in grey) are the “not tested” points, for which no SMS constraints exist.[]{data-label="fig:summary"}](plots/summary_mC1_snutau1.png "fig:"){width="44.00000%"} ![Scatter plots of points for which SMS results apply. The top row shows the ${\tilde{\nu}}_{\tau_1}$ vs. $\tilde g$, the bottom row the ${\tilde{\nu}}_{\tau_1}$ vs. ${\widetilde{\chi}^\pm}_1$ mass plane. In the panels on the left, the points excluded by the SMS constraints (red) are plotted on top of those which are not excluded (blue); in panels on the right this plotting order is inverted. Also shown (in grey) are the “not tested” points, for which no SMS constraints exist.[]{data-label="fig:summary"}](plots/summary_mC1_snutau1_inverted.png "fig:"){width="44.00000%"} Also noteworthy is the fact that most of the excluded points in Figs. \[fig:breakdown-snu-gluino\] and \[fig:breakdown-snu-c1\] have some grey points lying below them, which are not excluded or not tested at all. This is corroborated in Fig. \[fig:summary\], where we present the summary of not tested, allowed and excluded points in the ${\tilde{\nu}}_{\tau_1}$ versus $\tilde g$ and ${\tilde{\nu}}_{\tau_1}$ versus ${\widetilde{\chi}^\pm}_1$ mass planes. In the plots on the left, the excluded points (red) are plotted on on top of the allowed points (blue), while in the plots on the right this plotting order is inverted. As can be seen, only a small part of the parameter space can genuinely be excluded by the SMS results—over most of the regions where the SMS results are valid, there are almost always parameter combinations such that the limits can be avoided. For the dilepton signature originating from chargino-pair production, the chargino mixing plays an important rôle: wino-like charginos have a higher production cross section, and a higher branching fraction into $l{\tilde{\nu}}_{l1}$. The limits from $l^+l^-+{\rm MET}$ searches therefore mostly affect scenarios with wino-like ${\widetilde{\chi}^\pm}_1$, while higgsino scenarios are much less constrained. For illustration see Fig. \[fig:charginomixing\], which shows the SMS-allowed points in the ${\tilde{\nu}}_{\tau_1}$ versus ${\widetilde{\chi}^\pm}_1$ mass plane—here the color map gives the size of the $U_{11}$ entry of the chargino mixing matrix, indicating to the wino/higgsino content of the ${\widetilde{\chi}^\pm}_1$. As can be seen, in the region that is in principle constrained by the SMS results the surviving points feature ${\widetilde{\chi}^\pm}_1$s that have a large higgsino admixture ($|U_{11}|\lsim0.5$). These points have a lower ${\widetilde{\chi}^+}_1{\widetilde{\chi}^-}_1$ production cross section and the ${\widetilde{\chi}^\pm}_1$ decays preferably into $\tau{\tilde{\nu}}_{\tau_1}$ since the higgsino decay to $e,\,\mu$ is Yukawa suppressed; $\tau^+\tau^-+{\rm MET}$ is however a more difficult signature experimentally and thus only constrains a small strip at low ${\tilde{\nu}}$ mass, cf. the purple points in Fig. \[fig:breakdown-snu-c1\]. ![Allowed points in the ${\tilde{\nu}}_{\tau_1}$ vs. ${\widetilde{\chi}^\pm}_1$ mass plane, with the color code indicating the wino/higgsino content of the ${\widetilde{\chi}^\pm}_1$ ($|U_{11}|=1$ means a pure wino while $|U_{11}|=0$ means a pure higgsino). []{data-label="fig:charginomixing"}](plots/allowed_U11_mC1_snutau1.png){height="7cm"} Missing topologies {#missing-topologies .unnumbered} ------------------ The next question to ask is which are the most important signatures not covered by SMS results. Such information can be used to improve on the interpretation of the LHC searches for new physics. We call these uncovered signatures “missing topologies”. For any point passed through , we keep up to ten missing topologies sorted by their $\sigma \times {\mbox{\ensuremath{\mathcal{B}}}}$. To avoid double counting, missing topologies are evaluated after mass and invisible compressions. The total weight is computed by summing over all diagrams giving the same topology,  ignoring the mass vector of the SUSY states involved. Moreover, $l=e, \mu$ lepton flavors appearing in the final state are summed over (light quark flavors are always summed over). In the following, we only consider MSSM+RN scan points which are not excluded, and we demand that missing topologies have $\sigma \times {\mbox{\ensuremath{\mathcal{B}}}}\ge 1$ fb. The results can be presented in two ways, either by showing the most frequent missing topologies in a certain parameter space, or by selecting for each parameter point the missing topology with the highest cross section. ![Missing topologies with highest $\sigma \times {\mbox{\ensuremath{\mathcal{B}}}}$ in the ${\tilde{\nu}}_{\tau_1}$ vs. $\tilde g$ mass plane.[]{data-label="fig:missingtopos-gluino"}](plots/missing_mgluino_snutau1_1fb_no_WW.png){height="7cm"} We choose the latter approach to show in Fig. \[fig:missingtopos-gluino\] the missing topologies in the sneutrino- vs. gluino-mass plane. The various processes are denoted in the bracket notation of , explained in [@Kraml:2013mwa]. The structure is for the decay chains (“branches”) of the two initially produced SUSY particles; each branch contains inner brackets for each vertex, containing in turn the lists of outgoing standard model particles. Thus, denotes gluino-pair production with one gluino decaying into $b\bar b$ + MET (via $\tilde g\to b\bar b{\widetilde{\chi}^0}$, ${\widetilde{\chi}^0}\to\nu{\tilde{\nu}}$) and the other one into $q\bar q$ + MET. Likewise, denotes gluino-pair production with the first gluino decaying via $\tilde g\to q\bar q{\widetilde{\chi}^0}$, ${\widetilde{\chi}^0}\to\nu{\tilde{\nu}}$ and the other one via $\tilde g\to q\bar q'{\widetilde{\chi}^\pm}$, ${\widetilde{\chi}^\pm}\to l{\tilde{\nu}}$.[^5] It is apparent that many points with gluino masses below about $1.2$ TeV, for which the LHC searches should have good sensitivity, are not excluded by the SMS results because they feature “mixed topologies”, where the two pair-produced gluinos undergo different decays ( one gluino decaying into $b\bar b$ and the other one into light jets). Since the SMS results for pair-produced sparticles always assume two identical branches, these cases cannot be constrained by . Moreover, hadronic final states with additional leptons, as they arise from gluino decays into charginos and the chargino decaying further into a charged lepton ($e,\mu$ or $\tau$) plus the LSP, do not have any SMS equivalent. Finally, there are no SMS results available for $\tilde g \to tb {\widetilde{\chi}^\pm}_j$, no matter of whether the chargino has any visible decays. It is also worth noting that over a large part of the parameter space single lepton + MET ( in bracket notation) is the most important missing topology. This signature arises from ${\widetilde{\chi}^0}_i{\widetilde{\chi}^\pm}_j$ production; its importance is corroborated in Fig. \[fig:missingtopos-chargino\], where one can see that it is indeed dominating the whole sneutrino- vs. chargino-mass plane. (There are also cases where single $W$ + MET is dominant.) The cross section for single lepton + MET production, shown in Fig. \[fig:singlelepton\], can be very large and should give important additional constraints on the model. While searches for single lepton + MET were performed by both ATLAS [@ATLAS:2014wra] and CMS [@Khachatryan:2014tva], unfortunately no suitable SMS interpretation exists for these analyses. It would be extremely interesting if the experimental collaborations provided upper limit maps and/or efficiency maps for their single lepton + MET analyses in the context of a chargino–sneutrino simplified model. Having both light EW-inos and light staus can generate decay chains with more ‘exotic’ signatures, in particular ${\widetilde{\chi}^\pm}_i{\widetilde{\chi}^0}_j$ followed by ${\widetilde{\chi}^\pm}_i \to \nu \tilde{\tau}^{\pm} \to \nu W^{\pm}{\tilde{\nu}}_{\tau_1}$ and ${\widetilde{\chi}^0}_j \to \tau^{\pm} \tilde{\tau}^{\mp} \to \tau^{\pm} W^{\mp}{\tilde{\nu}}_{\tau_1}$. This appears as (yellow points) in Fig. \[fig:missingtopos-chargino\] and is interesting because the ${\widetilde{\chi}^0}_j$ decay produces with the same rate $\tau^+W^-$ and $\tau^-W^+$: together with the chargino decay this gives rise to a same-sign $W$ signature, $W^\pm W^\pm \tau^\mp$ + MET. ![Missing topologies with highest $\sigma \times {\mbox{\ensuremath{\mathcal{B}}}}$ in the ${\tilde{\nu}}_{\tau_1}$ vs. ${\widetilde{\chi}^\pm}_1$ mass plane.[]{data-label="fig:missingtopos-chargino"}](plots/missing_mC1_snutau1_1fb_no_WW.png){height="7cm"} ![Cross sections $\sigma \times {\mbox{\ensuremath{\mathcal{B}}}}$ for the single lepton + MET missing topology for not excluded or not tested points.[]{data-label="fig:singlelepton"}](plots/Single_lepton_xsec_mC1_snutau1.png){height="7cm"} ![Missing topologies with $\sigma \times {\mbox{\ensuremath{\mathcal{B}}}}\ge 1$ fb in the sneutrino- vs. chargino-mass plane ordered by frequency of occurrence. The ordering is from top to bottom in the legend, with single tau being the most frequent missing topology, followed by single lepton ($l=e,\mu$), lepton–tau, and so on. “Other topologies” are shown on top of the legend without considering their total count (however, each single one of them is less frequent than any of the topologies denoted explicitly).[]{data-label="fig:missingtopos-EW"}](plots/missing_mC1_snutau1_1fb_freq.png){height="7cm"} Before proceeding it is instructive to take another look at the missing topologies arising from EW-ino and slepton production, but this time ordered by their frequency of occurrence. This is done in Fig. \[fig:missingtopos-EW\]. Not surprisingly we see that besides single lepton ($e$ or $\mu$), single $\tau$ is an important signature. Although it is less clean experimentally, the relative weight of single $e,\mu$ or $\tau$ + MET might potentially give information on the mass pattern of the mostly RH sneutrinos. Another important class of “missing topologies” are different-flavor dileptons ( and ). Different-flavor dileptons + MET have in principle been considered by ATLAS and CMS in the context of chargino-pair production in the MSSM with the charginos decaying either into $W^{(*)}{\widetilde{\chi}^0}_1$ [@Aad:2014vma] or into $l\nu{\widetilde{\chi}^0}_1$ via on-shell sleptons/sneutrinos [@Aad:2014vma; @Khachatryan:2014qwa]. However, the associated SMS limits do not apply to the sneutrino LSP case for various reasons. For example, the leptons from ${\widetilde{\chi}^\pm}_1\to W^{(*)}{\widetilde{\chi}^0}_1$ are generally softer than those from ${\widetilde{\chi}^\pm}_1\to l^\pm{\tilde{\nu}}_l$ decays (for the same ${\widetilde{\chi}^\pm}_1$ and LSP masses) because of the additional neutrinos in the $W$ decay. The limits for the ${\widetilde{\chi}^+}_1{\widetilde{\chi}^-}_1\to 2\times \tilde l\nu ({\rm or}~\tilde\nu l)\to 2\times l\nu{\widetilde{\chi}^0}_1$ simplified model are also not applicable because they involve an additional intermediate mass scale. Finally, the topology again gives rise to same-sign $W$’s, see the red triangles in Fig. \[fig:missingtopos-EW\]. Similarly it is possible to have same sign $\tau$’s arising from (black stars). In this case, after ${\widetilde{\chi}^0}_i{\widetilde{\chi}^\pm}_j$ production, the decay chain is ${\widetilde{\chi}^0}_i \to W^{\mp} {\widetilde{\chi}^\pm}_k \to \ W^{\mp} \tau^{\pm}{\tilde{\nu}}_{\tau_1}$ and ${\widetilde{\chi}^\pm}_j \to \tau^{\pm} {\tilde{\nu}}_{\tau_1}$. Complementarity with direct DM searches {#complementarity-with-direct-dm-searches .unnumbered} --------------------------------------- ![Complementarity of LHC and direct DM detection experiments. The panel on the left shows SMS allowed, excluded and not tested points in the plane of $\sigma^{\rm SI}_n$ vs. $m_{{\tilde{\nu}}_{\tau_1}}$. The panel on the right shows the breakdown of most constraining analyses for the points that are excluded by the SMS limits (for the sake of comparison, the allowed points are shown in grey). In both panels, the solid magenta lines and the dashed blue lines are the current exclusion limit by LUX and the forecasted sensitivity of XENON1T experiment respectively, while the dashed light green line corresponds to the predicted neutrino coherent scattering on nuclei.[]{data-label="fig:complementarity"}](plots/complementarity_summary_msigma.png "fig:"){height="7cm"}![Complementarity of LHC and direct DM detection experiments. The panel on the left shows SMS allowed, excluded and not tested points in the plane of $\sigma^{\rm SI}_n$ vs. $m_{{\tilde{\nu}}_{\tau_1}}$. The panel on the right shows the breakdown of most constraining analyses for the points that are excluded by the SMS limits (for the sake of comparison, the allowed points are shown in grey). In both panels, the solid magenta lines and the dashed blue lines are the current exclusion limit by LUX and the forecasted sensitivity of XENON1T experiment respectively, while the dashed light green line corresponds to the predicted neutrino coherent scattering on nuclei.[]{data-label="fig:complementarity"}](plots/complementarity_breakdown_msigma.png "fig:"){height="7cm"} ![On the left allowed (grey), excluded (red) and not tested (blue and cyan) points are shown in the plane of sneutrino mass versus mixing angle. The subset of points with exceedingly small $\sigma\times{\mbox{\ensuremath{\mathcal{B}}}}$ at the LHC but in the reach of XENON1T is visualised in blue. On the right, we show the SMS allowed points in the sneutrino mass versus mixing angle plane, subdivided in blue points, which are in the reach of XENON1T and in light green (light grey) points, which are above (below) the neutrino background.[]{data-label="fig:complementarity1"}](plots/complementarity_summary_msinth.png "fig:"){height="7cm"}![On the left allowed (grey), excluded (red) and not tested (blue and cyan) points are shown in the plane of sneutrino mass versus mixing angle. The subset of points with exceedingly small $\sigma\times{\mbox{\ensuremath{\mathcal{B}}}}$ at the LHC but in the reach of XENON1T is visualised in blue. On the right, we show the SMS allowed points in the sneutrino mass versus mixing angle plane, subdivided in blue points, which are in the reach of XENON1T and in light green (light grey) points, which are above (below) the neutrino background.[]{data-label="fig:complementarity1"}](plots/complementarity_msinthsigma.png "fig:"){height="7cm"} Let us finally turn to the complementarity of LHC and direct DM searches—recall that all points in our scans are consistent with DM constraints, as described in Table \[tab:co\]. In Fig. \[fig:complementarity\], left panel, we plot the allowed (gray), excluded (red) and not tested points (cyan) as a function of the sneutrino mass and the SI scattering cross section. In the same plot we also show the forecasted sensitivity of XENON1T after two years of scientific run [@Aprile:2012zx] and the predicted value for neutrino coherent scattering on nuclei [@Billard:2013qya], which can be an irreducible background for direct detection experiments. From this plot, the complementarity between the two type of searches is striking. Points with a SI elastic cross section well below the neutrino background, and hence not detectable by direct detection experiments, are already excluded by SMS results. On the other hand, a bulk of points allowed (or even more interestingly, not tested) by SMS results is well in the reach of XENON1T, expected to start running in 2015. Notice however that there still exist combinations of parameters that allow sneutrino DM to escape both direct detection and LHC searches, represented by the cyan points below the neutrino background curve. In the MSSM+RN, DM direct searches are basically sensitive to the mass of the LSP and its couplings with the Higgs and $Z$ bosons. The rest of the SUSY mass spectrum is not relevant. This is different with respect to the MSSM with the neutralino LSP, where the interaction with the quarks is mediated as well by squarks on $t$-channel. This is clearly visible in the right panel of Fig. \[fig:complementarity\], which shows the most constraining SMS analyses. In Figs. \[fig:breakdown-snu-gluino\] and \[fig:breakdown-snu-c1\] these SMS analyses are typically correlated with the gluino or chargino mass, while now they are scattered all over the $\sigma^{\rm SI}_n$ versus $m_{{\tilde{\nu}}_{\tau_1}}$ plane. The same set of allowed, excluded and not tested points are plotted as a function of the sneutrino mixing angle in the left panel of Fig. \[fig:complementarity1\]. The bulk of not tested points in the reach of XENON1T (dark blue points) has, as expected, relatively large mixing angles, corresponding to sizeable contributions from $Z$ boson exchange to the SI scattering cross section. Excluded red points are scattered everywhere in the $\sin\theta_{{\tilde{\nu}}}$ vs. $m_{{\tilde{\nu}}_{\tau_1}}$ plane and probe also very RH sneutrinos. In the right panel of Fig. \[fig:complementarity1\] we see that among the allowed points, XENON1T can constrain a large portion of the sneutrino parameter space, while the very RH sneutrinos will remain inaccessible to future direct detection detectors. In general the points with negligible mixing angles have ${\tilde{\nu}}_{\tau_1}$ as LSP and the neutralino as NLSP, which tends to be almost degenerate with chargino. The relic density is then actually achieved by co-annihilation of neutralino-chargino and then communicated to the mostly sterile LSP (see [@Arina:2013zca] for details). Such scenarios are very difficult to test. Conclusions {#sec:concl} =========== Scenarios with a sneutrino as the LSP are an interesting alternative to MSSM models with neutralino LSPs. Indeed in SUSY models with a RH neutrino superfield (MSSM+RN) the fermionic field contributes to neutrino masses while the scalar field contributes to the DM candidate, which is a mixed, however mostly RH, sneutrino. The collider phenomenology of the MSSM+RN can be quite different from the typical MSSM case. It is therefore interesting and relevant to ask how the SUSY search results from Run 1 of the LHC, which were mostly designed with the MSSM in mind, constrain sneutrino LSP scenarios. To address this question, we used  for testing the MSSM+RN against more than 60 results from CMS and ATLAS searches in the context of so-called Simplified Model Spectra (SMS). More precisely, by considering the model parameter space where the sneutrino is a good DM candidate compatible with all current constraints, we assessed 1.) the constraining power of the current SMS results on such scenarios and 2.) the most relevant signatures not covered by the SMS approach. Concerning point 1.), we found that the dilepton + MET searches are among the most relevant ones, constraining sneutrino masses up to about 210 GeV and mostly wino-like charginos up to $m_{{\widetilde{\chi}^\pm}_1}\approx 440$ GeV. It is important to note here that this amounts to re-interpreting the ATLAS and CMS searches for $pp\to \tilde l^+\tilde l^-\to l^+ l^-{\widetilde{\chi}^0}_1{\widetilde{\chi}^0}_1$ in terms of $pp\to{\widetilde{\chi}^+}_1{\widetilde{\chi}^-}_1\to l^+ l^-{\tilde{\nu}}_l{\tilde{\nu}}_l$ (the validity of this is discussed in Appendix \[sec:efficiency\]). Hadronic SUSY searches exclude gluinos masses up to $m_{\tilde{g}}\approx1200$ GeV and LSP masses up to $m_{{\tilde{\nu}}_1}\approx 500$ GeV. Nonetheless in general we find that only a very limited portion of the parameter space can be properly excluded by SMS results. For most points in the $(m_{\tilde{g}}, m_{{\tilde{\nu}}_1})$ or $(m_{{\widetilde{\chi}^\pm}_1}, m_{{\tilde{\nu}}_1})$ planes there exist parameter combinations that allow to avoid all limits. Indeed, most of the parameter space is either allowed (SMS constraints exist for the specific topologies of the point but all $\sigma\times{\mbox{\ensuremath{\mathcal{B}}}}$ of these topologies are below their 95% CL upper limits) or not tested at all (there are no existing SMS constraints for the specific topologies of the point or each topology has a $\sigma\times{\mbox{\ensuremath{\mathcal{B}}}}$ which is smaller than 1 event at LHC Run 1). Direct DM searches are complementary to the SMS constraints: many points that are not tested by SMS results can potentially be excluded by XENON1T. Vice versa, points well below the neutrino background, hence not reachable by future DM detectors, are already excluded by SMS results. The second main result of this paper concerns point 2.),  the study of the allowed points in terms of missing topologies. In the hadronic sector, pair-produced gluinos with masses well in the reach of LHC Run 1 are not constrained because they feature one or more of the following: - additional leptons: since the gluino cannot directly decay into the sneutrino LSP, the hadronic final state is often accompanied by leptons; - mixed topologies: each of the pair-produced gluinos undergoes a different decay; - the gluinos decay into $tb$ final states. None of these possibilities are covered by the current SMS results. Note here that the last two items are also common in the MSSM, as described in [@Kraml:2013mwa]. For EW production, missing topologies include: - single leptons; - single $W$s; - different-flavour opposite-sign leptons; - same-sign $W$’s or same-sign taus (accompanied respectively by additional leptons/taus, or $W$s). While such signatures have been searched for by the SUSY and/or exotics groups in ATLAS and CMS, the results do not exist in terms of appropriate SMS interpretations. Such an SMS interpretation would be very interesting in particular for the mono-lepton + MET case, which promises to have a considerable impact for constraining the MSSM+RN model.[^6] A final comment is in order. While the SMS approach is very convenient for the characterisation of new physics signatures and vast surveys of parameter spaces, it clearly has its limitations. Given the high interest in non-standard SUSY (and other new physics) scenarios, we urge the experimental collaborations to document their analyses in a way that they can conveniently be re-casted in public simulation frameworks like [CheckMATE]{} [@Drees:2013wra] or the [MadAnalysis]{}5 PAD [@Dumont:2014tja]. (See also the recommendations in [@Dumont:2014tja] and [@Kraml:2012sg] in this context). This would allow to go beyond the limitations of SMS approach and give a much more rigorous assessment of the constraints in a large variety of new physics models, including the sneutrino DM scenario discussed in this paper. Unfortunately we are still a long way from this. Acknowledgements {#acknowledgements .unnumbered} ================ SK, SuK, and UL thank their colleagues from the [SModelS]{} developer team, Andre Lessa, Veronica and Wolfgang Magerl, Michael Traub and Wolfgang Waltenberger for many helpful discussions. We also thank B. Dumont, B. Fuks, E. Conte and D. Sengupta for helpful discussions on recasting with [MadAnalysis]{}5. SuK moreover acknowledges discussions with R. Schoeffbeck. This work was supported in part by the ANR project [DMAstroLHC]{} and the “Investissements d’avenir, Labex ENIGMASS”. CA is supported by the ERC project 267117 hosted by UPMC-Paris 6, PI J. Silk. MECC is supported by Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP). SuK is supported by the “New Frontiers” program of the Austrian Academy of Sciences. UL is supported by the Labex ENIGMASS; she also gratefully acknowledges the hospitality of LPSC Grenoble during research visits prior to starting her PhD thesis at the LPSC. Validity of slepton search results for chargino-pair production with decay into lepton+sneutrino {#sec:efficiency} ================================================================================================ In the spirit of the  philosophy, we apply SMS constraints for the ${{l}}^+{{l}}^- + {\rm MET}$ topology, obtained in the context of pair production of charged sleptons, $pp\to\tilde{{l}}^+\tilde{{l}}^-$ followed by $\tilde{{l}}^{\pm}\to {{l}}^\pm\tilde\chi^0_1$ to the case of chargino-pair production $pp\to\tilde\chi_1^+\tilde\chi_1^-$ followed by ${\widetilde{\chi}^\pm}_1\to {{l}}^\pm\tilde\nu_{{{l}}1}$, despite the opposite spin configuration. This can only be valid if the signal selection efficiencies in both scenarios are comparable. To test this assumption, we use the recast code [@PAD-code-ATLAS-SUSY-2013-11] for ATLAS search in final states with two leptons and missing transverse momentum, ATLAS-SUSY-2013-11 [@Aad:2014vma], which is available in the framework of the [MadAnalysis]{}5 “Public Analysis Database” [@Dumont:2014tja]. We consider two benchmark scenarios in the simplified-model spirit, an MSSM one with $(m_{\tilde{{l}}^{\pm}},\ m_{\tilde\chi^0_1})=(270,\, 100)$ GeV and an MSSM+RN one with $(m_{{\widetilde{\chi}^\pm}_1},\ m_{\tilde\nu_1})=(270,\, 100)$ GeV. Events are generated with [MadGraph]{} 5 [@Alwall:2011uj; @Alwall:2014hca] and [Pythia]{} 6.4 [@Sjostrand:2006za] and then passed through [Delphes]{} 3 [@deFavereau:2013fsa] for the simulation of the detector effects.[^7] For simplicity, in the following we restrict our study to pair-production of selectrons for the MSSM case, and pair-production of charginos decaying exclusively via electrons in the MSSM+RN case. The event selection requires two opposite sign (OS), same flavor (SF) leptons with high transverse momentum, concretely $p_T > 35$ GeV and $p_T > 20$ GeV.[^8] Figure \[fig:leptonpT\] compares the $p_T$ distributions in the two benchmark scenarios, in the left panel for the harder electron, $e_1$, in the right panel for the second electron, $e_2$. The bin sizes are chosen such that the first bin corresponds to the events that do not pass the $p_T > 35$ GeV (left panel) or $p_T > 20$ GeV (right panel) requirement. We see that the electrons originating from selectron-pair production tend to be harder than those originating from chargino-pair production. ![Comparison of the $p_T$ distributions of electrons originating from selectron decays in the MSSM and from chargino decays in MSSM+RN, at the level of reconstructed events. The benchmark scenarios used are $(m_{\tilde{{l}}^{\pm}},\ m_{\tilde\chi^0_1})=(270,\, 100)$ GeV for the MSSM case and $(m_{{\widetilde{\chi}^\pm}_1},\ m_{\tilde\nu_1})=(270,\, 100)$ GeV for the MSSM+RN case. See text for details.[]{data-label="fig:leptonpT"}](plots/pt_e1_reconstructed.png "fig:"){width="50.00000%"} ![Comparison of the $p_T$ distributions of electrons originating from selectron decays in the MSSM and from chargino decays in MSSM+RN, at the level of reconstructed events. The benchmark scenarios used are $(m_{\tilde{{l}}^{\pm}},\ m_{\tilde\chi^0_1})=(270,\, 100)$ GeV for the MSSM case and $(m_{{\widetilde{\chi}^\pm}_1},\ m_{\tilde\nu_1})=(270,\, 100)$ GeV for the MSSM+RN case. See text for details.[]{data-label="fig:leptonpT"}](plots/pt_e2_reconstructed.png "fig:"){width="50.00000%"} The analysis further requires the invariant mass of the lepton pair to be outside the $Z$ window, and $\tau$s and jets are vetoed. Finally, three signal regions are defined by thresholds on the $m_{T2}$ (“stransverse mass”) variable [@Lester:1999tx; @Cheng:2008hk] that is used for reducing the $t\bar t$ and $Wt$ backgrounds: $m_{T2} > 90$, $> 120$ and $> 150$ GeV. The $m_{T2}$ distributions after the preselection cuts are shown in Fig. \[fig:mT2\]. It can be seen that the distributions intersect around the minimum required value of $m_{T2} = 90$ GeV; events with electrons originating from chargino decays are more likely to pass this cut. ![Comparison of the $m_{T2}$ distributions for the two benchmark scenarios after all preselection cuts.[]{data-label="fig:mT2"}](plots/mt2.png){width="50.00000%"} To see the net effect on the signal efficiencies, Table \[tab:cutflow\] shows the complete cut-flow comparison for the two benchmark scenarios. As expected, differences arise in the first cut, selecting high $p_T$ OS lepton pairs, and when applying the lower bounds for $m_{T2}$. Because of the softer $p_T$ distribution in case of chargino production+decay, there are fewer events passing the first cut for this scenario. However, the opposite is true for the $m_{T2}$ cut. Ultimately, the efficiencies are comparable in all signal regions, and even somewhat higher for the MSSM+RN scenario. To check that this is still true closer to the kinematic edge, we reproduce the cut-flows for a second set of benchmark scenarios with an LSP mass of 200 GeV. As can be seen in Table \[tab:cutflow2\], we find a similar behaviour in this case. We conclude that we can safely apply the SMS upper limits given by the experimental collaborations in the context of slepton-pair production in the MSSM to constrain chargino-pair production followed by decays into $l{\tilde{\nu}}_l$ in the MSSM+RN. [| l | c | c |]{} Cut & Slepton production & Chargino production\ \ Initial number of events & 50000 & 50000\ 2 OS leptons & 35133 & 33464\ $m_{ll} > 20$ GeV & 35038 & 33337\ $\tau$ veto & 35007 & 33318\ $ee$ leptons & 35007 & 33318\ jet veto & 20176 & 19942\ $Z$ veto & 19380 & 18984\ \ $m_{T2} > 90$ GeV & 11346 & 11594\ $m_{T2} > 120$ GeV & 8520 & 8828\ $m_{T2} > 150$ GeV & 5723 & 5926\ [| l | c | c |]{} Cut & Slepton production & Chargino production\ \ Initial number of events & 50000 & 50000\ 2 OS leptons & 29291 & 27244\ $m_{ll} > 20$ GeV & 29082 & 26964\ $\tau$ veto & 29050 & 26956\ $ee$ leptons & 29050 & 26956\ jet veto & 16834 & 16114\ $Z$ veto & 15281 & 14025\ \ $m_{T2} > 90$ GeV & 3028 & 3198\ $m_{T2} > 120$ GeV & 85 & 140\ $m_{T2} > 150$ GeV & 0 & 0\ Lifetimes of long-lived particles {#sec:lifetime} ================================= As mentioned in Sec. \[sec:sms\], a considerable number of the scan points comprise long-lived sparticles. These occur mostly when enforcing light gluinos or squarks; in this case about 30 % of the points feature long-lived particles, while the fraction is below 1 % without this constraint. The long-lived particles are predominantly gluinos (85 %), mostly in the case where it is the NLSP, and in a few points where $\tilde{\chi}^0_1$ is slightly (up to about 50 GeV) lighter than the gluino. Apart from that we find points with long-lived stops or staus in case they are the NLSP, as well as single points with long-lived charginos. Here we will focus on the long-lived gluinos and stops, long-lived staus have been discussed before in [@Arina:2013zca].\ ![Lifetimes $c\tau$ in \[m\] for long-lived gluinos, the color code indicates the LSP mass (left) and the sneutrino mixing angle (right).[]{data-label="fig:gluinoLT"}](plots/ctau_gluino_sneutrino.png "fig:"){width="52.00000%"} ![Lifetimes $c\tau$ in \[m\] for long-lived gluinos, the color code indicates the LSP mass (left) and the sneutrino mixing angle (right).[]{data-label="fig:gluinoLT"}](plots/ctau_gluino_theta.png "fig:"){width="52.00000%"} In the MSSM long-lived gluinos appear when all squarks are extremely heavy, e.g. in split-SUSY scenarios. In case of the MSSM+RN with a sneutrino LSP additional causes come into play. If the gluino is the NLSP, its decay will proceed only via virtual squarks and gauginos, yielding an effective four body decay, $\tilde{g} \rightarrow q q \nu \tilde{\nu}$ (virtual $\tilde{q}$ and $\tilde{\chi}^0$) or $\tilde{g} \rightarrow q q' l \tilde{\nu}$ (virtual $\tilde{q}$ and $\tilde{\chi}^{\pm}$). The gluino lifetime will therefore depend not only on the squark mass, but also on the gaugino masses and mixings, as well as the sneutrino mixing angle. Meta-stable gluinos can thus appear even if the squarks are not completely decoupled. The gluino lifetime as a function of its mass is shown in Fig. \[fig:gluinoLT\]. The left plot illustrates the depencence on the sneutrino mass, the right plot the dependence on the sneutrino mixing. We can distinguish two general regions. First, we observe an exponential dependence of the lifetime on the gluino mass for decay lengths of $10$ mm up to $10^4$ m. Here the lifetime is largely independent of the sneutrino mass. Moreover lifetimes at constant gluino masses are longer for heavier squarks and gauginos. In this region we generally find large mixing angles $\sin\theta_{\tilde{\nu}}$, but heavy gauginos and squarks. Points with very small mixing angles may also appear in this region, in the case that the mass of the lightest neutralino is below the gluino mass. The second region, with lifetimes longer than $10^4$ m, and up to $10^{17}$ m, shows a very different behaviour. We can see a clear correlation between gluino and sneutrino masses in this region, with longer lifetimes found for smaller mass splittings. The lifetimes moreover increase when going to very small sneutrino mixing angles, with the maximum lifetimes achieved for $\sin\theta_{\tilde{\nu}}$ going to zero. Likewise, if the stop is the NLSP[^9] and has a small mass difference with the sneutrino, it can be long-lived, see Fig. \[fig:lifetime\]. As seen for the gluinos, the lifetime depends strongly on the sneutrino mixing. ![Lifetimes $c\tau$ in \[m\] for long-lived stops, the color code indicates the LSP mass (left) and the sneutrino mixing angle (right).[]{data-label="fig:lifetime"}](plots/ctau_stop_sneutrino.png "fig:"){width="52.00000%"} ![Lifetimes $c\tau$ in \[m\] for long-lived stops, the color code indicates the LSP mass (left) and the sneutrino mixing angle (right).[]{data-label="fig:lifetime"}](plots/ctau_stop_theta.png "fig:"){width="52.00000%"} Both long lived gluinos and long lived stops can be constrained by searches for R-hadrons, see [@Chatrchyan:2013oca; @ATLAS:2014fka] for R-hadrons escaping the detector, [@Aad:2013gva] for stopped R-hadrons, or [@ATLAS-CONF-2014-037] for metastable gluinos decaying in flight inside the detector. However, large uncertainties arise from modeling both the hadronisation and the strong interaction of the R-hadron with the detector. Therefore no collider constraints on long-lived sparticles have been included.\ Additionally, cosmological constraints become important for gluino lifetimes of about $100$ s ($10^{10}$ m) [@Arvanitaki:2005fa]. Lifetimes of that order would affect the fraction of heavy nuclei produced during the Big Bang nucleosynthesis. Longer lifetimes can further be constrained by searches for diffuse gamma ray background, distortions in the CMBR and heavy isotopes. [^1]: Pure right-handed sterile sneutrinos can also be viable (non-thermal, depending on the model) DM candidates, as discussed [*e.g.*]{} in [@Asaka:2005cn; @Deppisch:2008bp; @Cerdeno:2009dv; @Khalil:2011tb; @Choi:2012ap]. [^2]: Simplified Models are effective-Lagrangian descriptions involving only a small number of new particles. They were designed as a useful tool for the characterization of new physics, see  [@Alwall:2008ag; @Alves:2011wf]. [^3]: The [**Tx**]{} names are explained in the SMS dictionary on http://smodels.hephy.at/wiki/SmsDictionary [^4]: This occurs if no simplified model result exists for the signal topologies of the point considered, but also if the mass vector of a topology lies outside that of the experimental constraint. Moreover, we include here also the points for which all signal topologies are discarded because of ${\mbox{\ensuremath{\sigma\times\mathcal{B}}}\xspace}< \sigma_{\rm cut}$. [^5]: More generically, denotes production of $XY$ with $X$ undergoing a 1-step decay chain, $X\to b\bar b$ + MET ([branch1=\[vertex1\]=\[\[b,b\]\]]{}) and $Y$ undergoing a 2-step decay chain, $Y\to q\bar q+Z \to q\bar q+l+{\rm MET}$ ([branch2=\[vertex1,vertex2\]=\[\[jet,jet\],\[l\]\]]{}); $X$ can be different from $Y$ or both can be the same. [^6]: This could be done analogous to the existing ${\widetilde{\chi}^\pm}_1{\widetilde{\chi}^0}_2$ (${\widetilde{\chi}^\pm}_1\to W^\pm{\widetilde{\chi}^0}_1$, ${\widetilde{\chi}^0}_2\to Z^0{\widetilde{\chi}^0}_1$) simplified models that are already assessed by the ATLAS and CMS SUSY groups, but with the chargino decaying to 100% into $l^\pm{\tilde{\nu}}_l$ and the neutralino decaying 100% into $\nu{\tilde{\nu}}_l$. However, since the chargino and neutralino masses need not be degenerate, we propose to consider as a first step ${\widetilde{\chi}^\pm}_1{\widetilde{\chi}^0}_1$ production followed by ${\widetilde{\chi}^\pm}_1\to l^\pm{\tilde{\nu}}_l$ and ${\widetilde{\chi}^0}_1\to \nu{\tilde{\nu}}_l$. The cross section upper limits should be provided in the chargino- versus sneutrino mass plane for different neutralino masses, for the cases $l=e,\mu$ and $l=\tau$, and if computationally feasible also for $l=e,\mu,\tau$ assuming equal rates. [^7]: Note that for the reconstruction of events with a sneutrino LSP it is necessary to define the sneutrino as MET, by adding a corresponding [EnergyFraction]{} entry in the [Delphes]{} card. [^8]: We consider here only the part of the analysis that is relevant for the SMS result used to constrain the sneutrino LSP scenario in Section 4. [^9]: If the stop mass is close to the gluino mass, both stop and gluino may be long-lived.
{ "pile_set_name": "ArXiv" }
--- abstract: 'In this article, a survey of several important equilibrium concepts for decentralized networks is presented. The term decentralized is used here to refer to scenarios where decisions (e.g., choosing a power allocation policy) are taken autonomously by devices interacting with each other (e.g., through mutual interference). The iterative long-term interaction is characterized by stable points of the wireless network called equilibria. The interest in these equilibria stems from the relevance of network stability and the fact that they can be achieved by letting radio devices to repeatedly interact over time. To achieve these equilibria, several learning techniques, namely, the best response dynamics, fictitious play, smoothed fictitious play, reinforcement learning algorithms, and regret matching, are discussed in terms of information requirements and convergence properties. Most of the notions introduced here, for both equilibria and learning schemes, are illustrated by a simple case study, namely, an interference channel with two transmitter-receiver pairs.' author: - 'L. Rose, S. M. Perlaza, S. Lasaulce  and M. Debbah [^1] [^2][^3] [^4]' bibliography: - 'GT.bib' title: Learning Equilibria with Partial Information in Decentralized Wireless Networks --- Introduction {#SecIntroduction} ============ The notion of cognitive radio (CR) has gained momentum in recent years to build flexible and efficient networks. Indeed, CRs are nowadays widely accepted as a suitable solution to rationally exploit shared spectral resources and increase spectral efficiency. The main idea behind CR relies on the capability of a given radio device to self-configure its own communication parameters in an intelligent, autonomous and decentralized manner, as a result of its interaction with the environment. In this context, the choice of a particular communication configuration by a given CR is highly influenced by the choice of all other radio devices. Within this framework, non-cooperative game theory appears as a suitable paradigm to study and analyse such scenarios. Therefore, the idea of equilibrium, namely, Nash equilibrium (NE), becomes particularly relevant. Indeed, at the NE, the transmit configuration of each CR in the network is optimal with respect to the configuration of all its counterparts. Interestingly, in some cases, an equilibrium can be reached by using particular iterative procedures similar to learning processes [@Fudenberg-98]. In this article, we first present an overview of various equilibrium concepts. Later, we introduce a set of learning algorithms particularly relevant to achieving equilibrium in wireless networks. For each algorithm, we discuss the required information that each CR must possess at each iteration and the convergence properties. The rest of the paper is organized as follows. In Sec. \[def\_abb\], we briefly present the notations adopted in this paper, as well as usual game-theoretic terminology. In Sec. \[eq\_conc\], we present and discuss several important solution concepts for games, namely, the coarse correlated equilibrium, the correlated equilibrium, the Nash equilibrium, and the $\epsilon-$Nash equilibrium. In Sec. \[learning\], we discuss important learning algorithms which can, under certain conditions, converge to one of the aforementioned solutions. In Sec. \[IC\], we present an illustrative case study: the $2 \times 2$ interference channel, a simple, though very important, communication scenario, which we use as a test-bench for comparing the aforementioned algorithms. Definitions and abbreviations {#def_abb} ============================= In game theory, the normal form is a convenient mathematical representation of a game. Basically, it consists of a triplet: the set of players $\mathcal{K}=\lbrace 1,2,...,K \rbrace$, the set of actions $\mathcal{A}_k=\lbrace A_k^{(1)},...,A_k^{(N_{k})} \rbrace$, $\forall k\in\mathcal{K}$, and the utility functions $u_k({\boldsymbol}{a})$, where ${\boldsymbol}{a} \in \mathcal{A} = \mathcal{A}_1\times\mathcal{A}_2\times...\times\mathcal{A}_K$ is an action profile/vector. With a slight abuse of notation, we denote by ${\boldsymbol}{a}_{-k} \in \mathcal{A}_{-k}$ the vector of actions of all players except the $k$-th player and we write the vector ${\boldsymbol}{a}$ as $(a_k,{\boldsymbol}{a}_{-k})$ to stress the $k$-th component. For instance, the set of players can consist of the set of wireless terminals present in the network, the action set can be any feasible vector of transmit powers, and the utility function can be the spectral efficiency. Other components are also possible for the game representation and they depend on the scope and purpose of the network design. We denote by $\triangle\left( \mathcal{A}\right)$ the set of all possible probability distributions over the whole set of actions $\mathcal{A}$, and by $\triangle(\mathcal{A}_k)$ the set of all possible probability distributions of user $k$ over its action set. We refer to the elements of the set $\mathcal{A}_k$ as *actions* of player $k$ and those of the set $\triangle(\mathcal{A}_k)$ as *strategies* of player $k$. A given strategy of player $k$ is denoted by ${\boldsymbol}{\pi_{k}}=(\pi_{k,A_k^{(1)}},...,\pi_{k,A_k^{(N_{k})}}) \in \triangle\left(\mathcal{A}_k\right)$, where $\pi_{k,A_k^{(n_{k})}}$ represents the probability that player $k$ plays action $A_K^{(n_{k})}$. We indicate by ${\boldsymbol}{\phi}=(\phi_{A^{(1)}},...,\phi_{A^{(N)}}) \in \triangle\left( \mathcal{A} \right)$, with $N=\prod_{j=1}^{K}N_j$, a given joint probability distribution over the set $\mathcal{A}$, with $\phi_{A^{(n)}}$ being the probability of observing $A^{(n)}$ as an outcome of the game. From coarse correlated equilibria to Nash equilibria {#eq_conc} ==================================================== The most general type of equilibria we use in this paper is the so called *coarse correlated equilibrium* (CCE) [@Young-2004]. The idea behind CCE is that actions chosen by the players of a game may be correlated. For instance, correlation may appear when a common broadcast signal is observed by several transmitters choosing their transmit configuration, e.g., a power control policy. We call the signals received by the players recommendations. In such a context, a CCE is a probability distribution ${\boldsymbol}{\phi} \in \triangle\left( \mathcal{A} \right)$ over the set of action profiles of the game from which no player has interest in unilaterally deviating. The realizations of this joint distribution ${\boldsymbol}{\phi} $ are the recommendations. Mathematically, this can be written as follows. \[DefCCE\]*A joint probability distribution ${\boldsymbol}{\phi} \in \triangle\left(\mathcal{A}\right)$ is a CCE if $\forall k \in\mathcal{K}$ and $\forall a'_{k} \in \mathcal{A}_k$ it holds that $$\label{CCE_eq} \sum_{{\boldsymbol}{a}\in\mathcal{A}}u_k({\boldsymbol}{a})\phi_{{\boldsymbol}{a}} \geq \sum_{a_{-k}\in\mathcal{A}_{-k}}u_{k}(a'_{k},{\boldsymbol}{a}_{-k})\phi_{-k,{\boldsymbol}{a}_{-k}},$$ where $\phi_{-k,a_{-k}}=\sum_{a_k \in \mathcal{A}_k} {\boldsymbol}{\phi}_{\left(a_k,{\boldsymbol}{a}_{-k}\right)}$ is the marginal probability distribution w.r.t. $a_k$.* An important remark is that, following the notion of CCE, players are assumed to decide, *before* receiving the recommendation, whether to commit to follow it or not. At a CCE, all players are willing to commit to follow the recommendation given that all the others also choose to commit. That is, if a single player decides not to commit to follow the recommendations, it experiences a lower (expected) utility. A special case of CCE is the *correlated equilibrium* (CE, [@Young-2004]). The difference between the CCE and the CE is that, in the latter, players choose whether to follow or not a given recommendation, *after* it has been received. Therefore, there is not *a priori* commitment. It follows, in particular, that every CE is a CCE [@Young-2004]. Now, if the players choose their strategy following independent individual probability distributions ${\boldsymbol}{\pi}_k \in \triangle\left( \mathcal{A}_k \right)$, i.e., $\phi_{a} = \prod_{j=1}^{K} \pi_{j,a_j}$ in , we obtain from Def. \[DefCCE\], the definition of *mixed Nash equilibrium* (MNE); the MNE is, clearly, a special case of CE and thus, a special case of CCE. A MNE is, therefore, a vector of individual probability distributions ${\boldsymbol}{\pi} = \left({\boldsymbol}{\pi}_1, \ldots, {\boldsymbol}{\pi}_K \right)$ which is stable to unilateral deviations, i.e., if player $k$ decides to use a different probability distribution from the corresponding ${\boldsymbol}{\pi}_k$, then it observes a lower (expected) utility. As shown in [@Fudenberg-Tirole-1991], this type of equilibria always exists in games with finite number of players and finite action sets. For more results on the existence and multiplicity of NE, the reader is referred to [@Lasaulce-Tutorial-09]. The finiteness assumption is especially relevant when a wireless terminal has to select a given communication setting, e.g., a channel, a constellation size, or a transmit power level. We further introduce the concept of *pure NE* (PNE). A PNE is obtained by restricting the players to deterministically choose one of their actions instead of choosing it by following a probability distribution. A PNE is therefore a special case of MNE where the individual probability distribution is a Dirac delta function over a given action. Thus, a PNE is a vector of actions ${\boldsymbol}{a} = \left( a_1, \ldots, a_K \right)$ stable to unilateral deviations, i.e., if player $k$ uses a different action from its corresponding $a_k$, while the others keep their equilibrium action, player $k$ observes a lower (instantaneous) utility. As a last notion of equilibrium, we introduce the idea of $\epsilon-$*equilibrium* [@Young-2004]. An $\epsilon-$equilibrium is a mixed strategy profile ${\boldsymbol}{\pi} = \left({\boldsymbol}{\pi}_1, \ldots, {\boldsymbol}{\pi}_K \right)\in\triangle\left( \mathcal{A}_1\right)\times\ldots\times\triangle\left( \mathcal{A}_K\right)$ such that if only one player $k$ uses a different strategy from its corresponding ${\boldsymbol}{\pi}_k$, it does not observe a utility improvement greater than $\epsilon>0$. An instance of $\epsilon-$NE [@Young-2004] is the logit equilibrium [@Young-2004]. In what follows, some learning algorithms which may converge to CCE, CE, MNE, PNE, or $\epsilon-$equilibrium are provided. Learning Equilibria {#learning} =================== The process of learning equilibria is basically an iterative process. Each iteration of the learning process can be broadly divided into three phases: $(i)$ the observation of the environment at iteration $n$, which gives an idea to the players how well they played in the previous iteration; $(ii)$ the improvement of the strategy ${\boldsymbol}{\pi}_k(n)$ based on the current observation and $(iii)$ the selection of the action $a_k(n)$ according to the strategy ${\boldsymbol}{\pi}_k(n)$. Hence, we say that players learn to play an equilibrium, if after a given number of iterations, the strategy profile ${\boldsymbol}{\pi}(n) = \left({\boldsymbol}{\pi}_1(n), \ldots, {\boldsymbol}{\pi}_{K} (n)\right) \in \triangle\left( \mathcal{A}_1\right)\times\ldots\times \triangle\left( \mathcal{A}_K\right)$ converges to an equilibrium strategy. The purpose of this section and Sec. \[IC\] is, under the space limitations for this survey, to introduce the following set of learning algorithms: best response dynamics (BRD), fictitious play (FP), smooth fictitious play (SFP), regret matching (RM), reinforcement learning (RL) and the *joint utility and strategy learning estimation* reinforcement learning (JUSTE-RL). Then, in Sec. \[sec:discussion\], we compare such algorithms in terms of relevant features in the context of wireless communications, for instance, type of observations, type of action sets, convergence time, nature of the steady state achieved when convergence is observed and conditions for convergence. Informal definition of the learning algorithms under consideration {#sec:algo-definitions} ------------------------------------------------------------------ ### Best Response Dynamics In its most basic form, the best response dynamics relies on the following assumption: at each game stage $n \in \mathds{N}$, every player $k$ plays the action $a_k(n)$ which optimizes its own utility given the actions played by the other players. When all players play simultaneously at each stage (simultaneous-BRD), the optimization of player $k$ is done with respect to the action profile ${\boldsymbol}{a}_{-k}(n-1)$. When players play sequentially, only one player at each stage (sequential-BRD) updates its action $a_k(n)$, optimizing it with respect to the action profile $\left(a_{1}(n),\ldots, a_{k-1}(n), a_{k+1}(n-1),\ldots, a_{K}(n-1)\right)$. ### Fictitious Play The fictitious play relies on the assumption that at each stage $n$, each player $k$ knows all the past actions of all the other players, i.e., $a_j(0), \ldots, a_j(n-1)$, $\forall j \in \mathcal{K}\setminus\lbrace k \rbrace$. Based on such observations, player $k$ calculates the empirical frequencies with which each player plays its corresponding actions. We refer to these empirical frequencies as beliefs. Let us denote the belief that player $k \neq j$ has on player $j$ by the vector ${\boldsymbol}{f}_j(n) = \left(f_{j,A_j^{(1)}}(n), \ldots, f_{j,A_j^{(N_j)}}(n)\right) \in \triangle\left( \mathcal{A}_j\right)$. At each stage, all players (simultaneously or sequentially, as in the BRD) choose their current action by optimizing their expected utility with respect to the beliefs on all the other players, i.e., $a_k(n) \in {\mathrm{argmax}}_{a_k \in \mathcal{A}_k} {\mathds{E}_{{\boldsymbol}{f}(n)}\left[ u_k\left(a_k,{\boldsymbol}{a}_{-k}\right) \right]}$, where ${\boldsymbol}{f}(n) = \left({\boldsymbol}{f}_{1}(n), \ldots, {\boldsymbol}{f}_{K}(n)\right)$. ### Smooth Fictitious Play (SFP) The convergence of FP is not ensured in games with cycles and its ability to explore the whole action set is highly constrained. To overcome these issues, a simple variation of the FP has been proposed under the name of smooth fictitious play (SFP). The assumptions on which SFP relies are the same as FP and actions can be updated either simultaneously or sequentially. The main difference between SFP and FP is that, at each stage $n$, player $k$ does not choose a deterministic action. It rather builds a probability distribution ${\boldsymbol}{\pi}_k(n) \in \triangle\left( \mathcal{A}_k \right)$ to choose its action $a_k(n)$. Such a probability distribution can be interpreted as the one that maximizes a weighted sum of the original expected utility and other continuous strictly concave function. For instance, if such a function is the *entropy function* [@Young-2004], the resulting probability distribution is given by the logit probability distribution. ### Regret Matching (RM) Contrary to the case of BRD, FP and SFP, where players determine whether to play or not a particular action based on the idea of utility maximization, in RM, such a decision is made considering the notion of regret minimization. The regret that player $k$ associates with action $A_k^{(n_k)}$ is defined as the difference between the average utility the player would have obtained by always playing $A_k^{(n_k)}$ and the average utility actually achieved with the current strategy, i.e., $$r_{k,A_k^{(n_k)}}(n)=\frac{1}{n-1}\sum_{t = 1}^{n-1}(u_k(A_k^{(n_k)},{\boldsymbol}{a}_{-k}(t)) - u_k(a_k(t), {\boldsymbol}{a}_{-k}(t))). \label{regr}$$ RM relies on the assumptions that at every stage $n$, player $k$ is able to both evaluate its own utility, i.e., to calculate $u_k(a_k(n),{\boldsymbol}{a}_{-k}(n))$ and compute the utility it would have obtained if it had played any other action $a_k'$, i.e. $u_k(a_k',{\boldsymbol}{a}_{-k}(n))$. Finally, the action to be played at stage $n$ is taken following the probability distribution ${\boldsymbol}{\pi}_k(n)$, which is obtained by normalizing to one the regret vector ${\boldsymbol}{r}_k(n) = \left( r_{k,A_k^{(1)}}(n), \ldots, r_{k,A_k^{(N_k)}}(n)\right)$. ### Reinforcement Learning (RL) In the case of reinforcement learning (RL), players are modelled as automata that implement a given behavioural rule. In general, RL techniques rely on the following two conditions: $(i)$ for each player $k$, the action set $\mathcal{A}_k$ is finite and for all action profiles ${\boldsymbol}{a} \in \mathcal{A}$, the achieved utility $u_k\left( a_k, {\boldsymbol}{a}_{-k}\right)$ is bounded; $(ii)$ each player is able to periodically observe its own achieved utility. Intuitively, the idea behind CRL is that actions leading to higher utility observations in stage $n$ are granted with higher probabilities in the game stage $n+1$, and vice versa. ### Joint Utility and Strategy Estimation based - Reinforcement Learning (JUSTE-RL) A variant of the former algorithm has been proposed in [@Perlaza-Spawc2010]. The joint utility and strategy estimation behavioural rule relies on the same assumptions as the classical RL. The main difference between classical RL and JUSTE-RL is that, in the former, the observation $\tilde{u}_k(n)$ of the utility of player $k$ is used to directly modify the probability distribution ${\boldsymbol}{\pi}_k(n)$; in the latter, such an observation is used to build an estimation of the expected utility for each of the actions. Such utility estimates are, then, used in the same iteration to finally build a probability distribution ${\boldsymbol}{\pi}_k(n)$ from which action $a_k(n)$ will be drawn. Thus, each player always possesses an estimation of the expected utility it obtains by playing each of its actions. Discussion {#sec:discussion} ---------- The purpose of this section is to provide additional insights about the performance and pertinence of the learning algorithms described above in the context of decentralized wireless networks. In the following, we compare the algorithms in terms of several fundamental features. We summarize this discussion in Table \[FigTable\]. ### Observations At each iteration of a given learning algorithm, each player must obtain some information about how the other players are reacting to its current action, in order to update their strategy and choose the following action. Broadly speaking, in algorithms such as BRD, FP, SFP and RM, players must observe the actions played by all the other players. This implies that a large amount of additional signaling is required to broadcast such information in wireless networks. In some particular cases, this condition can be relaxed and less information is required [@Larsson-2009; @Leshem-Zehavi-2009]. However, this is highly dependent on the topology of the network and the explicit form of the utility function [@Scutari-Algorithms-2008]. Other algorithms, such as RL and JUSTE-RL, only require that each player observes its corresponding achieved utility at each iteration. This is in fact, their main advantage, since such information requires a simple feedback message from the receiver to the corresponding transmitters [@sastry-1994; @Perlaza-Spawc2010]. ### Knowledge and Calculation Capabilities Learning algorithms such as BRD, FP, SFP and RM involve an optimization problem at each iteration [@Fudenberg-98], that is, either the maximization of the (expected or instantaneous) utility or minimization of the regret. Therefore, generally, highly demanding calculation capabilities are required to implement them. More importantly, solving such optimization requires the knowledge of a closed-form expression of the utility function. This implies that each player knows the structure of the game, i.e., set of players, action sets, current strategies, channel realizations, etc. In this respect, RL and JUSTE-RL algorithms are more attractive since only algebraic operations are required to update the strategies. In terms of knowledge, in both RL and JUSTE-RL, players are only required to know, at each iteration, the action they actually played and the corresponding achieved utility. Indeed, one can say that players are not even aware of the presence of other players. ### Nature of the Action Sets The nature of the action sets of the game plays an important role. The BRD can be used for both continuous and discrete action sets, whereas in their standard versions FP, SFP, RM, CRL, and JUSTE-RL are designed for discrete action sets. For instance, action sets are discrete in problems where a channel, constellation size or discrete power levels must be selected, whereas continuous sets are more common in power allocation problems [@Lasaulce-Tutorial-09]. ### Steady State When a steady state is achieved by one of the algorithms under consideration, such state may correspond to one of the equilibrium notions presented in Sec. \[eq\_conc\]. In particular, when BRD and FP converge, the strategy of the players at the steady state is a NE. In the case of the RM, it converges to an element of the set of CCE. Here, we highlight the fact that, even though the notion of CCE relies on the idea of the recommendations studied in Sec. \[SecIntroduction\], it does not require the existence of recommendations to converge to an element of the set of CCE. When SFP or JUSTE-RL achieve a steady state, it corresponds to an $\epsilon$-NE. On the contrary, in the case of RL, a steady state not necessarily corresponds to a particular notion of equilibrium. ### Convergence Conditions Regarding the conditions for convergence, only sufficient conditions are available. As shown in Table \[FigTable\], the considered algorithms typically converge in certain classes of games [@Young-2004] such as dominance solvable games (DSG), potential games (PG), super-modular games (SMG), $2\times N$ non-degenerated games (NDG) or zero-sum games (ZSG). ### Synchronization In the particular case of algorithms where each player must observe the actions of the others, e.g., BRD, FP, SFP and RM, certain synchronization is required in order to allow players to know when to play and when to observe the actions of the others. In wireless communications, this requirement implies the existence of a given protocol for signalling messages exchange. Conversely, when players require only an observation of their individual utility, such a synchronization between all the players becomes irrelevant. Here, only a feedback message from the receiver to the corresponding transmitters per learning iteration is sufficient. ### Environment Learning techniques such as the BRD are highly constrained for real system implementations since they require the network to be static during the whole learning processes. On the contrary, all the other techniques allow the dynamics of the network to be captured by their statistics as long as they are stationary. This is basically because, contrary to BRD, all the other techniques determine whether to play or not a particular action based on the expected utility rather than instantaneous utility. ### Convergence Speed The speed of convergence (when it is observed) is highly influenced by the amount of information available for the players. For instance, FP, SFP and RM converge faster than JUSTE-RL since the formers calculate the expected utility relaying on a closed form expression. Conversely, the latter calculates it as the time-average of the instantaneous observations of the achieved utility. This requires a large number of observations to obtain a reliable approximation to the expected utility. We do not state any particular comment on the speed of convergence of BRD and RL since, in the former the scenario is considered fixed and the latter, it does not necessarily converge to an equilibrium strategy. However, conclusions for a particular case are stated in the following section. Case Study: The Parallel Interference Channel {#IC} ============================================= In this section, we introduce a simple but insightful example, which we use as a test-bench to compare the learning algorithms described above. Consider a parallel interference channel, that is, a set of $2$ transmitter-receiver pairs sharing a set of $S$ non-overlapping frequency bands. For the ease of presentation, assume that channel gains are time invariant during the whole transmission duration. Each transmitter chooses a single frequency band to transmit aiming to maximize its individual spectral efficiency, i.e., the ratio between the individual Shannon rate and available bandwidth. This problem has been analysed in the context of compact and convex sets of actions in [@Scutari-Algorithms-2008] and in discrete and finite sets in [@Rose-Perlaza-2011], which is the case of this section. In Figure \[fig:2x2perf\_vs\_iter\], we plot the average spectral efficiency of the system as a function of the SNR, in the case where only $2$ orthogonal channels are available. Here, all the algorithms iterate the same number of times ($40$ iterations). In Figure \[fig:2x2perf\_vs\_iter\], it is interesting to note how algorithms such as FP, SFP and RM converge always very close to the best NE, i.e, the NE associated with the highest network spectral efficiency. Nonetheless, this performance is achieved at the cost of a lot of information about the game. In particular, note that RL and JUSTE-RL are less performing, but at the same time, less demanding in terms of information. Interestingly, the BRD demands the same information assumptions than FP, SFP and RM. However, the performance is even worse that RL. This is due to the fact that BRD does not necessarily converge to a NE in this particular game. In Figure \[fig:2x2perf\_vs\_SNR\], we plot the network spectral efficiency of the algorithms as a function of the number of iterations for the case of two channels. Here, RM and BRD appear to be the best performing and worst performing algorithms, respectively. With respect to the BRD, such a performance is due to a *ping-pong* effect between two particular action profiles. In detail, since players are simultaneously selecting the channels with the highest gain, it may happen that the best channel is the same for both players. Thus, for instance, at odd iterations they both share the same channel and in the next one, they both select different channels. This effect will continue at the infinite preventing the algorithm to converge. In Figure \[fig:2x4\_vs\_iter\], we show how the algorithms perform when a higher number of channels is available, i.e, $4$ channels. BRD improves its performance, with respect to the other algorithms. This is mainly because the higher number of channel reduces the probability of the ping-pong effect described above. In Figure \[fig:channels\_varying\], we plot the network spectral efficiency as a function of the number of available channels. Here, the negative slot is due to the fact that we increase the number of available channels but transmitters remain subject to use a single channel. Thus, since $S > K$, there always exist a number of unused channels. The main observation in this figure is the following, the BRD becomes a very efficient solution when the number of channels is high enough to make the bouncing effect a very improbable event. Conversely, JUSTE-RL exhibits a lower performance when the number of possible actions increases. This is basically because, in JUSTE, all players play all their actions with non-zero probability in order to improve their utility estimation. Thus, this immediately implies that increasing the number of actions, increases the time that players are playing actions different from the optimal actions. Finally, in Figure \[fig:Trajectories\], we show for the $2$-players $2$-channel case, the trajectories of the algorithm during the transient phase. In this realization, BRD it can be observed that BRD does not converge. The two transmitters repeatedly select synchronously the same channel. FP and SFP converge to the best performing NE while CRL converges fast to a steady point with no game theoretical meaning. In the trajectory of JUSTE, it is possible to see that it converges to the best performing NE, for that particular channel realization. Similarly, RM also converge very fast to the best NE. Conclusion ========== In this paper, we have presented several notions of equilibrium and several learning dynamics that allow wireless networks to achieve such equilibria. In particular, we have described a general notion of equilibrium, namely, the coarse correlated equilibrium (CCE). Then, we introduced some particular cases of CCE, such as correlated equilibrium (CE) and Nash equilibrium (NE), are also analysed. Regarding the learning dynamics, we have presented the best response dynamics (BRD), fictitious play (FP), smooth fictitious play (SFP), regret matching (RM), reinforcement learning (RL) and joint utility and strategy estimation based reinforcement learning (JUSTE-RL). We have identified the pertinence of these algorithms for wireless communications in terms of system constraints (continuous/discrete actions, required information, synchronization, signalling, etc.) and the performance criteria (utility achieved at the steady state, convergence speed, etc.). As further work in this direction, we point out that existing results regarding the analysis of equilibrium in wireless networks strongly depend on the topology of the network. Indeed, a general framework for the analysis of equilibria and learning dynamics adapted to time-varying topology networks is still an open problem. Moreover, we must consider that some equilibrium notions, e.g., NE and $\epsilon$-NE, might be inefficient from a global point of view. Thus, learning algorithms to achieve Pareto optimal solutions with partial information is a further direction of research. ![Average system spectral efficiency \[bps/Hz\] as a function of signal to noise ratio (SNR) with $40$ iterations for the $2$ players and $2$ channel case.[]{data-label="fig:2x2perf_vs_iter"}](cm_figure/2x2_vs_snr_rev.pdf){width="100.00000%"} ![Average system spectral efficiency \[bps/Hz\] as a function of the number of iterations at a fixed SNR of $10$ dB for the $2$ players and $2$ channel case.[]{data-label="fig:2x2perf_vs_SNR"}](cm_figure/vs_iteration_def3.pdf){width="100.00000%"} ![Average system spectral efficiency \[bps/Hz\] as a function of the number of iterations at a fixed SNR of $10$ dB for the $2$ players and $4$ channel case.[]{data-label="fig:2x4_vs_iter"}](cm_figure/2x4_vs_iter.pdf){width="100.00000%"} ![Average system spectral efficiency as a function of the number of channels, with SNR=$10$dB and $40$ iterations.[]{data-label="fig:channels_varying"}](cm_figure/cv_notitle_rev.pdf){width="100.00000%"} ![Example of trajectories. BRD bounces between unstable solution; FP and SFP converge close to the best NE; CRL converges to a low performing NE, JUSTE-RL converges close to the best NE, RM converges close to the best NE[]{data-label="fig:Trajectories"}](cm_figure/Trajectories.pdf){width="100.00000%"} BRD FP SFP RM RL JUSTE-RL ----------------------------- ---------------------------- ------------------------------- ---------------------------- ---------------------------- --------------------- ------------------------- -- Observations ${\boldsymbol}{a}_{-k}(t)$ ${\boldsymbol}{a}_{-k}(t)$ ${\boldsymbol}{a}_{-k}(t)$ ${\boldsymbol}{a}_{-k}(t)$ $\tilde{u}_k(t)$ $\tilde{u}_k(t)$ Closed Expression for $u_k$ Yes Yes Yes Yes No No Computation complexity Optimization Optimization Optimization Optimization Algebraic Operation Algebraic Operation Steady State NE NE $\epsilon$-NE CCE $--$ $\epsilon$-NE Condition for Convergence DSG,PG, SMG DSG, PG, ZSG, $2\times N$-NDG DSG,PG,ZSG NDG $--$ DSG, $2-$player ZSG, PG Synchronization to Play Yes Yes Yes Yes No No Environment Static Stationary Stationary Stationary Stationary Stationary : Benchmark of Learning Algorithms.[]{data-label="FigTable"} [^1]: L. Rose is with Thales Communication, 160 Boulevard de Valmy, 92700 Colombes, France (e-mail: [email protected]) [^2]: S. M. Perlaza is with the Alcatel-Lucent Chair in Flexible Radio at SUPELEC. $3$ rue Joliot-Curie, $91192$, Gif-sur-Yvette, cedex. France. ([email protected]) [^3]: S. Lasaulce is with the Laboratoire des Signaux et Systèmes (LSS) at SUPELEC. $3$ rue Joliot-Curie, $91192$, Gif-sur-Yvette, cedex. France. ([email protected]) [^4]: M. Debbah is with the Alcatel-Lucent Chair in Flexible Radio at SUPELEC. $3$ rue Joliot-Curie, $91192$, Gif-sur-Yvette, cedex. France. ([email protected])
{ "pile_set_name": "ArXiv" }
--- abstract: 'We consider a problem posed by Shparlinski, of giving nontrivial bounds for rational exponential sums over the arithmetic function $\tau(n)$, counting the number of divisors of $n$. This is done using some ideas of Sathe concerning the distribution in residue classes of the function $\omega(n)$, counting the number of prime factors of $n$, to bring the problem into a form where, for general modulus, we may apply a bound of Bourgain concerning exponential sums over subgroups of finite abelian groups and for prime modulus some results of Korobov and Shkredov.' author: - | [Bryce Kerr]{}\ [Department of Computing, Macquarie University]{}\ [Sydney, NSW 2109, Australia]{}\ [[email protected]]{} date: title: Rational exponential sums over the divisor function --- Introduction ============ We consider a problem posed by Shparlinski [@Sp Problem 3.27] of bounding rational exponential sums over the divisor function. More specifically, for integers $a,m$ with $(a,m)=1$ and $m$ odd we consider the sums $$\label{main sums1} T_{a,m}(N)=\sum_{n=1}^{N}e_m({a\tau(n)}),$$ where $e_m(z)=e^{2\pi i z/m}$ and $\tau(n)=\sum_{d|n}1$ counts the number of divisors of $n$. Arithmetic properties of the divisor function have been considered in a number of works, see for example  [@DeIw; @ErdMir; @Hb; @LuSh], although we are concerned mainly with congruence properties of the divisor function, which have also been considered in [@Co; @Nz; @Se]. Exponential sums over some other arithmetic functions have been considered in [@BHS; @BS]. Our first step in bounding the sums  is to give a sharper version of a result of Sathe [@Se Lemma 1] concerning the distribution of the function $\omega(n)$ in residue classes, where $\omega(n)$ counts the number of distinct prime factors of $n$. This allows us to reduce the problem of bounding  to bounding sums of the form $$\label{S} S_m(r)=\sum_{n=1}^{t}e_m(r2^n),$$ where $t$ denotes the order of $2 \pmod m$ and we may not necessarily have $(r,m)=1$. For arbitrary $m$ we deal with these sums using a bound of Bourgain [@Bou] and when $m$ is prime we obtain sharper bounds using results of Korobov [@Kor] when the order of $2 \pmod m$ is not to small. For smaller values we use results of Shkredov [@Sk], which are based on previous results of Heath-Brown and Konyagin [@HbKg]. Notation ======== We use the notation $f(x) \ll g(x)$ and $f(x)=O(g(x))$ to mean there exists some absolutle constant $C$ such that $f(x)\le Cg(x)$ and we use $f(x)=o(g(x))$ to mean that $f(x)\le \varepsilon g(x)$ for any $\varepsilon>0$ and sufficiently large $x$. If $p|n$ and $\theta$ is the largest power of $p$ dividing $n$, we write $p^{\theta}||n$. We let $\mathcal{S}$ denote the set of all square-free integers, $\mathcal{M}_m$ the set of integers which are perfect $m$-th powers, $\mathcal{Q}_m$ the set of integers $n$, such that if $p^{\theta}||n$ then $2\le \theta \le m-1$ and $\mathcal{K}$ the set of integers $n$ such that if $p^{\theta}||n$ then $\theta\ge 2$. Given an arbitrary set of integers $\mathcal{A}$, we let $\mathcal{A}(x)$ count the number of integers in $\mathcal{A}$ less than $x$, so that $$\mathcal{Q}_m(x)\le \mathcal{K}(x) \ll x^{1/2},$$ hence the sums $$\label{H def} H(r,m)=\displaystyle\sum_{\substack{q \in \mathcal{Q}_m \\ \tau(q) \equiv r \pmod m}}\frac{h(q)}{q}, \quad \quad h(q)=\prod_{p|q}\left(1+\frac{1}{p}\right)^{-1},$$ converge. We let $\zeta(s)$ denote the Riemann-zeta function, $$\zeta(s)=\sum_{n=1}^{\infty}\frac{1}{n^s}, \quad \Re(s)>1,$$ and $\Gamma(s)$ the Gamma function, $$\Gamma(s)=\int_{0}^{\infty}x^{s-1}e^{-x}dx, \quad \Re(s)>0.$$ For odd integer $m$ we let $t$ denote the order of $2 \pmod m$ and define $$\label{alpha def} \alpha_t=1-\cos(2\pi/t).$$ Main results ============ \[thm:t1\] Suppose $m$ is odd and sufficiently large. Then with notation as in , ,  and  we have $$T_{a,m}(N)=\frac{\zeta{(m)}}{t}\frac{6}{\pi^2}\left(\displaystyle\sum_{r=0}^{m-1}H(r,m)S_{m}(ar)\right)N+O\left(tN(\log{N})^{-\alpha_t}\right).$$ When $m=p$ is prime we use a different approach to save an extra power of $\log{N}$ in the asymptotic formula above, although our bound is worse in the $t$ aspect. \[thm:t2\] Suppose $p>2$ is prime, then $$\begin{aligned} T_{a,m}(N)&=\frac{\zeta{(p)}}{t}\frac{6}{\pi^2}\left(\displaystyle\sum_{r=0}^{p-1}H(r,p)S_{p}(ar)\right)N+O\left(pN(\log{N})^{-(\alpha_t+1)}\right).\end{aligned}$$ Combining Theorem \[thm:t1\] with the main result from [@Bou] we obtain a bound which is nontrivial for $N \ge e^{ct^{1/\alpha_t}}$ for some fixed constant $c$. \[ub m\] Suppose $m$ is odd and sufficiently large, then for all $\varepsilon>0$ there exists $\delta>0$ such that if $t>m^{\varepsilon}$ then $$\max_{(a,m)=1}|T_{a,m}(N)| \ll \left(\frac{1}{m^{\delta}}+t(\log{N})^{-\alpha_t}\right)N.$$ Combining Theorem \[thm:t2\] with results from [@Kor] and [@Sk] we get, \[ub p\] Suppose $p>2$ is prime and let $$A(t)=\begin{cases} p^{1/8}t^{-7/18}(\log{p})^{7/6}, \quad \ \ \ \ t \le p^{1/2}, \\ p^{1/4}t^{-23/36}(\log{p})^{7/6}, \quad \ \ p^{1/2}<t \le p^{3/5}(\log{p})^{-6/5}, \\ p^{1/6}t^{-1/2}(\log{p})^{4/3}, \quad \ \ \ \ \ p^{3/5}<t \le p^{2/3}(\log{p})^{-2/3}, \\ p^{1/2}t^{-1}\log{p}, \quad \quad \quad \ \ \ \ \ \ \ t>p^{2/3}(\log{p})^{-2/3}, \end{cases}$$ then we have $$\max_{(a,p)=1} |T_{a,p}(N)|\ll \left(A(t)+p(\log{N})^{-(\alpha_t+1)}\right)N.$$ Preliminary results =================== We use the decomposition of integers as in [@Se]. \[decom\] For integer $m$, any $n \in \mathbb{N}$ may be written uniquely in the form $$n=sqk$$ with $s\in \mathcal{S}$, $q\in \mathcal{Q}_m$, $k \in \mathcal{M}_m$ and  $\gcd(q,s)=1$. For such a representation, we have $$\tau(n)\equiv \tau(s)\tau(q) \pmod m.$$ We first fix an integer $m$. Given any integer $n$, let $n=p_1^{\alpha_1}\dots p_{j}^{\alpha_j}$ be the prime factorisation of $n$, so that $$\label{tau m} \tau(n)=(\alpha_1+1)\dots (\alpha_j+1).$$ Let $\beta_i$ be the remainder when $\alpha_i$ is divided by $m$. Then for some $k\in \mathcal{M}_m$ we have $$n=k p_1^{\beta_1}\dots b_j^{\beta_j}=k\displaystyle\prod_{\beta_i=1}p_i^{\beta_i}\displaystyle\prod_{\beta_{i}\neq 1}p_i^{\beta_i}=ksq,$$ with $s\in \mathcal{S}$, $q\in \mathcal{Q}_m$ and $\gcd(q,s)=1$. Finally, we have from  $$\tau(n)\equiv(\beta_1+1)\dots(\beta_j+1)\equiv \tau(qs) \equiv \tau(q)\tau(s) \pmod m,$$ since $\gcd(q,s)=1$. Given integer $k$, we let $\omega(k)$ denote the number of distinct prime factors of $k$. The proof of the following Lemma is well known (see [@Gr] and references therein for sharper results and generalizations). We provide a standard proof. \[finite primes\] Suppose $q$ is squarefree and let $A_q(X)$ count the number of integers $n\le X$ such that any prime dividing $n$ also divides $q$, then $$A_q(X)\ll \frac{1}{\omega(q)!}\left(\log{X}+2\omega(q)^{1/2}\log{q}\right)^{\omega(q)}.$$ Suppose $p_1,\dots p_N$ are the distinct primes dividing $q$. Let $\langle .\ , . \rangle$ denote the standard inner product on $\mathbb{R}^N$, $||.||$ the Euclidian norm and let $\mathbb{R}_{+}^N\subset \mathbb{R}^N$ be the set of all points with nonnegative coordinates. Let $P=(\log{p_1},\dots,\log{p_N})$ and $$\mathcal{P}(Y)=\{ \mathbf{x}\in \mathbb{R}_{+}^N : \langle \mathbf{x},P\rangle \le Y \},$$ so that $$\label{A intersect} A_q(X)= \#(\mathbb{Z}^N \cap \mathcal{P}(\log{X})).$$ Let $\mathcal{C}$ denote the set of cubes of the form $$\left[j_1, j_1+1\right] \times \dots \times \left[j_N, j_N+1\right], \quad j_1,\dots,j_N \in \mathbb{Z},$$ which intersect $\mathcal{P}(\log{X})$, so that by  we have $$\label{Aq} A_q(X)\le \# \mathcal{C}.$$ Suppose $\mathcal{B}\in \mathcal{C}$, then for some $\mathbf{a}\in \mathbb{R}^N$ independent of $\mathcal{B}$ we have $$\label{subset inc} \mathcal{B}\subset \mathcal{P}(\log{X}+2N^{1/2}||P||)+\mathbf{a}.$$ Since choosing $\mathbf{x}_0\in \mathcal{B} \cap \mathcal{P}(\log{X})$ we may write any $\mathbf{x}\in \mathcal{B}$ as $\mathbf{x}=\mathbf{x}_0+\mathbf{x}'$ with $||\mathbf{x}'||\le N^{1/2}.$ Hence by the Cauchy-Schwarz inequality and the assumption $\mathbf{x}_0 \in \mathcal{B}$ we have $$\langle \mathbf{x},P\rangle = \langle \mathbf{x}_0,P\rangle+\langle \mathbf{x}',P\rangle \le \log{X}+N^{1/2}||P||,$$ $$\langle \mathbf{x},P\rangle \ge -|\langle \mathbf{x'},P\rangle|\ge -N^{1/2}||P||,$$ so that  holds with $\mathbf{a}=-(N^{1/2}||P||,\dots,N^{1/2}||P||)$. Hence from  $$\begin{aligned} A_q(X)\le \# \mathcal{C}\le \displaystyle\int_{\substack{\langle \mathbf{x},P \rangle \le \log{X}+2N^{1/2}||P|| \\ \mathbf{x}\in \mathbb{R}_+^N}}1 \ d\mathbf{x}= \frac{(\log{X}+2N^{1/2}||P||)^{N}}{N!\log{p_1}\dots \log{p_N}},\end{aligned}$$ and the result follows since $N=\omega(q)$ and $||P||\le \log{p_1}+\dots+\log{p_N}=\log{q}$ since $q$ is squarefree. We use the following result of Selberg [@Se1], for related and more precise results see  [@Te II.6]. \[selberg\] For any $z \in \mathbb{C}$, $$\displaystyle\sum_{\substack{n\leq x \\ n\in \mathcal{S}}}z^{\omega(n)}=G(z)x(\log{x})^{z-1}+O\left(x(\log{x})^{\Re({z})-2}\right),$$ with $$G(z)=\frac{1}{\Gamma(z)}\displaystyle\prod_{p}\left(1+\frac{z}{p}\right)\left(1-\frac{1}{p}\right)^{z},$$ and the implied constant is uniform for all $|z|=1.$ We combine Lemma \[finite primes\] and Lemma \[selberg\] to get a sharper version of  [@Se Lemma 1]. \[Mq\] For integers $q,r,t$ let $$M(x,q,r,t)=\#\{ \ n\leq x \ : \ n\in \mathcal{S}, \quad \omega(n) \equiv r\pmod t, \quad (n,q)=1 \}.$$ Then for $x\ge q$ we have $$M(x,q,r,t)=\frac{6h(q)}{\pi^2t}x+O\left(x^{1/2}(e^4\log{x})^{\omega(q)}\right)+O\left( x (\log{x})^{-\alpha_t}\log \log{q}\right).$$ Suppose first $q$ is squarefree and let $$S(a,x)=\displaystyle\sum_{\substack{n\leq x \\ n\in \mathcal{S}}}e_t{(a\omega(n))},$$ and $$S_1(a,q,x)=\displaystyle\sum_{\substack{n\leq x \\ n\in \mathcal{S} \\ (n,q)=1}}e_t{(a\omega(n))}.$$ Since the numbers $e_m(a\omega(n))$ with $(n,q)=1$ and $n\in \mathcal{S}$ are the coefficients of the Dirichlet series $$\displaystyle\prod_{p \nmid q}\left(1+\frac{e_t(a)}{p^{s}}\right)=\displaystyle\prod_{p |q}\frac{1}{\left(1+\frac{e_t(a)}{p^s}\right)}\displaystyle\prod_{p}\left(1+\frac{e_t(a)}{p^{s}}\right),$$ we let the numbers $a_n$ and $b_n$ be defined by $$\begin{aligned} \displaystyle\prod_{p |q}\frac{1}{\left(1+\frac{e_t(a)}{p^s}\right)}&=\displaystyle\sum_{n=1}^{\infty}\frac{a_n}{n^s}, \\ \displaystyle\prod_{p}\left(1+\frac{e_t(a)}{p^{s}}\right)&=\displaystyle\sum_{n=1}^{\infty}\frac{b_n}{n^s},\end{aligned}$$ so that $$S(a,x)=\sum_{n\le x}b_n,$$ and $$\begin{aligned} S_1(a,q,x)=\displaystyle\sum_{n\leq x}\displaystyle\sum_{d_1d_2=n}b_{d_1}a_{d_2}=\displaystyle\sum_{n\le x}a_n S(a,x/n). \end{aligned}$$ Consider when $a\neq 0$, by Lemma \[selberg\] $$\begin{aligned} \displaystyle\sum_{n\le x}a_nS(a,x/n)&=G(e_t(a))x\displaystyle\sum_{n\le x}\frac{a_n}{n}(\log{(x/n)})^{e_t(a)-1}\\ & \quad +O\left(\displaystyle\sum_{n\le x}|a_n|\frac{x}{n}(\log{x/n})^{\cos{(2\pi /t)}-2}\right) \\ &\ll x (\log{x})^{-(1-\cos(2\pi /t))}\sum_{n\le x}\frac{|a_n|}{n} \\ &\ll x (\log{x})^{-\alpha_t}\prod_{p|q}\left(1-\frac{1}{p}\right)^{-1},\end{aligned}$$ and since $$\displaystyle\prod_{p|q}\left(1-\frac{1}{p}\right)^{-1}=\frac{q}{\phi{(q)}}\ll \log \log {q},$$ where $\phi$ is Euler’s totient function, we get $$\label{sa} S_1(a,q,x) \ll x (\log{x})^{-\alpha_t}\log \log {q}.$$ For $a=0$, by [@HardyWright Theorem 334] $$\begin{aligned} S_1(0,q,x)&=\displaystyle\sum_{n\le x}a_n S(0,x/n) \\ &=\frac{6x}{\pi^2}\sum_{n\le x}\frac{a_n}{n}+O\left(x^{1/2}\sum_{n\le x}\frac{|a_n|}{n^{1/2}} \right) \\ &=\frac{6x}{\pi^2}\displaystyle\prod_{p|q}\left(1+\frac{1}{p}\right)^{-1}+O \left(x\sum_{n\ge x}\frac{|a_n|}{n}\right) +O\left(x^{1/2}\sum_{n\le x}\frac{|a_n|}{n^{1/2}} \right) .\end{aligned}$$ For the first error term, with notation as in Lemma \[finite primes\], we have $$\sum_{n\le t}|a_n|=A_q(t),$$ so that $$\begin{aligned} \label{step 1} \sum_{n\ge x}\frac{|a_n|}{n}&\ll \displaystyle\int_{x}^{\infty}\frac{A_q(t)}{t^2}dt \nonumber \\ &\ll \frac{1}{\omega(q)!}\displaystyle\int_{x}^{\infty}\frac{\left(\log{t}+2\omega(q)^{1/2}\log{q}\right)^{\omega(q)}}{t^2}dt,\end{aligned}$$ and $$\begin{aligned} \label{binom expand} \displaystyle\int_{x}^{\infty}\frac{\left(\log{t}+2\omega(q)^{1/2}\log{q}\right)^{\omega(q)}}{t^2}dt &= \displaystyle\sum_{n=0}^{\omega(q)}\binom{\omega(q)}{n}\left(2\omega(q)^{1/2}\log{q}\right)^{\omega(q)-n} \displaystyle\int_{x}^{\infty}\frac{\left(\log{t}\right)^{n}}{t^2}dt. \end{aligned}$$ The integral $$\displaystyle\int_{x}^{\infty}\frac{\left(\log{t}\right)^{n}}{t^2}dt,$$ is the $n$-th derivative of the function $$H(z)=\displaystyle\int_{x}^{\infty}t^{z-2}dz=\frac{x^{z-1}}{1-z},$$ evaluated at $z=0$. Hence by Cauchy’s Theorem, letting $\gamma \subset \mathbb{C}$ be the circle centered at 0 with radius $1/\log{x}$ we have $$\begin{aligned} \displaystyle\int_{x}^{\infty}\frac{\left(\log{t}\right)^{n}}{t^2}dt=\frac{n!}{2\pi i}\displaystyle\int_{\gamma}\frac{x^{z-1}}{1-z}\frac{1}{z^{n+1}}dz\ll \frac{n!(\log{x})^n}{x}.\end{aligned}$$ Hence by  and  $$\begin{aligned} \sum_{n\ge x}\frac{|a_n|}{n}\ll \frac{1}{x}\frac{\left(2\omega(q)^{1/2}\log{q}\right)^{\omega{(q)}}}{\omega(q)!}\sum_{n=0}^{\omega(q)}\binom{\omega(q)}{n}n!\left(\frac{\log{x}}{\omega(q)^{1/2}\log{q}}\right)^n,\end{aligned}$$ and by Stirling’s formula [@MgVu Equation B.26] $$\begin{aligned} \sum_{n=0}^{\omega(q)}\binom{\omega(q)}{n}n!\left(\frac{\log{x}}{\omega(q)^{1/2}\log{q}}\right)^n&\ll \sum_{n=0}^{\omega(q)}\binom{\omega(q)}{n}n^{1/2}\left(\frac{n}{e}\right)^n\left(\frac{\log{x}}{\omega(q)^{1/2}\log{q}}\right)^n \\ &\le \omega(q)^{1/2}\sum_{n=0}^{\omega(q)}\binom{\omega(q)}{n}\left(\frac{\omega{(q)}^{1/2}\log{x}}{e\log{q}}\right)^n \\ &\ll \omega(q)^{1/2}\left(\frac{\omega(q)^{1/2}\log{x}}{e\log{q}}+1\right)^{\omega{(q)}},\end{aligned}$$ so that $$\begin{aligned} \sum_{n\ge x}\frac{|a_n|}{n}&\ll\frac{1}{x}\frac{\omega(q)^{1/2}\left(2\omega(q)^{1/2}\log{q}\right)^{\omega{(q)}}}{\omega(q)!} \left(\frac{\omega(q)^{1/2}\log{x}}{e\log{q}}+1\right)^{\omega{(q)}}.\end{aligned}$$ By another application of Stirling’s formula, $$\begin{aligned} \sum_{n\ge x}\frac{|a_n|}{n}&\ll \frac{2^{\omega(q)}}{x}\left(\frac{e\log{q}}{\omega{(q)}^{1/2}}\right)^{\omega(q)}\left(\frac{\omega(q)^{1/2}\log{x}}{e\log{q}}+1\right)^{\omega{(q)}} \\ &\ll \frac{2^{\omega(q)}}{x}\left(\log{x}+\frac{e\log{q}}{\omega{(q)}^{1/2}}\right)^{\omega(q)} \\ &\ll2^{\omega(q)}\left(1+\frac{3}{\omega(q)^{1/2}}\right)^{\omega(q)}\frac{(\log{x})^{\omega(q)}}{x} \\ &\ll 2^{\omega(q)}e^{3\omega(q)^{1/2}}\frac{(\log{x})^{\omega(q)}}{x}\ll e^{4\omega(q)}\frac{(\log{x})^{\omega(q)}}{x},\end{aligned}$$ which gives $$\begin{aligned} S_1(0,q,x)=\frac{6h(q)}{\pi^2}x+O\left((e^4\log{x})^{\omega(q)}\right)+O\left(x^{1/2}\sum_{n\le x}\frac{|a_n|}{n^{1/2}} \right).\end{aligned}$$ For the last term, $$\begin{aligned} \sum_{n\le x}\frac{|a_n|}{n^{1/2}}\le \prod_{p|q}\left(1-p^{-1/2}\right)^{-1}\le \prod_{p|q}(e^4 \log{x})=(e^4\log{x})^{\omega(q)},\end{aligned}$$ so that $$\label{s1} S_1(0,q,x)=\frac{6h(q)}{\pi^2}x+O\left(x^{1/2}(e^4\log{x})^{\omega(q)}\right).$$ Since $$\begin{aligned} M(x,q,r,t)&=\frac{1}{t}\displaystyle\sum_{a=0}^{t-1}e_t(-ar)S_1(a,q,x) \\ &=\frac{1}{t}S_1(0,q,x)+\frac{1}{t}\displaystyle\sum_{a=1}^{t-1}e_t(-ar)S_1(a,q,x),\end{aligned}$$ we have from  and  $$M(x,q,r,t)=\frac{6h(q)}{\pi^2t}x+O\left(x^{1/2}(e^4\log{x})^{\omega(q)}\right)+O\left( x (\log{x})^{-\alpha_t}\log \log{q}\right).$$ If $q$ is not squarefree, repeating the above argument with $q$ replaced by its squarefree part gives the general case since the error term is increasing with $q$. For complex $s$ we write $s=\sigma+it$ with both $\sigma$ and $t$ real. \[tau and zeta\] Let $m$ be odd and $\chi$ a multiplicative character$\pmod m$. Let $$L(s,\chi,\tau)=\displaystyle\sum_{n=1}^{\infty}\frac{\chi(\tau(n))}{n^s},$$ then for $\sigma >1$ we have $$L(s,\chi,\tau)=\zeta(s)^{\chi(2)}F(s,\chi),$$ with $F(1,\chi)\neq 0$ and $$F(s,\chi)=\displaystyle\sum_{n=1}^{\infty}\frac{b(\chi,n)}{n^s},$$ for some constants $b(\chi,n)$ satisfying $$\begin{aligned} \displaystyle\sum_{n=1}^{\infty}\frac{|b(\chi,n)|(\log{n})^3}{n}=O(1),\end{aligned}$$ uniformly over all characters $\chi$. Since both $\chi$ and $\tau$ are multiplicative we have for $\sigma>1$, $$\begin{aligned} L(s,\chi,\tau)&=\displaystyle\prod_{p}\left(1+\displaystyle\sum_{n=1}^{\infty}\frac{\chi(\tau(p^n))}{p^{ns}}\right) \\ &=\zeta(s)^{\chi(2)}F(s,\chi),\end{aligned}$$ with $$\begin{aligned} F(s,\chi)=\displaystyle\prod_{p}\left(1+\displaystyle\sum_{n=1}^{\infty}\frac{\chi(n+1)}{p^{ns}}\right)\left(1-\frac{1}{p^{s}}\right)^{\chi(2)}.\end{aligned}$$ We have $$\begin{aligned} F(s,\chi)&=\displaystyle\prod_{p}\left(1-\frac{\chi(2)}{p^{s}}\right) \left(1+\frac{\chi(2)}{p^{s}}+\displaystyle\sum_{n=2}^{\infty}\frac{\chi(n+1)}{p^{ns}}\right) \times \\ & \qquad \qquad\qquad \qquad\qquad \displaystyle\prod_{p}\left(1-\frac{\chi(2)}{p^{s}}\right)^{-1} \left(1-\frac{1}{p^{s}}\right)^{\chi(2)} \\ &=F_1(s,\chi)F_2(s,\chi),\end{aligned}$$ where $$\begin{aligned} \label{F_1} F_1(s,\chi)&=\displaystyle\prod_{p}\left(1-\frac{\chi(2)}{p^{s}}\right) \left(1+\frac{\chi(2)}{p^{s}}+\displaystyle\sum_{n=2}^{\infty}\frac{\chi(n+1)}{p^{ns}}\right) \nonumber \\ &= \displaystyle\prod_{p}\left(1+\displaystyle\sum_{n=2}^{\infty}\frac{\chi(n+1)-\chi(2n)}{p^{ns}}\right),\end{aligned}$$ and $$\begin{aligned} F_2(s,\chi)&=\displaystyle\prod_{p}\left(1-\frac{\chi(2)}{p^{s}}\right)^{-1} \left(1-\frac{1}{p^{s}}\right)^{\chi(2)}.\end{aligned}$$ Considering $F_2(s,\chi)$, we have for $\sigma>1$ $$\begin{aligned} \label{F_2} \log{F_2(s,\chi)}&=\displaystyle\sum_{p}\displaystyle\sum_{n=1}^{\infty} \frac{1}{n}\left(\frac{\chi(2^n)} {p^{ns}}-\frac{\chi(2)}{p^{ns}}\right) \nonumber \\ &=\displaystyle\sum_{p}\displaystyle\sum_{n=2}^{\infty}\frac{\chi(2^n)-\chi(2)}{n}\frac{1}{p^{ns}}.\end{aligned}$$ In the equations  and , the product and the series converge absolutely on $\sigma=1$ so that $F(1,\chi)\neq 0$. Also since $|\chi(j)-\chi(k)|\leq 2$ for all integers $k,j$ we see that the coefficents $b(\chi,n)$ in $$F(s,\chi)=\displaystyle\sum_{n=1}^{\infty}\frac{b(\chi,n)}{n^s},$$ satisfy $$|b(\chi,n)|\leq c_n,$$ where the numbers $c_n$ are defined by $$\displaystyle\prod_{p}\left(1+\displaystyle\sum_{n=2}^{\infty}\frac{2}{p^{ns}}\right)\exp\left(\displaystyle\sum_{p}\displaystyle\sum_{n=2}^{\infty}\frac{2}{n}\frac{1}{p^{ns}}\right)=\displaystyle\sum_{n=1}^{\infty}\frac{c_n}{n^s}.$$ The function defined by the above formula converges uniformly in any halfplane $\sigma\geq \sigma_0 >1/2$, so that $$\displaystyle\sum_{n\leq X}c_n=O(X^{1/2+\varepsilon})$$ and the last statement of the Lemma follows by partial summation. The following is [@MgVu Theorem 7.18]. \[Selberg\] Suppose for each complex $z$ we have a sequence $(b_z(n))_{n=1}^{\infty}$ such that the sum $$\displaystyle\sum_{n=1}^{\infty}\frac{|b_z(n)|(\log{n})^{2R+1}}{n},$$ is uniformly bounded for $|z|\leq R$ and for $\sigma\ge 1$ let $$F(s,z)=\displaystyle\sum_{n=1}^{\infty}\frac{b_z(m)}{m^s}.$$ Suppose for $\sigma > 1$ we have $$\zeta(s)^zF(s,z)=\displaystyle\sum_{n=1}^{\infty}\frac{a_z(n)}{n^s},$$ for some $a_z(n)$ and let $S_z(x)=\displaystyle\sum_{n\leq x}a_z(n).$ Then for $x\ge 2$, uniformly over all $|z|\leq R$, $$S_z(x)=\frac{F(1,z)}{\Gamma(z)}x(\log{x})^{z-1}+O(x(\log{x})^{\Re(z)-2}).$$ Combining Lemma \[tau and zeta\] and Lemma \[Selberg\] gives \[c tau\] For integer $m$ let $\chi$ be a multiplicative character $\pmod m$ and let $$G(\chi)=\frac{1}{\Gamma(\chi(2))}\displaystyle\prod_{p}\left(\displaystyle\sum_{n=0}^{\infty}\frac{\chi(n+1)}{p^{n}}\right)\left(1-\frac{1}{p}\right)^{\chi(2)}.$$ Then uniformly over all characters $\chi$, $$\label{neq} \displaystyle\sum_{n\leq x}\chi(\tau(n))=G(\chi)x(\log{x})^{\chi(2)-1}+O\left(x(\log{x})^{\Re(\chi(2))-2}\right).$$ \[tau cong\] For any integer $m$, $$\displaystyle\sum_{\substack{q \in \mathcal{K} \\ \tau(q) \equiv 0 \pmod{m}}}\frac{1}{q} \ll \frac{1}{m^{\log{2}/2}},$$ and if $p$ is prime $$\displaystyle\sum_{\substack{q \in \mathcal{K} \\ \tau(q) \equiv 0 \pmod{p}}}\frac{1}{q} \ll \frac{1}{2^{p/2}}.$$ Suppose $\tau(q) \equiv 0 \pmod{m}$ and let $q=p_1^{\alpha_1}\dots p_k^{\alpha_k}$ be the prime factorization of $q$, so that $$\tau(q)=(\alpha_1+1)\dots(\alpha_k+1)\ge m.$$ By the arithmetic-geometric mean inequality, $$q \ge 2^{(\alpha_1+1)+ \dots +(\alpha_k+1)-k}\ge 2^{k(\tau(q)^{1/k}-1)}\ge 2^{k(m^{1/k}-1)}\ge 2^{\log{m}}=m^{\log{2}},$$ and since $\mathcal{K}(x)\ll x^{1/2}$ we get $$\displaystyle\sum_{\substack{q \in \mathcal{K} \\ \tau(q) \equiv 0 \pmod{m}}}\frac{1}{q} \le \displaystyle\sum_{\substack{q \in \mathcal{K} \\ q \ge m^{\log{2}}}}\frac{1}{q}\ll \displaystyle\int_{m^{\log{2}}}^{\infty}\frac{\mathcal{K}(x)}{x^2}dx\ll \frac{1}{m^{\log{2}/2}}.$$ Suppose $p$ is prime, if $\tau(n)\equiv 0 \pmod p$ then $n\ge 2^{p-1}$. As before we get $$\displaystyle\sum_{\substack{q \in \mathcal{K} \\ \tau(q) \equiv 0 \pmod{p}}}\frac{1}{q}\ll \frac{1}{2^{p/2}}.$$ Proof of Theorem \[thm:t1\] =========================== By Lemma \[decom\] we have $$\begin{aligned} \displaystyle\sum_{n=1}^{N}e_m(a\tau(n))&=\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}} \displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le N/k}} \displaystyle\sum_{\substack{s\in \mathcal{S}\\ \gcd(s,q)=1 \\ s\le N/qk}} e_m(a\tau(kqs)) \\ &=\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}} \displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le N/k}} \displaystyle\sum_{\substack{s\in \mathcal{S}\\ \gcd(s,q)=1 \\ s\le N/qk}} e_m(a\tau(q)\tau(s)) \\ &=\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}} \displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le N/k}} \displaystyle\sum_{\substack{s\in \mathcal{S}\\ \gcd(s,q)=1 \\ s\le N/qk}} e_m(a\tau(q)2^{\omega(s)}).\end{aligned}$$ Let $K=N^{1/2}$ and grouping together values of $2^{\omega(s)}$ in the same residue class$\pmod m$ we have, $$\begin{aligned} \displaystyle\sum_{n=1}^{N}e_m(a\tau(n)) &=\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}} \displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le K/k}}\displaystyle\sum_{r=1}^{t}M(N/qk,q,r,t)e_m(a\tau(q)2^r) \\ &\quad \quad \quad+\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}} \displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ K/k<q\le N/k}}\displaystyle\sum_{r=1}^{t}M(N/qk,q,r,t)e_m(a\tau(q)2^r). \\\end{aligned}$$ By choice of $K$, we have $N/qk\ge q$   when   $q\le K/k$. Hence we may apply Lemma \[Mq\] to the first sum above, $$\begin{aligned} \displaystyle\sum_{n=1}^{N}e_m(a\tau(n))&=\frac{6}{\pi^2}\frac{N}{t}\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le N/k}}\displaystyle\sum_{r=1}^{t}\frac{h(q)}{qk}e_m(a\tau(q)2^r) \\ &+O\left(\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le K/k}}\frac{Nt\log \log {q}}{qk}(\log{(N/qk)})^{-\alpha_t}\right)+O\left(\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}} \displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ K/k<q\le N/k}}\frac{tN}{qk} \right) \\ &\quad \quad \quad+O\left(\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le K/k}}(e^4\log{(N/kq)})^{\omega(q)}\left(\frac{N}{kq}\right)^{1/2}\right).\end{aligned}$$ Considering the first two error terms, $$\begin{aligned} \label{bound 1} \displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}} \displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ K/k<q\le N/k}}\frac{tN}{qk} \ll tN \int_{K}^{N}\frac{\mathcal{K}(x)}{x^2}dx \ll \frac{tN}{K^{1/2}},\end{aligned}$$ and since the sum $$\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le K/k}}\frac{\log \log{q}}{qk},$$ is bounded uniformly in $m$ as $K,N \rightarrow \infty$, we get $$\begin{aligned} \label{bound 2} \displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le K/k}}\frac{Nt\log{q}}{qk}(\log{(N/qk)})^{-\alpha_t}&\ll Nt\log (N/K)^{-\alpha_t} \displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le K/k}}\frac{\log \log{q}}{qk} \nonumber \\ &\ll Nt(\log(N/K))^{-\alpha_t}.\end{aligned}$$ For the last term, $$\begin{aligned} &\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le K/k}}(e^{4}\log{(N/kq)})^{\omega(q)}\left(\frac{N}{kq}\right)^{1/2} \le N^{2/3}\displaystyle\sum_{\substack{n \in \mathcal{K} \\ n\le N }}\left(\frac{1}{n}\right)^{2/3}(e^4\log{N})^{\omega(n)},\end{aligned}$$ and since $$\omega(n)\le (1+o(1))\frac{\log{n}}{\log\log{n}},$$ we get $$\begin{aligned} N^{2/3}\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le K/k}}(e^{4}\log{N})^{\omega(q)}\left(\frac{1}{kq}\right)^{2/3} \le N^{5/6+o(1)}\displaystyle\sum_{\substack{n \in \mathcal{K} \\ n\le N }}\left(\frac{1}{n}\right)^{2/3}(e^4(\log{N})^{5/6})^{\omega(n)}.\end{aligned}$$ We may bound the sum on the right by noting $$\begin{aligned} \displaystyle\sum_{\substack{n \in \mathcal{K} \\ n\le N }}\left(\frac{1}{n}\right)^{2/3}(e^4(\log{N})^{5/6})^{\omega(n)} &\le \displaystyle\prod_{p}\left(1+e^4(\log{N})^{5/6}\displaystyle\sum_{k=2}^{\infty}\frac{1}{p^{2k/3}}\right),\end{aligned}$$ taking logarithms we see that $$\begin{aligned} \log\left(\displaystyle\prod_{p}\left(1+e^4(\log{N})^{5/6}\displaystyle\sum_{k=2}^{\infty}\frac{1}{p^{2k/3}}\right)\right)&= \displaystyle\sum_{p}\log\left(\left(1+\frac{e^4(\log{N})^{5/6}}{p^{4/3}}\frac{p^{2/3}}{p^{2/3}-1}\right)\right) \\ &\le e^4(\log{N})^{5/6}\sum_{p}\frac{1}{p^{4/3}}\frac{p^{2/3}}{p^{2/3}-1} & \\ &\ll (\log N)^{5/6},\end{aligned}$$ hence we have for some absolute constant $c$ $$\label{bound 3} \displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le K/k}}(e^{4}\log{(N/kq)})^{\omega(q)}\left(\frac{N}{kq}\right)^{1/2}\ll N^{5/6+o(1)}e^{c(\log{N})^{5/6}}.$$ Combining ,  and  gives $$\begin{aligned} \displaystyle\sum_{n=1}^{N}e_m(a\tau(n))&=\frac{6}{\pi^2}\frac{N}{t}\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\frac{1}{k}\displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le N/k}}\frac{h(q)}{q}\displaystyle\sum_{r=1}^{t}e_m(a\tau(q)2^r)+O\left(Nt(\log(N/K))^{-\alpha_t}\right) \\ & \quad +O\left(\frac{tN}{K^{1/2}}\right)+O\left(N^{5/6+o(1)}e^{c(\log{N})^{5/6}}\right).\end{aligned}$$ Recalling the choice of $K$ we get $$\begin{aligned} \label{almost finished} \displaystyle\sum_{n=1}^{N}e_m(a\tau(n))=\frac{6}{\pi^2}\frac{N}{t}\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\frac{1}{k}\displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le N/k}}\frac{h(q)}{q}\displaystyle\sum_{r=1}^{t}e_m(a\tau(q)2^r)+O\left(Nt(\log{N})^{-\alpha_t}\right).\end{aligned}$$ For the main term, $$\begin{aligned} \displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\frac{1}{k}\displaystyle\sum_{\substack{q\in \mathcal{Q}_m \\ q\le N/k}}\frac{h(q)}{q}\displaystyle\sum_{r=1}^{t}e_m(a\tau(q)2^r)&=\displaystyle\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\frac{1}{k}\displaystyle\sum_{q\in \mathcal{Q}_m}\frac{h(q)}{q}\displaystyle\sum_{r=1}^{t}e_m(a\tau(q)2^r) \\ &+O\left(t\sum_{\substack{k\in \mathcal{M}_m \\ k \le N}}\frac{1}{k}\left(\frac{k}{N}\right)^{1/2}\right) \\ &=\displaystyle\sum_{k\in \mathcal{M}_m }\frac{1}{k}\displaystyle\sum_{q\in \mathcal{Q}_m}\frac{h(q)}{q}\displaystyle\sum_{r=1}^{t}e_m(a\tau(q)2^r) \\ &+O\left(\frac{t}{N^{1/2}}\right) \\ &=\zeta(m)\displaystyle\sum_{q\in \mathcal{Q}_m}\frac{h(q)}{q}\displaystyle\sum_{r=1}^{t}e_m(a\tau(q)2^r)+O\left(\frac{t} {N^{1/2}}\right).\end{aligned}$$ Hence we have $$\displaystyle\sum_{n=1}^{N}e_m(a\tau(n))=\frac{\zeta{(m)}}{t}\frac{6}{\pi^2}\left(\displaystyle\sum_{r \pmod m}H(r,m)S_m(ar)\right)N+O\left(tN(\log{N})^{-\alpha_t}\right).$$ Proof of Theorem \[thm:t2\] =========================== Let $$C(p,r,N)= \# \{ n\le N : \tau(n) \equiv r \pmod p \},$$ so that $$\label{con exp} \displaystyle\sum_{n=1}^{N}e_p{(a\tau(n))}=\displaystyle\sum_{r=0}^{p-1}C(p,r,N)e_m(an).$$ Suppose $(r,p)=1$, using orthogonality of characters and Lemma \[c tau\], $$\begin{aligned} C(p,r,N)&=\frac{1}{p-1}\sum_{n=1}^{N}\displaystyle\sum_{\chi (\text{mod} \ p)}\overline\chi(r)\displaystyle\chi(\tau(n)) \\ &=\frac{1}{p-1}\sum_{\chi(2)=1}\overline\chi(r)G(\chi)N +\frac{1}{p-1}\sum_{\chi(2)\neq 1}\overline\chi(r)G(\chi)N(\log{N})^{\chi(2)-1} \\ & \quad +O\left(N (\log{N})^{-(\alpha_p+1)}\right),\end{aligned}$$ and for $r=0$, we have by another applications of Lemma \[c tau\] $$C(p,0,N)=\displaystyle\sum_{n=1}^{N}(1-\chi_0(\tau(n)))=C_pN+O\left(N (\log{N})^{-2}\right),$$ for some constant $C_p$. Hence from \[con exp\] $$\begin{aligned} \displaystyle\sum_{n=1}^{N}e_p{(a\tau(n))}&=C_pN+O\left(pN(\log{N})^{-(\alpha_t+1)}\right) \\ &+ \frac{N}{p-1}\sum_{r=1}^{p-1}\left(\sum_{\chi(2)=1}\overline\chi(r)G(\chi)e_p(ar) +\sum_{\chi(2)\neq 1}\overline\chi(r)e_p(ar)G(\chi)(\log{N})^{\chi(2)-1}\right) \\ &=A_pN+\frac{N}{p-1}\sum_{\chi(2) \neq 1}G(\chi)(\log{N})^{\chi(2)-1}\sum_{r=1}^{p-1}\overline\chi(r)e_p(ar) \\ & \quad +O\left(pN(\log{N})^{-(\alpha_t+1)}\right),\end{aligned}$$ for some constant $A_p$. If $\chi(2)\neq 1$ then we have $$\left|\displaystyle\sum_{r=1}^{p-1}\overline\chi(r)e_p(ar)\right|=p^{1/2},$$ so that $$\begin{aligned} \label{AAAAAAAAAA} \displaystyle\sum_{n=1}^{N}e_p{(a\tau(n))}&=A_p N +O\left(p^{1/2}N(\log{N})^{-\alpha_t}\right)+O\left(pN(\log{N})^{-(\alpha_t+1)}\right) \nonumber \\ &=A_p N +O\left(pN(\log{N})^{-(\alpha_t+1)}\right),\end{aligned}$$ since if $$p^{1/2}N(\log{N})^{-\alpha_t}\le pN(\log{N})^{-(\alpha_t+1)} \quad \text{then} \quad N\le pN(\log{N})^{-(\alpha_t+1)}.$$ Finally, comparing  with the leading term in the asymptotic formula from Theorem \[thm:t1\], we see that $$A_p=\frac{\zeta{(p)}}{t}\frac{6}{\pi^2}\left(\displaystyle\sum_{r=0}^{p-1}H(r,p)S_{p}(ar)\right).$$ Proof of Theorem \[ub m\] ========================= Considering the main term in Theorem \[thm:t1\] $$\begin{aligned} \left|\displaystyle\sum_{r=0}^{m-1}H(r,m)S_{m}(ar)\right|&\le\sum_{d | m}\left|\sum_{\substack{r=0 \\ \gcd(r,m)=d}}^{m-1}H(r,m)S_{m}(ar)\right| \\ &\le \sum_{d | m}\left(\sum_{\substack{r=0 \\ \gcd(r,m)=d}}^{m-1}H(r,m)\right)\max_{\gcd(\lambda,m)=d}|S_{m}(\lambda)|.\end{aligned}$$ Writing $c=\log{2}/2$, by Lemma \[tau cong\] $$\begin{aligned} \sum_{\substack{r=0 \\ \gcd(r,m)=d}}^{m-1}H(r,m)&=\sum_{\substack{r=0 \\ \gcd(r,m)=d}}^{m-1} \displaystyle\sum_{\substack{q \in \mathcal{Q}_m \\ \tau(q) \equiv r \pmod m}}\frac{h(q)}{q} \\ &\le \displaystyle\sum_{\substack{ q \in \mathcal{K} \\ \tau(q) \equiv 0 \pmod{d}}}\frac{1}{q} \ll \dfrac{1}{d^{c}},\end{aligned}$$ so that $$\label{coefficient bound} \displaystyle\sum_{r=0}^{m-1}H(r,m)S_{m}(ar)\ll \sum_{d | m}\frac{1}{d^c}\max_{\gcd(\lambda,m)=d}|S_{m}(\lambda)|.$$ Suppose $\gcd(\lambda,m)=d$, so that $\lambda=d\lambda'$ and $m=dm'$ for some $\lambda'$ and $m'$ with $\gcd(\lambda',m')=1$. Let $t_d$ denote the order of $2 \pmod {m'}$. Then we have $$S_{m}(\lambda)=\sum_{n=1}^{t}e_m(\lambda 2^n)=\frac{t}{t_d}\sum_{n=1}^{t_d}e_{m'}(\lambda' 2^n).$$ By the main result of [@Bou], if $t_d \ge (m/d)^{\varepsilon}$ then for some $\delta>0$, $$\label{coefficient bound 1} S_{m}(\lambda)\ll \left(\frac{d}{m}\right)^{\delta}t.$$ Suppose $t\ge m^{\varepsilon}$, then since $$t_d \ge \dfrac{t}{d}\ge \frac{m^{\varepsilon/2}}{d}m^{\varepsilon/2},$$ if $d \le m^{\varepsilon/2}$ then we have $t_d \ge (m/d)^{\varepsilon/2}$. Hence by  $$S_{m}(\lambda)\ll \left(\frac{d}{m}\right)^{\delta}t\ll \frac{t}{m^{\delta_0}}.$$ Hence by , for some $\delta_1>0$ $$\begin{aligned} \label{coefficient final bound} \displaystyle\sum_{r=0}^{m-1}H(r,m)S_{m}(ar)&\ll \sum_{\substack{d | m \\ d\le m^{\varepsilon/2}}}\frac{1}{d^c}\max_{\gcd(\lambda,m)=d}|S_{m}(\lambda)| +\sum_{\substack{d | m \\ d \ge m^{\varepsilon/2}}}\frac{1}{d^c}\max_{\gcd(\lambda,m)=d}\left|S_{m}(\lambda)\right| \nonumber \\ & \ll \sum_{\substack{d | m \\ d\le m^{\varepsilon/2}}}\frac{t}{m^{\delta_1}} +\sum_{\substack{d | m \\ d \ge m^{\varepsilon/2}}}\frac{1}{m^{\delta_1}}=\frac{\tau(m)}{m^{\delta_1}}t,\end{aligned}$$ and the result follows combining  with Theorem \[thm:t1\]. Proof of Theorem \[ub p\] ========================= By Lemma \[tau cong\] $$\begin{aligned} \displaystyle\sum_{r=0}^{p-1}H(r,p)S_{p}(ar,t)&=\displaystyle\sum_{r=1}^{p-1}H(r,p)S_{p}(ar)+tH(0,p) \\ &\ll \frac{t}{2^{p/2}}+\left(\displaystyle\sum_{r=1}^{p-1}H(r,p) \right)\max_{\gcd(\lambda,p)=1}\left|S_{p}(\lambda)\right| \\ & \ll \frac{t}{2^{p/2}}+\max_{\gcd(\lambda,p)=1}\left|S_{p}(\lambda)\right|. \end{aligned}$$ In [@Kerr] it is shown the following bound is a consequence of [@Kor] and [@Sk], $$\max_{\gcd(\lambda,p)=1}|S_{p}(\lambda)| \ll \begin{cases} p^{1/8}t^{22/36}(\log{p})^{7/6}, \quad \ \ \ \ t \le p^{1/2}, \\ p^{1/4}t^{13/36}(\log{p})^{7/6}, \quad \ \ p^{1/2}<t \le p^{3/5}(\log{p})^{-6/5}, \\ p^{1/6}t^{1/2}(\log{p})^{4/3}, \quad \ \ \ \ \ p^{3/5}<t \le p^{2/3}(\log{p})^{-2/3}, \\ p^{1/2}\log{p}, \quad \quad \quad \ \ \ \ \ \ \ \ \ \ t>p^{2/3}(\log{p})^{-2/3}, \end{cases}$$ and the result follows by Theorem \[thm:t2\]. [12]{} W. D. Banks, G. Harman and I. E. Shparlinski, ‘Distributional properties of the largest prime factor’, [ *Michigan Math. J*]{}., [**53**]{} (2005), 665–681. W. Banks and I. E. Shparlinski, ‘Congruences and rational exponential sums with the Euler function’,[ *Rocky Mountain J. Math*]{}., [**36** ]{} (2006), 1415–1426. J. Bourgain, ‘Exponential sum estimates on subgroups of $\mathbb{Z}_q$, $q$ arbitrary’, [*J. Anal. Math.*]{}, [**97**]{} (2005), 317–355. E. Cohen, ‘Arithmetical notes, V. A divisibility property of the divisor function’, [*Amer. Jour. Math.*]{}, [**83**]{}, (1961), 693–697. J. M. Deshouillers and H. Iwaniec, ‘An additive divisor problem’, [*J. London Math. Soc.*]{}, [**26**]{}, (1982), 1–14. P. Erd[ő]{}s and L. Mirsky, ‘The distribution of values of the divisor function $d(n)$’, [*Proc. London Math. Soc.*]{}, [**2**]{}, (1952), 257–271. A. Granville, ‘The lattice points of an n-dimensional tetrahedron’,[ *Aequationes Math.*]{}, [**41**]{}, (1991), 234–241. G. H. Hardy and E. M. Wright, [*An introduction to the theory of numbers*]{}, Oxford Univ. Press, Oxford, 1979. D. R. Heath-Brown, ‘The divisor function at consecutive integers’, [*Mathematika*]{}, [**31**]{} (1984), 141-149. D. R. Heath–Brown and S. V. Konyagin, ‘New bounds for Gauss sums derived from kth powers, and for Heilbronn’s exponential sum’, [ *Quart. J. Math.*]{}, [**51**]{}, (2000), 221–235. B. Kerr, ‘Incomplete Exponential sums over exponential functions’, arXiv:1302.4170. N. M. Korobov, ‘On the distribution of digits in periodic fractions’, [*Matem. Sbornik*]{}, [**89**]{}, (1972), 654–670 (in Russian). F. Luca and I. E. Shparlinski, ‘On the values of the divisor function’, [*Monatsh. Math.*]{} [**154**]{}, (2008), 59–69. H. L. Montgomery and R. C. Vaughan, [*Multiplicative number theory I. Classical Theory*]{}, Cambridge University Press, 2007. W. Narkiewicz, [*Uniform Distribution of Sequences of Integers in Residue Classes*]{}, Springer-Verlag, 1984. L. G. Sathe, ‘On a Congruence Property of the Divisor Function’, [*Am. Jour. Math.*]{}, [**67**]{}, (1945), 397–406. A. Selberg, ‘Note on a paper by L.G. Sathe’, [ *J. Indian Math. Soc.*]{}, [**18**]{}, (1954), 83–87. I. D. Shkredov, ‘Some new inequalities in additive cominatorics’, arXiv:1208.2344, v3 I. E. Shparlinski, ‘Open problems on exponential and character sums’, [*Number Theory: Proc. 5th China-Japan Seminar, Osaka, 2008*]{}, World Scientific, 2010, 222–242. G. Tenenbaum, [*Introduction to Analytic and Probabilistic Number Theory*]{}, Cambrdge University Press, 1995
{ "pile_set_name": "ArXiv" }
--- abstract: | We consider a [conditioned [Galton–Watson]{} tree]{} and prove an estimate of the number of pairs of vertices with a given distance, or, equivalently, the number of paths of a given length. We give two proofs of this result, one probabilistic and the other using generating functions and singularity analysis. Moreover, the second proof yields a more general estimate for generating functions, which is used to prove a conjecture by Bousquet–Mélou and Janson [@SJ185], saying that the vertical profile of a randomly labelled [conditioned [Galton–Watson]{} tree]{} converges in distribution, after suitable normalization, to the density of ISE (Integrated Superbrownian Excursion). address: - 'School of Computer Science, McGill University, 3480 University Street, Montreal, Canada H3A 2K6' - 'Department of Mathematics, Uppsala University, PO Box 480, SE-751 06 Uppsala, Sweden' author: - Luc Devroye - Svante Janson date: 'December 17, 2008' title: 'Distances between pairs of vertices and vertical profile in conditioned Galton–Watson trees' --- [^1] [^2] Introduction and results {#S:intro} ======================== Let ${T_n}$ be a [conditioned [Galton–Watson]{} tree]{}, [i.e.=1000]{}, the random rooted tree ${{\mathcal T}}$ obtained as the family tree of a [[Galton–Watson]{} process]{} with some given offspring distribution $\xi$, conditioned on the number of vertices $|{{\mathcal T}}|=n$. We will always assume that $$\begin{aligned} \label{Axi} {\operatorname{\mathbb E{}}}\xi=1 \qquad \text{and} \qquad 0<{\sigma^2}{:=}{\operatorname{Var}}\xi<\infty\end{aligned}$$ In other words, the [[Galton–Watson]{} process]{} is critical and with finite variance, and ${\operatorname{\mathbb P{}}}(\xi=1)<1$. (Note that this entails $0<{\operatorname{\mathbb P{}}}(\xi=0)<1$.) It is well-known that this assumption is without essential loss of generality, and that the resulting random trees are essentially the same as the simply generated families of trees introduced by Meir and Moon [@MM]. The importance of this construction lies in that many combinatorially interesting random trees are of this type, for example random plane (= ordered) trees, random unordered labelled trees (Cayley trees), random binary trees, and (more generally) random $d$-ary trees. For further examples see [e.g.=1000]{} Aldous [@AldousII] and Devroye [@Devroye]. We consider only $n$ such that ${T_n}$ exists, [i.e.=1000]{}, such that ${\operatorname{\mathbb P{}}}(|{{\mathcal T}}|=n)>0$. The *span* of $\xi$ is defined to be the largest integer $d$ such that $\xi\in d{\mathbb Z}$ a.s. If the span of $\xi$ is $d$, then ${T_n}$ exists only for $n\equiv 1 \pmod d$, and it exists for all large such $n$. We consider in this paper two types of properties of ${T_n}$ that turn out to have proofs using a common argument. First, for an arbitrary rooted tree $\tau$, let $P_k(\tau)$, $k\ge1$, be the number of (unordered) pairs of vertices [$\{v,w\}$]{} in $\tau$ such that the distance $d(v,w)=k$; equivalently, $P_k(\tau)$ is the number of paths of length $k$ in $\tau$. Our first result is an estimate, uniform in all $k$ and $n$, of the expectation of this number $P_k({T_n})$ for a conditioned Galton–Watson tree ${T_n}$. We let in this paper $C_1,C_2$ and $c_1,c_2$ denote various positive constants that may depend on (the distribution of) $\xi$, and sometimes later $\eta$ introduced below, but not on $n$, $k$ and other variables unless explicitly stated. Recall that we tacitly assume . \[T1\] There exists a constant ${\stepcounter{CC}{C_{\arabic{CC}}}}$ such that for all $k\ge1$ and $n\ge1$, ${\operatorname{\mathbb E{}}}P_k({T_n})\le {C_{\arabic{CC}}}nk$. One way to interpret this result is that the expected number of vertices of distance $k$ from a randomly chosen vertex in ${T_n}$ is $O(k)$. In other words, if ${{T_n}^*}$ is ${T_n}$ randomly rerooted, and ${Z}_k(\tau)$ is the number of vertices of distance $k$ from the root in a rooted tree $\tau$, then the following holds. \[C1\] ${\operatorname{\mathbb E{}}}{Z}_k({{T_n}^*})=O(k)$, uniformly in all $k\ge1$ and $n\ge1$. This can be compared to [@SJ167 Theorem 1.13], which shows that $$\label{wk} {\operatorname{\mathbb E{}}}{Z}_k({T_n})=O(k),$$ again uniformly in $k$ and $n$. Note that in the special case when ${T_n}$ is a random (unordered) labelled tree, ${{T_n}^*}$ has the same distribution, so [Corollary \[C1\]]{} reduces to . However, in general, a randomly rerooted [conditioned [Galton–Watson]{} tree]{} is not a [conditioned [Galton–Watson]{} tree]{}. The emphasis is on uniformity in both $k$ and $n$. If we, on the contrary, fix $k$ and consider limits as [${n\to\infty}$]{}, we have ${\operatorname{\mathbb E{}}}{Z}_k({T_n})\to1+k{\sigma^2}$, see Meir and Moon [@MM] and Janson [@SJ167; @SJ188]. It is shown in [@SJ188] that the sequence ${\operatorname{\mathbb E{}}}{Z}_k({T_n})$ is not always monotone in $n$. We give a probabilistic proof of [Theorem \[T1\]]{}, and thus of [Corollary \[C1\]]{} too, in [Section \[Spf1\]]{}. We also give a second proof by first proving a corresponding estimate for the generating function. (We present two different proofs, since we find both methods interesting, and both methods yield as intermediary steps in the proofs other results that we find interesting.) Let ${f_n}(z)$ be the generating function defined by $${f_n}(z) {:=}\sum_{k=1}^\infty {\operatorname{\mathbb E{}}}P_k({T_n})\,z^k . $$ We will use standard singularity analysis, see [e.g.=1000]{} Flajolet and Sedgewick [@FS], and define the domain, for $0<{\beta}<\pi/2$ and ${\delta}>0$, $${{{\Delta}({\beta},{\delta})}}{:=}{\ensuremath{\{z\in{\mathbb C}:|z|<1+{\delta},\, z\neq1,\, |\arg(z-1)|>\pi/2-{\beta}\}}}.$$ Note that $|\arg(z-1)|>\pi/2-{\beta}$ is equivalent to $|\arg(1-z)|<\pi/2+{\beta}$. \[Tgen1\] For every $\xi$ there exist positive constants ${\stepcounter{CC}{C_{\arabic{CC}}}}$, ${\beta}$, ${\delta}$ such that for all $n\ge1$, ${f_n}$ extends to an analytic function in ${{{\Delta}({\beta},{\delta})}}$ with $$\label{fn} |{f_n}(z)| \le {C_{\arabic{CC}}}n|1-z|{^{-2}}, \qquad z\in{{{\Delta}({\beta},{\delta})}}.$$ By standard singularity analysis ([i.e.=1000]{}, estimate of the Taylor coefficients of $f_n(z)$ using Cauchy’s formula and a suitable contour in ${{{\Delta}({\beta},{\delta})}}$), implies ${\operatorname{\mathbb E{}}}P_k(T_n)=O(n k)$, see Flajolet and Sedgewick [@FS], Theorem VI.3 and (for the uniformity in $n$) Lemma IX.2 (applied to the family [$\{f_n(z)/n\}$]{}). Hence, [Theorem \[T1\]]{} follows from [Theorem \[Tgen1\]]{}. For each pair of vertices $v,w$ in a rooted tree, the path from $v$ to $w$ consists of two (possibly empty) parts, one going from $v$ towards the root, ending at the last common ancestor ${v{\wedge}w}$ of $v$ and $w$, and another part going from ${v{\wedge}w}$ to $w$ in the direction away from the root. We will also prove extensions of the results above for ${T_n}$, where we consider separately the lengths of these two parts. Define the corresponding bivariate generating function (now considering ordered pairs $v,w$) $$\label{hn} {h_n}(x,y) {:=}{\operatorname{\mathbb E{}}}{\sum_{v,w\in{T_n}}}x^{d(v,{v{\wedge}w})} y^{d(w,{v{\wedge}w})}.$$ \[Tgen2\] For every $\xi$ there exist positive constants ${\stepcounter{CC}{C_{\arabic{CC}}}}{\xdef\CCgenii{{C_{\arabic{CC}}}}}$, ${\beta}$, ${\delta}$ such that for all $n\ge1$, $$|{h_n}(x,y)| \le {C_{\arabic{CC}}}n|1-x|{^{-1}}|1-y|{^{-1}}, \qquad x,y\in{{{\Delta}({\beta},{\delta})}}.$$ Note that, by and $${h_n}(z,z)={\operatorname{\mathbb E{}}}{\sum_{v,w\in{T_n}}}z^{d(v,w)} = n+2{f_n}(z).$$ Hence [Theorem \[Tgen1\]]{} follows from [Theorem \[Tgen2\]]{}. We prove [Theorem \[Tgen2\]]{} in [Section \[Sgen2\]]{}. If we define ${{\widetilde P_{\ell,m}}}(\tau){:=}\#{\ensuremath{\{(v,w)\in \tau:d(v,{v{\wedge}w})=\ell,\,d(w,{v{\wedge}w})=m\}}}$, then singularity analysis as above (but twice) shows that [Theorem \[Tgen2\]]{} implies the following. (Since $P_k=\frac12\sum_{\ell=0}^k {\widetilde P_{\ell,k-\ell}}$, this too implies [Theorem \[T1\]]{}.) \[T11\] There exists a constant ${\stepcounter{CC}{C_{\arabic{CC}}}}$ such that for all $\ell,m\ge0$ and $n\ge1$, ${\operatorname{\mathbb E{}}}{{\widetilde P_{\ell,m}}}({T_n})\le {C_{\arabic{CC}}}n$. One motivation for these results is that they (more precisely, [Theorem \[Tgen2\]]{}) are used to prove the second type of result in this paper. For this, we assume that we are given a further random variable $\eta$. Given a rooted tree $\tau$, we take an independent copy ${\eta_e}$ of $\eta$ for every edge $e\in \tau$. We give each vertex $v$ the label ${{L_{v}}}$ obtained by summing ${\eta_e}$ for all $e$ in the path from the root ${o}$ to $v$. (Thus, ${L_{{o}}}=0$.) We assume that $$\begin{aligned} \label{Aeta} {\operatorname{\mathbb E{}}}\eta=0 \qquad \text{and} \qquad 0<{{\sigma^2}_\eta}{:=}{\operatorname{Var}}\eta<\infty.\end{aligned}$$ We further assume that $\eta$ is integer valued and with span $1$; thus all labels are integers, and all integers are possible labels. We let $X(j;\tau)$ be the number of vertices in $\tau$ with label $j$; the sequence $(X(j;\tau))_{j=-\infty}^{\infty}$ is the *vertical profile* of the labelled tree. For the random tree ${T_n}$, we assume that the variables ${\eta_e}$ are independent of ${T_n}$. The vertical profile $X(j;{T_n})$ then is a random function defined for $j\in{\mathbb Z}$; we write ${{X_n}}(j){:=}X(j;{T_n})$ and extend the domain of ${{X_n}}$ to ${\mathbb R}$ by linear interpolation between the integer points. Our next theorem says that this function ${{X_n}}$, suitable normalized, converges in distribution in the space ${C_0({\mathbb R})}$ of continuous functions on ${\mathbb R}$ that tend to 0 at $\pm\infty$; we equip ${C_0({\mathbb R})}$ with the usual uniform topology defined by the supremum norm. Let, further, ${f{_{\text{\sc ise}}}}$ denote the density of the random measure ISE introduced by Aldous [@AldousISE]; ${f{_{\text{\sc ise}}}}$ is a random continuous function with (random) compact support, see Bousquet–Mélou and Janson [@SJ185 Theorem 2.1]. \[T2\] With the assumptions above, including and , let ${\gamma}{:=}{{\sigma}_\eta}{^{-1}}{\sigma}{^{1/2}}$. Then, as [${n\to\infty}$]{}, $$\label{t2a} \frac1n {\gamma}{^{-1}}n{^{1/4}}{{X_n}}{\bigl({\gamma}{^{-1}}n{^{1/4}}x\bigr)} {\overset{\mathrm{d}}{{\longrightarrow}}}{f{_{\text{\sc ise}}}}(x),$$ in the space ${C_0({\mathbb R})}$ with the usual uniform topology. Equivalently, $$\label{t2b} n{^{-3/4}}{{X_n}}{\bigl(n{^{1/4}}x\bigr)} {\overset{\mathrm{d}}{{\longrightarrow}}}{\gamma}{f{_{\text{\sc ise}}}}({\gamma}x).$$ Note that the random functions on the left and [right hand side]{}[s]{} of and are density functions, [i.e.=1000]{}, non-negative functions with integral 1. \[C2\] If [${n\to\infty}$]{} and $j_n/n{^{1/4}}\to x$, where $-\infty<x<\infty$, then $n{^{-3/4}}X(j_n;{T_n}){\overset{\mathrm{d}}{{\longrightarrow}}}{\gamma}{f{_{\text{\sc ise}}}}({\gamma}x)$. The limit law is characterized in [@BM] by a formula for its Laplace transform. [Theorem \[T2\]]{} was conjectured in [@SJ185], and proved there in two special cases, [viz.=1000]{} when $\xi$ has the Geometric distribution ${\operatorname{Ge}}(1/2)$ and thus ${T_n}$ is a random ordered tree, and $\eta$ is uniformly distributed on either [$\{-1,1\}$]{} or [$\{-1,0,1\}$]{}. Moreover, it was shown there [@SJ185 Remark 3.7] that the proof given in [@SJ185] applies generally under the assumptions above, provided the following estimate holds. \[L0\] Under the assumptions above, there exists a constant ${\stepcounter{CC}{C_{\arabic{CC}}}}$ such that for all $n\ge1$ and $t\in[-\pi,\pi]$, $$\label{l2} {\operatorname{\mathbb E{}}}{\biggl|\frac1n\sum_j X(j;{T_n})e^{{\mathrm{i}}jt}\biggr|}^2 \le \frac{{C_{\arabic{CC}}}}{1+nt^4}.$$ We prove [Lemma \[L0\]]{}, and thus [Theorem \[T2\]]{}, in [Section \[Sl2\]]{}, assuming [Theorem \[Tgen2\]]{}. Finally, we prove [Theorem \[Tgen2\]]{}, using singularity analysis again, in [Section \[Sgen2\]]{}, which completes the proof of all other results. This research was mainly done during a workshop at Bellairs Research Institute in Barbados, March 2006, and completed during a visit of SJ to Centre de recherches mathématiques, Université de Montréal, October 2008. First proof of [Theorem \[T1\]]{} {#Spf1} ================================= In a rooted tree $\tau$, let $Q_k(\tau)$, $k\ge1$, denote the number of (unordered) pairs of vertices at path distance $k$ from each other such that the path between them visits the root, and let $Q'_k(\tau)$ be the number of such pairs where the root cannot be one of the two vertices in the pair; thus $Q_k(\tau)=Q'_k(\tau) + Z_k(\tau)$. Then, in the Galton–Watson tree ${{\mathcal T}}$, if $\xi$ is the number of children of the root, and the subtrees rooted at these children are denoted ${{\mathcal T}}_1,\dots,{{\mathcal T}}_\xi$, $$\label{qkt} Q'_k({{\mathcal T}}) = \sum_{(r,s): 1 \le r < s \le \xi} \sum_{j=0}^{k-2} Z_j ({{\mathcal T}}_r) Z_{k-2-j} ({{\mathcal T}}_s)$$ and thus, since we assume ${{\mathcal T}}$ to be critical, [i.e.=1000]{}, ${\operatorname{\mathbb E{}}}\xi=1$, so ${\operatorname{\mathbb E{}}}Z_k({{\mathcal T}})=1$ for every $k$, $$\label{qk} {\operatorname{\mathbb E{}}}Q_k({{\mathcal T}})={\operatorname{\mathbb E{}}}Z_k({{\mathcal T}})+{\operatorname{\mathbb E{}}}Q'_k({{\mathcal T}}) = 1+ {\operatorname{\mathbb E{}}}\frac{\xi(\xi-1)}2 (k-1) =1+(k-1)\frac{{\sigma^2}}2.$$ Let ${{\widehat T}_n}$ denote the random subtree of $T_n$ rooted at a uniformly selected random vertex. (Note the difference from ${{T_n}^*}$ in [Corollary \[C1\]]{}; in ${{T_n}^*}$ we keep all $n$ vertices, but in ${{\widehat T}_n}$ we keep only the vertices below the new root.) Then, clearly, $${{\operatorname{\mathbb E{}}}}\{ P_{k}(T_n) \} = n {{\operatorname{\mathbb E{}}}}\{ Q_k ({\widehat T}_n) \}.$$ Consequently, [Theorem \[T1\]]{} is equivalent to: \[TP1\] There exists a constant ${\stepcounter{CC}{C_{\arabic{CC}}}}$ such that for all $k\ge1$ and $n\ge1$, ${\operatorname{\mathbb E{}}}Q_k({{\widehat T}_n})\le {C_{\arabic{CC}}}k$. In order to prove this, we will need a related, but different, estimate for the [conditioned [Galton–Watson]{} tree]{} $T_n$. \[TQ\] There exists a constant ${\stepcounter{CC}{C_{\arabic{CC}}}}{\xdef\CCTQ{{C_{\arabic{CC}}}}}$ such that for all $k\ge1$ and $n\ge1$, ${\operatorname{\mathbb E{}}}Q_k({T_n})\le {C_{\arabic{CC}}}k\sqrt n$. It is easy to see ${\operatorname{\mathbb E{}}}Q_k({T_n})\ge {\stepcounter{cc}{c_{\arabic{cc}}}}n^{3/2}$ when $k\sim\sqrt n$, so the estimate in [Theorem \[TQ\]]{} then is of the right order; in particular, the estimate in [Theorem \[TP1\]]{} for ${{\widehat T}_n}$ does *not* hold for ${T_n}$. To prove these theorems we use a few more or less standard estimates. \[L2\] Assume, as above, , and let $d$ be the span of $\xi$. Let $S_n{:=}\sum_{i=1}^n\xi_i$, where $\xi_i$ are independent copies of $\xi$. Then, for $n\equiv 1\pmod d$, $$\label{tail} {\operatorname{\mathbb P{}}}(|{{\mathcal T}}|=n)=\frac1n{\operatorname{\mathbb P{}}}{(S_n=n-1)} \sim \frac{d}{\sigma \sqrt{2 \pi}\, n^{3/2}} \quad\text{as }{\ensuremath{{n\to\infty}}}.$$ More generally, let $W_\ell{:=}\sum_{i=1}^\ell|{{\mathcal T}}_i|$ be the size of the union of $\ell$ independent copies of ${{\mathcal T}}$, or equivalently, the total progeny of a Galton–Watson process started with $\ell$ individuals, with offspring distribution $\xi$. Then, for all $\ell\ge1$ and $n\ge1$, $$\label{l2b} {\operatorname{\mathbb P{}}}(W_\ell=n)=\frac{\ell}n{\operatorname{\mathbb P{}}}(S_{n}=n-\ell) \le {\stepcounter{CC}{C_{\arabic{CC}}}}\ell n{^{-3/2}}\exp(-{\stepcounter{cc}{c_{\arabic{cc}}}}\ell^2/n). {\xdef\CClb{{C_{\arabic{CC}}}}} {\xdef\cclb{{c_{\arabic{cc}}}}}$$ In particular, $$\label{l2c} {\operatorname{\mathbb P{}}}(W_\ell=n)\le {\stepcounter{CC}{C_{\arabic{CC}}}}n{^{-1}}.$$ The identity in is well-known, see [e.g.=1000]{}, Dwass [@Dwass], Kolchin [@Kolchin Lemma 2.1.3, p. 105] and @Pitman:enum. The identity in is the special case $\ell=1$, and the well-known tail estimate in then follows by the local central limit theorem, see, [e.g.=1000]{}, Kolchin [@Kolchin Lemma 2.1.4, p. 105]. Similarly, the inequality in follows by the estimate ${\operatorname{\mathbb P{}}}(S_n=n-\ell)\le \CClb n{^{-1/2}}\exp(-\cclb\ell^2/n)$ from [@SJ167 Lemma 2.1]. The inequality $e^{-x}\le x{^{-1/2}}$ yields . \[L1a\] For every $r>0$ there is a constant ${\stepcounter{CC}{C_{\arabic{CC}}}}(r){\xdef\CCr{{C_{\arabic{CC}}}}}$ such that for all $k\ge0$ and $n\ge1$, ${\operatorname{\mathbb E{}}}Z_k(T_n)^r\le \CCr(r) n^{r/2}$. For any rooted tree ${T}$, let ${T}{^{(k)}}$ be the tree pruned at height $k$, i.e., the subtree consisting of all vertices of distance at most $k$ from the root. Let $\tau$ be a given rooted tree of height $k$, and let $m{:=}Z_k(\tau)$, the number of leaves at maximal depth. Note that if $\tau= {T}{^{(k)}}$ for some tree ${T}$, then $|{T}|=n$ if and only if ${T}$ has $n-|\tau|$ vertices at greater depth than $k$, and thus $N{:=}n-|\tau|+m$ vertices at depth $k$ or greater. Hence, with $W_m$ as in [Lemma \[L2\]]{} and using and , for any $r>0$ and assuming $N>0$ (otherwise the probability is 0), $$\begin{split} {\operatorname{\mathbb P{}}}{\bigl(T_n{^{(k)}}=\tau\bigr)} &= \frac{{\operatorname{\mathbb P{}}}{\bigl({{\mathcal T}}{^{(k)}}=\tau,\,|{{\mathcal T}}|=n\bigr)}}{{\operatorname{\mathbb P{}}}(|{{\mathcal T}}|=n)} =\frac{{\operatorname{\mathbb P{}}}({{\mathcal T}}{^{(k)}}=\tau){\operatorname{\mathbb P{}}}(W_m=N)}{{\operatorname{\mathbb P{}}}(|{{\mathcal T}}|=n)} \\& \le {\stepcounter{CC}{C_{\arabic{CC}}}}n{^{3/2}}{\operatorname{\mathbb P{}}}({{\mathcal T}}{^{(k)}}=\tau)m N{^{-3/2}}e^{-\cclb m^2/N} \\& \le {\stepcounter{CC}{C_{\arabic{CC}}}}(r) n{^{3/2}}{\operatorname{\mathbb P{}}}({{\mathcal T}}{^{(k)}}=\tau)m N{^{-3/2}}(m^2/N)^{-r/2} \\& = {C_{\arabic{CC}}}(r) n{^{3/2}}N^{r/2-3/2} m^{1-r}{\operatorname{\mathbb P{}}}({{\mathcal T}}{^{(k)}}=\tau). \end{split}$$ If $r\ge3$, this yields, since $N\le n$, the estimate $${\operatorname{\mathbb P{}}}(T_n{^{(k)}}=\tau)\le {C_{\arabic{CC}}}(r) n^{r/2} m^{1-r}{\operatorname{\mathbb P{}}}({{\mathcal T}}{^{(k)}}=\tau),$$ and summing over all $\tau$ of height $k$ with $Z_k(\tau)=m$ we obtain $${\operatorname{\mathbb P{}}}(Z_k(T_n)=m)\le {C_{\arabic{CC}}}(r) n^{r/2} m^{1-r}{\operatorname{\mathbb P{}}}(Z_k({{\mathcal T}})=m).$$ Consequently, $$\begin{split} {\operatorname{\mathbb E{}}}Z_k(T_n)^r &={\sum_{m=1}^\infty}m^r {\operatorname{\mathbb P{}}}(Z_k(T_n)=m) \\& \le {C_{\arabic{CC}}}(r) n^{r/2} {\sum_{m=1}^\infty}m{\operatorname{\mathbb P{}}}(Z_k({{\mathcal T}})=m) \\& = {C_{\arabic{CC}}}(r) n^{r/2} {\operatorname{\mathbb E{}}}Z_k({{\mathcal T}}) = {C_{\arabic{CC}}}(r) n^{r/2}. \end{split}$$ This proves the result for $r\ge3$, and the result for $0<r<3$ follows by Lyapounov’s (or Hölder’s) inequality. \[L1b\] For all $k\ge1$ and $n\ge1$, ${\operatorname{\mathbb E{}}}Z_k(T_n)\le {\stepcounter{CC}{C_{\arabic{CC}}}}(k{\wedge}\sqrt n) {\xdef\CCLi{{C_{\arabic{CC}}}}}$. Equivalently, for all $k\ge0$ and $n\ge1$, ${\operatorname{\mathbb E{}}}Z_k(T_n)\le {\stepcounter{CC}{C_{\arabic{CC}}}}((k+1){\wedge}\sqrt n) {\xdef\CCLia{{C_{\arabic{CC}}}}}$. The estimate ${\operatorname{\mathbb E{}}}Z_k(T_n)\le {\stepcounter{CC}{C_{\arabic{CC}}}}k$ is , which is proved in [@SJ167 Theorem 1.13]. The estimate ${\operatorname{\mathbb E{}}}Z_k(T_n)\le {\stepcounter{CC}{C_{\arabic{CC}}}}\sqrt n$ is proved in [Lemma \[L1a\]]{}. The estimate ${\operatorname{\mathbb E{}}}Z_k(T_n)\le {\stepcounter{CC}{C_{\arabic{CC}}}}\sqrt n$ was proved by Drmota and Gittenberger [@DrmotaG:width], assuming that $\xi$ has an exponential moment; in fact, they then prove the stronger bound ${\operatorname{\mathbb E{}}}Z_k(T_n)\le {\stepcounter{CC}{C_{\arabic{CC}}}}\sqrt n \exp(-{\stepcounter{cc}{c_{\arabic{cc}}}}k/\sqrt n)$. The bound in [Lemma \[L1b\]]{} can be further improved to ${\operatorname{\mathbb E{}}}Z_k(T_n)\le {\stepcounter{CC}{C_{\arabic{CC}}}}k \exp(-{\stepcounter{cc}{c_{\arabic{cc}}}}k^2/ n)$, but we do not know a reference for this estimate. (Details may appear elsewhere.) Note that [Lemma \[L1a\]]{} yields an estimate $O(n^{r/2})$ of the $r$th moment of $Z_k(T_n)$ for an arbitrary $r$ assuming only a second moment of $\xi$. This is in contrast to the estimate , where the corresponding estimate ${\operatorname{\mathbb E{}}}Z_k(T_n)^r = O(k^r)$ is valid (for integer $r\ge1$ at least) if $\xi$ has a finite $r+1$:th moment, but not otherwise (not even for a fixed $k\ge2$); one direction is by Theorem 1.13 in [@SJ167], and the converse follows from the discussion after Lemma 2.1 in [@SJ167]. We have $Q_k (T_n)=Q'_k (T_n)+Z_k (T_n)$, and ${\operatorname{\mathbb E{}}}Z_k (T_n)\le \CCLi k$ by [Lemma \[L1b\]]{}, so it suffices to show that ${\operatorname{\mathbb E{}}}Q'_k (T_n)\le {\stepcounter{CC}{C_{\arabic{CC}}}}k\sqrt n$. We use , condition on $|{{\mathcal T}}|=n$ and take expectations. Using the symmetry and recalling that ${{\mathcal T}}_1,\dots,{{\mathcal T}}_\xi$ are independent and $({{\mathcal T}}_i\mid |{{\mathcal T}}_i|=n_i){\overset{\mathrm{d}}{=}}({{\mathcal T}}\mid |{{\mathcal T}}|=n_i){\overset{\mathrm{d}}{=}}T_{n_i}$ for any $n_i$, we obtain, with $p_\ell{:=}{\operatorname{\mathbb P{}}}(\xi=\ell)$ and $q_m{:=}{\operatorname{\mathbb P{}}}(|{{\mathcal T}}|=m)$, $$\begin{split} {{\operatorname{\mathbb E{}}}}\left\{ Q'_k (T_n) \right\} &= \frac{ {{\operatorname{\mathbb E{}}}}\left\{ {\boldsymbol1_{[\xi \ge 2, |{{\mathcal T}}|=n ]}} \sum_{1 \le r<s \le \xi} \sum_{j=0}^{k-2} Z_{j} ({{\mathcal T}}_r) Z_{k-2-j} ({{\mathcal T}}_s) \right\}} { {{\operatorname{\mathbb P{}}}}\{ |{{\mathcal T}}| = n \} } \\ &= \frac{ {{\operatorname{\mathbb E{}}}}\left\{ {\boldsymbol1_{[|{{\mathcal T}}|=n ]}}\binom{\xi}2 \sum_{j=0}^{k-2} Z_{j} ({{\mathcal T}}_1) Z_{k-2-j} ({{\mathcal T}}_2) \right\}} {{{\operatorname{\mathbb P{}}}}\{ |{{\mathcal T}}| = n \} } \\ &= q_n{^{-1}}\sum_{\ell=2}^\infty p_\ell \binom{\ell}2 \sum_{n_1,n_2\ge1} q_{n_1} q_{n_2} {\operatorname{\mathbb P{}}}{\left(\sum_{i=3}^\ell|{{\mathcal T}}_i|=n-1-n_1-n_2\right)} \\&\qquad\qquad\times \sum_{j=0}^{k-2} {{\operatorname{\mathbb E{}}}}\left\{ Z_{j}(T_{n_1}) \right\} {{\operatorname{\mathbb E{}}}}\left\{ Z_{k-2-j}(T_{n_2}) \right\}. \end{split}$$ We begin with the inner sum over $j$, ${\Sigma_1(n_1,n_2)}$ say. By symmetry, we consider only $n_1\le n_2$, and then we obtain from [Lemma \[L1b\]]{} the estimates ${\operatorname{\mathbb E{}}}Z_{k-2-j}(T_{n_2})\le \CCLia((k-1-j){\wedge}n_2{^{1/2}})\le \CCLia(k{\wedge}n_2{^{1/2}})$ and $$\sum_{j=0}^{k-2} {{\operatorname{\mathbb E{}}}}\left\{ Z_{j}(T_{n_1}) \right\} \le \begin{cases} \sum_{j=0}^{k-2} \CCLia(j+1) \le \CCLia k^2 , \\ {\operatorname{\mathbb E{}}}\sum_{j=0}^{\infty} Z_{j}(T_{n_1}) =n_1. \end{cases}$$ Hence $$\label{sa} {\Sigma_1(n_1,n_2)}\le {\stepcounter{CC}{C_{\arabic{CC}}}}(k^2{\wedge}n_1)(k{\wedge}n_2{^{1/2}}).$$ Let $m{:=}n_1+n_2$ and sum over $n_1,n_2$ with a given sum $m$. We have by and , $$\label{sbm} \begin{split} {\Sigma_2(m)}&:= \sum_{n_1+n_2=m} q_{n_1}q_{n_2}{\Sigma_1(n_1,n_2)}\\& \le 2\sum_{n_1=1}^{m/2} q_{n_1}q_{m-n_1}{C_{\arabic{CC}}}(k^2{\wedge}n_1)(k{\wedge}(m-n_1){^{1/2}}) \\ &\le {\stepcounter{CC}{C_{\arabic{CC}}}}\sum_{n_1=1}^{m/2} n_1{^{-3/2}}(m-n_1){^{-3/2}}(k^2{\wedge}n_1) (k{\wedge}(m-n_1){^{1/2}}) \\ &\le {\stepcounter{CC}{C_{\arabic{CC}}}}\frac{k{\wedge}m{^{1/2}}}{m{^{3/2}}} \sum_{n_1=1}^{m/2} \frac{k^2{\wedge}n_1}{n_1{^{3/2}}} \\ &\le {\stepcounter{CC}{C_{\arabic{CC}}}}\frac{k{\wedge}m{^{1/2}}}{m{^{3/2}}} (k{\wedge}m{^{1/2}}) = {C_{\arabic{CC}}}\frac{k^2{\wedge}m}{m{^{3/2}}}. \end{split}$$ We define further $$\label{scl} {\Sigma_3(\ell)}{:=}\sum_{m=2}^{n-1} {\Sigma_2(m)}{\operatorname{\mathbb P{}}}{\Bigl(\sum_{i=3}^\ell|{{\mathcal T}}_i|=n-1-m\Bigr)};$$ thus $$\label{qg} {\operatorname{\mathbb E{}}}Q_k'(T_n) = q_n{^{-1}}\sum_{\ell=2}^\infty p_\ell \binom{\ell}2 {\Sigma_3(\ell)}\le {\stepcounter{CC}{C_{\arabic{CC}}}}n{^{3/2}}\sum_{\ell=2}^\infty p_\ell {\ell}^2 {\Sigma_3(\ell)}.$$ We will show that ${\Sigma_3(\ell)}\le {\stepcounter{CC}{C_{\arabic{CC}}}}k/n$, uniformly in $\ell\ge2$, and the result follows by , recalling that $\sum_\ell p_\ell\ell^2={\operatorname{\mathbb E{}}}\xi^2<\infty$. (The proof can be simplified in the case ${\operatorname{\mathbb E{}}}\xi^3<\infty$, when it suffices to show that ${\Sigma_3(\ell)}\le {\stepcounter{CC}{C_{\arabic{CC}}}}k\ell/n$.) First, if $\ell=2$, the only non-zero term in is for $m=n-1$, which yields, by , $${\Sigma_3(2)}={\Sigma_2(n-1)}\le {\stepcounter{CC}{C_{\arabic{CC}}}}\frac{k^2{\wedge}n}{n{^{3/2}}} \le {C_{\arabic{CC}}}\frac{k\sqrt n}{n{^{3/2}}}.$$ For $\ell>2$, we split the sum in into two parts, with $m\le n/2$ and $m>n/2$. We have, by , $$\begin{split} \sum_{m=n/2}^{n-1} {\Sigma_2(m)}{\operatorname{\mathbb P{}}}{\Bigl(\sum_{i=3}^\ell|{{\mathcal T}}_i|=n-1-m\Bigr)} &\le {\stepcounter{CC}{C_{\arabic{CC}}}}\frac{k^2{\wedge}n}{n{^{3/2}}} {\operatorname{\mathbb P{}}}{\Bigl(\sum_{i=3}^\ell|{{\mathcal T}}_i|\le n/2\Bigr)} \\ \le {C_{\arabic{CC}}}\frac{k^2{\wedge}n}{n{^{3/2}}} \le {C_{\arabic{CC}}}\frac k n. \end{split}$$ Similarly, using and (with $\ell$ replaced by $\ell-2$), $$\begin{split} \sum_{m=1}^{n/2} {\Sigma_2(m)}{\operatorname{\mathbb P{}}}{\Bigl(\sum_{i=3}^\ell|{{\mathcal T}}_i|=n-1-m\Bigr)} &\le {\stepcounter{CC}{C_{\arabic{CC}}}}\sum_{m=1}^{n/2} \frac{k^2{\wedge}m}{m{^{3/2}}}\cdot \frac 1n \\ &\le \frac {C_{\arabic{CC}}}n \sum_{m=1}^{\infty} \frac{k^2{\wedge}m}{m{^{3/2}}} \le {\stepcounter{CC}{C_{\arabic{CC}}}}\frac k n. \end{split}$$ Thus ${\Sigma_3(\ell)}\le {\stepcounter{CC}{C_{\arabic{CC}}}}k/n$, and the theorem follows by . Aldous [@Aldous:fringe] has studied the behavior of a random subtree ${{\widehat T}_n}$ in a conditional Galton–Watson tree ${T_n}$. In particular, he has the following identity [@Aldous:fringe p. 242], for any fixed ordered tree $\tau$ of order at most $n$ (provided that the probabilities in the denominators are nonzero): $$\frac{{{\operatorname{\mathbb P{}}}}\{ {\widehat T}_n = \tau \} }{ {{\operatorname{\mathbb P{}}}}\{ {{\mathcal T}}= \tau \} } = \frac{(n-|\tau|+1) {{\operatorname{\mathbb P{}}}}\{ |{{\mathcal T}}| = n - |\tau| +1 \} } { n {{\operatorname{\mathbb P{}}}}\{ |{{\mathcal T}}| = n \} } \cdot \frac{\gamma }{ p_0},$$ where $\gamma$ is the expected proportion of leaves in ${\widehat T}_{n-|\tau|+1}$ and $p_0 = {{\operatorname{\mathbb P{}}}}\{ \xi = 0 \}$. We will simply bound $\gamma \le 1$, but it is well-known that as $n-|\tau|+1 \to \infty$, $\gamma \to p_0$, see [e.g.=1000]{} Kolchin [@Kolchin Theorem 2.3.1, p. 113]. Thus, using the well-known tail estimate , for all (permitted) $n$ and $\tau$ $$\begin{split} \frac{{{\operatorname{\mathbb P{}}}}\{ {\widehat T}_n = \tau \} }{ {{\operatorname{\mathbb P{}}}}\{ {{\mathcal T}}= \tau \} } &\le {\stepcounter{CC}{C_{\arabic{CC}}}}\frac{(n-|\tau|+1) {{\operatorname{\mathbb P{}}}}\{ |{{\mathcal T}}| = n - |\tau| +1 \} } { n {{\operatorname{\mathbb P{}}}}\{|{{\mathcal T}}| = n \} } \le {\stepcounter{CC}{C_{\arabic{CC}}}}{\xdef\CCb{{C_{\arabic{CC}}}}} \sqrt\frac{ n }{ n-|\tau|+1} . \end{split}$$ Hence, $$\begin{split} {{\operatorname{\mathbb E{}}}}\{ Q_k ({\widehat T}_n) \} &= \sum_{\tau} Q_k (\tau) {{\operatorname{\mathbb P{}}}}\{ {\widehat T}_n = \tau \} \\ &\le \CCb\sum_{\tau} Q_k (\tau) \sqrt\frac{ n }{ n-|\tau|+1} {{\operatorname{\mathbb P{}}}}\{ {{\mathcal T}}= \tau \}\\ &= \CCb{\operatorname{\mathbb E{}}}{\left(Q_k ({{\mathcal T}}) \sqrt\frac{ n }{ n-|{{\mathcal T}}|+1}\,\right)} \\ &\le {\stepcounter{CC}{C_{\arabic{CC}}}}{{\operatorname{\mathbb E{}}}}\{ Q_k ({{\mathcal T}}) \} + \CCb \sum_{n \ge \ell > n/2} \sqrt\frac{ n }{ n-\ell+1} \, {{\operatorname{\mathbb E{}}}}\left\{ {\boldsymbol1_{[|{{\mathcal T}}|=\ell]}} Q_k ({{\mathcal T}}) \right\}. \end{split}$$ We have ${\operatorname{\mathbb E{}}}{\{Q_k({{\mathcal T}})\}}\le{\stepcounter{CC}{C_{\arabic{CC}}}}k{\xdef\CCk{{C_{\arabic{CC}}}}}$ by , and, using and [Theorem \[TQ\]]{}, $${{\operatorname{\mathbb E{}}}}\left\{ {\boldsymbol1_{[|{{\mathcal T}}|=\ell]}} Q_k ({{\mathcal T}}) \right\} ={\operatorname{\mathbb P{}}}{\{|{{\mathcal T}}|=\ell\}} {{\operatorname{\mathbb E{}}}}\left\{ Q_k (T_\ell) \right\} \le {\stepcounter{CC}{C_{\arabic{CC}}}}\ell{^{-3/2}}k\ell{^{1/2}}={C_{\arabic{CC}}}k/\ell.$$ Consequently, $$\begin{split} {{\operatorname{\mathbb E{}}}}\{ Q_k ({\widehat T}_n) \} &\le \CCk k+{\stepcounter{CC}{C_{\arabic{CC}}}}\sum_{n \ge \ell > n/2} \frac{k }{ \ell} \, \sqrt\frac{ n }{ n-\ell+1} \\ &\le \CCk k+{\stepcounter{CC}{C_{\arabic{CC}}}}\frac kn \sum_{j=1}^n \sqrt\frac{n}{ j} \\ &\le {\stepcounter{CC}{C_{\arabic{CC}}}}k. \\ \end{split}$$ This proves [Theorem \[TP1\]]{}, which by the argument at the beginning of the section yields [Theorem \[T1\]]{}. Proof of [Lemma \[L0\]]{} and [Theorem \[T2\]]{} {#Sl2} ================================================ Denote the [left hand side]{} of by ${\Psi(n,t)}$. Since $\sum_j X(j;{T_n})e^{{\mathrm{i}}jt} = {\sum_{v\in{T_n}}}e^{{\mathrm{i}}t L_v} $,we have $$\label{b1} {\Psi(n,t)}=n{^{-2}}{\operatorname{\mathbb E{}}}\,{\biggl|{\sum_{v\in{T_n}}}e^{{\mathrm{i}}t L_v}\biggr|}^2 =n{^{-2}}{\operatorname{\mathbb E{}}}{\sum_{v,w\in{T_n}}}e^{{\mathrm{i}}t (L_v-L_w)}.$$ Condition on $T_n$ and consider two vertices $v$ and $w$ in ${T_n}$. If ${v{\wedge}w}$ is the last common ancestor of $v$ and $w$, then $L_v-L_{{v{\wedge}w}}$ and $L_w-L_{{v{\wedge}w}}$ are (conditionally, given ${T_n}$) independent sums of $d(v,{v{\wedge}w})$ and $d(w,{v{\wedge}w})$ copies of $\eta$, respectively. Consequently, letting ${{\varphi}_\eta}(t){:=}{\operatorname{\mathbb E{}}}e^{{\mathrm{i}}t\eta}$ be the characteristic function of $\eta$, $$\begin{split} {\operatorname{\mathbb E{}}}{\bigl(e^{{\mathrm{i}}t (L_v-L_w)}\mid{T_n}\bigr)} &= {\operatorname{\mathbb E{}}}{\bigl(e^{{\mathrm{i}}t (L_v-L_{{v{\wedge}w}})}\mid{T_n}\bigr)} {\operatorname{\mathbb E{}}}{\bigl(e^{-{\mathrm{i}}t (L_w-L_{{v{\wedge}w}})}\mid{T_n}\bigr)} \\& ={{\varphi}_\eta}(t)^{d(v,{v{\wedge}w})}\,\overline{{{\varphi}_\eta}(t)}^{d(w,{v{\wedge}w})}. \end{split}$$ Hence, by and , $${\Psi(n,t)}= n{^{-2}}{\operatorname{\mathbb E{}}}{\sum_{v,w\in{T_n}}}{{\varphi}_\eta}(t)^{d(v,{v{\wedge}w})}\overline{{{\varphi}_\eta}(t)}^{d(w,{v{\wedge}w})} =n{^{-2}}{h_n}{\bigl({{\varphi}_\eta}(t),\overline{{{\varphi}_\eta}(t)}\bigr)}$$ and [Theorem \[Tgen2\]]{} yields $$\label{b2} {\Psi(n,t)}\le \CCgenii n{^{-1}}|1-{{\varphi}_\eta}(t)|{^{-2}}.$$ Since ${\operatorname{\mathbb E{}}}\eta=0$ and ${\operatorname{\mathbb E{}}}\eta^2={{\sigma^2}_\eta}<\infty$, we have ${{\varphi}_\eta}(t)=\exp{\bigl(-\tfrac12{{\sigma^2}_\eta}t^2+o(t^2)\bigr)}$ for small $|t|$; moreover, since $\eta$ has span 1, ${{\varphi}_\eta}(t)\neq1$ for $0<|t|\le\pi$. It follows that $\psi(t){:=}(1-{{\varphi}_\eta}(t))/t^2$ is a continuous non-zero function on $[-\pi,\pi]$ (with $\psi(0){:=}\frac12{{\sigma^2}_\eta}$); hence, by compactness, $|\psi(t)|\ge {\stepcounter{cc}{c_{\arabic{cc}}}}$ for some ${\stepcounter{cc}{c_{\arabic{cc}}}}>0$, and thus $$|1-{{\varphi}_\eta}(t)|\ge {c_{\arabic{cc}}}t^2, \qquad |t|\le\pi.$$ It now follows from , and the obvious fact that ${\Psi(n,t)}\le1$, that $$(1+nt^4){\Psi(n,t)}\le 1+nt^4{\Psi(n,t)}\le 1+ \CCgenii \frac{t^4}{|1-{{\varphi}_\eta}(t)|^2} \le {\stepcounter{CC}{C_{\arabic{CC}}}}.$$ This proves [Lemma \[L0\]]{}, which as remarked in [Section \[S:intro\]]{} implies [Theorem \[T2\]]{} by [@SJ185 Remark 3.7]. Proof of [Theorem \[Tgen2\]]{} {#Sgen2} ============================== We use some further generating functions. Recall that ${{\mathcal T}}$ is the (unconditioned) [[Galton–Watson]{} tree]{} with offspring distribution $\xi$, and define $$\begin{aligned} \Phi(z) &{:=}{\operatorname{\mathbb E{}}}z^\xi, \\ F(z) &{:=}{\operatorname{\mathbb E{}}}z^{|{{\mathcal T}}|}, \\ G(z,x) &{:=}{\operatorname{\mathbb E{}}}{\Bigl(z^{|{{\mathcal T}}|}{\sum_{v\in{{\mathcal T}}}}x^{d(v,{o})}\Bigr)}, \\ H(z,x,y) &{:=}{\operatorname{\mathbb E{}}}{\Bigl(z^{|{{\mathcal T}}|}{\sum_{v,w\in{{\mathcal T}}}}x^{d(v,{v{\wedge}w})}y^{d(w,{v{\wedge}w})}\Bigr)} =\sum_{n=1}^\infty {\operatorname{\mathbb P{}}}(|{{\mathcal T}}|=n){h_n}(x,y)z^n.\end{aligned}$$ These functions are defined and analytic at least for $|z|,|x|,|y|<1$. Let us condition on the degree ${d_{o}}$ of the root of ${{\mathcal T}}$, recalling that ${d_{o}}{\overset{\mathrm{d}}{=}}\xi$. If ${d_{o}}=\ell$, then ${{\mathcal T}}$ has $\ell$ subtrees ${{\mathcal T}}_1,\dots,{{\mathcal T}}_\ell$ at the root ${o}$, and conditioned on ${d_{o}}=\ell$, these are independent and with the same distribution as ${{\mathcal T}}$; we denote their roots (the neighbours of ${o}$), by ${o}_1,\dots,{o}_\ell$. Assume ${d_{o}}=\ell$, and let $|z|,|x|,|y|<1$. First, $|{{\mathcal T}}|=1+{\sum_{i=1}^\ell}|{{\mathcal T}}_i|$ and thus $z^{|{{\mathcal T}}|}=z{\prod_{i=1}^\ell}z^{|{{\mathcal T}}_i|}$. Taking the expectation, we obtain, as is well-known, first $${\operatorname{\mathbb E{}}}{\bigl(z^{|{{\mathcal T}}|}\mid{d_{o}}=\ell\bigr)} = z{\operatorname{\mathbb E{}}}{\prod_{i=1}^\ell}z^{|{{\mathcal T}}_i|} = z F(z)^\ell,$$ and then $$F(z)={\operatorname{\mathbb E{}}}{\bigl(z^{|{{\mathcal T}}|}\bigr)} = z {\sum_{\ell=0}^\infty}{\operatorname{\mathbb P{}}}(\xi=\ell)F(z)^\ell =z{\Phi{(F(z))}}.$$ Similarly, separating the cases $v\in{{\mathcal T}}_i$, $i=1,\dots,\ell$, and $v={o}$, $${\sum_{v\in{{\mathcal T}}}}x^{d(v,{o})} = {\sum_{i=1}^\ell}\sum_{v\in{{\mathcal T}}_i} x^{d(v,{o}_i)+1} +1.$$ Hence, $$\begin{aligned} {\operatorname{\mathbb E{}}}{\bigl(z^{|{{\mathcal T}}|}{\sum_{v\in{{\mathcal T}}}}x^{d(v,{o})}\mid{d_{o}}=\ell\bigr)} &= {\operatorname{\mathbb E{}}}{\sum_{i=1}^\ell}z z^{|{{\mathcal T}}_i|}\sum_{v\in{{\mathcal T}}_i} x^{d(v,{o}_i)+1} \prod_{j\neq i} z^{|{{\mathcal T}}_j|} +{\operatorname{\mathbb E{}}}{\Bigl(z{\prod_{i=1}^\ell}z^{|{{\mathcal T}}_i|}\Bigr)} \\ &= \ell zx G(z,x)F(z)^{\ell-1}+ z F(z)^\ell\end{aligned}$$ and $$G(z,x)=\sum_{\ell=0}^\infty{\operatorname{\mathbb P{}}}(\xi=\ell)\ell zxG(z,x)F(z)^{\ell-1}+F(z) =zx\Phi'(F(z))G(z,x)+F(z)$$ which gives $$\label{yg} G(z,x)=\frac{F(z)}{1-zx\Phi'(F(z))}.$$ Similarly, $$\begin{aligned} {\operatorname{\mathbb E{}}}{\Bigl(z^{|{{\mathcal T}}|}&{\sum_{v,w\in{{\mathcal T}}}}x^{d(v,{v{\wedge}w})}y^{d(w,{v{\wedge}w})}\mid{d_{o}}=\ell\Bigr)} = \\ &{\operatorname{\mathbb E{}}}{\sum_{i=1}^\ell}z z^{|{{\mathcal T}}_i|}\sum_{v,w\in{{\mathcal T}}_i} x^{d(v,{v{\wedge}w})}y^{d(w,{v{\wedge}w})} \prod_{j\neq i} z^{|{{\mathcal T}}_j|} \\& \quad+ {\operatorname{\mathbb E{}}}\sum_{i\neq j} z z^{|{{\mathcal T}}_i|}\sum_{v\in{{\mathcal T}}_i} x^{d(v,{o}_i)+1} z^{|{{\mathcal T}}_j|}\sum_{w\in{{\mathcal T}}_j} y^{d(w,{o}_j)+1} \prod_{k\neq i,j} z^{|{{\mathcal T}}_k|} \\& \quad+ {\operatorname{\mathbb E{}}}{\sum_{i=1}^\ell}z z^{|{{\mathcal T}}_i|}\sum_{v\in{{\mathcal T}}_i} x^{d(v,{o}_i)+1} \prod_{k\neq i} z^{|{{\mathcal T}}_k|} \\& \quad+ {\operatorname{\mathbb E{}}}{\sum_{j=1}^\ell}z z^{|{{\mathcal T}}_j|}\sum_{w\in{{\mathcal T}}_j} y^{d(w,{o}_j)+1} \prod_{k\neq j} z^{|{{\mathcal T}}_k|} +{\operatorname{\mathbb E{}}}{\Bigl(z{\prod_{i=1}^\ell}z^{|{{\mathcal T}}_i|}\Bigr)}\end{aligned}$$ leading to $$\begin{gathered} H(z,x,y) = z\Phi'(F(z))H(z,x,y) + zxy\Phi''(F(z))G(z,x)G(z,y)\\ + zx\Phi'(F(z))G(z,x) + zy\Phi'(F(z))G(z,y) +F(z)\end{gathered}$$ which gives $$\begin{gathered} \label{yh} H(z,x,y) =\\ \frac{zxy\Phi''(F(z))G(z,x)G(z,y) + z\Phi'(F(z)){\bigl(xG(z,x)+ yG(z,y)\bigr)} +F(z)} {1-z\Phi'(F(z))} $$ Assume now for simplicity that $\xi$ has span 1. (The case when the span is $d>1$ is treated similarly with the standard modification that we have to give special treatment to neighbourhoods of the $d$:th unit roots.) Then, by [@SJ167 Lemma A.2], for some ${\delta}>0$ and ${\beta}\le\pi/4$, $F$ extends to an analytic function in ${{{\Delta}({\beta},{\delta})}}$ with $|F(z)|<1$ for $z\in{{{\Delta}({\beta},{\delta})}}$ and $$\label{a5} F(z)=1-\sqrt2{\sigma}{^{-1}}\sqrt{1-z}+{o{\bigl(|z-1|^{1/2}\bigr)}}, \qquad \text{as $z\to1$ with } z\in{{{\Delta}({\beta},{\delta})}}.$$ We will prove the following companion results. \[Lx\] If $\xi$ has span $1$, then there exists ${\beta},{\delta}>0$ such that $F$ extends to an analytic function in ${{{\Delta}({\beta},{\delta})}}$ and, for some ${\stepcounter{cc}{c_{\arabic{cc}}}}{\xdef\cclxa{{c_{\arabic{cc}}}}},{\stepcounter{cc}{c_{\arabic{cc}}}}{\xdef\ccj{{c_{\arabic{cc}}}}}>0$, if $x,z\in{{{\Delta}({\beta},{\delta})}}$, then $$\begin{aligned} \label{sofie} |1-z\Phi'(F(z))|&\ge \cclxa|1-z|{^{1/2}}, \\ |1-xz\Phi'(F(z))|&\ge \ccj|1-x|.\label{julie}\end{aligned}$$ Consequently, $G(z,x)$ and $H(z,x,y)$ extend to analytic functions of $x,y,z\in{{{\Delta}({\beta},{\delta})}}$, and, for all $x,y,z\in{{{\Delta}({\beta},{\delta})}}$, $$\begin{aligned} |G(z,x)| &\le {\stepcounter{CC}{C_{\arabic{CC}}}}|1-x|{^{-1}}, \label{erika} \\ |H(z,x,y)| &\le {\stepcounter{CC}{C_{\arabic{CC}}}}|1-z|{^{-1/2}}|1-x|{^{-1}}|1-y|{^{-1}}. \label{magnus}\end{aligned}$$ Standard singularity analysis [@FS Lemma IX.2] applied to yields $$\begin{aligned} |{\operatorname{\mathbb P{}}}(|{{\mathcal T}}|=n){h_n}(x,y)| & \le {\stepcounter{CC}{C_{\arabic{CC}}}}n {^{-1/2}}|1-x|{^{-1}}|1-y|{^{-1}}, \qquad x,y\in{{{\Delta}({\beta},{\delta})}},\end{aligned}$$ which proves [Theorem \[Tgen2\]]{} because, as is well known, a singularity analysis of yields $$\begin{aligned} {\operatorname{\mathbb P{}}}(|{{\mathcal T}}|=n)\sim {\stepcounter{cc}{c_{\arabic{cc}}}}n^{-3/2}.\end{aligned}$$ It thus remains only to prove [Lemma \[Lx\]]{}. Since ${\operatorname{\mathbb E{}}}\xi^2<\infty$, $\Phi'$ and $\Phi''$ extend to continuous functions on the closed unit disc with $\Phi'(1)={\operatorname{\mathbb E{}}}\xi=1$ and $\Phi''(1)={\operatorname{\mathbb E{}}}\xi(\xi-1)={\sigma^2}$. Hence, yields, for $z\in{{{\Delta}({\beta},{\delta})}}$, $$ \begin{split} \Phi'(F(z))&=\Phi'(1)+\Phi''(1){\bigl(F(z)-1\bigr)}+o(|F(z)-1|) \\& =1-\sqrt2{\sigma}\sqrt{1-z}+{o{\bigl(|z-1|^{1/2}\bigr)}}\end{split}$$ and $$\label{david} z\Phi'(F(z)) =\Phi'(F(z))+O(|z-1|) =1-\sqrt2{\sigma}\sqrt{1-z}+{o{\bigl(|z-1|^{1/2}\bigr)}}.$$ Let ${B(1,{\varepsilon})}{:=}{\ensuremath{\{z:|z-1|<{\varepsilon}\}}}$, and take ${\beta}<\pi/4$. Since $z\in{{\overline{{{{\Delta}({\beta},{\delta})}}}}\setminus{\ensuremath{\{1\}}}}$ entails $|\arg(1-z)|\le\pi/2+{\beta}$ and thus $|\arg\sqrt{1-z}|\le\pi/4+{\beta}/2$, it follows from that, for some small ${\varepsilon}>0$, if $z\in\overline{{{{{\Delta}({\beta},{\delta})}}\cap{B(1,{\varepsilon})}}}$ with $z\neq1$, then holds, $\bigl|{z\Phi'(F(z))-1}\bigr|=O({\varepsilon}{^{1/2}})$, $$\label{emma} \begin{split} \bigl|\arg{\bigl(z\Phi'(F(z))-1\bigr)}\bigr| & >|\arg(-\sqrt{1-z})|-{\beta}/2 \\& \ge \pi-(\pi/4+{\beta}/2)-{\beta}/2 =3\pi/4-{\beta}, \end{split}$$ and consequently, since $3\pi/4-{\beta}>\pi/2$, if ${\varepsilon}$ is small enough, $$\label{samuel} |z\Phi'(F(z))|<1.$$ Similarly, if $x\in{{{\Delta}({\beta},{\delta})}}$, then $|\arg(1-x)|<\pi/2+{\beta}$ and $$x{^{-1}}={\bigl(1-(1-x)\bigr)}{^{-1}}= 1+(1-x)+o(|1-x|), \qquad x\to1,$$ so if ${\varepsilon}>0$ is small enough, then, for $x\in{{{\Delta}({\beta},{\delta})}}\cap{B(1,{\varepsilon})}$, $$\label{jesper} |\arg(x{^{-1}}-1)|<\pi/2+2{\beta}.$$ If we choose ${\beta}\le\pi/16$, it follows from and that the triangle with vertices in $1$, $x{^{-1}}$ and $z\Phi'(F(z))$ has an angle at least $\pi/4-3{\beta}\ge\pi/16$ at 1, and thus by elementary trigonometry (the sine theorem), $$|x{^{-1}}- z\Phi'(F(z))| \ge {\stepcounter{cc}{c_{\arabic{cc}}}}|x{^{-1}}-1|,$$ and so holds, when $z,x\in{{{\Delta}({\beta},{\delta})}}\cap{B(1,{\varepsilon})}$, provided ${\beta},{\delta},{\varepsilon}$ are small enough. It remains to treat the case when $x$ or $z$ does not belong to ${B(1,{\varepsilon})}$, [i.e.=1000]{}, $|x-1|\ge{\varepsilon}$ or $|z-1|\ge{\varepsilon}$. We do this by compactness arguments. First, let $$\begin{aligned} A&{:=}{\ensuremath{\{z\Phi'(F(z)):z\in{{{\Delta}({\beta},{\delta})}}\cap{B(1,{\varepsilon})}\}}} \\ {B_\rho}&{:=}{\ensuremath{\{x{^{-1}}:x\in\overline{{{{\Delta}({\beta},\rho)}}}\setminus{B(1,{\varepsilon})},\, |x|\ge1/2\}}}.\end{aligned}$$ Then $B{:=}\bigcap_{\rho>0}{B_\rho}\subset{\ensuremath{\{\zeta:|\zeta|\ge1\}}}\setminus{\ensuremath{\{1\}}}$, and it follows from that ${\overline{A}}\cap B=\emptyset$. Since ${\overline{A}}$ and all ${B_\rho}$ are compact, it follows that ${\overline{A}}\cap{B_\rho}=\emptyset$ for some $\rho>0$, and thus, if $z\in{{{\Delta}({\beta},{\delta})}}\cap{B(1,{\varepsilon})}$ and $x\in{{{\Delta}({\beta},\rho)}}\setminus{B(1,{\varepsilon})}$ with $|x|\ge1/2$, then $|x{^{-1}}-z\Phi'(F(z))|\ge {\stepcounter{cc}{c_{\arabic{cc}}}}$ for some ${c_{\arabic{cc}}}>0$, which implies for such $z$ and $x$. Moreover, if $z\in{{{{\Delta}({\beta},{\delta})}}\cap{B(1,{\varepsilon})}}$ and $|x|<1/2$, shows that $|1-xz\Phi'(F(z))|\ge 1-|x|\ge1/2$, so then holds if $\ccj\le1/3$. Finally, if $z\in{{{\Delta}({\beta},{\delta})}}$, then $|F(z)|<1$ [@SJ167 Lemma A.2] as stated above, and thus $|\Phi'(F(z))|<1$. If $0<{\beta}_1<{\beta}$ and $0<{\delta}_1<{\delta}$, then $\overline{{{{\Delta}({\beta}_1,{\delta}_1)}}}\subset{{{\Delta}({\beta},{\delta})}}\cup{\ensuremath{\{1\}}}$, and thus by compactness $${C_{\varepsilon}}{:=}\sup{\ensuremath{\bigl\{|\Phi'(F(z))|:z\in\overline{{{{\Delta}({\beta}_1,{\delta}_1)}}}\setminus{B(1,{\varepsilon})}\bigr\}}}<1.$$ Consequently, if ${\delta}_2\le{\delta}_1$ is small enough and $x,z\in{{{\Delta}({\beta}_1,{\delta}_2)}}$ with $|z-1|\ge{\varepsilon}$, then $$|xz\Phi'(F(z))|\le(1+{\delta}_2)^2{C_{\varepsilon}}<1.$$ Hence holds in this case too for some $\ccj>0$, and similarly holds for $z\in{{{\Delta}({\beta}_1,{\delta}_2)}}\setminus{B(1,{\varepsilon})}$. This completes the proof of and , for some new ${\beta},{\delta}>0$ ([viz.=1000]{}, ${\beta}_1$ and $\min({\delta}_2,\rho)$). $G(z,x)$ now can be defined for all $x,z\in{{{\Delta}({\beta},{\delta})}}$ by , and holds by . Similarly, $H(z,x,y)$ can be defined for all $x,y,z\in{{{\Delta}({\beta},{\delta})}}$ by , and holds by , , and the fact that $\Phi'$ and $\Phi''$ are bounded on the unit disc. (Recall that $|F(z)|<1$ for $z\in{{{\Delta}({\beta},{\delta})}}$.) This completes the proof of [Lemma \[Lx\]]{}, and thus of [Theorem \[Tgen2\]]{} and of all results in this paper. \#1\#2,[\#2, no. \#1,]{} \#1 [99]{} D. Aldous, The continuum random tree II: an overview. *Stochastic Analysis (Proc., Durham, 1990)*, 23–70, London Math. Soc. Lecture Note Ser. 167, Cambridge Univ. Press, Cambridge, 1991. D. Aldous, Asymptotic fringe distributions for general families of random trees. [*Ann. Appl. Probab.* ]{}1, no. 2, 228–266 (1991). D. Aldous, Tree-based models for random distribution of mass. [J. Statist. Phys.]{} [73]{}, 625–641 (1993). M. Bousquet-Mélou, Limit laws for embedded trees. Applications to the integrated superBrownian excursion. [*Random Struct. Alg.* ]{}[29]{}, no. 4, 475–523 (2006). M. Bousquet-Mélou & S. Janson, The density of the ISE and local limit laws for embedded trees. [*Ann. Appl. Probab.* ]{}[16]{}, no. 3, 1597–1632 (2006). L. Devroye, Branching processes and their applications in the analysis of tree structures and tree algorithms. [Probabilistic Methods for Algorithmic Discrete Mathematics]{}, 249–314, eds. M. Habib et al., Algorithms Combin. 16, [Springer-Verlag]{}, Berlin, 1998. M. Drmota & B. Gittenberger, The width of Galton–Watson trees conditioned by the size. [Discrete Mathematics and Theoretical Computer Science]{}, 6, 387–400 (2004). M. Dwass, The total progeny in a branching process and a related random walk. [*J. Appl. Probab.* ]{}[6]{} (1969), 682–686. P. Flajolet & R. Sedgewick, [Analytic Combinatorics]{}. Cambridge Univ. Press, 2008, to appear. Web edition (available from the authors’ web sites). S. Janson, Random cutting and records in deterministic and random trees. [*Random Struct. Alg.* ]{}[29]{}, no. 2, 139–179 (2006). S. Janson, Conditioned Galton–Watson trees do not grow. [Proceedings, Fourth Colloquium on Mathematics and Computer Science Algorithms, Trees, Combinatorics and Probabilities (Nancy, 2006)]{}, DMTCS Proceedings AG, 331–334 (2006). V.F. Kolchin, [Random Mappings]{}. Optimization Software, New York, 1986. A. Meir & J.W. Moon, On the altitude of nodes in random trees. *Canad. J. Math.* [30]{}, 997–1015 (1978). J. Pitman, Enumerations of trees and forests related to branching processes and random walks. *Microsurveys in Discrete Probability (Princeton, NJ, 1997)*, 163–180, DIMACS Ser. Discrete Math. Theoret. Comput. Sci., 41, [Amer. Math. Soc.]{}, Providence, RI, 1998. [^1]: *CR Categories.* 3.74, 5.25, 5.5. [^2]: The first author’s research was sponsored by NSERC Grant A3456 and FQRNT Grant 90-ER-0291
{ "pile_set_name": "ArXiv" }
--- address: | Observatoire de Strasbourg, 11, rue de l’Université, 67000 Strasbourg, France\ $^t$ present address: Astronomy Centre, University of Sussex, Falmer, Brighton BN2 1TN, UK author: - 'J.G. Bartlett, A. Blanchard, M. Le Dour, M. Douspis & D. Barbosa$^t$' title: Constraints on Cosmological Parameters from Existing CMB Data --- Introduction ============ It is already possible, with existing data on the fluctuations of the cosmic microwave background (CMB), to constrain some cosmological parameters (Rocha & Hancock 1996; Bond & Jaffe 1996; Lineweaver et al. 1997). Of course, the actual data is far from providing the kind of constraints eagerly awaited from the next generation of experiments (Jungman et al. 1996; Bond et al. 1997), but there are nevertheless more than 10 different experiments detecting fluctuations over a large range of angular scales. To place our present work in context, consider the general problem of analyzing the results (say, a map) of a CMB experiment. If the true sky fluctuations and, as is fortunately often the case, the instrumental noise are gaussian, then the probability of obtaining any particular set of $N$ pixel values is a multivariate gaussian in the pixel vector $\vec{d}$ (a one–dimensional listing of the possible pixel values): $${\cal L}(\vec{p}) = {\cal P}(\vec{d}/\vec{p}) = \frac{1}{(2\pi)^{N/2} |C|^{1/2}} e^{-(\vec{d})^t \cdot C^{-1} \cdot (\vec{d})/2}$$ The cosmological model is completely embodied (as well as the noise) in the covariance matrix of the pixels, $C \equiv C^T(\vec{p}) + C(noise)$, where we denote the underlying model parameters by the vector $\vec{p}$. Thus, we have a likelihood function for $\vec{p}$, given a particular $\vec{d}$, i.e., a map of the sky. It is important to note that even if the pixels are gaussian, the likelihood function ${\cal L}(\vec{p})$ is NOT so, in general, because the parameters enter into the [*covariance*]{} matrix and not as linear combinations of the pixel values themselves. Maximization of this likelihood function is one way to estimate cosmological parameters. Another approach is to work in Fourier space with estimates of the power spectrum, i.e., the $C_l$. These are nothing more than the variances of the individual spherical harmonic coefficients of the temperature fluctuations. We are really doing the same thing as before, just working with the covariances of Fourier coefficients instead of the measured pixel values. A practical reason for using this method is that band–power estimates and their uncertainties are readily available in the literature, this being the most common way of concisely reporting the results of an experiment. In the present contribution, we assemble various band–power estimates (e.g., $C_l$ intergrated over experimental window functions) and constrain certain cosmological parameters by fitting to these power spectrum data points. We must mention several caveats applying to the results given here: Firstly, the data set is certainly not homogeneous, if for no other reason than for the different ways that the so–called cosmic variance has been estimated and included in the quoted error bars. Secondly, we have not yet fully accounted for the experimental window functions. A third important point is that the $C_l$ are NOT gaussian random variables, unlike the individual pixel values (or Fourier coefficients). This is clear – the $C_l$ represent the variance of gaussian random variables, and thus are themselves distributed according to a $\chi^2$ distribution. We have, for the work presented in this contribution, effectively assumed gaussianity in the sense that we apply a $\chi^2$ minimization for the fitting. In principle, this could lead to biased best–estimates and incorrect confidence intervals. We are currently working on improving these aspects of our calculations. The results presented here should therefore be taken as indicative, but perhaps relatively good indications, all the same, given the current status of affairs (Jaffe & Bond 1997).\ [|c||c|c||c|c|]{}\ parameter & range & step & range & step\ $\Omega$ & 0 – 0.95 & 0.05 & 0 – 1 & 0.05\ $H_o$ & 15 – 100 km/s/Mpc & 5 km/s/Mpc & 15 – 100 km/s/Mpc & 5 km/s/Mpc\ $n$ & 0.5 – 1.5 & 0.03 & 0.55 – 1.45 & 0.09\ $Q$ & 14 – 20 $\mu$K & 1 $\mu$K & 14 – 20 $\mu$K & 1 $\mu$K \ $\Omega_B h^2$ & 0.015 & —- & 0.006 – 0.030 & 0.002\ The models ========== We consider two broad class of inflation–based models (e.g., gaussian fluctuations): Open and flat with a non–zero cosmological constant. For the [**open**]{} models, we explored the four–dimensional parameter space of [$\Omega,H_o,n,Q$]{}, where $n$ is the spectral index of an assumed power–law power spectrum and $Q$ is the normalization expressed as the CMB quadrupole; the baryon density was fixed at its nucleosythesis predicted value of $\Omega_b h^2 = 0.015$. For the [**flat**]{} models, we varied the baryon density in addition to these four parameters, thereby exploring a five–dimensional space. This is all summarized in the Table, where we also provide the individual step sizes on each parameter and the range covered. The calculations were performed with CMBFAST (Seljak & Zaldarriaga 1996). To cover the parameter space indicated, the slower open model calculations required a couple of months of uninterrupted computing time. Results ======= Our results in the $(\Omega,H_o)$–plane for both types of model are shown in Figure 1. The light solid contours are defined by $\chi^2_{\rm min} + 2.3$ and $+6.17$, which would correspond to $1\sigma$ and $2\sigma$ limits in this plane for a gaussian likelihood function (of the [*parameters*]{}), which, as we have mentioned, is not really the case. Nevertheless, this gives some idea of the region favored by the data: in general, we find that it corresponds to high $\Omega$ and low values of $H_o$. The dashed lines show the contours whose projection onto the axes give the confidence range on each parameter individually. Finally, for comparison, the dark solid lines (outer contours) are contours of constant [*goodness–of–fit*]{} for, if all were normal, 68% and 95% probability. The most interesting aspect of these results concerns the open models. We see from Figure 1 that the contours place a lower limit on $\Omega$; values as low as $\Omega=0.2$ are strongly disfavored for any $H_o$. The reason for this is clear from the left–hand–side of Figure 2. The “Doppler Peak” is displaced too far to the right, due to the focusing of light rays in an open geometry (Hu & White 1996). Thus we find that a “comfortable” model with $\Omega=0.2$ and $H_o=60$ km/s/Mpc has [*very serious difficulty*]{} with the CMB data. How about the cosmological constant? Here, the constraints on $\Omega$ are much less stringent, as can be seen from the contours shown in the right–hand panel of Figure 1. In contrast to the purely open model with $\Omega=0.2$, the corresponding flat model, with $\Lambda=0.8$, is acceptable and provides a good–looking fit to the power spectrum (right–hand–side of Figure 2). Conclusions =========== It is a little early to draw detailed conclusions from the CMB concerning the cosmological parameters, but it is worth noting that some conclusions are already possible with existing data; and the next round of data releases will not be long in coming. Despite the various caveats of the method used here, it seems that one result should remain rather robust - that low–$\Omega$, purely open models are strongly disfavored, for the clear and simple reason that the “Doppler Peak” is just not in the correct place due to the large angular distance to the surface of last scattering in such models. Acknowledgements ================ D.B acknowledges support by a Ph.D. Praxix XXI grant attributed by JNICT/FDC, Portugal. References {#references .unnumbered} ========== [99]{} Bond J.R., Efstathiou G. & Tegmark M. 1997, MNRAS 291, L33 Bond J.R. & Jaffe A.H. 1996, in “Microwave Background Anisotropies” (proceedings of the XVIth Moriond Astrophysics meeting), F.R. Bouchet, R. Gispert et al. (Eds), Editions Frontières (Gif–sur–Yvette) p. 197 Hu W. & White M. 1996, ApJ 471, 30 Jaffe A.H. & Bond J.R. 1997, proceedings of the 18th Texas Symposium, astro-ph/9702109 Jungman G., Kamionkowski M., Kosowsky A. & Spergel D. 1996, PhyRev D54, 1332 Lineweaver C.H., Barbosa D., Blanchard A. & Bartlett J.G. 1997, A&A 322, 365 Rocha G. & Hancock S. 1996, in “Microwave Background Anisotropies” (proceedings of the XVIth Moriond Astrophysics meeting), F.R. Bouchet, R. Gispert et al. (Eds), Editions Frontières (Gif–sur–Yvette) p. 189 Seljak U. & Zaldarriaga M. 1996, ApJ 469, 437
{ "pile_set_name": "ArXiv" }
--- abstract: 'The study of inelastic scattering and multi-nucleon transfer reactions was performed by bombarding a $^{9}$Be target with a $^3$He beam at an incident energy of 30 MeV. Angular distributions for $^9$Be($^3$He,$^3$He)$^{9}$Be, $^9$Be($^3$He,$^4$He)$^{8}$Be, $^9$Be($^3$He,$^5$He)$^{7}$Be, $^9$Be($^3$He,$^6$Li)$^6$Li and $^9$Be($^3$He,$^5$Li)$^7$Li reaction channels were measured. Experimental angular distributions for the corresponding ground states (g.s.) were analysed within the framework of the optical model, the coupled-channel approach and the distorted-wave Born approximation. Cross sections for channels leading to unbound $^5$He$_{g.s.}$, $^5$Li$_{g.s.}$ and $^8$Be systems were obtained from singles measurements where the relationship between the energy and the scattering angle of the observed stable ejectile is constrained by two-body kinematics. Information on the cluster structure of $^{9}$Be was obtained from the transfer channels. It was concluded that cluster transfer is an important mechanism in the investigated nuclear reactions. In the present work an attempt was made to estimate the relative strengths of the interesting $^8$Be+$n$ and $^5$He+$\alpha$ cluster configurations in $^9$Be. The branching ratios have been determined confirming that the $^5$He+$\alpha$ configuration plays an important role. The configuration of $^9$Be consisting of two bound helium clusters $^3$He+$^6$He is significantly suppressed, whereas the two-body configurations ${}^{8}$Be+$n$ and ${}^{5}$He+$\alpha$ including unbound $^8$Be and $^5$He are found more probable.' address: - '$^1$ Flerov Laboratory of Nuclear Reactions, Dubna, Russian Federation' - '$^2$ KVI-CART, University of Groningen, Groningen, Netherlands' - '$^3$ Nuclear Physics Institute, Řež, Czech Republic' - '$^4$ Department of Physics, University of Jyväskylä, Jyväskylä, Finland' - '$^5$ Khlopin Institute, St. Petersburg, Russian Federation' - '$^6$ National Research Nuclear University “MEPhI”, Moscow, Russian Federation' - '$^7$ Eurasian Gumilev University, Astana, Kazakhstan' - '$^8$ Nuclear Physics Institute, Almaty, Kazakhstan' author: - 'S M Lukyanov$^1$, M N Harakeh$^2$, M A Naumenko$^1$, Yi Xu$^3$, W H Trzaska$^4$, V Burjan$^3$, V Kroha$^3$, J Mrazek$^3$, V Glagolev$^3$, Š Piskoř$^3$, E I Voskoboynik$^1$, S V Khlebnikov$^{5}$, Yu E Penionzhkevich$^{1,6}$, N K Skobelev$^1$, Yu G Sobolev$^1$, G P Tyurin$^3$, K Kuterbekov$^{7}$ and Yu Tuleushev$^8$' title: Evidence of Cluster Structure of $^9$Be from $^3$He+$^9$Be Reaction --- Introduction ============ In recent years the study of light radioactive nuclei [@Ref1; @Ref2] has intensified due to the significant progress made with radioactive beam facilities. It has led to a decrease of interest in the study of light stable nuclei such as ${}^{6,7}$Li and ${}^{9}$Be. It has been shown that in light nuclei the nucleons tend to group into clusters, the relative motion of which defines to a large extent the properties of these nuclei. Consequently, the cluster structure of their ground as well as low-lying excited states became the focus of theoretical and experimental studies. For example, ${}^{6}$Li and ${}^{7}$Li nuclei are both well described by the two-body cluster models ($\alpha+d$ and $\alpha+t$, respectively). Due to its Borromean structure, a special attention has been focused on the ${}^{9}$Be nucleus which may be considered as a nuclear system with the following two-body configurations: ${}^{8}$Be+$n$ or ${}^{5}$He+$\alpha$. The breakup of $^9$Be via $^8$Be$_{g.s.}$ has been measured for many of the low-lying excited states of $^9$Be [@Brown; @Papka]. However, the breakup branching via the first-excited 2$^+$ state of $^8$Be and via $^5$He remained uncertain. The three-body configuration ($\alpha\alpha n$) of $^9$Be may have a significant astrophysical interest. The short lifetimes of $^8$Be$_{g.s.}$, the first-excited state of $^8$Be and $^5$He$_{g.s.}$ suggest that the sequential capture of a neutron or an $\alpha$ particle is very unlikely. The formalism used to derive the ($\alpha\alpha n$) rate by Grigorenko [@Grigorenko] also suggested that any broad intermediate resonances will have little effect on the ($\alpha\alpha n$) rate. However, according to another theoretical calculation [@Buchmann] the stellar reaction rate for this reaction proceeding via the $^5$He$_{g.s.}$ channel is significant for the formation of $^9$Be. In the measurement [@Soic] of the $^9$Be($^7$Li,$^7$Li$\alpha$)$\alpha- n$ reaction it was unambiguously established for the first time that the $^9$Be excited states around 6.5 and 11.3 MeV decay into the $\alpha+^5$He channel. In Ref. [@Charity], the structure of $^9$Be was discussed in the frame of isobaric analogue states of $^9$B. Charity et al. [@Charity] found that the decay of ${}^9$B has the dominant branch $\alpha$+${}^5$Li implying that “the corresponding mirror state in ${}^{9}$Be would be expected to decay through the mirror channel $\alpha$+${}^{5}$He, instead of through the $n$+${}^{8}$Be(2$^+$) channel The mirror state of $^9$Be at 2.429 MeV is also reported to decay by $\alpha$ emission". In this case it would decay to the unstable ground state of $^5$He producing the $n$+2$\alpha$ final state. On the contrary, another experimental work [@Papka] claimed that this state decays almost exclusively by $n$ emission to the unstable first-excited state of $^8$Be. The excited states of $^9$Be have been populated in various ways including among others $\beta$-decay [@Nyman]. Most experiments confirm that the 2.429 MeV state has a branching ratio to the $^8$Be$_{g.s.}$+$n$ channel of only $ \sim$7% [@Brown], but could not determine whether the remaining strength was in the $^8$Be($2^+$)+$n$ or the $^5$He+$\alpha$ channel. However, it was reported in Ref. [@Nyman] that the ratio 2:1 could be assigned for the two channels, respectively. The results of Refs. [@Brown; @Papka; @Grigorenko; @Soic; @Charity] along with the qualitative breakup data discussed above suggest the necessity of obtaining quantitative branching-ratio data for the low-lying states in the Borromean nucleus $^9$Be. The present experiment was designed to study the breakup of $^9$Be in the attempt to determine the contribution of $^5$He+$\alpha$ and $n$+$^8$Be channels. Our data are based on inclusive measurements, whereas the experimental results of Refs. [@Brown; @Papka] were obtained in exclusive measurements. The exclusive measurements require fully defined kinematics with complicated detector systems allowing the complete reconstruction and identification of the breakup event. The inclusive measurements, in spite of their simplicity, may also be useful in the study of clustering phenomena. For example, in our recent paper [@luk] the obtained large value of the deformation parameter might be considered as the confirmation of the cluster structure of the low-lying states of $^9$Be. However, the inclusive measurement did not allow us to give unambiguous preference to one of the possible configurations, e.g. ($\alpha$+$\alpha$+$n$) or ($\alpha$+${}^{5}$He). Another purpose of the present study was the attempt to find the cluster structure (for instance, $^5$He) and how the cluster structure influences the nuclear reaction mechanism. Indeed, Détraz et al. [@Detraz; @Detraz2] argued that multi-particle-multi-hole structures are expected to occur at rather low excitation energies in nuclei. Therefore, four-nucleon transfer reactions have been extensively studied. It is hoped that the major features of such reactions, in spite of the a priori complexity of such a transfer, may be understood assuming that the nucleons are transferred as a whole, i.e. strongly correlated in the cluster which has the internal quantum numbers of a free $\alpha$-particle. Whether one or several internal configurations of the transferred fragments contribute to the transition amplitudes of these transfer reactions is both a crucial and difficult question to address. In the present study, we tried to examine how, for instance, two protons and two neutrons merged into an $\alpha$-cluster are transferred. The angular distributions of the $^9$Be($^3$He,$^7$Be)$^5$He and $^9$Be($^3$He,$^6$Li)$^6$Li reaction channels were scrupulously measured at the energy of 60 MeV for the transitions to the ground states of $^5$He and $^6$Li by Rudchik et al. [@Rudchik]. The experimental data were analysed using the coupled reaction channels (CRC) model including one- and two-step cluster transfer and cluster spectroscopic amplitudes calculated in the framework of the translation-invariant shell model. It was concluded that the cluster transfer is unimportant: in the $^9$Be($^3$He,$^7$Be)$^5$He reaction channel the $\alpha$-transfer dominates only at small angles while the transfer of two neutrons dominates at large angles. Nevertheless, the conclusion of Ref. [@Rudchik] in general supports the idea of cluster transfer. The second motivation for studying again the $^3$He+$^9$Be reaction was the attempt to find evidence for the cluster structure of these nuclear systems by the analysis of the reaction products. Experimental procedure ====================== The $^3$He+$^9$Be experiments were performed at the K130 Cyclotron facility of the Accelerator Laboratory of the Physics Department of Jyväskylä University and at the Nuclear Physics Institute (NPI), Řež, Czech Republic. The $^3$He beam energy was 30 MeV. The average beam current during the experiments was maintained at 10 nA. The self-supporting Be target was prepared from a 99% pure thin foil of beryllium. The target thickness was 12 $\mu$m. Peaks due to carbon and oxygen contaminations were not observed in the energy spectra. ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------ ![ Particle identification plots for the products of the $^{3}$He+$^9$Be reaction: a) $p$, $d$, $t$ and $^{3,4}$He; b) $^{6}$He, $^{6,7}$Li and $^{7,9}$Be. $\Delta$E is the energy loss and E$_r$ is the residual energy. The loci for the products are indicated. []{data-label="fig1"}](pido_aa.eps "fig:"){width="90mm"} ![ Particle identification plots for the products of the $^{3}$He+$^9$Be reaction: a) $p$, $d$, $t$ and $^{3,4}$He; b) $^{6}$He, $^{6,7}$Li and $^{7,9}$Be. $\Delta$E is the energy loss and E$_r$ is the residual energy. The loci for the products are indicated. []{data-label="fig1"}](pido_bb.eps "fig:"){width="85mm"} ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------ \[figure\] To measure (in)elastically scattered ions four Si-Si(Li) telescopes each consisting of $\Delta$E$_0$, $\Delta$E and E$_r$ detectors with thicknesses 10 $\mu$m, 100 $\mu$m and 3 mm, respectively, were used. Each telescope was mounted at a distance of $\sim$45 cm from the target. Particle identification was based on the measurements of energy-loss $\Delta$E and residual energy E$_r$ (the so-called $\Delta$E-E method). The telescopes were mounted on rotating supports, which allowed to obtain data from $\theta_{lab}$ = 20$^\circ$ to $\theta_{lab}$ = 107$^\circ$ in steps of 1-2$^\circ$. The overall energy resolution of the telescopes was $\sigma \sim$200 keV. The two-dimensional plots energy loss $\Delta$E vs. residual energy E$_r$ and $\Delta$E$_0$ vs. $\Delta$E are shown in Fig. \[fig1\]a and Fig. \[fig1\]b, respectively. The excellent energy resolutions of both $\Delta$E and E$_r$ detectors allowed unambiguous identification of $A$ and $Z$ of each product. The panel (a) is mainly related to the transfer from the projectile to the target, whereas the panel (b) is completely related to the transfer from the target to the projectile. ![ Spectra of the total deposited energy for the detected $^6$Li (a), $^7$Li (b), $^4$He (c) and $^7$Be (d) measured at 18$^\circ$ in the laboratory system for the reaction $^3$He(30 MeV)+$^9$Be. []{data-label="fig2"}](Fig_2_double2.eps){width="160mm"} Fig. 2 shows experimental spectra of the total deposited energy for the detected $^6$Li, $^7$Li, $^4$He and $^7$Be measured at 18$^\circ$ in the laboratory system for the reaction $^3$He(30 MeV)+$^9$Be. The plotted total energies were calculated as the sum of the calibrated energy losses $\Delta$Es and the residual energy E$_r$. The ground and the most populated excited states of $^6$Li and $^8$Be as well as the ground states populated in the reaction channels $^9$Be($^3$He,$^5$Li)$^7$Li and $^9$Be($^3$He,$^5$He)$^7$Be were unambiguously identified. The population of the excited states of $^5$He was studied in the reaction channel $^9$Be($^3$He,$^5$He)$^7$Be together with the population of the excited states of $^6$Li in the reaction channel $^9$Be($^3$He,$^6$Li)$^6$Li as a reference case. For the case of $^6$Li one may see the known levels of this nucleus, and the peak width has value $\sigma$=200 keV equal to the detector resolution. Since $^5$He, $^5$Li and $^8$Be are unbound nuclei and decay into $^4$He+$n$, $^4$He+$p$ and $^4$He+$^4$He, respectively, we calculated the positions in the measured spectra (Fig. 2) of the maximum energies of $^7$Li, $^4$He and $^7$Be assuming either two bodies or three bodies in the exit channels. Calculations were performed with the aid of the NRV server [@NRV]. The solid arrows show the maximum values of the kinetic energies of the detected $^6$Li, $^7$Li, $^4$He and $^7$Be which correspond in the case of $^7$Li, $^4$He and $^7$Be to the two-body character of the reaction in the exit channels: $^9$Be($^3$He,$^5$Li)$^7$Li, $^9$Be($^3$He,$^8$Be)$^4$He and $^9$Be($^3$He,$^5$He)$^7$Be, respectively. The maximum values of the kinetic energies in the case of three-body final states in the reaction channels $^9$Be($^3$He,$^4$He+$p$)$^7$Li, $^9$Be($^3$He,$^4$He+$^4$He)$^4$He and $^9$Be($^3$He,$^4$He+$n$)$^7$Be are shown by the broken arrows. As can be seen from Fig. 2, the local maxima at the g.s. are close to the values corresponding to the two-body exit channels in all studied reaction channels. This indicates that the two-body reaction channels dominate leading to the unbound $^5$He, $^5$Li and $^8$Be in their ground states. Therefore, it may certainly be concluded that one-step multi-nucleon transfer is the main reaction mechanism leading to the binary exit channels. It is expected that the experimental information provided at the end of the de-excitation process shows the strongest fingerprints of the last stages, while the influence of earlier stages and the situation just after the first reaction step are strongly masked. We can even assume that in multi-step reactions there are no fingerprints of the intermediate products in the energy distribution. All multi-nucleon transfer processes will be described as a cluster transfer. The nucleons participating in the transfer are transferred tightly bound together, in one step. The question as to whether the transfer of so many nucleons is adequately described by the cluster model of the reaction is left open. In the spectrum, the broad ground state of $^5$He is overlapping with the ground and first-excited states of $^7$Be, which explains why the peak is rather broad (FWHM is $\sim-$1.5 MeV). The real peak was not observed at large angles and no states other than those discussed above were observed. In the case of $^5$Li (Fig. 2b) the ground state in not really prominent showing a larger value of FWHM than in the case of $^5$He. It might indicate that $^5$Li is a more loosely bound nucleus than $^5$He due to the Coulomb repulsion of the additional proton from the $\alpha$ core. Detection of unbound nuclei such as $^5$He and $^5$Li provides the opportunity to estimate the time for energy equilibration which should be less than the lifetime of the shortest-lived nuclei, $i.e.$ $^5$Li or $^5$He with the lifetime of about a few units of 10$^{-22}$ s. --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- ![ Angular distributions for a) $^3$He + $^9$Be $\rightarrow$ $^5$He + $^7$Be ($\blacklozenge$), $^3$He + $^9$Be $\rightarrow$ $^5$Li + $^7$Li$_{g.s.}$ ($\Box$), $^3$He + $^9$Be $\rightarrow$ $^6$Li + $^6$Li$_{g.s.}$ ($\blacktriangle$), $^3$He + $^9$Be $\rightarrow$ $^4$He + $^8$Be$_{g.s.}$ ($\triangle$), $^3$He + $^9$Be $\rightarrow$ $^6$He + $^6$Be$_{g.s.}$ ($\bullet$); b) $^3$He + $^9$Be $\rightarrow$ $^4$He + $^8$Be$_{g.s.}$ ($\triangle$), 3.0 MeV ($\blacksquare$) and 16.6 MeV ($\ast$) excited states. []{data-label="fig3"}](Figure3_a.eps "fig:"){width="140mm"} ![ Angular distributions for a) $^3$He + $^9$Be $\rightarrow$ $^5$He + $^7$Be ($\blacklozenge$), $^3$He + $^9$Be $\rightarrow$ $^5$Li + $^7$Li$_{g.s.}$ ($\Box$), $^3$He + $^9$Be $\rightarrow$ $^6$Li + $^6$Li$_{g.s.}$ ($\blacktriangle$), $^3$He + $^9$Be $\rightarrow$ $^4$He + $^8$Be$_{g.s.}$ ($\triangle$), $^3$He + $^9$Be $\rightarrow$ $^6$He + $^6$Be$_{g.s.}$ ($\bullet$); b) $^3$He + $^9$Be $\rightarrow$ $^4$He + $^8$Be$_{g.s.}$ ($\triangle$), 3.0 MeV ($\blacksquare$) and 16.6 MeV ($\ast$) excited states. []{data-label="fig3"}](Figure3_bb.eps "fig:"){width="148mm"} --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- \[figure\] Results ======= Elastic and inelastic scattering -------------------------------- The measured angular distributions for the reaction channels $^3$He + $^9$Be $\rightarrow$ $^5$He + $^7$Be ($\blacklozenge$), $^3$He + $^9$Be $\rightarrow$ $^5$Li + $^7$Li$_{g.s.}$ ($\Box$), $^3$He + $^9$Be $\rightarrow$ $^6$Li + $^6$Li$_{g.s.}$ ($\blacktriangle$), $^3$He + $^9$Be $\rightarrow$ $^4$He + $^8$Be$_{g.s.}$ ($\triangle$), $^3$He + $^9$Be $\rightarrow$ $^6$He + $^6$Be$_{g.s.}$ ($\bullet$) are shown in Fig. 3a. First of all, one may notice that the cross section for the case of $^5$He+$^7$Be ($\blacklozenge$) is much larger than for the other reaction channels shown in Fig. 3. The $^7$Be energy spectrum (Fig. 2d) indicates the large probability of transitions to the ground state of the $^5$He nucleus in conjunction with the 3$^-$ ground state and the $1/2^-$ first-excited (0.429 MeV) state of $^7$Be (unresolved in the present experiment). Therefore, the angular distribution for the $^5$He+$^7$Be ($\blacklozenge$) channel corresponds to the ground state and the first-excited state of $^7$Be. As it was mentioned, this exit channel of the reaction has the largest cross section among the channels shown in Fig. 3a. The calculated Q-values of the investigated channels are given in the inset of Fig. 3a. It may be seen that the behaviour of the cross sections does not strictly follow the famous Q-reaction systematics. For instance, the measured differential cross section for the channel $^4$He+$^8$Be$_{g.s.}$ with the maximum value of $Q=+19$ MeV is much less than for the channel $^5$He+$^7$Be ($Q=-0.7$ MeV). The lowest cross section was observed for $^6$He+$^6$Be$_{g.s.}$ ($Q=-9.7$ MeV). Fig. 3b shows the angular distributions of the differential cross sections for the $^3$He(30 MeV)+$^9$Be $\rightarrow$ $^4$He+$^8$Be reaction populating the ${g.s.}$, and the 3.0 MeV and around 16.6 MeV excited states. It seems that the high positive Q-value of the reaction (18.9 MeV) for this channel allows to excite very high-lying levels in the unbound $^8$Be nucleus. It is clearly visible from Fig. 3b that in the reaction $^3$He+$^9$Be $\rightarrow$ $^4$He+$^8$Be the cross section for 16.6 MeV level dominates over the other reaction channels. In order to describe the various reaction channels in the framework of the coupled-channel (CC) framework a special attention was paid to the elastic scattering to obtain the optical model (OM) potentials. The OM potential was chosen in the usual Woods-Saxon form $$V(r) = - {V_0}f(r,{R_v},{a_v}) - i{W_0}f(r,{R_w},{a_w}),$$ where the function $f(r,R,a) = {(1 + {e^{(r - R)/a}})^{ - 1}}$ and radii $R_i = r_i A^{1/3}$ depend on the mass of the heavier fragment $A$ and the corresponding reduced radius $r_i$. In addition, a spin-orbit potential with a Woods-Saxon form has been used. In the first step, the experimental data for $^9$Be+$^3$He elastic scattering were fitted. The obtained OM potential parameters are listed in Table 1 in comparison with the parameters of Ref. [@Rudchik]. The results of fitting are shown in Figs. 4 and 5 together with the experimental data. The difference between the obtained parameters and those of Ref. [@Rudchik] is probably due to the different projectile energies: our data were obtained at 30 MeV and the data of Rudchik et al. [@Rudchik] were obtained at 60 MeV. For instance, the depths of the real and imaginary parts are smaller at the lower incident energy, whereas the radii are larger than in Ref. [@Rudchik]. Concerning the $^9$Be($^3$He,$^5$He)$^7$Be reaction channel, the OM parameters of both sets do not reflect any possible peculiarities due to the unbound $^5$He$_{g.s.}$, which might otherwise be expected (for example, a larger radius of this unbound nucleus considered as the candidate for one-neutron halo structure). In the second step, of the data analysis, the cross section for the transfer channel $^9$Be($^3$He,$^4$He)$^8$Be$_{g.s.}$ was fitted and the OM potential parameters for the exit channel $^4$He+$^8$Be were obtained. In the fitting process, the OM potential parameters for the entrance channel $^9$Be+$^3$He obtained in the first step were used. ![The ratio of the measured elastic scattering cross section to the Rutherford cross section for $^3$He+$^9$Be at the incident energy 30 MeV in comparison with the OM fit.[]{data-label="fig4"}](Fig_4.eps){width="100mm"} [ccccccccccc]{} reaction & $V_0$ & $r_v$ & $a_v$ & $W_0$ & $r_w$ & $a_w$ & V$_{so}$ & $r_{so}$ & $a_{so}$\ channel & \[MeV\] & \[fm\] & \[fm\] & \[MeV\] & \[fm\] & \[fm\] & \[MeV\] & \[fm\] & \[fm\]\ $^3$He+$^9$Be & 108.5 & 1.123 & 0.54 & 15.69 & 1.15 & 0.855 & 13.63 & 1.83 & 0.4\ $^3$He+$^9$Be [@Rudchik]& 143.4 & 1.02 & 1.10 & 38.3 & 1.405 & 1.17 & & &\ $\alpha$+$^8$Be & 90.98 & 1.382 & 0.404 & 12.36 & 1.01 & 0.4& $-$ & $-$ & $-$\ $\alpha$+$^8$Be [@Rudchik]& 143.4 & 1.024 & 1.35 & 38.3 & 2.539 & 1.67 & & &\ $^5$He+$^7$Be & 60.88 & 1.25 & 0.65 & 2.8 & 1.25 & 0.65 & 8.93 & 1.28 & 0.65\ $^5$He+$^7$Be [@Rudchik]& 122.5 & 0.9768 & 0.61 & 51.71 & 0.914 & 1.178 & & &\ $^6$Li+$^6$Li & 110.95 & 1.307 & 0.621 & 2.48 & 1.25 & 0.65 & 1.05 & 1.25 & 0.65\ $^6$Li+$^6$Li [@Rudchik]& 122.5 & 0.979 & 0.905 & 51.71 & 0.914 & 1.178 & & &\ ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- ![ The angular distributions for a) $^3$He + $^9$Be $\rightarrow$ $^4$He + $^8$Be$_{g.s.}$ ($\triangle$), b) $^3$He + $^9$Be $\rightarrow$ $^5$He + $^7$Be ($\blacklozenge$) and c) $^3$He + $^9$Be $\rightarrow$ $^6$Li + $^6$Li$_{g.s.}$ ($\blacktriangle$). The curves are the results of the optical-model and the coupled-channel calculations (see text). []{data-label="fig5"}](Fig_5a.eps "fig:"){width="90mm"} ![ The angular distributions for a) $^3$He + $^9$Be $\rightarrow$ $^4$He + $^8$Be$_{g.s.}$ ($\triangle$), b) $^3$He + $^9$Be $\rightarrow$ $^5$He + $^7$Be ($\blacklozenge$) and c) $^3$He + $^9$Be $\rightarrow$ $^6$Li + $^6$Li$_{g.s.}$ ($\blacktriangle$). The curves are the results of the optical-model and the coupled-channel calculations (see text). []{data-label="fig5"}](Fig5_b.eps "fig:"){width="90mm"} ![ The angular distributions for a) $^3$He + $^9$Be $\rightarrow$ $^4$He + $^8$Be$_{g.s.}$ ($\triangle$), b) $^3$He + $^9$Be $\rightarrow$ $^5$He + $^7$Be ($\blacklozenge$) and c) $^3$He + $^9$Be $\rightarrow$ $^6$Li + $^6$Li$_{g.s.}$ ($\blacktriangle$). The curves are the results of the optical-model and the coupled-channel calculations (see text). []{data-label="fig5"}](Fig5_c.eps "fig:"){width="90mm"} ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- \[figure\] In the third step, the cross sections for the transfer reactions $^9$Be($^3$He,$^7$Be)$^5$He and $^9$Be($^3$He,$^5$He)$^7$Be were fitted simultaneously and the OM potential parameters for the exit channel $^5$He+$^7$Be were obtained (Table 1). In the fitting process, the OM potential parameters for the entrance channel $^9$Be+$^3$He obtained in the first step were used. Later in the calculations of the total cross section both transition amplitudes corresponding to the alpha-transfer mechanism (the $^9$Be($^3$He,$^7$Be)$^5$He channel) and the 2$n$-transfer mechanism (the $^9$Be($^3$He,$^5$He)$^7$Be channel) have been taken into account. All calculations and fitting have been performed using the FRESCO code [@Fresco] in the framework of the CC method. By fitting the experimental angular distributions, the attempt was made to determine which reaction mechanisms are the most important for the measured distributions. The results are shown in Fig. 5 for the following channels: a) $^3$He + $^9$Be $\rightarrow$ $^4$He + $^8$Be$_{g.s.}$, b) $^3$He + $^9$Be $\rightarrow$ $^5$He + $^7$Be and c) $^3$He + $^9$Be $\rightarrow$ $^6$Li + $^6$Li$_{g.s.}$. It is clear that in the case of $^3$He+$^9$Be $\rightarrow$ $^7$Be+$^5$He reaction channel the $\alpha$-transfer (long-dashed curve) dominates at forward angles, whereas the transfer of the two neutrons (short-dashed curve) dominates at backward angles. The fit for $^3$He+$^9$Be $\rightarrow$ $^6$Li+$^6$Li (solid curve) was obtained assuming $t$ transfer from the target nucleus to the projectile. A good agreement between the experimental data and the calculations is observed. A similar situation may be expected for the case of $^3$He+$^9$Be $\rightarrow$ $^5$Li+$^7$Li where $d$ transfer dominates. As could be seen from Fig. 5, the main processes describing the angular distributions of the exit channels $^4$He+$^8$Be$_{g.s.}$, $^5$He+$^7$Be and $^6$Li+$^6$Li$_{g.s.}$ are, respectively, the neutron, $\alpha$ and $t$ transfer from the target nucleus to the projectile. Cluster transfer in $^3$He+$^9$Be reaction ------------------------------------------ Due to the Borromean structure of $^9$Be the breakup of this nucleus may, in principle, proceed directly into two $\alpha$ particles plus a neutron or via one of the unstable intermediate nuclei: $^8$Be or $^5$He. The attempt was made to get some insights into $n$+$^8$Be, $^4$He+$^5$He, $d$+$^7$Li and $t$+$^6$Li cluster configurations inside of $^9$Be and especially to estimate the relative strengths of the most interesting $^8$Be+$n$ and $^5$He+$\alpha$ cluster configurations in $^9$Be. Most likely the studied reaction channels result from the cluster transfer: $^3$He + $^9$Be $\rightarrow$ $^3$He + ($n$ + $^8$Be) $\rightarrow$ ($^3$He + $n$)+ $^8$Be $\rightarrow$ $^4$He + $^8$Be $^3$He + $^9$Be $\rightarrow$ $^3$He + ($\alpha$ + $^5$He) $\rightarrow$ ($^3$He + $\alpha$) + $^5$He $\rightarrow$ $^7$Be + $^5$He $^3$He + $^9$Be $\rightarrow$ $^3$He + ($t$ + $^6$Li) $\rightarrow$ ($^3$He + $t$) + $^6$Li $\rightarrow$ $^6$Li + $^6$Li $^3$He + $^9$Be $\rightarrow$ $^3$He + ($d$ + $^7$Li) $\rightarrow$ ($^3$He + $d$) + $^7$Li $\rightarrow$ $^5$Li + $^7$Li $^3$He + $^9$Be $\rightarrow$ $^3$He + ($^3$He + $^6$He) $\rightarrow$ ($^3$He + $^3$He) + $^6$He $\rightarrow$ $^6$Be + $^6$He These reaction channels are of specific interest since they provide direct information on the cluster structure of the initial and final states. Based on the angular distributions shown in Fig. 3, one may estimate the probability of cluster configurations in $^9$Be. To obtain the branching ratios (BRs) for the observed channels, the angular distributions were integrated over the solid angle. The probability was calculated as weighted integrals among the channels listed above. The obtained values of the BRs are given in Table 2. [lll]{} Exit channel & $^9$Be cluster configuration & Branching ratio (%)\ $^4$He + $^8$Be & $n$ + $^8$Be & $\le$68.7 $\pm 10$\ $^5$He + $^7$Be & $\alpha$ + $^5$He & $\le$25.1 $\pm 5$\ $^6$Li + $^6$Li$_{g.s.}$ & $t$ + $^6$Li & $\ge$3.3 $\pm 2$\ $^5$Li + $^7$Li$_{g.s.}$ & $d$ + $^7$Li & $\ge$2.7 $\pm 2$\ $^6$Be + $^6$He$_{g.s.}$ & $^3$He + $^6$He & $\ge$0.2 $\pm 0.7$\ The ratio of $^8$Be+$n$ to $^5$He+$\alpha$ is 2.7, which is close to the ratio 2:1 given in Ref. [@Nyman] and contradicts the conclusion of Ref. [@Grigorenko] where it was reported that the contribution of the $^5$He+$\alpha$ configuration is negligible. The obtained result is partially consistent with the results of Ref. [@Charity] confirming that decay of ${}^9$B has the dominant branch $\alpha$+${}^5$Li implying that “the corresponding mirror state in ${}^{9}$Be would be expected to decay through the mirror channel $\alpha$+${}^{5}$He, instead of through the $n$+${}^{8}$Be(2$^+$) channel”. As for the conclusion of the more detailed experiment Refs. [@Brown; @Papka] where measurements showed that the $^5$He+$\alpha$ channel was not important for the low-lying excited states of $^9$Be, the discrepancy may be explained from the experimental point of view. Our data are based on inclusive measurements, whereas the experimental results of Refs. [@Brown; @Papka] are based on exclusive measurements. The latter always requires fully defined kinematics allowing the complete reconstruction and identification of events. The values of the BR given in Ref. [@Brown] were corrected by the geometrical detector efficiency obtained using a special simulating code. Additionally, it should be noted that the values of the branching ratios given in Ref. [@Brown] are assigned especially to the excited states, whereas our values of the branching ratios are associated entirely with the overall nuclear system $^9$Be. This clearly calls for theoretical calculations of the $^8$Be+$n$ and $\alpha$+$^4$He reaction rate, which at the moment represents a great theoretical challenge. The direct triton pick-up process requires that $^9$Be has the ($^6$Li+$t$) cluster configuration, which seems highly unlikely compared to the ($\alpha$+$\alpha$+$n$) configuration. Also, the interpretation of any ($^3$He,$^6$Li) reaction in terms of triton pick-up requires that $^6$Li has the ($^3$He +$t$) cluster configuration, which seems equally unlikely compared to the ($\alpha$ + $d$) configuration. It has been shown in Ref. [@Young] that for $^6$Li the ($^3$He+$t$) configuration is only slightly less probable than the ($\alpha$+$d$) configuration. Since such cluster configurations are not mutually exclusive, one cannot judge upon the relative importance of the ($^6$Li+$t$) cluster configuration of $^9$Be without making careful cluster structure calculations. Indeed, the differential cross section for the $^7$Li in the outgoing channel is higher than for the channel leading to $^6$Li. Another objective in measuring the ($^3$He,$^6$Li) channel was to evaluate the strength of the triton clustering. The present measurements imply that $^9$Be contains a significant fraction of the ($^6$Li+$t$) cluster configuration. The same is true for the $^3$He-clustering related to the reaction channel $^3$He+$^9$Be$\rightarrow$$^6$He+$^6$Be, i.e. $^6$He+$^3$He clustering in $^9$Be. Similarly, the $^3$He+$^3$He cluster structure could be assigned to $^6$Be, $t$+$^3$He to $^6$Li, $^3$He+$^4$He to $^7$Be as well as $^5$He+$^4$He and $^3$He+$^6$He to $^9$Be. As estimated above, the time for the energy equilibration is rather short. Therefore, it could be assumed that the sequential multi-step reactions cannot not take place and all transfers must happen rather suddenly. Conclusions =========== The angular distributions of the $^9$Be($^3$He,$^3$He)$^9$Be, $^9$Be($^3$He,$^5$He)$^{7}$Be, $^9$Be($^3$He,$^5$Li)$^7$Li, $^9$Be($^3$He,$^6$Be)$^6$He and $^9$Be($^3$He,$^6$Li)$^{6}$Li reaction channels were measured and described within the framework of the optical model, the coupled-channel approach and the distorted-wave Born approximation. The performed analysis of the experimental data shows that the potential parameters are quite sensitive to the exit channel and hence to the cluster structure of the populated states, which allows to make general observations and conclusions regarding the internal structure of the target and residual nuclei. In the first step, the experimental data for $^9$Be+$^3$He elastic scattering were fitted. The fitting parameters were obtained and were compared with the previous measurements. In the second step, the cross sections for the transfer reactions $^9$Be($^3$He,$^4$He)$^8$Be$_{g.s.}$ and $^9$Be($^3$He,$^5$He)$^7$Be were fitted. The fitting result does not reflect any possible peculiarities in the OM parameters in the case of the unbound $^5$He nucleus. The cross sections for $n$, $d$, $t$, and $^3$He transfer channels were measured in the $^3$He+$^9$Be collision at the incident energy 30 MeV. The proposed OM potential provides a good fit of the elastic scattering in the entrance channels. The DWBA calculations agree well with the transfer reaction data. The spectroscopic factors needed to fit the transfer channels are close to unity, which confirms the significant contribution of the considered cluster configurations to the structure of the ground states. We also show the possibility to extract the structural information from the comparative analysis of the $^9$Be($^3$He,$^6$Li)$^6$Li and $^9$Be($^3$He,$^6$Be)$^6$He reaction channels. The experiment was designed to study the breakup of $^9$Be in an attempt to determine the contribution of the $^8$Be+$n$ and $^5$He+$\alpha$ channels in the inclusive measurements. The ratio 2.7:1 may be assigned for the contributions of these two channels, respectively. The determined branching ratios confirm that the $^5$He+$\alpha$ breakup channel plays a significant role. The inclusive experimental method used has the advantage of its experimental simplicity in comparison with the exclusive one  [@Brown; @Papka]. Acknowledgments {#acknowledgments .unnumbered} =============== We would like to thank the JYFL Accelerator Laboratory and NPI (Řež) for giving us the opportunity to perform this study as well as the cyclotron staff of both institutes for the excellent beam quality. This work was supported in part by the Russian Foundation for Basic Research (project numbers: 13-02-00533 and 14-02-91053), the CANAM (IPN ASCR) and by the grants to JINR (Dubna) from the Czech Republic, the Republic of Poland and the mobility grant from the Academy of Finland. References {#references .unnumbered} ========== [20]{} Yu E Penionzhkevich, S M Lukyanov, R A Astabatyan, N A Demekhina, M P Ivanov, R Kalpakchieva, A A Kulko, E R Markaryan, V A Maslov, Yu A Muzychka, R V Revenko, N K Skobelev, V I Smirnov, and Yu G Sobolev, , 025104 (2009). N K Skobelev, N A Demekhina, R Kalpakchieva, A A Kulko, S M Lukyanov, Yu A Muzychka, Yu E Penionzhkevich, and T V  Chuvilskaya, , 208 (2009). T A D Brown, P Papka, B R Fulton, D L Watson, S P Fox, D Groombridge, M Freer, N M Clarke, N I Ashwood, N Curtis, V  Ziman, P McEwan, S Ahmed, W N Catford, D Mahboub, C N Timis, T D Baldwin, D C Weisser, [*Phys. Rev. C*]{} [**76**]{}, 054605 (2007). P Papka, T A D Brown, B R Fulton, D L Watson, S P Fox, D Groombridge, M Freer, N M Clarke, N I Ashwood, N Curtis, V Ziman, P McEwan, S Ahmed, W N Catford, D Mahboub, C N Timis, T D Baldwin, D C Weisser, [*Phys. Rev. C*]{} [**75**]{}, 045803 (2007). L V Grigorenko and M V Zhukov, [*Phys. Rev. C*]{} [**72**]{}, 015803 (2005). L Buchmann, E Gete, J C Chow, J D King, and D F Measday, [*Phys. Rev. C*]{} [**63**]{}, 034303 (2001). N Soić, D Cali, S Cherubini [*et al.*]{}, [*Europhys. Lett.*]{} [**41**]{}, 489 (1998). R J Charity, T D Wiser, K Mercurio, R Shane, L G Sobotka, A H Wuosmaa, A Banu, L Trache, and R E Tribble, [*Phys. Rev. C*]{} [**80**]{}, 024306 (2009). G Nyman, R E Azuma, P G Hansen, B Jonson, P O Larsson, S Mattsson, A Richter, K Riisager, O Tengblad, K Wilhelmsen, and the ISOLDE Collaboration [*et al.*]{}, [*Nucl. Phys. A*]{} [**510**]{}, 189 (1990). S M Lukyanov, A S Denikin, E I Voskoboynik, S V Khlebnikov, M N Harakeh, V A Maslov, Yu E Penionzhkevich, Yu G Sobolev, W H Trzaska, G P Tyurin and K A Kuterbekov, [*J. Phys. G*]{} [**41**]{}, 035103 (2014). C Détraz, H H Duhm and H Hafner, [*Nucl. Phys. A*]{} [**147**]{}, 488 (1970). C Détraz, F Pougheon, M Bernas, M Langevin, P Roussel and J Vernotte, [*Nucl. Phys. A*]{} [**228**]{}, 39 (1974). A T Rudchik, E I Koshhchy, A Budzanovsky, R Siudak, A Szczurek, I Skwirczynska, Yu Mashkarov, L Glowalska, J Turkiewicz, I Zalyubovsky, V Ziman, N Burtebaev, A Duyselbaev, V Adodin, [*Nucl. Phys. A*]{} [**609**]{}, 147 (1996). V Zagrebaev, A Denikin and A Alekseev, $\mathrm{DWBA}$ for nucleon transfer reactions http://nrv.jinr.ru/nrv/webnrv/transfer/ (2009), Nuclear Reaction Video Project. I Thompson, http://www.fresco.org.uk/, http://www.ianthompson.org/surrey/. F C Young, and A R Knudson, [*Nucl. Phys. A*]{} [**184**]{}, 563 (1972).
{ "pile_set_name": "ArXiv" }
--- abstract: | The standard resolution of the twin paradox in relativity states that the age of the inertial twin ‘jumps’ when the traveling twin undergoes his turn-around acceleration. This resolution is based on the use of the equivalent gravitational shift in the frame of the accelerating twin. We show that this is an incorrect and untested use of the equivalence principle. We also propose an unambiguous test of the standard resolution based on the Pound-Rebka experiment.\ [**Keywords: Twin paradox, Gravitational redshift, Equivalence principle**]{} author: - Vasant Natarajan title: The twin paradox in relativity revisited --- The twin paradox in relativity has a history almost as long as the theory of relativity itself. It was originally proposed by Einstein as a [*gedanken*]{} experiment to highlight the fact that an observer sees a moving clock going at a slower rate than a clock at rest. Because this is an entirely symmetric effect (in that each clock sees the other clock ticking slower), the paradox arises when two identical synchronized clocks are temporarily separated into different Lorentz frames and then brought back together. Applied to twins, the paradox starts with identical twins on Earth. One of the twins then accelerates away on a rocket, moves away from Earth at a constant velocity $u$ for a time $T/2$, fires rockets to accelerate again so that his velocity changes to $-u$, moves at the constant velocity $u$ towards the Earth for a time $T/2$, and decelerates to a stop on reaching the Earth again. The acceleration times are assumed to be negligible compared to $T$. The paradox arises because both twins observe the other to be aging slower during the period of uniform relative motion. Are they the same age when they meet again or is one of them younger, and if so, which one and by how much? Over the years, the paradox has been discussed extensively in many books and articles. The American Journal of Physics has carried a large number of articles on the paradox, [@GRE72; @GIG79; @BOU89] highlighting the difficulty in conveying this concept to first-time students of relativity. It still remains one of the most puzzling aspects of the theory of relativity. In the standard resolution, as presented in many textbooks on relativity, both twins conclude that the traveling twin who [*accelerated*]{} is younger. The argument proceeds as follows. The Earth-bound twin always remains in an inertial frame and therefore his observation that the other twin is aging slower is [*correct*]{}. On the other hand, the rocket-bound twin sees his brother age slowly during the time when the relative velocity is constant, but sees a [*sudden jump*]{} in his brother’s age during the short acceleration phase when he is not in an inertial frame. Thus, the change of inertial frames results in a jump in age. The use of the word sudden, which is standard terminology in these discussions, is not meant to imply that the change is discontinuous but only to mean that the change happens during the relatively short acceleration phase. Since the paradox involves an acceleration phase for at least one of the twins, it is useful to bring in some of the framework of general relativity to analyze this problem. A mathematical analysis along these lines is presented by R. C. Tolman in his book [*Relativity Thermodynamics and Cosmology*]{}.[@TOL87] His main argument is that the jump in age seen by the rocket-bound twin during the acceleration phase is due to the [*equivalent gravitational shift*]{} between clocks placed at different points in a uniform gravitational field. By the equivalence principle, the acceleration can be viewed equally well as arising due to the turning on of a uniform gravitation field. If the distance between the twins is $D$, the acceleration of the rocket is $g$, and the duration of acceleration (in the rocket frame) is $\tau_R$, then in this time the accelerated twin will perceive his Earth-bound brother to age by an amount $$\tau_E \approx \tau_R \left( 1 + \frac{g D}{c^2} \right) \, ,$$ where the $\approx$ sign indicates that the relation is correct to first order. This change can be very large compared to $\tau_R$ for large values of $D$. In other words, the short acceleration time in the rocket frame is equal to a large time in the Earth frame. Using this idea, we can calculate the ages of the twins from both viewpoints. For the Earth-bound twin, if the time of uniform relative motion is $T_E$ and the three acceleration phases have negligible duration, then his age has increased by $T_E$, while his brother’s age $T_R$ is shorter and related to $T_E$ by: $$T_E = \frac{T_R}{\sqrt{1-u^2/c^2}} = T_R \left(1 + u^2/2c^2 + \ldots \right) . \label{e1}$$ From the viewpoint of the rocket-bound twin, the equivalent gravitational shifts during the three acceleration phases should also be taken into account. During the initial acceleration away from Earth and the final deceleration upon returning to Earth, the twins are almost at the same gravitational potential and the differential shift is negligible. However, during the intermediate turn-around acceleration, the twins are widely separated and the shift is large. If the travel time (in the rocket frame) before turn around is $T_R/2$, then their separation is $ D = u T_R /2 $. Since the velocity changes by $2u$ in a time $\tau_R$, the acceleration is $ g = 2u / \tau_R $. Therefore, the age of the Earth-bound twin as seen by his brother is: $$T_E = T_R \sqrt{1-u^2/c^2} + \tau_R \left( 1+ \frac{T_R}{\tau_R} \frac{u^2}{c^2} \right) . \label{gr1}$$ Neglecting the term proportional to $\tau_R$, this is the same relation as Eq. (2), and thus both twins agree that the rocket-bound twin is younger, and by the same amount (at least to leading order). We thus see that the successful resolution of the paradox from the viewpoint of the traveling twin rests on the statement that the accelerated clock will see the far-away clock go at a slower rate due to the equivalent gravitational shift predicted by general relativity. In the words of Tolman:[@TOL87]\ [*The solution thus provided for the well-known clock paradox of the special theory $\ldots$ has been made possible by the adoption of the general theory of relativity*]{}. The above analysis also shows why the following statement from Ref. 3 is incorrect:\ [*It has often been pointed out that while the acceleration of one twin is the key to the resolution of the paradox, it is wrong to suppose that reduced aging is a direct result of acceleration. The age difference of the twins is proportional to the length of the trip while the period of acceleration is determined only by how long it takes to turn around and is independent of the length of the trip and, hence, the final age difference of the twins.*]{}\ Eq. \[gr1\] shows clearly that the age difference is indeed proportional to the length of the trip ($T_R$) while the period of acceleration ($\tau_R$) cancels out. In other words, changing the period of acceleration changes the value of $g$ exactly by the amount required to produce an age difference proportional to $T_R$. Some authors argue that there is no need to bring in the framework of general relativity to resolve the paradox. However, even when the analysis stays within the domain of special relativity, these authors conclude that there is a sudden jump in age of the far-away twin during the acceleration phase. To quote from the classic textbook [*Spacetime Physics: Introduction to Special Relativity*]{} by Taylor and Wheeler (Ref. 5, page 130):\ [*This ‘jump’ $\ldots$ is the result of the traveler changing frames. And notice that the traveler is unique in the experience of changing frames; only the traveler suffers the terrible jolt of reversing direction of motion. In contrast, the Earth observer stays relaxed and comfortable in the same frame during the astronaut’s entire trip.*]{} Despite the above arguments, Tolman’s general relativistic analysis has two features that will prove useful to us: The equivalent gravitational shift when the two clocks are nearby is negligible, so that the two observers agree on the duration of the acceleration times in the first and third phases. In other words, a local inertial observer can calculate the acceleration time in the non-inertial frame. The equivalent gravitational shift when the two clocks are widely separated provides a way of quantitatively calculating the ‘jump’ due to the change in frames. Furthermore, Tolman’s analysis shows that it is not the acceleration (or the terrible jolt of changing frames) [*per se*]{} but only the acceleration which happens when the twins are separated that is important for the resolution of the paradox. In fact, this shows that a further simplification is possible. The accelerating twin need not really complete the return journey. It is sufficient if he just decelerates to come to rest with respect to his far-away twin, since their simultaneous ages can now be compared without ambiguity, although they will be at different locations. The initial acceleration which both twins experienced (but at the same location) is not relevant. We next consider a case discussed widely in the American Journal of Physics, [@BOU89] namely one in which both twins experience a [*symmetric*]{} acceleration phase but one that still results in asymmetric aging because they are spatially separated. Consider twins who are moved to widely different locations in the Earth frame after birth. Then they get on to spaceships and accelerate identically by firing identical rockets. After the acceleration phase, they are both moving in the same direction with identical velocity of $u$. Thus they are at rest with respect to each other and their (simultaneous) ages can be compared without ambiguity. There is no doubting the fact that they have gone through identical experiences with respect to the acceleration phase. However, according to the earlier equivalence-principle analysis, the twin on the left would see the one on the right to be older after the acceleration phase. The only asymmetry is that the direction of acceleration is towards the other twin, so that the left twin would be older if they both accelerated to the left instead. In the language of [*Spacetime Physics*]{} (Ref. 5, page 117), the twins are both jumping on to a moving frame at different locations. The analysis using the equivalence principle proceeds as follows. Let the duration of the uniform acceleration phase in the frame of the left twin be $\tau_L$. Then the acceleration is $g = u/ \tau_L$, so that the equivalent duration for the right twin at a distance $x_0$ away is $$\tau_R = \tau_L \left( 1 + \frac{u x_0}{\tau_L c^2} \right) .$$ To give some concrete numerical values, let us take $\tau_L$ to be 1 yr and $u x_0 / c^2$ also to be 1 yr. Thus the left twin concludes that his age after the acceleration phase has increased by 1 yr while that of his twin has increased by 2 yrs. From the point of view of the right twin, the acceleration is $g = u/ \tau_R$, but now he is at the higher gravitational potential so that $$\tau_L = \tau_R \left( 1 + \frac{u x_0}{\tau_R c^2} \right) ^{-1} .$$ Hence he will reach [*the same conclusion*]{} that the twin on the left is younger. But, since the duration of the acceleration phase [*in the rest frame*]{} is the same for both twins, $\tau_R$ is 1 yr and the age of the twin on the left has increased only by 0.5 yr. In other words, while the twins will agree that the twin on the right is older, they will disagree on the amounts of aging. The fact that the acceleration phase lasts for the same (proper) time seems correct because it depends on the amount of fuel and the rate of burning, which is taken to be identical for the two rockets. This is also consistent with point (i) of Tolman’s argument, namely that a locally inertial observer can verify that the duration of the acceleration phases in the non-inertial frame are identical. The above analysis shows that the use of the equivalence principle in a standard variation of the twin paradox, namely the case of the identically accelerated twins, results in an inconsistent answer. It therefore begs the question as to whether the equivalence principle, which is the bedrock of general relativity and undoubtedly correct, is being used properly in this case. Before we answer the question, let us consider the statement of the principle from Weinberg’s book on Relativity and Cosmology:[@WEI72]\ [*at every space-time point in an arbitrary gravitational field it is possible to choose a “locally inertial coordinate system” such that, within a sufficiently small region of the point in question, the laws of nature take the same form as in unaccelerated Cartesian coordinate systems in the absence of gravitation.*]{}\ The key word in the above statement is [*locally*]{}, defining a region which is “sufficiently small” so that the gravitational field is constant and tidal effects can be ignored. On the other hand, the gravitational redshift is an explicitly [*nonlocal*]{} phenomenon, relating to the relative rates of two clocks placed at different potentials in a gravitational field. Therefore, the use of the equivalent gravitational shift in an accelerated frame requires adequate care. To see how to derive this, we follow the procedure given by Tolman in his book.[@TOL87] Consider two identical clocks, placed left and right and separated by a distance $D$, in a frame being uniformly accelerated to the right at a rate $g$. Let the first clock emit a photon at time $t=0$. Traveling at $c$, the photon reaches the second clock after a time $D/c$. In this time, the second clock has picked up an additional speed of $gD/c$, so that the first-order Doppler effect yields for the relative rates $$\frac{\tau_R}{\tau_L} = 1 + \frac{g D}{c^2} \, ,$$ exactly as expected for the gravitational redshift in a uniform gravitational field. Thus, the physical basis for the shift in an accelerated frame is the normal Doppler effect. But this derivation is valid in a frame undergoing constant acceleration that lasts forever, or at least as long as it takes for a photon to reach from one clock to the other. This is explicitly not the case in the twin-paradox experiment, where the duration of the acceleration phase is negligibly small and certainly not long enough for a photon to cover the distance between the twins. The increased relative velocity can at most be $u$ (depending on the instant at which the photon was emitted), and definitely not $gD/c$. This shows clearly why the equivalent shift cannot be applied to the short acceleration phase of the twin-paradox kind. This also shows that the use of the equivalence principle in the standard resolution to the twin paradox [*remains untested*]{}. We therefore propose a straightforward and unambiguous test of the jump in age predicted by the twin-paradox effect with the following experiment. We propose a repeat of the Pound-Rebka measurement of the gravitational redshift using recoilless gamma rays emitted by iron nuclei,[@POR60b] but this time with the source and absorber at the same level and [*with the absorber accelerated towards the source at a constant rate $g$*]{}. Since the source and absorber are at the same gravitational potential, there will be no frequency shift due to the gravitational redshift. However, the twin-paradox effect predicts a shift of $gD/c^2$, where $D$ is the distance between the two. This is an exact reproduction of the twin-paradox experiment, and should provide conclusive results one way or the other. There are several attractive features of the proposed experiment. First is that, since the sign of the shift depends on the direction of acceleration, several sources of systematic error can be eliminated by looking for a difference between acceleration to the right and acceleration to the left. Secondly, the shift depends linearly on the distance $D$. Therefore, one can increase the sensitivity of the experiment by increasing $D$. Note that the original Pound-Rebka experiment was done with a height difference of 74 ft, while on the ground the distance could be increased by a factor of 2 or more. Finally, there will be a small shift from the first-order Doppler effect depending on the exact velocity of the source at the time of emission, but there are two experimental handles to address this error: (i) the shift will be limited in magnitude by the final velocity of the absorber which can be made small, and (ii) it will be independent of $D$. Let us also consider here an experiment [@POR60a] that is often quoted as an experimental verification of asymmetric aging in the twin paradox (for example in Ref. 5, page 134). This is also an experiment by Pound and Rebka, but where the frequency of the recoilless gamma rays were studied as a function of temperature. This experiment was done primarily to check for sources of systematic error in their gravitational redshift experiment. They did find a shift in the frequency of the gamma ray as the temperature was increased. However, this is not a test of the twin-paradox effect. On the contrary, this is only a measurement of the second-order Doppler effect, because the increased temperature results in an increased average value of $u^2$. Indeed, if we could transform to the frame of the oscillating iron nuclei, the laboratory clocks would appear to go slower! This is a manifestation of time dilation similar to the [*apparently*]{} long lifetime of particles decaying in flight as observed from a stationary frame. Nobody will argue that particles in flight age slower because of this apparent dilation. We conclude by discussing what, in the author’s opinion, is the logical fallacy in the standard resolution to the twin paradox. The fallacy is to ascribe a [*real*]{} change to the [*apparent*]{} effect of time dilation between relatively moving frames. This is akin to saying that the well-known phenomenon of Lorentz contraction is real and causes a real change in the length of a moving rod. The length of a rod is determined by interatomic forces and is not dependent on its motion with respect to some arbitrary frame, while the Lorentz contraction is an apparent effect due to the relativity of simultaneity. Similarly, the age of a person (or clock) is determined by physical processes that control decay rates. Moving one person temporarily to another Lorentz frame cannot change this rate. Indeed, it can be shown that the differential aging in the twin-paradox effect arises from the fact that the distance to the turn-around point appears Lorentz contracted in the Rocket-bound frame and hence takes a shorter time to travel. This leads to the unlikely conclusion that travel to a distant star is possible if we travel close to the speed of light. The nearest star (outside the solar system) is 4.5 light years away, so that an astronaut traveling at $0.9c$ would take 5 yrs (in the Earth frame) to get there. But the twin-paradox effect predicts that the astronaut would have aged only by $5 \sqrt{1-0.9^2}=2.18$ yrs (Eq. \[e1\]). From the Lorentz contraction point of view, the distance to the star appears shorter by the same factor, and therefore takes a shorter time to travel. This suggests that the distance to any faraway object can be made arbitrarily small by traveling sufficiently close to the speed of light. Not very likely! It is also important to realize that the gravitational redshift is a real effect on the relative rates of clocks placed at different gravitational potentials. The curvature of spacetime in the presence of a gravitational field has a physical effect on natural processes so that there is a difference in clock rates when one clock is compared to the other, though no local measurements on one clock will show any change because all processes are affected the same way. Therefore, if one twin climbs up to the top of a mountain (i.e. to a higher gravitational potential) while his brother stays at the base, lives atop for one year, and then comes back down to rejoin his brother, he will find that his age is slightly higher. But this is because the curved spacetime near the Earth’s surface causes a real change in the relative heart rates of the two brothers. But mere acceleration in flat spacetime for a short time (short enough so that the normal Doppler effect can be neglected) can not cause such a change, as required in the twin-paradox effect. We have therefore proposed a simple experiment to test this hypothesis. The author thanks Krishna Parat for introducing him to the twin paradox and Wolfgang Ketterle for useful discussions. [1]{} D. M. Greenberger, “The reality of the twin paradox effect,” Am. J. Phys. **40**, 750–754 (1972). C. Giannoni and O. Gron, “Rigidly connected accelerated clocks,” Am. J. Phys. **47**, 431–435 (1979). S. P. Boughn, “The case of the identically accelerated twins,” Am. J. Phys. **57**, 791–793 (1989). R. C. Tolman, *Relativity Thermodynamics and Cosmology* (Dover Publications Inc., New York, 1987). E. Taylor and J. A. Wheeler, *Spacetime Physics: Introduction to Special Relativity* (W. H. Freeman and Company, New York, 1992), 2nd ed. S. Weinberg, *Gravitation and cosmology: Principles and applications of the general theory of relativity* (John Wiley and Sons, New York). R. V. Pound and J. G. V. Rebka, “Apparent weight of photons,” Phys. Rev. Lett. **4**, 337–341 (1960). R. V. Pound and J. G. V. Rebka, “Variation with temperature of the energy of recoil-free gamma rays from solids,” Phys. Rev. Lett. **4**, 274–275 (1960).
{ "pile_set_name": "ArXiv" }
--- author: - 'T. Shigehara$^{1}$, H. Mizoguchi$^{1}$, T. Mishima$^{1}$ and Taksu Cheon$^{2}$' title: Wave Chaos in Quantum Pseudointegrable Billiards --- We clarify from a general perspective, the condition for the appearance of chaotic energy spectrum in quantum pseudointegrable billiards with a point scatterer inside. Introduction ============ The quantum billiard with point scatterers is a quasi-exactly solvable model which is closely related to real systems. The billiards are a natural idealization of the particle motion in bounded systems. The single-electron problem in mesoscopic structures is now a possible setting, owing to the rapid progress in the mesoscopic technology. Real systems are, however, not free from impurities which affect the particle motion inside. In the presence of a small amount of contamination, even a single-electron problem becomes unmanageable. The modeling of the impurities with point scatterers makes the problem easy to handle without changing essential dynamics. A fundamental problem for the billiards considered here is to understand global behavior of the energy spectrum in the parameter space of particle energy and the strength of the scatterers. In particular, statistical properties of the spectrum are important because they reflect the degree of complexity of underlying dynamics. It is widely believed that integrable systems obeys Poisson statistics, while the predictions of the Gaussian Orthogonal Ensembles (GOE) describe chaotic systems [@BO89]. In this paper, we discuss the spectral properties of quantum billiards with a single point scatterer. It should be noticed that the nature of classical motion in such systems is pseudointegrable [@RB81] in the sense that unstable trajectories, which hit the point scatterer, are of measure zero. As a result, it is expected that the energy spectrum of a quantum analogue does not substantially differ from Poisson statistics. However, as shown in this paper, quantization induces the chaotic spectrum under a certain condition. This phenomenon might be called [*wave chaos*]{} [@AS91] because its origin is the wavelike nature of quantum dynamics. After deriving the eigenvalue equation with the help of self-adjoint extension theory in functional analysis, we discuss the statistical properties of the energy spectrum from a general perspective and deduce the condition for the appearance of chaotic spectrum in the quantum pseudointegrable system. Formulation =========== We start from an empty billiard in spatial dimension $d$, $d=1, 2, 3$. Let us consider a quantum point particle of mass $M$ moving freely in a bounded region $\Omega^{(d)}$. We impose the Dirichlet boundary condition such that wave functions vanish on the boundary of $\Omega^{(d)}$. The eigenvalues and the corresponding normalized eigenfunctions are denoted by $E_n$ and $\varphi_n(\vec{x})$ respectively; $$\begin{aligned} \label{eq2-1} H_0 \varphi _n({\vec x}) \equiv -{\nabla^2 \over {2M}} \varphi _n({\vec x}) = E _n \varphi _n({\vec x}). \end{aligned}$$ The Hamiltonian $H_0$ is the kinetic operator in $L^2(\Omega^{(d)})$ with domain $D(H_0)=H^2(\Omega^{(d)})\cap H^1_0(\Omega^{(d)})$ in terms of the Sobolev spaces. The Green’s function of $H_0$ is given by $$\begin{aligned} \label{eq2-2} G^{(0)}({\vec x},{\vec x'};\omega ) = \sum\limits_{n=1}^\infty {{{\varphi _n({\vec x})\varphi _n({\vec x'})} \over {\omega -E_n}}}.\end{aligned}$$ The average level density is given by $$\begin{aligned} \label{eq2-3} \rho^{(d)}_{av}(\omega) = \left\{ \begin{array}{ll} \frac{M^{1/2}\Omega^{(1)}}{2^{1/2}\pi}\frac{1}{\sqrt{\omega}}, & \ \ \ d=1, \\ \frac{M\Omega^{(2)}}{2\pi}, & \ \ \ d=2, \\ \frac{M^{3/2}\Omega^{(3)}}{2^{1/2}\pi^2}\sqrt{\omega}, & \ \ \ d=3, \end{array} \right. \end{aligned}$$ where we denote the “volume” of $\Omega^{(d)}$ by the same symbol. (For example, the area of a two-dimensional region $\Omega^{(2)}$ is simply denoted by $\Omega^{(2)}$.) We now place a single point scatterer at ${\vec x_0}$ inside the billiard. Naively, one defines the scatterer in terms of the $d$-dimensional Dirac’s delta function; $$\begin{aligned} \label{eq2-4} H=H_0+ v \delta^{(d)} ({\vec x}-{\vec x_0}).\end{aligned}$$ However, the Hamiltonian $H$ is not mathematically sound for $d \geq 2$. It is easy to see that the eigenvalue equation of $H$ is reduced to $$\begin{aligned} \label{eq2-5} \sum_{n=1}^{\infty} \frac{\varphi_n(\vec{x}_0)^2}{\omega-E_n}=v^{-1}. \end{aligned}$$ Keeping Eq.(\[eq2-3\]) in mind, we realize that the infinite series in Eq.(\[eq2-5\]) does not converge for $d \geq 2$, since the average of the numerator (among many $n$) is energy-independent; $\left< \varphi_n (\vec{x}_0)^2 \right> \simeq 1/\Omega^{(d)}$. To handle the divergence, a scheme for renormalization is called for. One of the most satisfying schemes is given by the self-adjoint extension theory of functional analysis [@AG88]. We first consider in $L^2(\Omega^{(d)})$ the nonnegative operator $$\begin{aligned} \label{eq2-7} H_{\vec{x}_0} = - \left. {\nabla^2 \over {2M}} \right|_{C^{\infty}_0 (\Omega^{(d)}-\vec{x}_0)} \end{aligned}$$ with its closure ${\bar H_{\vec{x}_0}}$ in $L^2(\Omega^{(d)})$. Namely, we restrict $D(H_0)$ to the functions which vanish at the location of the point scatterer. By using integration by parts, it is easy to prove that the operator ${\bar H_{\vec{x}_0}}$ is symmetric (Hermitian). But it is not self-adjoint. Indeed, the equation $$\begin{aligned} \label{eq2-8} {\bar H_{\vec{x}_0}}^* \psi_{\omega}({\vec x}) = \omega \psi_{\omega}({\vec x}), \ \ \ \psi \in D({\bar H_{\vec{x}_0}}^*) \end{aligned}$$ has a solution for $Im \ \omega \neq 0$ [@ZO80]; $$\begin{aligned} \label{eq2-9} \psi_{\omega}({\vec x}) = G^{(0)}({\vec x},{\vec x_0};\omega ), \ \ {\vec x} \in \Omega^{(d)}-\vec{x}_0 \end{aligned}$$ indicating $$\begin{aligned} \label{eq2-10} D(\bar{H}_{\vec{x}_0}^*) & \! \! \! \! \! = & \! \! \! \! \! D(\bar{H}_{\vec{x}_0}) \oplus N(\bar{H}_{\vec{x}_0}^* -\omega) \oplus N(\bar{H}_{\vec{x}_0}^* -\bar{\omega}) \nonumber \\ & \! \! \! \! \! \neq & \! \! \! \! \! D(\bar{H}_{\vec{x}_0}), \end{aligned}$$ where $N(A)$ is the kernel of an operator $A$. Since ${\bar H_{\vec{x}_0}}$ has the deficiency indices $(1,1)$, ${\bar H_{\vec{x}_0}}$ has one-parameter family of self-adjoint extensions $H_{\theta}$ [@RS75]; $$\begin{aligned} \label{eq2-11} D(H_{\theta}) & \! \! \! \! \! = & \! \! \! \! \! \{ f \vert f = \varphi + c ( \psi_{i\Lambda} - e^{i\theta} \psi_{-i\Lambda} ); \nonumber \\ && \varphi\in D(\bar{H}_{\vec{x}_0}), c\in {\bf C}, 0 \leq \theta < 2\pi \}, \nonumber \\ H_{\theta} f & \! \! \! \! \! = & \! \! \! \! \! \bar{H}_{\vec{x}_0} \varphi+i \Lambda c ( \psi_{i\Lambda} + e^{i\theta} \psi_{-i\Lambda} ), \end{aligned}$$ where $\Lambda >0$ is an arbitrary mass scale. With the aid of Krein’s formula, we can write down the Green’s function for the Hamiltonian $H_{\theta}$ as $$\begin{aligned} \label{eq2-12} G_{\theta}(\vec{x},\vec{x}';\omega) & \! \! \! \! \! = & \! \! \! \! \! G^{(0)}(\vec{x},\vec{x}';\omega) \nonumber \\ & \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! + & \! \! \! \! \! \! \! \! \! \! \! \! \! \! G^{(0)}(\vec{x},\vec{x}_{0};\omega) T_{\theta}(\omega) G^{(0)}(\vec{x}_{0},\vec{x}';\omega). \end{aligned}$$ The transition matrix (T-matrix) $T_{\theta}$ is calculated by $$\begin{aligned} \label{eq2-13} T_{\theta}(\omega) =\frac{1-e^{i\theta}} {(\omega-i\Lambda) c_{i\Lambda}(\omega)- e^{i\theta}(\omega+i\Lambda) c_{-i\Lambda}(\omega)}, \end{aligned}$$ where $$\begin{aligned} \label{eq2-14} c_{\pm i\Lambda}(\omega)= \int_{\Omega^{(d)}} \! \! \! \! \! G^{(0)}(\vec{x},\vec{x}_{0};\omega) G^{(0)}(\vec{x},\vec{x}_{0};\pm i\Lambda)d\vec{x}. \end{aligned}$$ Using the resolvent equation, we have $$\begin{aligned} \label{eq2-15} T_{\theta}(\omega)=(v_{\theta}^{-1}-G(\omega))^{-1}, \end{aligned}$$ where $$\begin{aligned} \label{eq2-16} v_{\theta}^{-1}=\Lambda \cot \frac{\theta}{2} \sum_{n=1}^{\infty} \frac{\varphi_{n}(\vec{x}_{0})^{2}}{E_{n}^2+\Lambda^{2}}, \end{aligned}$$ $$\begin{aligned} \label{eq2-17} G(\omega)=\sum_{n=1}^{\infty} \varphi_{n}(\vec{x}_{0})^{2} (\frac{1}{\omega-E_{n}}+\frac{E_{n}} {E_{n}^2+\Lambda^{2}}).\end{aligned}$$ The constant $v_{\theta}$ is a coupling constant of the point scatterer, the value of which ranges over the whole real number as one varies $0\leq\theta <2\pi$. It follows from Eq.(\[eq2-15\]) that the eigenvalues of $H_{\theta}$ are determined by $$\begin{aligned} \label{eq2-18} G(\omega)=v_{\theta}^{-1}.\end{aligned}$$ On any interval $(E_n,E_{n+1})$, $G$ is a monotonically decreasing function of $\omega$ that ranges over the whole real number. This means that Eq.(\[eq2-18\]) has a single solution on each interval for any $v_{\theta}$. The eigenfunction of $H_{\theta}$ corresponding to an eigenvalue $\omega_n$ is given by $$\begin{aligned} \label{eq2-19} \psi_n (\vec{x}) \propto G^{(0)}(\vec{x},\vec{x}_0;\omega_n). \end{aligned}$$ Condition for Strong Coupling ============================= $d=1$ ----- In one dimension, each infinite series on RHS in Eq.(\[eq2-17\]) converges separately. Thus, if we set $$\begin{aligned} \label{eq3a-1} v^{-1}=v_{\theta}^{-1}- \sum_{n=1}^{\infty} \varphi_{n}(\vec{x}_{0})^{2}\frac{E_{n}} {E_{n}^2+\Lambda^{2}}, \end{aligned}$$ Eq.(\[eq2-18\]) is reduced to Eq.(\[eq2-5\]); A somewhat complicated argument above leads us to the well-known result with Dirac’s delta. In the following, we consider Eq.(\[eq2-5\]), the LHS of which is denoted by $G^{(1)}(\omega)$. To obtain each solution of Eq.(\[eq2-5\]), a numerical task is needed in general. However, since our main purpose lies in the statistical properties of spectrum, we proceed further by introducing some approximations without losing the essence, while still keeping loss of generality minimal [@SC97]. The first is that the value of $\varphi_n (\vec{x}_0)^2$ is replaced by its average among many $n$; [^1] $$\begin{aligned} \label{eq3a-2} \varphi_n (\vec{x}_0)^2 \simeq \langle \varphi_n (\vec{x}_0)^2 \rangle =1/\Omega^{(d)}. \end{aligned}$$ Since the statistics is taken within a large number of, sometimes thousands of eigenstates, Eq.(\[eq3a-2\]) is quite satisfactory. Keeping $\left| \varphi_n(\vec{x}_0) \right| \simeq 1/\sqrt{\Omega^{(d)}}$ in mind, we recognize from Eq.(\[eq2-19\]) with Eq.(\[eq2-2\]) that if $\omega_m \simeq E_m$ (or $E_{m+1}$) for some $m$, then $\psi_m \simeq \varphi_m$ (or $\varphi_{m+1}$), implying that only $\psi_m$ with an eigenvalue $\omega_m \simeq (E_{m}+E_{m+1})/2$ is distorted by a point scatterer. For such $\omega_m$, since the contributions on the summation of $G^{(d)}$ from the terms with $n \simeq m$ cancel each other, $G^{(d)}$ can be estimated by a principal integral; $$\begin{aligned} \label{eq3a-3} G^{(d)}(\omega_m) \simeq g^{(d)}(\omega_m), \ \ \ \omega_m \simeq \frac{E_{m}+E_{m+1}}{2}, \end{aligned}$$ $$\begin{aligned} \label{eq3a-4} g^{(d)}(\omega) = \langle \varphi_{n}(\vec{x}_{0})^{2} \rangle P \int_{0}^{\infty} \frac{\rho_{av}^{(d)}(E)}{\omega-E}dE, & d=1, \end{aligned}$$ where we have defined a continuous function $g^{(d)}$ of $\omega$ which behaves like an interpolation of the inflection points of $G^{(d)}$. Inserting Eq.(\[eq2-3\]) into Eq.(\[eq3a-4\]) and using the elementary indefinite integral $$\begin{aligned} \label{eq3a-5} \int \frac{dE}{(\omega-E)\sqrt{E}} = \frac{1}{\sqrt{\omega}} \ln \left| \frac{\sqrt{\omega}+\sqrt{E}}{\sqrt{\omega}-\sqrt{E}} \right| \end{aligned}$$ for $\omega>0$, we obtain for $\omega_m \simeq (E_{m}+E_{m+1})/2$ $$\begin{aligned} \label{eq3a-6} G^{(d)}(\omega_m) \simeq 0, & d=1, \end{aligned}$$ indicating that the maximal coupling of a point scatterer is attained when the strength $v$ satisfies $$\begin{aligned} \label{eq3a-7} v^{-1} \simeq 0, & d=1, \end{aligned}$$ which is going well with our intuition. The “width” of the strong coupling region (allowable error of $v^{-1}$ in Eq.(\[eq3a-7\])) is estimated by considering a linearized eigenvalue equation. Expanding $G^{(d)}$ at $\omega_m \simeq (E_{m}+E_{m+1})/2$, we can rewrite the eigenvalue equation as $$\begin{aligned} \label{eq3a-8} G^{(d)'}(\omega_m)(\omega-\omega_m) \simeq v^{-1}-G^{(d)}(\omega_m). \end{aligned}$$ In order that Eq.(\[eq3a-8\]) has a solution $\omega\simeq \omega_m$, the range of RHS has to be restricted to $$\begin{aligned} \label{eq3a-9} \left| v^{-1} - G^{(d)}(\omega_m) \right| \ \ {\displaystyle{\mathop{<}_{\sim}}}\ \ \frac{\Delta^{(d)}(\omega_m)}{2}, \end{aligned}$$ $$\begin{aligned} \label{eq3a-10} \Delta^{(d)}(\omega_m) \equiv \left| G^{(d)'}(\omega_m) \right| \rho_{av}^{(d)}(\omega_m)^{-1}, \end{aligned}$$ where we have defined the width $\Delta^{(d)}$ which is nothing but the average variance of the linearized $G^{(d)}$ on the interval $(E_m,E_{m+1})$. Using the approximation (\[eq3a-2\]), the value of $\left| G^{(d)'}(\omega_m) \right|$ can be estimated as follows; $$\begin{aligned} \label{eq3a-11} \left| G^{(d)'}(\omega_m) \right| & = & \sum_{n=1}^{\infty} \left( \frac{\varphi_{n}(\vec{x}_{0})} {\omega_m-E_{n}} \right)^2 \nonumber \\ & \simeq & \langle \varphi_{n}(\vec{x}_{0})^{2} \rangle \sum_{n=1}^{\infty} \frac{2} {\{(n-\frac{1}{2})\rho_{av}^{(d)}(\omega_m)^{-1} \}^2} \nonumber \\ & = & 8 \langle \varphi_{n}(\vec{x}_{0})^{2} \rangle \rho_{av}^{(d)}(\omega_m)^2 \sum_{n=1}^{\infty} \frac{1}{(2n-1)^2} \nonumber \\ & = & \pi^{2} \langle \varphi_{n}(\vec{x}_{0})^{2} \rangle \rho_{av}^{(d)}(\omega_m)^2. \end{aligned}$$ The second equality follows from the approximation that the unperturbed eigenvalues are distributed with a mean interval $\rho_{av}^{(d)}(\omega_m)^{-1}$ in the whole energy region. This assumption is quite satisfactory, since the denominator of $G^{(d)'}(\omega)$ is of the order of $(\omega-E_n)^2$, indicating that the summation in Eq.(\[eq3a-11\]) converges rapidly. From Eqs.(\[eq2-3\]) and (\[eq3a-2\]) for $d=1$, we obtain $$\begin{aligned} \label{eq3a-12} \Delta^{(1)}(\omega_m) \simeq \frac{\pi M^{1/2}}{2^{1/2}}\frac{1}{\sqrt{\omega_m}}, \end{aligned}$$ which is inversely proportional to square root of the energy $\omega$. We can summarize the finding in one dimension as follows; The effect of a point scatterer of coupling strength $v$ is substantial only in the eigenstates with eigenvalue $\omega$ such that $$\begin{aligned} \label{eq3a-13} \left| v^{-1} \right| {\displaystyle{\mathop{<}_{\sim}}}\frac{\Delta^{(1)}(\omega)}{2} \simeq \frac{\pi M^{1/2}}{2^{3/2}}\frac{1}{\sqrt{\omega}}. \end{aligned}$$ $d=2, 3$ -------- It is easy to apply the argument above to higher dimensions. Since each series on RHS in Eq.(\[eq2-17\]) diverges when summed separately for $d=2,3$, we have to resort to Eq.(\[eq2-18\]). As a result, Eq.(\[eq3a-4\]) is replaced by $$\begin{aligned} \label{eq3b-1} g^{(d)}(\omega) = \langle \varphi_{n}(\vec{x}_{0})^{2} \rangle P \int_{0}^{\infty} \left( \frac{1}{\omega-E}+\frac{E}{E^{2}+\Lambda^2} \right) \nonumber\end{aligned}$$ $$\begin{aligned} \times \rho_{av}^{(d)} (E)dE, & d=2,3. \end{aligned}$$ Using the elementary indefinite integrals, $$\begin{aligned} \label{eq3b-2} \int \left( \frac{1}{\omega-E}+\frac{E}{E^{2}+\Lambda^2} \right) dE = -\ln \frac{ | \omega-E | } {\sqrt{ E^2 + \Lambda^2}}, \end{aligned}$$ $$\begin{aligned} \int \left( \frac{1}{\omega-E}+\frac{E}{E^{2}+\Lambda^2} \right) \sqrt{E} dE = \sqrt{\omega} \ln \left| \frac{\sqrt{\omega}+\sqrt{E}} {\sqrt{\omega}-\sqrt{E}} \right| \nonumber \end{aligned}$$ $$\begin{aligned} - \frac{1}{2}\sqrt{\frac{\Lambda}{2}} \ln \left( \frac{E+\sqrt{2\Lambda E}+\Lambda}{E-\sqrt{2\Lambda E}+\Lambda} \right) \nonumber \end{aligned}$$ $$\begin{aligned} - \sqrt{\frac{\Lambda}{2}} \left\{ \arctan(\sqrt{\frac{2E}{\Lambda}}+1)+ \arctan(\sqrt{\frac{2E}{\Lambda}}-1) \right\}, \nonumber \end{aligned}$$ $$\begin{aligned} \label{eq3b-3} {\rm for \ \ } \omega>0, \end{aligned}$$ we obtain for $\omega_m \simeq (E_m + E_{m+1})/2$ $$\begin{aligned} \label{eq3b-4} G^{(d)}(\omega_m) \simeq \left\{ \begin{array}{ll} \frac{M}{2\pi} \ln \frac{\omega_m}{\Lambda}, & d=2, \\ -\frac{M^{3/2}\Lambda^{1/2}}{2\pi}, & d=3. \end{array} \right. \end{aligned}$$ The width $\Delta^{(d)}$ is estimated by Eq.(\[eq3a-10\]) with Eq.(\[eq3a-11\]). Using Eq.(\[eq2-3\]), we find $$\begin{aligned} \label{eq3b-5} \Delta^{(d)}(\omega_m) \simeq \left\{ \begin{array}{ll} \frac{\pi M}{2}, & d=2, \\ \frac{M^{3/2}}{2^{1/2}}\sqrt{\omega_m}, & d=3. \end{array} \right. \end{aligned}$$ From Eq.(\[eq3a-9\]) with $v$ replaced by $v_{\theta}$, we realize that the strong coupling is attained under the condition $$\begin{aligned} \label{eq3b-6} \left| v_{\theta}^{-1} - \frac{M}{2\pi} \ln \frac{\omega}{\Lambda} \right| {\displaystyle{\mathop{<}_{\sim}}}& \! \! \! \! \! \! \! \! \! \! \displaystyle \frac{\pi M}{4}, & d=2, \\ \label{eq3b-7} \left| v_{\theta}^{-1} + \frac{M^{3/2}\Lambda^{1/2}}{2\pi} \right| {\displaystyle{\mathop{<}_{\sim}}}& \displaystyle \frac{M^{3/2}}{2^{3/2}}\sqrt{\omega}, & d=3. \end{aligned}$$ Numerical Example ================= We consider a quantum particle with mass $M=1/2$ moving in a three-dimensional rectangular box with side-lengths $(l_{x},l_{y},l_{z})=(1.047, 1.186, 0.8049)$ and hence $\Omega^{(3)}=1$. The mass scale is set to $\Lambda=1$. A single point scatterer is placed at the center of the billiard. We take into account only the states with even parity in each direction, since the others are unaffected by the scatterer in this case. Fig.1 shows, for $v_{\theta}^{-1}=0, 10$ and $30$, the nearest-neighbor level spacing distribution $P(S)$, which is defined such that $P(S)dS$ is the probability to find the spacing between any two neighboring eigenstates in the interval $(S,S+dS)$. (The average spacing is normalized to one.) Generic integrable systems obey the Poisson distribution $P(S)=\exp (-S)$ (dotted line), while chaotic systems are described by the GOE prediction $P(S)=\pi S /2 \times \exp(-\pi S^2/4)$ (solid line). The statistics is taken within $\omega_{100} \sim \omega_{3100}$ in Fig.1 ($z_{1600}=8304$). According to Eq.(\[eq3b-7\]), the most strong coupling is attained for $v_{\theta}^{-1}=-M^{3/2}\Lambda^{1/2}/2\pi = -0.05626$. As $v_{\theta}^{-1}$ increases, $P(S)$ tends to approach the Poisson prediction. For $v_{\theta}^{-1} \simeq \Delta^{(3)}(\omega_{1600})/2 =11.3$, $P(S)$ is expected to be intermediate in shape between Poisson and GOE. The value can be considered as the upper bound of $v_{\theta}^{-1}$ for inducing a GOE-like shape in this energy region. With $v_{\theta}^{-1}$ beyond the bound, the system is not substantially different from the empty billiard, and as a result, $P(S)$ resembles to the Poisson distribution. These features are successfully reproduced in Fig.1. Conclusion ========== We have discussed the condition for the appearance of chaotic spectrum in the quantum pseudointegrable billiards with a single point scatterer inside. Chaotic spectrum appears if the conditions (\[eq3a-13\]), (\[eq3b-6\]) and (\[eq3b-7\]) are satisfied for dimension $d=1,2,3$, respectively. In two dimension, the condition is described by a logarithmically energy-dependent strip with a constant width, indicating that the system recovers integrability in the high energy limit for any $v_{\theta}$. In three dimension, the spectrum shows GOE-like nature when $v_{\theta}^{-1}$ is within a band whose width increases parabolically as a function of the energy. This implies that the spectrum becomes chaotic at high energy for any $v_{\theta}$, which makes a clear contrast with two dimension. [99]{} O. Bohigas, [*Random Matrix Theories and Chaotic Dynamics*]{}, Les Houches Summer School Proceedings Session 52, edited by M.-J. Giannoni, A. Voros and J. Zinn-Justin (North-Holland, Amsterdam, 1989), p.87 and references therein. P. J. Richens and M. V. Berry, Physica D [**2**]{}, 495 (1981); A. N. Zemlyakov and A. B. Katok, Math. Notes [**18**]{}, 291 (1975). S. Albeverio and P. Šeba, J. Stat. Phys. [**64**]{}, 369 (1991). S. Albeverio, F. Gesztesy, R. H[ø]{}egh-Krohn and H. Holden, [*Solvable Models in Quantum Mechanics*]{} (Springer-Verlag, New York, 1988). J. Zorbas, J. Math. Phys. [**21**]{}, 840 (1980). M. Reed and B. Simon, [*Methods of Modern Mathematical Physics*]{}, Vol. 2, [*Fourier Analysis, Self-adjointness*]{} (Academic Press, New York, 1975). T. Shigehara and T. Cheon, Phys. Rev. E [**55**]{}, 6832 (1997). [**Fig.1.**]{} $P(S)$ for $v_{\theta}^{-1}=0, 10, 30$. The solid (dotted) line is GOE prediction (Poisson distribution). [^1]: With some minor modifications if necessary, a major part of the argument here is applicable to higher dimensions. We thus explicitly signify “$d=1$” only for the assertions specific to one dimension.
{ "pile_set_name": "ArXiv" }
[Notes on Strings and Higher Spins]{}\ \ [*Scuola Normale Superiore and INFN\ Piazza dei Cavalieri, 7\ I-56126 Pisa  ITALY\ e-mail: [*[email protected]*]{}*]{} [Abstract]{} These notes are devoted to the intriguing and still largely unexplored links between String Theory and Higher Spins, the types of excitations that lie behind its most cherished properties. A closer look at higher–spin fields provides some further clues that String Theory describes a broken phase of a Higher–Spin Gauge Theory. Conversely, string amplitudes contain a wealth of information on higher–spin interactions that can clarify long–standing issues related to their infrared behavior. 3 truecm [*Based on the lectures presented at the International School for Subnuclear Physics “Searching for the Unexpected at LHC and Status of Our Knowledge” (Erice, June 24–July 3  2011), and on the talks presented at “Strings, Branes and Supergravity” (Istanbul, July 31–Aug 5  2011), at “Quantum Theory and Symmetries (QTS7)” (Prague, Aug. 7–13  2011) and at “Fundamental Fields and Particles (FFP12)” (Udine, Nov 21–23  2011). To be reprinted in a special J. Phys. A issue devoted to Higher–Spin Theory, eds. M. Gaberdiel and M.A. Vasiliev.*]{} \[sec:Introduction\] More than forty years ago, Veneziano [@veneziano] gave birth to String Theory [@stringtheory] relating Euler’s Beta–function $B(u,v)$ to tree–level S–matrix amplitudes. Together with remarkable subsequent developments, this linked its two arguments $u$ and $v$ to $\a^\prime$, a parameter reflecting a string tension, and to the Mandelstam variables $s$ and $t$ of four–particle amplitudes, in what eventually became tachyon amplitudes of the 26–dimensional bosonic string. This model, however, also contains infinitely many massive excitations, and among them infinitely many higher–spin (HS) ones [@solvay] that correspond to multi–symmetric tensors of the type $\vf_{\mu_1 \ldots \, \mu_{s_1}; \, \nu_1 \ldots \, \nu_{s_2};\, \ldots}$. These massive HS excitations are crucial in order to guarantee “planar duality”, the manifest symmetry of $B(u,v)$ under the interchange of its two arguments, despite their different origins in diagrammatic constructions. For instance, the amplitude built exhibiting *only* $s$–channel poles also possesses $t$–channel poles that can be recovered by resumming polynomial residues, and clearly their infinite number is instrumental to this end. One can argue, in similar naive terms, that other key properties of String Theory, such as modular invariance or open–closed duality, rest on the presence of infinitely many HS modes. Yet, HS amplitudes have not attracted much attention over the years, possibly because they appear somewhat remote from applications, but certainly because they are rather unwieldy in general. Perhaps this has some bearing on the present, not fully satisfactory, state of String Theory. No neat geometrical principles underlying it have emerged over the years, and this incomplete grasp of the foundations goes along with a lack of universal agreement on the actual lessons that it has in store. Could it be, then, that the massive excitations hold the key for penetrating more deeply into the whole setup? And if this were the case, could the key for understanding the massive excitations lie in the breaking of HS symmetries? This picture, in fact, has long been in the minds of many authors, at least since the 1980’s, when String Field Theory [@sftheory] displayed Stueckelberg–like kinetic terms for massive string modes [^1]. Stueckelberg symmetries are indeed deformations of ordinary massless gauge symmetries induced by “debris” that is brought about by interactions in the presence of symmetry breaking. As we shall see in the final part of these notes, a closer look at string amplitudes provides further evidence to this effect, although the order parameter, related to $\a^\prime$, lacks a dynamical origin in the current formulation of String Theory. In mentioning HS fields, I have stressed that they are very complicated in general. As we shall see in the next Section, they predated String Theory by a few decades, since their history started in the 1930’s [@majorana; @dfp]. Yet, they have long been lagging somewhat behind, so much so that basic ingredients like their free Lagrangian formulations could still receive new inputs during the last decade, while both the nature of their interactions and their actual meaning, let alone their subtle behavior when infra–red cutoffs (masses and/or a cosmological constant) are removed, remain largely mysterious. On the other hand, this framework can exhibit a signpost, the Vasiliev system [@vasiliev], which sits in some respects slightly ahead of those laid by String Theory. Lack of space forces me to leave it out of these notes, so that I can only mention that it describes the mutual interactions of infinitely many symmetric tensors in the presence of a parameter that cuts them off in the infra–red. The actual meaning of this infra–red cutoff is quite interesting in its own right, as stressed in [@fms]. The Vasiliev system has the looks of an effective description for the first Regge trajectory of the open bosonic string, in a regime where it should dominate the low–energy dynamics and in a language that extends the frame–like formulation of gravity. The two main sections that follow are devoted almost exclusively to free and interacting massless HS fields around flat space, the starting point for building massive counterparts in the presence of symmetry breaking. They focus on an extension of the metric–like formulation of gravity that allows a direct comparison with String Theory and begin with some cursory historical remarks that are tailored to the ensuing discussions. Sections \[sec:freesymmetric\] and \[sec:triplets\] discuss free irreducible HS fields and their reducible counterparts that emerge from String Theory in the $\a^{\, \prime} \to \infty$ limit, while Section \[sec:adsdeformations\] discusses free fields in (A)dS backgrounds, stressing role and meaning of their mass–like terms. Section \[sec:extcurrents\] discusses external currents and the van Dam–Veltman–Zakharov discontinuity, while Section \[sec:cubicstringlessons\] illustrates some definite lessons that String Theory already provides for HS interactions. Section \[sec:beyondcubic\], strictly speaking, would not belong to the Erice course, but it was part of the subsequent talks and I took the freedom to include it since it contains, in my opinion, very interesting clues on the behavior of HS interactions around flat space. \[sec:free\] This Section is devoted to some properties of free symmetric HS tensors, fields of the form $\varphi_{\mu_1 \ldots \, \mu_s}$. Lack of space forces me to confine my attention to Bose fields and to leave out altogether HS fields of mixed symmetry [@old; @labastida; @nlocmixed; @mixed], although they constitute the vast majority of string excitations. The material is meant to prepare the grounds for the discussion of string interactions of Section 3, so that I review Fronsdal’s constrained formulation of irreducible HS fields [@ff] and its minimal unconstrained extension [@fs1], their links with the geometry of free HS fields and the reducible “triplets” [@triplet] that emerge from String Theory in the $\a^\prime \to \infty$ limit. \[sec:freesymmetric\] It is commonly stated that the theory of HS fields originates from the works of Dirac, Fierz and Pauli [@dfp]. Their key results were the elegant DFP physical state conditions ( + m\^2 ) \_[\_1 … \_s]{} &= 0  ,\ \^[ \_1]{} \_[\_1 \_2 … \_s]{} &= 0  , \[DFP1\]\ \^[ \_1\_2]{} \_[\_1 \_2 … \_s]{} &= 0 for symmetric tensors $\varphi_{\mu_1 \ldots \ \mu_s}$ of arbitrary rank and their counterparts for symmetric tensor–spinors $\psi_{\mu_1 \ldots \ \mu_s}$ of arbitrary rank ( - m ) \_[\_1 … \_s]{} &= 0  ,\ \^[ \_1]{} \_[\_1 \_2 … \_s]{} &= 0  , \[DFP2\]\ \^[ \_1]{} \_[\_1 … \_s]{} &= 0  , whose first members are the Klein–Gordon and Dirac equations, and the first evidence that a naive coupling to Electromagnetism faces unexpected problems. For any given $s$, the first member of the sets and defines the mass–shell, the second eliminates unwanted “time” components, and finally the last confines the available excitations to irreducible multiplets. Equations of this type lie at the heart of Quantum Field Theory and are key building blocks for most current and past constructions. The logic behind their prominence surfaced long ago in Wigner’s work [@wigner]: eqs.  and , while based on *finite–dimensional non–unitary* representations of the Lorentz group, lead upon quantization to one–particle states filling *unitary* irreducible representations of the Poincaré group. The emergence of higher spins, however, dates back to an older and long neglected paper of Majorana [@majorana]. Although this work lies somewhat outside the scope of these notes, I would like to comment briefly on its striking content, since after all the Erice School is rightfully named after this great Italian scientist. Majorana tried to formulate a Dirac–like equation free of the then disturbing negative–energy solutions, and in doing so was led to consider *infinite–dimensional unitary* representations of the Lorentz group SO(3,1). Let me stress that in 1932 this subject was largely unfamiliar in Mathematics, let alone in Theoretical Physics. Majorana readily found out that, in the massive case, his equation was not describing a single type of particle, but rather a whole “trajectory” of states with M(s) \~   . He was thus anticipating, to some extent, both Regge’s idea [@regge] and its eventual realization in the Veneziano amplitude [@veneziano], and thus in String Theory altogether [@stringtheory], by over thirty years! On the mathematical side, he had come across an oscillator realization of the $so(3,1)$ algebra, well before Mathematics touched upon a tool that only in the 1970’s became relatively familiar in Theoretical Physics [@gunaydin]. The paper is striking, since its techniques are a sublimation of elementary facts of angular momentum theory that were then common tools in Atomic Physics. His analysis rests on oscillator generalizations of the Dirac $\gamma$–matrices, but it was perhaps too early even for him to appreciate that he was actually building, at no extra cost, unitary representations of the larger $so(3,2)$ algebra. The extension reflects, in modern terms, a familiar fact: if one combines $\gamma^\mu$’s and $\gamma^{\mu\nu}$’s, the resulting commutators generate indeed $so(3,2)$. In fact, without being fully aware Majorana had stumbled upon Dirac’s “singletons” [@diracsing], which lie at the heart of beautiful sequels by Flato, Fronsdal and others [@flatofronsdal], and more recently of both the AdS/CFT correspondence [@adscft] and Vasiliev’s theory [@vasiliev]. I will end here these parenthetic remarks, referring the interested reader to [@majorana] and to some interesting recent papers [@casalbuoni] on Majorana’s 1932 contribution, where the subsequent history of the subject is also addressed. The discovery of the Rarita–Schwinger theory [@RS], a development that played a key role in Theoretical Physics toward the end of the last Century as a key building block of supergravity [@supergravity], followed rather closely the Schrödinger–like DFP conditions and , but World War II forced the scientific community to leave aside these themes for a while. As a result, a systematic search for action principles, the next natural task along this line, had to await about four decades, since in the 1950’s and 1960’s, when a sizable activity on fundamental questions resumed, the focus was rather on the $S$–matrix, with an eye on important issues in current algebra. Rather than spurring progress in a constructive sense, key developments on HS that took place in the 1960’s actually spotted further unexpected difficulties. In retrospect, these results were unveiling important information on the nature of HS interactions, but for a while their net effect was to force most of the community away from these themes. I am referring, in particular, to Weinberg’s 1964 argument on soft HS particles and the $S$ matrix [@weinb64], to the Coleman–Mandula theorem [@cm] and to the Velo–Zwanziger violation of causality in the presence of electromagnetic interactions [@vz]. These were then followed, after a few years, by the Aragone–Deser field–theory argument [@ad] on the conflict between “minimal”, two–derivative, gravitational couplings and HS fields around flat space, and by the Weinberg–Witten theorem [@ww; @porratiww], a result that points in the same direction but is again inspired by $S$–matrix techniques. I will return to these issues in Section \[sec:interacting\]. Some major developments that took place in Field Theory during the 1970’s were certainly not foreign to the renewed interest that ultimately led to action principles for massive symmetric Bose and Fermi HS fields [@SH]. These results were the starting point for the subsequent work by Fronsdal and Fang and Fronsdal [@ff], where gauge–invariant actions were finally recovered in the massless limit. Fronsdal arrived at a natural generalization of the Maxwell and linearized Einstein equations. He extended the familiar $s=1,2$ cases & A\_ - \_ A = 0  ,\ & \_ - ( \_ \_ + \_ \_ ) + \_ \_ \^[ ]{} = 0  , \[s12\] where the “prime” denotes a trace, that are invariant under the gauge transformations & A\_ = \_  ,\ & \_[ ]{} = \_ \_ + \_ \_ , \[gauge12\] to symmetric tensors $\varphi_{\mu_1 \ldots \, \mu_s}$ of arbitrary rank, arriving at the second–order equations \_[ \_1 … \_s]{} \_[ \_1 … \_s]{} - ( \_[\_1]{} \_[ \_2 … \_s]{} + … ) + ( \_[\_1]{} \_[\_2]{} \^[ ]{}\_[ \_3 … \_s]{} + … ) = 0  , \[ss\] where the combinations ${\cal F}_{\, \mu_1 \ldots \, \mu_s\,}$ are often referred to as Fronsdal’s tensors. As these are natural generalizations of eqs. , one would expect that eqs.  leave way to the natural spin–$s$ gauge transformations \_[ \_1 … \_s]{} = \_[\_1]{} \_[ \_2 … \_s]{} + … , \[gaugess\] but as we shall see shortly this is not quite the case. In eqs.  and , a number of terms needed to complete the symmetrizations are inevitably left implicit. In order to streamline the ensuing discussion, it is very convenient to resort to a notation introduced in [@fs1] that may be regarded as a counterpart, for these symmetric tensors, of the language of forms. Lorentz labels are thus eliminated altogether, so that a generic spin–$s$ field is simply denoted by $\varphi$, and in this fashion eqs.  and take the simpler forms - + \^[ 2]{} \^[ ]{} = 0  , \[ssred\] and =  . \[gaugessred\] Let me stress that this choice is not merely a shorthand. Not only does it simplify greatly all expressions at hand, in fact, but it also embodies the key properties that are needed to manipulate them at will. To this end, one only needs a handful of rules, that are discussed in some detail, for instance, in [@fms]: ( \^[ p]{} )\^[ ]{} & =   \^[ p-2]{}  + 2 \^[ p-1]{}  + \^[ p]{} \^[ ]{} ,\ \^[ p]{} \^[ q]{} & =    \^[ p+q]{}  ,\ ( \^[ p]{}  ) & =  \^[ p-1]{}   + \^[ p]{}   ,\ (\^[ k]{}) & =   \^[ k-1]{} +  \^[ k]{} ( )  ,\ ( \^k )\^[ ]{} & =   \^[ k-1]{}  +  \^k \^[ ]{} . \[etak\] Suffice it to say, here, that in this notation every implicit symmetrization is meant to involve the *least* number of terms that are needed to effect it, and that the overall coefficients are always one. One can thus understand, for instance, why $\partial \, \partial = 2\, \partial^{\, 2}$, the counterpart of the vanishing of the squared exterior derivative: since derivatives commute, a naive symmetrization would overcount the terms actually needed by a factor two. Returning to the Fronsdal equations , their structure is clearly quite natural in view of eqs. , and yet their behavior under gauge transformations entails a little surprise, since [F]{} = 3 \^[ 3]{} \^[ ]{}  . \[gauge\_unconstr\] In other words, gauge invariance demands that $\Lambda^{\, \prime} = 0$, *i.e.* that the parameters be *traceless*. This condition first shows up for $s=3$ and is often referred to as *Fronsdal’s first constraint*. A second, subtler constraint, is also needed in order that the natural counterpart of the free Einstein Lagrangian, that up to partial integrations can be presented in the form = ( [F]{} -   [F]{}\^[ ]{} )  , be gauge invariant, even with traceless parameters. As we have seen, in this case ${\cal F}$ is gauge invariant, while up to total derivatives the variation of the “naked” $\varphi$ yields = - s ( -  \^[ ]{} )  , \[lags\] an expression that would vanish identically if a HS counterpart of the familiar Noether or Bianchi identity for the linearized version of Einstein’s gravity, \^ [R]{}\_ -  \_ [R]{} = 0  , were to hold. As it turns out, however, this is not the case, since in general -   [F]{}\^[ ]{} = -  \^[ 3]{} \^[ ]{} , so that for $s\geq4$ gauge invariant action principles obtain only if the gauge fields are subject to *Fronsdal’s second constraint*, $\varphi^{\, \prime\prime} = 0$, the condition that $\varphi$ be *doubly traceless*. The need for these constraints is a sign that Fronsdal’s formulation is incomplete, since eqs.  and do not accommodate the natural gauge symmetry , despite the fact that algebraic constraints certainly do not raise doubts on the nature of the actual propagating degrees of freedom. This problem was first bypassed in the 1990’s by the Dubna group [@dubna], adapting to free higher spins the BRST construction [@brst] of free String Field Theory [@sftheory], but at the price of introducing ${\cal O}(s)$ different fields for the case of a rank–$s$ tensor $\varphi_{\mu_1 \ldots \mu_s\,}$. A “minimal” solution was then obtained in [@fs1; @fms]. Its “compensator equations” [F]{} -  3 \^[ 3]{} = 0 \[ssredc\] involve, for $s \geq 3$, only one additional field, a spin–$(s-3)$ compensator $\alpha_{\mu_1 \ldots \, \mu_{s-3}}$, and the resulting extension ${\cal A}$ of Fronsdal’s ${\cal F}$ tensor is gauge invariant provided = \^[ ]{}  . \[deltacomp\] Actually, for $s=3$ the compensator made an early appearance in the work of Schwinger [@schwinger], as I was told some time ago by G. Savvidy. Schwinger’s books contain indeed a discussion, for $s=3$, of this additional field, which he also introduced in order to bypass Fronsdal’s first constraint but readily abandoned since it brings along higher derivatives. In retrospect, these higher derivatives are harmless since they affect non–propagating components, and we now understand how to trade them for a few more fields, whose number does not grow with the spin, as in [@lowder; @bgk]. An additional field, a Lagrange multiplier $\beta_{\mu_1 \ldots \, \mu_{s-4}}$ such that =  , is also present for $s \geq 4$ in the complete Lagrangians [@fs1; @fms] = ( [A]{} -   [A]{}\^) -   \^ + 3 [C]{}  , where = \^[ ]{} - 4 - \^[ ]{} is a gauge–invariant completion of the double trace of $\varphi$. The resulting Lagrangian equations for $\vf$ are \[boseeom\] & - 2 \^[ ]{} + \^[ 2]{} = 0 ,\ & \^[ ]{} - 4 - \^[ ]{} = 0 , where $\eta$ is the Minkowski metric and = - 2 (\^[ ]{} - 2 - ) is a gauge invariant completion of the Lagrange multiplier $\beta$. Let me stress that eqs.  generalize the linearized Lagrangian equation for gravity, \_ -  \_  [R]{} = 0  , and are thus more complicated than eq.  that is rather the counterpart for these systems of the simpler linearized non–Lagrangian equation \_ = 0  . A recursive argument [@fs1; @fms] shows that eqs.  can turned into = 0 = 0 \[nonlags\] once they are combined with their traces. Removing Fronsdal’s constraints via the compensator $\alpha$ (and the Lagrange multiplier $\beta$) is clearly eliminating an asymmetry between the first two cases, $s=1,2$, and the HS ones. Interestingly, this also opens a small and yet instructive window on the geometrical nature of HS constructions. After all, eqs.  admit a different but equally familiar presentation in terms of the Maxwell curvature $F_{\mu\nu}$ and, as we have already anticipated, of its counterpart ${\cal R}_{\mu\nu}$ for the linearized Einstein theory. There is a key difference between these objects, however: for $s=2$ one arrives at the linearized Riemann curvature tensor ${\cal R}_{\mu\nu\rho\sigma}$ via an intermediate object, the Christoffel connection, that in the linearized case reads \_ = ( \_ [\^]{}\_ + \_ [\^]{}\_ - \^ \_)  . As a result, while the Maxwell curvature contains one derivative of $A_\mu$, its $s=2$ counterpart is bound to contain two derivatives of $\varphi_{\mu\nu}$. It is convenient to focus here on a variant of the usual Riemann curvature that is *symmetric* under interchanges within its two groups of indices, although it is merely a combination of conventional Riemann tensors on account of the cyclic identity. Yet, when proceeding to arbitrary values of $s$, this choice has the virtue of suggesting a recursive construction where the linearized $\Gamma_{\mu, \, \nu\rho}$ leaves way to a tower of $s-1$ Christoffel–like connections, $\Gamma_{\mu_1 , \, \nu_1 \ldots \, \nu_s}$, …, $\Gamma_{\mu_1 \ldots \, \mu_{s-1} , \, \nu_1 \ldots \, \nu_s}$. Each of these quantities is obtained from combinations of derivatives of the previous member of the list in such a way that, in the gauge transformations, all $\mu$ indices fall on the gauge parameter. These expressions carry, by construction, 1, …, $s$-1 derivatives of the original field, and the next member of the sequence is a spin–$s$ analogue of the linearized $s=2$ curvature in its doubly–symmetric incarnation, a tensor ${\cal R}_{\mu_1 \ldots \, \mu_s , \, \nu_1 \ldots \, \nu_s}$ built from combinations of $s$ derivatives of the original $\varphi_{\mu_1 \ldots \mu_s\,}$ that is *gauge invariant* under transformations like , albeit with unconstrained parameters. One can also show that these curvature tensors possess the additional symmetry \_[\_1 … \_s , \_1 … \_s]{} = (-1)\^[ s]{}  [R]{}\_[\_1 … \_[s]{} , \_1 … \_s]{}  , which is already visible in the $s=1,2$ cases. The beautiful construction that I just sketched, due to de Wit and Freedman [@dwf], marks the real entry point of HS geometry into the game. Since, as I have stressed, the relevant HS curvatures are by construction invariant under unconstrained gauge transformations of the type , in this fashion the authors came short of bypassing Fronsdal’s constraints. They did not make this further step, however, since they insisted on the apparently natural condition that Fermi or Bose equations contain, respectively, one or two derivatives. As a result, while they could recognize that the Fronsdal tensor ${\cal F}$ is related to the second Christoffel–like connection according to \_[\_1 … \_s]{} \~ \^[\_1 \_2]{}  \_[\_1 \_2 , \_1 … \_s]{}  , all in all their HS geometry appeared for a while somewhat remote from the actual free dynamics. The compensator equations do better in this respect, as they should since they embody an unconstrained gauge symmetry. Francia and I obtained these equations directly [@fs1], before formulating the compensator theory, but it is instructive to retrace the argument in this fashion, starting from the compensator equations. Let us do it in some detail for the simplest case, $s=3$, in which the non–Lagrangian equations reduce to \_ = 3 \_\_\_  . One can formally solve this equation for $\partial_\mu \, \alpha$, obtaining \_ =    [[F]{}\^[ ]{}]{}\_  , \[nloc1\] or directly for $\alpha$, obtaining =      , \[nloc2\] where I have resorted again to the shorthand “prime” notation to indicate traces. This is tantamount to turning the compensator equation into a pair of distinct *non–local* equations for $\varphi_{\mu\nu\rho}$ alone. By construction, the end result is invariant under the unconstrained gauge transformations , but it is clearly *not* unique, since eq.  leads to \_ [F]{}\_ - ( \_ \_ [[F]{}\^[ ]{}]{}\_ + \_ \_ [[F]{}\^[ ]{}]{}\_ + \_ \_ [[F]{}\^[ ]{}]{}\_) = 0  , \[s3nl1\] while eq.  leads to the alternative gauge invariant non–Lagrangian equation \_ [F]{}\_ -  \_ \_ \_ [F]{}\^[ ]{} = 0  . \[s3nl2\] Having more than one option might seem problematic, but as we shall see it is actually instrumental for the very consistency on the non–local formulation, and to begin with the reader can verify that eqs.  and can be turned into one another, so that they are nicely equivalent in the absence of external sources. Interestingly, factoring out the inverse d’Alembertian present in eq.  according to \_ = 0  , \[s3nl12\] leaves room for a total of four derivatives within the square brackets, as many as the divergence of the spin–3 curvature would contain, according to the preceding discussion. All in all, one is thus led to the conclusion that the spin–3 non–local equations of motion, and thus a fortiori their local compensator counterparts, are equivalent to  \^[ \_1\_2]{} \^[ \_3]{}  [R]{}\_[\_1\_2\_3,\_1\_2\_3]{} = 0  , \[nlocs3\] with ${\cal R}$ the spin–3 de Wit–Freedman curvature, symmetric in its two sets of Lorentz labels and antisymmetric under an overall interchange of them. This can be simply verified, and clearly eq.  is a very natural and satisfactory spin–3 counterpart of the Maxwell equation, albeit *not* a Lagrangian equation. This pattern extends to arbitrary spin, as can be appreciated via an inductive argument that rests on the sequence of integro–differential operators \^[(n+1)]{} = [F]{}\^[(n)]{} +    [F]{}\^[(n)]{} -    \^[(n)]{}  , \[iteration\] whose Bianchi identities take the form \^[(n)]{}  -     [[F]{}\^[(n)]{}]{}[ ’]{}  =  - ( 1 + )   \^[\[n+1\]]{}  , \[bianchin\] where $\vf^{[n+1]}$ denotes the $(n+1)$–st trace of the HS field $\vf$. For these combinations, the original compensator equation translates into the sequence \^[(k)]{}  =  (2 k  +  1)    \^[\[k-1\]]{} , \[compensk\] where $\alpha^{[k-1]}$ denotes the $(k-1)$–fold trace of $\a$, whose members have thus the virtue of involving successive traces of the compensator. For any given $s$, these traces are simply not available for $k$ sufficiently large, so that fully gauge invariant equations obtain after completing $\left[\frac{s}{2}\right]$ iterations. Taking into account for any $s$ the first of these equations, one is finally led to an infinite family of fully gauge invariant *non–local* counterparts of the Maxwell and Einstein equations, & s=2n+1 :  \_\^[;\_1 …\_s]{} = 0  ,\ & s = 2n :      [R]{}\^[\[n\];\_1 …\_s]{} =0  . \[nloceqs\] While these non–Lagrangian equations involve natural geometric structures for these systems in the simplest combinations that one could a priori conceive, the corresponding Lagrangian equations rest not only on traces of the first gauge–invariant ${\cal F}^{(k)}$ but also on higher members of the set, obtained via further iterations of eq. . We shall return shortly to this point, after introducing some additional tools. Before closing this section, let me stress briefly that what I have said extends to Fermi fields carrying a set of totally symmetric vector labels, up to standard complications brought about by $\gamma$–matrices [@ff; @dubna; @fs1; @fms; @lowder]. Tensors and spinor–tensors of mixed symmetry, on the other hand, rest on a richer theory that originates from the work of Labastida and others [@old; @labastida] and would require a detailed discussion. I will content myself with stating that one can extend to this case the non–local equations [@nlocmixed] and, up to some subtleties that are discussed in detail in [@mixed], the compensator formulation as well. The subject is very interesting and is crucial in order to come eventually to terms with String Theory, but lack of space forces me to leave it out here. \[sec:triplets\] A key question that these notes are meant to address is what these notions about higher spins are teaching us about the massive string excitations. At the free level, the link between String Theory and HS is relatively simple, and yet it is interesting and quite instructive. In short, String Theory favors reducible representations, so that it drops somehow the last DFP condition of eq. . For symmetric Bose fields this entails the replacement of the Fronsdal Lagrangian or its compensator extensions with an interesting system of three fields that is often called a “triplet” [@triplet; @dubna; @fs1; @st1; @francia10] in the literature. The three fields are a rank–$s$ tensor $\varphi_{\mu_{\,1} \ldots \, \mu_{\,s}}$, the dynamical field in this context, and two additional tensors of ranks $s-1$ and $s-2$, $C_{\mu_{\,1} \ldots \, \mu_{\,s-1}}$ and $D_{\mu_{\,1} \ldots \, \mu_{\,s-2}}$ that disappear on shell but are nonetheless crucial to attain an *unconstrained* gauge symmetry. The $s=1$ case is degenerate, and the system reduces to the description of an electromagnetic potential in the Nakanishi–Lautrup presentation. In the index–free notation, the triplet equations read &  =   C  ,\ &  -   D  =  C  ,\ & D  =  C  , \[triplet\] and are invariant, as anticipated, under the *unconstrained* gauge transformations &  =    ,\ & C  =  ,\ & D  =   . If one is ready to take for granted that $C$ and $D$ disappear on shell, which can be shown in a few steps, the triplet equations clearly boil down to the first two DFP conditions . The triplet equations follow from the elegant BRST [@brst] formulation of String Field Theory [@sftheory]. Its field equation, | = 0  , rests on the BRST operator ${\cal Q}$ of world–sheet reparametrizations, whose nilpotency in the critical dimension brings about the unconstrained gauge symmetry | = [Q]{} |  , together with gauge–for–gauge counterparts. However, while the string BRST operator is only nilpotent in the critical dimension, its contracted version that emerges in the formal limit $\a^\prime \to \infty$ is *identically* nilpotent [@dubna], and as a result the triplet system is indeed gauge invariant in any number of dimensions. An alternative presentation of eqs.  obtains if $C$ is eliminated algebraically via the second of them, reducing the first to = \^[ 2]{} ( \^[ ]{} - 2 D )  . \[tripletred\] One can simply select the spin–$s$ excitations of this system [@fs1], adding to eqs.  the further condition \^[ ]{} - 2 D =   \[hstriplet\] that enforces on shell, in a gauge invariant fashion, the last DFP condition of eq. . This step brings about a compensator $\alpha$ with the gauge transformation of eq. , while eq.  turns readily into eq. . Following [@bgk], one can also perform this reduction at the Lagrangian level, at the price of a few additional fields whose number, however, does not grow with the spin. Another interesting question concerns the non–local formulation of the triplet systems, based on $\varphi$ alone. One can build it systematically [@francia10] and the end results,  \~ [R]{}\^[ \_1 … \_s ; \_1 … \_s]{}   [R]{}\_[\_1 … \_s ; \_1 … \_s]{}  , \[lagnlred\] are strikingly natural non–local counterparts of the Maxwell Lagrangian. \[sec:interacting\] We can now turn to HS interactions, and to begin with I would like to comment on some classic results of the 1960’s and 1970’s that revealed unexpected difficulties. \[soft\] The first and most striking of these results was obtained by S. Weinberg in 1964 [@weinb64]. It concerns the behavior that S–matrix amplitudes for a single *massless* spin–$s$ particle and a number of scalar particles would exhibit in the presence of long–range HS interactions. Referring to fig. 1, the “soft” limit $q \to 0$ of one such amplitude, here denoted by $A_s$, would expose long–range HS effects and would be dominated by poles proportional to $p_i \cdot q$ originating from scalar propagators close to their mass shells. Moreover, Lorentz–invariant vertices for two scalars and a spin–$s$ particle giving rise to long–range HS effects would lead, in the “soft” limit, to $s$ powers of $p_i$ contracting the HS polarization tensor. For $q \to 0$, an $N$–point $A_s$ amplitude would thus behave asymptotically as A\_s(1,…,N)  \~ A(1,…,N-1) \_[i=1]{}\^[N-1]{}  , \[wei64\] where $A(1,\ldots\,,N-1)$ denotes a more conventional $(N-1)$–point scalar amplitude, the residue of the pole. Weinberg’s key observation was that, although an *on–shell* Abelian gauge transformation (q) (q)  +  i q (q) of the spin–$s$ polarization tensor would seem to cause to no problems due the “soft” nature of the momentum $q$, the pole would make the end result *finite* in the limit. Consequently, HS amplitudes might well be *not* invariant under these shifts, with the disastrous consequence that unphysical polarizations might *not* decouple. Unless, of course, some conditions are met, and a clear way around does exists, but only for $s=1$ or $2$. In the former case, the limiting contribution would be indeed proportional to the total electric charge of the (ingoing – outgoing) scattered particles, which ought to vanish in $S$–matrix amplitudes, while in the latter it would be proportional to the sum of the scalar momenta weighted by the corresponding “gravitational charges”. Hence, with gravity coupling universally, for $s=2$ consistency would translate into momentum conservation and therefore would be automatically guaranteed. Conversely, the requirement of consistency for long–range gravitational interactions emerges, in this fashion, as a reason for their universality. On the other hand, for $s \geq 3$ eq.  would involve polynomials in the momenta, that clearly would not add up to zero in general. All in all, generic amplitudes with a single *massless* HS particle would thus seem inconsistent, but there are possible loopholes. First, as noticed in [@weinb64], if more powers of q were present in the vertices, they could compensate the pole, making the terms in eq.  vanish individually in the “soft” limit. Vertices of this type, however, would not mediate long–range HS effects, and actually can often be eliminated altogether by field redefinitions. Moreover, a recent analysis [@tar] that takes into account the explicit form of the cubic HS vertices of [@st; @mmr] revealed that the problem presents itself *only* when the exchanged particles have spins *lower* than the external ones. We shall return to this important point in Section \[sec:beyondcubic\]. However, the reader should already appreciate that, with these amplitudes alone, the possible long–range effects of HS particles would be confined to their own world, a scenario that was nicely foreseen by Porrati in [@porratiww]. The next classic result, the Coleman–Mandula theorem [@cm], is a formal $S$–matrix argument that has had a wide impact over the years. One of its key assumptions, however, the presence of finitely many HS excitations below any given mass level, makes it not directly applicable to the HS constructions one is after, whose gauge algebras bring about infinitely many massless gauge fields. The theorem does provide a rationale for the failure of naive attempts involving finitely many HS gauge fields, but I cannot help finding it less palatable and instructive than Weinberg’s analysis. The Velo–Zwanziger argument [@vz] can be regarded as a refinement of the old work of Fierz and Pauli [@dfp]. It concerns the loss of causality that HS fields tend to manifest in the presence of minimal couplings to uniform electromagnetic fields, and in this context String Theory has proved strikingly instructive. Indeed, Argyres and Nappi [@argnap] showed long ago that for the massive $s=2$ mode of the open string, that in some respects behaves like a HS field, the actual coupling to uniform electromagnetic fields, which rests on an exact conformal field theory and can thus be thoroughly analyzed, is free from such pathologies. The crux of the matter lies in the peculiar highly non–linear contact terms that are present in String Theory, and a recent extension of their analysis shows that all symmetric tensors of the first Regge trajectory work exactly in the same way, without any loss of causality [@prs]. However, the minimal couplings of these states to Electromagnetism fade out when the string tension or the HS masses tend to zero, so that HS interactions loose their infrared cutoff. The result resonates again with Weinberg’s argument, and reinforces the feeling that HS fields do not like to be involved in *long–range* interactions mediated by *low–spin* particles. The last classic results that I would like to describe point again in the same direction, albeit in different ways. They both indicate that HS fields *do not* interact with gravity around flat space with two–derivative minimal couplings as their low–spin counterparts. The Weinberg–Witten argument [@ww] compares two ways of determining the effect of rotations on matrix elements of the energy–momentum tensor $\langle f | T_{\mu\nu}(q) | i \rangle$ between two spin–$s$ states in three dimensions. If $T_{\mu\nu}$ is assumed to behave as a Lorentz tensor, a property that it possesses when it is gauge invariant, and to involve only two derivatives, the argument shows that the matrix element vanishes for any type of *massless* HS particles, be they fundamental or composite, and the authors present a similar argument for spin–one currents. The assumption is hardly a weak one, since already for $s=2$ the energy–momentum tensor $T_{\mu\nu}$ is *not* gauge invariant, while already for $s=1$ spin–one currents are also not gauge invariant. Still, a recent refinement of [@ww] took into account the subtle behavior of $T_{\mu\nu}$ but reached nonetheless the same conclusion [@porratiww]. The problem has to do, once more, with the number of derivatives in HS couplings: if more derivatives were present, the matrix elements would vanish as $q \to 0$ and the arguments of [@ww; @porratiww] would not apply. Finally, the Aragone–Deser argument [@ad] rests on an explicit attempt to construct a standard gravitational coupling for a spin–$\frac{5}{2}$ field. The attempt fails, since the computation generates Weyl tensors, which cannot be compensated by the Einstein–Hilbert terms and thus constitute a genuine obstruction to the gauge symmetry. Similar results obtain for $s> 5/2$ Fermi fields or for $s>2$ Bose fields. This, however, is only true around flat space, and remarkably in the 1980’s Fradkin and Vasiliev [@Fradkin] could show that the obstruction disappears in the presence of a cosmological term $\lambda$. In other words, in the presence of a cosmological term, and up to the cubic order, HS fields do allow conventional–looking gravitational couplings, albeit with an important proviso: they are inevitably accompanied by higher–derivative partners that, for dimensional reasons, bring along negative powers of $\lambda$. As a result, the solution to the Aragone–Deser problem found by Fradkin and Vasiliev has a singular behavior in the limit of a flat background. The Aragone–Deser argument has had a profound influence on subsequent developments, which have been strongly biased toward the search for “minimal–like” HS interactions in the presence of a cosmological term. The main outcome of this research, the Vasiliev system [@vasiliev], is remarkable and confusing at the same time. It is remarkable as a mathematical realization of free–differential algebras with non–polynomial scalar couplings and as an example of classically consistent HS interactions, and not surprisingly starts to play an important role in the AdS/CFT correspondence. It is confusing, however, since it points to intriguing dynamical effects. Vasiliev’s fields are in fact a scalar, a vector and infinitely many symmetric tensors, one for each rank $s\geq2$, all valued in (anti)symmetric representations of a Chan–Paton group [@cp]. Hence, they are in one–to–one correspondence with the first Regge trajectory of the *open* bosonic string. These fields have *minimal*–like interactions, albeit around (A)dS backgrounds, and moreover the spin–2 singlet is naturally associated with gravity, while in String Theory gravity has to do with *closed*, rather than open, strings. This circumstance would seem to suggest that a Cabibbo–like mixing takes place between closed and open fields, when they become degenerate in mass in the tensionless limit $\a^{\,\prime}\to \infty$. This point was raised in [@fms] and clearly deserves further attention. For all these reasons, and mainly to stress the key role of non–minimal interactions, both in the lectures that I have presented at Erice and in subsequent talks, I have kept a (slightly embellished) slide of my Strings 2009 presentation, where I had elaborated on the apparent tension between the Fradkin–Vasiliev minimal–like couplings and two recent results. One of these is due to Metsaev [@metsaev], who completed the analysis that in the 1980’s opened our first window on HS interactions [@goteborg], and concerns the light–cone classification of cubic HS vertices in flat space, while the other is the subsequent scaling analysis of Boulanger, Leclercq and Sundell [@bls]. Briefly stated, Metsaev’s analysis showed that cubic light–cone vertices exist around flat space in general, albeit with *lower bounds* on the number of derivatives that depend on the collection of fields involved. For example, the $s=2$ couplings of a spin–$s$ field involve at least $2s-2$ derivatives. Boulanger, Leclercq and Sundell proposed a scaling argument that pulls out of the Fradkin–Vasiliev couplings the dominant contributions (what I called “seeds” in [@rome09]), and in so doing recovered explicit examples of covariant (as opposed to light–cone, as in Metsaev’s analysis) cubic flat–space vertices. The existence of the scaling limit indicates that terms containing lower numbers of derivatives are really dressings due to the (A)dS background, no more no less than the mass–like terms that I shall revisit shortly. A later section will illustrate how String Theory sheds light on the issue. In the low–tension limit of its amplitudes, in fact, one can identify whole chains of covariant terms that fill Metsaev’s list, whose high–derivative couplings naturally forego the previous restrictions. Although this section is mostly devoted to the behavior around flat space, it actually begins with a brief discussion of how mass–like terms emerge in (A)dS backgrounds, starting from the free equations of Section \[sec:free\]. I then turn to external currents and the van Dam–Veltman–Zakharov discontinuity [@vdvz], whose origin can now be related to special types of HS gauge symmetries that are allowed in dS backgrounds. External currents are also a first step toward more general Lagrangian couplings: the HS currents of free scalar fields, for instance, embody key information on their possible HS interactions and on their behavior at high energies [@bjm]. Section \[sec:cubicstringlessons\] is then devoted to disk amplitudes for the open bosonic string and their neat lessons for the cubic interactions of symmetric Bose fields, and actually also of symmetric tensor–spinors [@st]. The end result of this analysis provides further clues that massive string spectra draw their origin from broken HS symmetries. The section ends with a brief review of some recent results of Taronna [@tar] that address explicitly the subtle behavior of the higher–point functions for massless HS fields in flat space. \[sec:adsdeformations\] As we have seen, the classic Aragone–Deser problem [@ad] is signalled by terms involving the Weyl tensor, which disappear in conformally flat backgrounds. This is an important exception that I cannot refrain from touching upon, since it underlies HS dynamics in the presence of a cosmological constant. The corresponding maximally symmetric backgrounds are (A)dS, rather than Minkowski, space times, where ordinary derivatives leave way to covariant derivatives such that V\_ = ( g\_ V\_ - g\_ V\_) . \[covariantdev\] For definiteness, here the sign is tailored to the dS case, $g$ is the background metric and $L$, the dS radius, is related to the cosmological constant in $D$ dimensions according to =  . \[adsradius\] Corresponding results for AdS obtain replacing $L^2$ with $-L^2$. The commutator brings about modifications of both ${\cal F}$ and its unconstrained completion ${\cal A}$ that make them compatible with the deformed gauge transformations =  . These modifications involve the replacement of ${\cal F}$ with \_L = [F]{} + {  +  2 g \^[’]{} } , where $D$ denotes the number of space-time dimensions and = - + \^[ 2]{} \^[ ]{} \[fl\] is now built in terms of the covariant derivatives . Notice that the deformed non–Lagrangian equations of motion, that are now $\cF_L=0$, involve *mass–like terms*. Mass–like terms, by no means ordinary mass terms, since they are precisely as needed for gauge invariance. At the same time the gauge transformations of $\alpha$ and $\beta$ become &= \^[ ]{}  ,\ &=  . Consequently, the $AdS$ deformations of $\cA$ and $\cC$ read \_L & = \_L - 3 \^[ 3]{} -  g ,\ \_L & = \^[ ]{} - 4 - \^[ ]{} , and are rather simple, while \_L & = - {2 \^[ ]{} - - 2\ & + ( 2 \^[ ]{} + 2 g \^[ ]{} + ) } is more involved. Let me stress that the actual form of the mass–like terms, which are mere curvature–induced dressings of the flat–space kinetic terms, is *not* unique, but depends on the ordering chosen for the covariant derivatives. This is strikingly clear in the $s=1$ case, since the Maxwell equation admits the two naively different presentations & A\_ - \_[ ]{} \^[ ]{} A\^[ ]{} = 0  ,\ & A\_ - \_[ ]{} A -    A\_[ ]{} = 0  , which are actually equivalent in view of the commutator . \[sec:extcurrents\] External currents are responsible for the simplest interactions in Field Theory, but much can be learned from them notwithstanding. Static sources give rise in this fashion to Coulomb’s law or its HS generalizations, while generic sources encode important information on the number of propagating degrees of freedom. One can start from the general Lagrangian field equations for symmetric tensors and extend them allowing for the presence of an external current ${\cal J}$, so that they become & [A]{} - [A]{}\^[ ]{} [ + \^2 [B]{}]{} = [J]{}  ,\ & \^[ ]{} [ - (2 + ) [B]{}]{} = 0  ,\ & \^ [ - 4 - \^[ ]{}]{} = 0  ,\[localspins\] where the unconstrained gauge symmetry demands that ${\cal J}$ be conserved. In the familiar $s=1$ case, in momentum space these equations reduce to -  p\^2 A\_ +  p\_ p A  =  [J]{}\_ , and the scalar product of this expression with the conserved current ${\cal J}_\mu$, \_0(D,1) - p\^2 [J]{}\^ A\_ =   [J]{}\^ [J]{}\_ ,\[exch1flat\] makes it possible, so to speak, to invert the kinetic operator without the need for gauge fixing, so that the “current–exchange amplitude” identifies the residue of the pole. With an eye to the ensuing discussion, the subscript of ${\cal E}$ is meant to emphasize that the amplitude is computed in flat space, while its two arguments reflect the space–time dimension and the spin of the exchanged field. Notice that the square of the conserved current in rests on $D-2$ independent contributions, as many as the independent polarizations of a massless spin–1 field in $D$ dimensions. Before discussing the actual solution of the system , let us elaborate briefly on the important lessons that this analysis has in store for the non–local formulation. In principle, the non–local formulation could be derived directly, integrating out both the compensator $\alpha$ and the Lagrange multiplier $\beta$ along the lines of [@francia10], but current exchange amplitudes provide an alternative path that is conceptually important. The issue is selecting a non–local Lagrangian equation that yields the same physical effects, and thus the same current exchange amplitudes, as the local system . Complete details can be found in [@fms], but here we can give some clues that a proper non–local Lagrangian does indeed exist and is actually unique by taking a closer look at the $s=3$ case. In Section \[sec:freesymmetric\] we identified two alternatives for the $s=3$ field, and now we can see explicitly how external currents can distinguish the pseudo–differential operators ${{\cal F}^{(1)}}$ and ${\widetilde{\cal F}^{(1)}}$ of eqs.  and . Their differences reflect themselves in their Bianchi identities, that in index–free notation read & -   [[F]{}\^[(1)]{}]{}\^ = 0  ,\ & \^[(1)]{} -   \^[(1) \^]{} = 0  , where the first is a special case of eq.  while the second can be computed directly and coincides with the Bianchi identity for the Fronsdal operator ${\cal F}$. As a result, in the two cases one is led to define the divergence–free Einstein–like tensors \^[(1)]{} = [[F]{}\^[(1)]{}]{} -   [[F]{}\^[(1)]{}]{}\^ \^[(1)]{} = [\^[(1)]{}]{} -   [\^[(1) \^]{}]{}  , and the corresponding Lagrangian equations, where they are sourced by conserved external currents, then yield the current exchange amplitudes \^[(1)]{} = [J]{} -  [J]{}\^\^and \_0(D,3) [J]{} \^[(1)]{} = [J]{} -  [J]{}\^\^ . $\widetilde{\cal G}^{(1)}$ is the correct Einstein–like tensor in this case, since one can verify that only the last expression identifies a traceless and transverse $s=3$ current. The detailed analysis in [@fms] provides a clear explanation of how the Bianchi identities for the ${\cal F}^{(n)}$ operators of Section \[sec:freesymmetric\] determine the proper non–local Einstein–like tensors. For any given $s$, for $n$ large enough these operators become gauge invariant, while their Bianchi identities loose memory of the term proportional to $\vf^{\,\prime\prime}$ and simply relate divergences to gradients of their traces. Therefore, combining traces of the first gauge–invariant ${\cal F}^{(n)}$ with corresponding powers of $\frac{\partial^{\,2}}{\Box}$ one ought to arrive at a non–local analogue of the gauge–invariant $\cA$ tensor of Section \[sec:freesymmetric\]. As shown in [@fms], this is indeed true, and moreover there is a unique choice to this effect. Consequently, the non–local Lagrangians make full use of the independent gauge–invariant operators that are built by eq. , as we had stated in Section \[sec:freesymmetric\], and the end results yield the same current exchange amplitudes as the local theories. Here we can content ourselves with a brief discussion of the final result, $$\begin{aligned} && {\cal E}_0(D,s) \, \equiv \, {\sum_{n=0}^{N}\ \rho_n(D-2,s) \ \frac{s!}{{n!\; (s-2n)!\; 2^n}}\ {\cal J}^{[n]}\cdot {\cal J}^{[n]}} \ , \label{exchangeD} \\ && { \rho_{n+1}(D,s)\ =\ -\ \frac{\rho_n(D,s)}{ D+2(s-n-2)}} \ , \qquad \rho_0(D,s)\, =\, 1 \ , \label{recursionD}\end{aligned}$$ which can be justified noticing that the recursive definition of the $\rho_n$ makes the sum transverse and traceless, as in the preceding $s=3$ example. The first non–trivial case corresponds to $s=2$, where eq.  reduces to \_0(D,2) = T\^ T\_  -    ( [T\^]{}\_)\^2 \[exchangeDs2\] where, abiding to standard conventions, we have called $T_{\mu\nu}$ the corresponding current. Let me stress that the amplitude ${\cal E}_0(D,s)$ depends on the spin $s$ of the HS field determining the exchange and on the number $D$ of space–time dimensions. If one insists on currents that are *conserved* in $D$ dimensions, it is possible to adapt the analysis to *massive* spin–$s$ fields. A harmonic dependence on an internal circle coordinate suffices in fact to introduce masses *à la* Stueckelberg, so that massive exchanges driven by currents that are still *conserved* in $D$ dimensions and lack internal components obtain replacing $D$ with $D+1$ in eq. . The difference between the exchanges for any given value of $s$ computed in $D+1$ and $D$ dimensions, \_0(D,s) = [\_[n=0]{}\^[N]{}  \^[\[n\]]{}\^[\[n\]]{}]{} , thus encodes the generalization of the discontinuity originally found by van Dam, Veltman and Zakharov [@vdvz] for $s=2$, that can be computed starting from eq. . This phenomenon has a very instructive counterpart in the presence of a cosmological constant $\lambda$ that deforms Minkowski space to (A)dS backgrounds depending on its (negative)positive sign. Starting from the deformed equations of Section \[sec:adsdeformations\] and focussing on the dS case, for $s=2$ the resulting massive exchange for a generic mass $M$, computed some time ago in [@hp], reads[^2] \_(D,2) = T\^ T\_ -  ([T\^]{}\_)\^2  , \[elambda2\] where the relation between $L$ and the cosmological constant is given in eq. . Notice that this exchange is a rational function of $ML$ with a simple pole determined by the condition (ML)\^2 = D-2  . \[pmass\] The pole in eq.  is a manifestation of an interesting phenomenon, a shortening that occurs in dS representations precisely for the value of $(ML)^2$ [@shortening]. Indeed, as first noticed in [@pgauges], when eq.  holds the theory acquires the “partial gauge symmetry” \_ = \_\_ +  g\_  , \[pgauget\] with $\zeta$ a scalar parameter. There is a marked difference between eq.  and the standard massless gauge symmetry \_ = \_ \_[ ]{} + \_ \_[ ]{}  , with $\xi_\mu$ a vector parameter, since the coupling to a generic conserved energy–momentum tensor is *not* compatible with the second term in eq.  unless the trace of $T_{\mu\nu}$ vanishes identically. This fact also brings about the vDVZ discontinuity, since the limiting values of the rational function in eq.  for $ML\to 0$, that identifies the flat massless case, and for $ML \to \infty$, that identifies the flat massive case, differ in compliance with Liouville’s theorem. In other words, the vDVZ discontinuity in flat space is somehow a reflection of the partial gauge symmetry that emerges, for real fine–tuned values of $(ML)^2$, in dS space. More details can be found in [@fms], where eq.  was extended to symmetric tensors of arbitrary rank. The first terms, drawn from [@fms] but adapted to the index–free notation of Section \[sec:freesymmetric\], read &[[E]{}\_(D,s)]{} [=]{} [J]{}\_s + [[g [J]{}\_s\^]{} 2(- )]{} [(ML)\^2+2(-)(ML)\^2-2(-3)]{} \ [+]{} &[[g\^2 [J]{}\_s\^]{} 8]{} [(ML)\^4+8(ML)\^2(-)+12([52]{}-)\_2 ([52]{}-)\_2\[(ML)\^2-2(-3)\]\[(ML)\^2-6(-4)\]]{} \[spinsds\]\ [+]{}  &[[g\^3 [J]{}\_s\^]{} 48]{} [(ML)\^6-(ML)\^4 (18-77)+92(ML)\^2([72]{}-)\_2+120([52]{}-)\_3([52]{}-)\_3\[(ML)\^2-2(-3)\]\[(ML)\^2-6(-4)\] \[(ML)\^2-10(-5)\]]{}\ [+]{} &[…]{}  , where ${\cal J}_s^{\;[2,3]}$ are higher traces of ${\cal J}_s$, $g$ is the background dS metric, $\zeta = \frac{D}{2}+s$ and (a)\_n=a(a+1)…(a+n-1) are Pochhammer symbols. In the dS case the poles lie precisely at real values of $ML$ where partially massless gauge transformations involving terms like the second in eq. , not protected by gradients and thus incompatible with generic conserved currents, emerge. For a rank–$s$ tensor there are in fact $s-1$ partially massless points, for which (ML)\^2=(s-1-r)(D+s+r-4) , where $r=0,\ldots,s-2$, while $r=s-1$ corresponds to the more familiar massless point. For all $s\geq 2$ a first pole thus appears at $r=s-2$, and for $s>3$ others lie at alternate values of $r$ below it. For instance, for $s=3$ the relevant value is $r=1$, since for $r=0$ all terms present in the partially massless gauge transformation contain gradients of the parameter and are thus manifestly compatible with conserved currents. One can derive eq.  starting from a $D+1$–dimensional flat Minkowski space and performing a radial reduction [@radial], a generalization of the polar decomposition of Euclidean space. Among the technical complications discussed in [@fms], I would like to mention here that the end result was recast in a form that depends manifestly on $(ML)^2$ via interesting identities for the generalized hypergeometric functions $_3F_2$, more complicated cousins of the familiar $_2F_1$ [@whiwat]. Let me stress that a close analogy exists between the singular behavior of dS exchanges at partially massless points and the massless limit of the more familiar flat–space exchanges in the presence of non–conserved currents. For instance, in the Proca theory a non–conserved current would lead to the exchange ( p\^[ 2]{} + m\^[ 2]{} ) J A = J J - (p J)\^2  , whose massless limit is singular precisely because a current that is not conserved conflicts with the emerging gauge symmetry. Following [@vdvz], we insisted on conserved currents also in the massive case, and this allowed a direct comparison between massive and massless exchange amplitudes. A further step along these lines was made by Bekaert, Joung and Mourad in [@bjm]. The key idea of their work was to apply the current–exchange formula to the *dynamical* Noether HS currents of a complex scalar field $\Phi$. These can be defined, as in [@bbvd], via the Wigner function J(x,v)= |(x+i v) (x - i v) , \[scalarHScurrent\] where the spin–$s$ term is combined with $s$ powers of the vector $v$. At the same time, the strength of the interactions with generic HS symmetric tensors can be characterized via a single dimensionful coupling, that by analogy with String Theory I will call $\alpha^{\, \prime}$ in the following, and via an infinite number of dimensionless couplings $a_k$, or rather via a *coupling function* a(z)=\_r  a\_r  , \[couplingfunction\] that in String Theory would be dominated by an exponential. Classic results in the theory of orthogonal polynomials allow one to combine the current–exchange amplitude , a pair of HS currents as in eq.  and the coupling function into a compact formula for the interaction between complex scalars and infinitely many massless HS fields. In four dimensions the result is strikingly neat, and for the $\vf+\vf \to \vf+\vf$ amplitude reads [A]{}\^[(s)]{} = - & \ & \_[1]{}(p\_[1]{}) \_[2]{}(p\_[2]{}) \_[3]{}(p\_[3]{}) \_[4]{}(p\_[4]{})  , \[bjmformula\] where $s$, $t$ and $u$ are the familiar Mandelstam variables, since the sum contributing to the current exchange can be related to the Chebyshev polynomials [@whiwat]. This beautiful expression has a number of interesting lessons in store. For one matter, it is a consistent four–scalar amplitude involving the exchange of infinitely many *massless* HS particles. Moreover, the detailed discussion in [@bjm] shows that, in principle, a soft behavior at high energies can be attained working only with (infinitely many) symmetric fields, provided the coupling function tends to zero for large negative real values of its argument. In String Theory the essential singularity of $a(z)$ may be held ultimately responsible for the presence of lower Regge trajectories, since a soft behavior for the conjugate amplitude $\vf+\bar{\vf} \to \vf+\bar{\vf}$ would also demand that $a(z)$ tend to zero for large positive real arguments. Therefore, as stressed in [@bjm], in the present setting a soft behavior for all conjugate amplitudes would require that the coupling function $a(z)$ tend to zero at infinity in the complex plane. This is a subtle condition, since Liouville’s theorem would then require the presence of singularities in the finite plane, which in their turn would signal in general an extended nature for the objects involved. There is clearly more to be understood here, and other intriguing properties will show up in the ensuing discussion. \[sec:cubicstringlessons\] I can now turn to the scattering amplitudes of the open bosonic string involving its massive HS excitations. As we shall see, they contain a wealth of information on HS couplings [@st], and this analysis provides some further evidence for the long–held, but never fully quantified, expectation that String Theory describes a broken phase for HS gauge symmetries. This brings about a natural link between cubic couplings and Noether currents, with the potential of clarifying also the nature of the latter. The situation is particularly manageable for the symmetric tensors of the first Regge trajectory, whose cubic disk amplitudes can be simply computed from the path integral for the free action S\_P\[X,\]=- \_M d\^[2]{} \^[ab]{}\_[a]{} X\^ \_[b]{} X\_\[Polyakov\] in the presence of special external currents. Eq.  is in fact the remnant of the bosonic string action after a complete gauge fixing in the critical dimension, while the vertex operators of the first Regge trajectory, combinations of an exponential with powers of $\partial X^{\, \mu}$, can all be recovered as Taylor coefficients in the “symbols" $\xi^{\, \mu}$ of (,p,) =  . \[vertex\_general\] Up to a measure factor that can be ascribed to ghost fields, tree–level correlation functions of these types of vertices on the disk are then tantamount to a gaussian path integral in the presence of the boundary currents $$J(\sigma, p, \xi)\,=\,\sum_{i\,=\,1}^N\left(p_{\,i}\,\delta^{\,2}(\s-\s_i)\, - \, \frac{{\xi}_{\,i}}{\sqrt{2\a^{\,\prime}}} \ \partial_\sigma \, \delta^{\,2}(\s-\s_i)\right)\ .\label{curr}$$ All these amplitudes, and three–point functions in particular, can be extracted from S\^[open]{}\_[j\_1j\_n]{}=\_[\^[n-3]{}]{} dy\_4&dy\_[n]{} |y\_[12]{}y\_[13]{}y\_[23]{}|  \ & \_[j\_1]{}(\_1)\_[j\_2]{}(\_2)\_[j\_3]{}(\_3)\_[j\_n]{}([y]{}\_n)Tr(Ł\^[a\_1]{}Ł\^[a\_n]{})  , \[diskamplitude\] where the trace is a Chan–Paton factor [@cp] and the integrals are computed along the real axis, so that the interior of the disk corresponds to the upper half–plane. For brevity, we have left implicit all momenta and $\xi$–variables for the $n$ vertex operators in . Standard field theory techniques then imply that correlation functions of the vertex operators, determined by the two–dimensional Green function, take the form \_[j\_1]{}(\_1) \_[j\_n]{}([y]{}\_n)& =\ &  ,\[Gen3\] where the $y_i$ denote the locations of the punctures along the real axis and $y_{ij}=y_i-y_j$. Eq.  would seem to violate basic tenets of String Theory, since the dependence on the location of the punctures ought to disappear from three–point functions, up to a measure factor compensating exactly the one introduced by ghosts in eq. . Of course, this property ought to hold solely for on–shell external states satisfying the Virasoro conditions, so that a closer look into the matter is necessary before coming to definite conclusions. The independent Virasoro conditions, (L\_0-1)|&= 0  ,\ L\_1|&= 0  ,\ L\_2|&= 0 translate directly into the DFP conditions of Section 2, as the reader can simply verify, and considering three external states such that $$\begin{aligned} -p_{\,1}^{\,2}\,=\,&\frac{n_1-1}{\a^{\,\prime}}\ ,& -p_{\,2}^{\,2}\,=\,&\frac{n_2-1}{\a^{\,\prime}}\ ,& -p_{\,3}^{\,2}\,=\,&\frac{n_3-1}{\a^{\,\prime}}\ ,\end{aligned}$$ the three–point correlation functions take the form ||\^[n\_1]{} | |\^[n\_2]{}||\^[n\_3]{}\  . \[cubicstring\] Apparently, this expression still depends on the location of the punctures. However, on account of the Virasoro conditions, the spins of the external states should be correlated to their mass levels, so that in eq.  one should only retain terms containing $n_1$ powers of $\xi_1$, $n_2$ powers of $\xi_2$ and $n_3$ powers of $\xi_3$. It is then pleasing to verify how, for this class of terms, the dependence on the $y_i$ disappears altogether, so that one is finally left with {  + } \[Zphys\]  . In this expression the brackets are merely signs: they reflect the flip symmetry [@cp] of string amplitudes, which is implemented by projective–disk amplitudes and plays a crucial role whenever external states of odd spin are present. Eq.  should be translated into the language of Field Theory, taking into account that $s$ powers of the $\xi_i$ are meant to correspond to a spin–$s$ field of the first Regge trajectory, a rank–$s$ symmetric tensor. Alternatively, one would like to associate to the external fields generic polarization tensors rather than simply powers of the $\xi_i$’s, and this is effected by a $\star$ – product, a convenient procedure to multiply pairs of series. Thus, if A()= \_[k=0]{}\^A\_[\_1 … \_k]{} \[Axi\] and B()= \_[k=0]{}\^B\_[\_1 … \_k]{}  , then A B = \_[k=0]{}\^  A\^[\_1 … \_k]{} B\_[\_1 … \_k]{} , which is indeed reconstructing generic polarization tensors from the special low–rank tensors determined by the generating function . The standard presentation of this $\star$ – product is A B = . ( ) A() B() |\_[==0]{}  , while a second presentation, A B =  (U)B(i U) , where $\tilde{A}$ denotes the Fourier transform of $S$ with respect to the symbols, follows inserting into the first the definition of $A$ via an inverse Fourier transform. More details on how these results lead directly to the general cubic couplings .\_[1]{}(p\_[1]{},[\_]{}p\_[31]{}) \_[2]{}(p\_[2]{},+[\_]{}p\_[23]{}) \_[3]{}(p\_[3]{},p\_[12]{}) |\_[=0]{} \[cubicHS\] can be found in [@st]. This expression is to be computed at $\xi=0$, $p_{ij}=p_i-p_j$ and the notation is as in eq. , so that for example the first factor is \_[k=0]{}\^\_[\_1 … \_k]{}(p\_[1]{})  , where derivatives act to the right. For traceless $\vf$ fields there is no ordering ambiguity. It is instructive to expand eq.  for the first few cases, [A]{}\^\_[0-0-s]{}&=()\^[s]{}\_[1]{}\_[2]{} \_[3]{}p\_[12]{}\^[s]{} ,\ [A]{}\^\_[1-1-s]{}&=()\^[s-2]{}s(s-1) A\_[1]{}A\_[2]{}\^[…]{}p\_[12]{}\^[s-2]{}\ &+()\^s\ &+()\^[s+2]{}A\_[1]{} p\_[23]{}A\_[2]{}p\_[31]{} p\_[12]{}\^[s]{} , \[11s\] the first of which corresponds to the Wigner currents used in [@bjm]. Notice that the $0-0-s$ amplitude in does not retain any memory, so to speak, of the fact that for $s>1$ the $\vf$ fields are *massive* in the open bosonic string. Indeed, if all momenta were those of massless fields it would be automatically invariant under the gauge transformation up to Virasoro or DFP conditions, which would imply $p_i \cdot p_j = 0$ for all $i$ and $j$. On the other hand, even with all massless momenta the first line of the $1-1-s$ amplitude, which contains its lowest–derivative terms, would be incompatible with the spin–$s$ gauge symmetry. This feature of cubic vertices is along the lines of what we have stressed in preceding sections: in the free string spectrum, mass generation *à la* Stueckelberg rests precisely on terms that interactions feed into the quadratic level, consistently with the fact that a conventional gauge symmetry is recovered if these are left aside. This is what happens in the limit $\a^\prime \to\infty$, *i.e.* in the limit of vanishing string tension. Hence, if a breaking phenomenon were really at work in String Theory, contributions from higher orders ought to pop up amidst cubic couplings. This phenomenon is strikingly visible at the cubic level: once the non–invariant terms, that carry along higher powers of the order parameter $\frac{1}{\sqrt{\a^\prime}}$ for the breaking, are taken apart, they leave behind a conventional gauge symmetry. Unfortunately, $\a^\prime$ lacks a dynamical origin in our present formulation of String Theory, while a minor subtlety is that all this occurs up to divergences and traces in these S–matrix amplitudes. Still, any gauge variations of the cubic vertices that are proportional to $p_{i}^{\,2}$ signal the mere need for deformed gauge transformations, up to trivial redefinitions, since any non–linear gauge variations of the kinetic terms that could in principle compensate them would be inevitably proportional to the free field equations. Let us therefore remove from the cubic vertices all terms like the first in eq.  and focus, for all values of $s_1$, $s_2$ and $s_3$, on contributions that become gauge invariant on shell when the masses of the external states are removed, The end result, rather rewarding and strikingly simple, is captured by the expression [@st] [^3] \^= e\^[ ]{}   \_[1]{}(p\_[1]{},\_[1]{}) \_[2]{}(p\_[2]{},\_[2]{}) \_[3]{}(p\_[3]{},\_[3]{})|\_[\_[i]{}=0]{}  , \[cubicmassless\] so that it rests on the operator = {&(\_[\_[3]{}]{}p\_[12]{})+(\_[\_[1]{}]{}p\_[23]{})\ & +(\_[\_[2]{}]{}p\_[31]{}) }  , \[Gamma\] which commutes with gauge transformations (in this notation, with terms of the form $p_i \cdot \xi_i$), up to Virasoro or DFP conditions. There is therefore a unique term containing $s_1+s_2+s_3$ powers of momenta, proportional to p\_[23]{}\^[s\_1]{} \_1(p\_1)  p\_[31]{}\^[s\_2]{} \_2(p\_2)  p\_[12]{}\^[s\_3]{} \_3(p\_3)  , together with other groups of terms containing lower numbers of derivatives, whose structure is best captured by an equivalent expression for $\cA$, \^=e\^[ ]{}  & \_[1]{}(p\_[1]{},\_[1]{} p\_[23]{}) \[cubicmassless2\]\ & . \_[2]{}(p\_[2]{} ,\_[2]{} p\_[31]{}) \_[3]{}(p\_[3]{} ,\_[3]{}p\_[12]{}) |\_[\_[i]{}=0]{} , where $p_{ij}=p_i-p_j$, the expression is to be evaluated at ${\xi_{\,i}\,=\,0}$ and =  . \[GammaG\] Notice that $\cG$ reduces indeed by two units the total number of momenta and removes at the same time three $\xi_i$’s, one from each of the fields. Therefore if $s_3$, say, is the smallest of the three spins, the operation can be repeated at most $s_3$ times, leading precisely to the removal of $2 \,s_3$ of the original momenta. In particular, for three spin–1 external states eqs.  and yield both the three–derivative vertex displayed in Schwarz’s 1982 review [@cp] and the cubic Yang–Mills vertex. Similarly, for three spin–2 external states they yield three groups of terms with six, four and two derivatives, the last of which is the standard cubic Einstein vertex. As a result, eqs.  and involve terms with overall powers of momenta ranging between $s_1+s_2+s_3-2 \,{\rm min} \{s_i\}$ and $s_1+s_2+s_3$, precisely as in Metsaev’s light–cone classification of cubic HS interactions [@metsaev].[^4] Infinitely many gauge–invariant three point functions are tantamount to infinitely many conserved currents that generalize the Wigner function of eq. , [J]{}\^(&x;)= (i \_ )  \[bosecurents\]\ &\_[1]{}(x\_1 i ,\_1 i \_[2]{}) \_[2]{}(x\_2 i , \_2 i \_[1]{})|\_[\_i=0]{} , where $\partial_1$, $\partial_2$ and $\partial_{12}$ are shorthands for $\partial_{x_1}$, $\partial_{x_2}$ and their difference, and where at the end of the computation one is to set $x_1$ and $x_2$ equal to $x$ and $\zeta_i=0$. In [@st], where the reader can find more details, we also presented an educated guess for the counterpart of eq.  for the superstring, where leaves way to bosonic and fermionic currents. Notice that the spin–2 currents built from a pair of spin–$s$ fields contain at least 2$s$-2 derivatives, and therefore *are not* the energy–momentum tensors of the spin–$s$ fields implied by the Lagrangians of Section \[sec:free\]. At this stage we have really left String Theory to return to a field theory framework, albeit in an apparently incomplete fashion, since the starting point was provided by cubic on–shell string amplitudes. Complete cubic vertices that are gauge invariant up to the Fronsdal equations or their compensator analogs can be recovered completing eq. , as in [@st; @mmr], with two types of terms that we ignored so far. These involve divergences and traces of the three external fields, and actually fall into the combinations \_[ij]{} = ( 1 + \_[\_[i]{}]{}\_[\_[j]{}]{} ) i [D]{}\_j - p\_j \_[\_[i]{}]{} \_[\_[j]{}]{}\_[\_[j]{}]{}  , with i [D]{}\_i = p\_i \_[\_[i]{}]{} -  p\_i \_i  \_[\_[i]{}]{} \_[\_[i]{}]{} the de Donder operator. In terms of the ${\cal H}_{ij}$ the complete cubic vertex in Fronsdal’s constrained form reads [^5] \^ &= e\^[ ]{} \_[1]{}(p\_[1]{},\_[1]{}) \_[2]{}(p\_[2]{},\_[2]{}) \_[3]{}(p\_[3]{},\_[3]{})  , where the normal ordering moves de Donder operators to the right in products with $\partial_{\xi_{\,i}}$’s. This is the point where the construction of [@st] meets the independent work of Manvelyan, Mkrtchyan and Ruhl [@mmr], who obtained the same cubic vertices from field theory considerations. Strictly speaking, however, in [@st] we were more precise with the Chan–Paton symmetry [@cp], which makes full use in general of the sign option in eqs.  or . \[sec:beyondcubic\] In [@tar] Taronna brought the discussion forward by some stimulating steps that can potentially lead to further progress. He moved from the observation that the operator of eq. , whose exponential generates the whole sequence of cubic vertices, comprises two different groups of familiar terms. The first, (\_[\_[1]{}]{}\_[\_[2]{}]{})(\_[\_[3]{}]{}p\_[12]{})+(\_[\_[2]{}]{}\_[\_[3]{}]{}) (\_[\_[1]{}]{}p\_[23]{})+(\_[\_[3]{}]{}\_[\_[1]{}]{})(\_[\_[2]{}]{}p\_[31]{})  , in the usual tensor notation would become the familiar cubic Yang–Mills vertex ([YM3]{})\_ = \_ p\_[12, ]{} + \_ p\_[23, ]{} + \_ p\_[31, ]{}  , \[YM3\] while the rest is a combination of scalar–scalar–vector vertices. Let me stress that the key property of $\Gamma$ is that it commutes with gauge transformations, up to Virasoro or DFP conditions. As we have seen, this makes it possible, for an action combining the quadratic terms of Section \[sec:freesymmetric\] with the vertex , to be gauge invariant under transformations that are generally deformed. While String Theory uses the exponential of $\Gamma$, it would seem possible to consider more general functions of this operator in HS constructions, altering the relative weights of terms involving different numbers of derivatives. This is another point that deserves further investigation. A key question addressed in [@tar] is what should be the counterparts, for four and higher–point functions, of the exponent $\Gamma$ that builds cubic HS vertices. For brevity, let me leave aside the scalar–scalar–vector terms present in $\Gamma$ to concentrate on the four–point functions built starting solely from . To begin with, these are four–point Yang–Mills S–matrix amplitudes, but interestingly they suffice to define a whole class of HS amplitudes. In order to display them, let us first recast the Yang–Mills amplitudes in terms of *independent* group–theory factors, or independent combinations of Chan–Paton factors, that a reasoning along the lines of [@cp] would guarantee to also allow the factorization of S–matrix amplitudes involving generic HS states. The end result combines exchange amplitudes in pairs of channels and quartic Yang–Mills interactions and thus, amusingly, has *planar duality*, albeit in a very simple form where the poles are manifest in pairs of channels. It is uniquely determined once, making use of the Jacobi identity, one sorts out the contributions to the Yang–Mills amplitude that carry along independent products of structure constants. One of these contributions takes the form A\_(p\_1) A\_(p\_2) A\_(p\_3) A\_(p\_4) \^  [tr]{} ()  , \[suamplitude\] where the dynamical amplitude has been combined with its Chan–Paton factors and, schematically, \^  =   ([YM3]{}   [YM3]{} + [YM4]{})\^  + ([YM3]{}   [YM3]{} + [YM4]{})\^  . \[su\] Here $s$ and $u$ are Mandelstam variables, while the ${\rm YM4}$’s are the usual quartic Yang–Mills couplings. Notice that the decoupling of longitudinal polarizations from translates into an invariance of under [on–shell Abelian gauge transformations]{} of the type A\_i(p\_i)   A\_i(p\_i)  +  i p\_i (p\_i)  . \[shifts\] This key property of Yang–Mills tree amplitudes must hold individually for the two groups of terms, since they carry along independent group theory factors, and in the complete amplitude the $(s,u)$ contribution is accompanied by a similar $(s,t)$ contribution, A\_(p\_1) A\_(p\_2) A\_(p\_3) A\_(p\_4) \^  [tr]{} ()  . \[stamplitude\] The spin–$s$ amplitudes proposed in [@tar] include \_[\_1 …\_s]{}(p\_1) \_[\_1 …\_s]{}(p\_2) \_[\_1 …\_s]{}(p\_3) \_[\_1 …\_s]{}(p\_4)  (su)\^[s-1]{} \_[k=1]{}\^s \^[\_k\_k\_k\_k]{}  , \[4sampl\] that should be properly dressed with the same Chan–Paton factors as in eq.  and then combined with corresponding $(s,t)$ contributions. Notice that these amplitudes are invariant by construction under the HS counterparts of eq. , Abelian on–shell transformations that in the index–free notation of Section \[sec:free\] would read (p\_i) (p\_i) + i p\_i (p\_i)  . Moreover, the crucial power of $su$ has a two–fold effect: it guarantees that only single poles in $s$ and $u$ be present and raises the spin of the exchanged particles to an *odd* value, $2s-1$, which is higher than the spin of the external states for $s>1$. Notice that the end result possesses the same symmetries as the Chan–Paton factor of eq. , which it should carry in a construction along the lines of [@cp], where the flip symmetries would be directly correlated to the spins of the massless external states. Tensoring the independent contributions to four–point Yang–Mills amplitudes has thus led to an infinite family of HS amplitudes that by construction do not suffer from Weinberg’s factorization problem. How can this be the case? In order to answer this question, Weinberg’s argument was extended in [@tar] to generic soft amplitudes, taking into account the explicit form of the vertex . The conclusion is that *no problems are encountered if the intermediate spins are not lower than the external ones*, precisely as was the case for eq. . The very existence of the amplitudes , however, raises another important question. Namely, once the cubic vertices are given, what prevents one from building with them “bad” amplitudes, with internal spins that are too low to comply with Weinberg’s factorization? The answer proposed in [@tar] is very interesting and controversial at the same time. It makes use of non–local quartic vertices, in fact the minimal set of them needed to barely remove the “bad” exchange amplitudes built from any given pair of vertices . In this fashion, one would be left with a subset of non–vanishing amplitudes, all invariant by construction under the Abelian gauge transformations that signal the decoupling of unphysical polarizations and thus complying, a fortiori, with the extension of Weinberg’s soft emission argument. As we have already mentioned, however, in this fashion external HS particles would interact only via exchanges of other HS particles with spins not lower than the external ones! This state of affairs brings to one’s mind the peculiar superstring states considered in [@dima], and it would be interesting to elaborate further on the possible links between the two situations. Unfortunately, as noticed in [@tar], with finitely many spins the removal of the unwanted exchanges is apparently in conflict with unitarity, which is somehow the Coleman–Mandula argument haunting back again. A proper understanding of the singular behavior of infinitely many HS fields once all infrared cutoffs, be they masses or a cosmological term, are removed, will thus require more effort. Clearly, further progress along these lines will be highly instructive. Acknowledgments {#acknowledgments .unnumbered} =============== This work was supported in part by Scuola Normale Superiore, by INFN, by the MIUR-PRIN contract 2009-KHZKRX and by the ERC Senior Grant n. 226455, “Supersymmetry, Quantum Gravity and Gauge Fields” (SUPERFIELDS). I am very grateful to Mirian Tsulaia and Jihad Mourad for pleasant and instructive collaborations on some of the topics reviewed in these notes, and to Dario Francia, Andrea Campoleoni and Massimo Taronna for long and fruitful collaborations and also for a number of useful suggestions on the manuscript, whose final form also benefitted from some input from Karapet Mkrtchyan. Finally, I would like to thank Prof. Noriaki Kitazawa of Tokyo Metropolitan University for a careful “non–expert reading” that helped me greatly to streamline the presentation. [99]{} G. Veneziano, Nuovo Cim.  [**A57** ]{} (1968) 190-197. See, for instance: M. B. Green, J. H. Schwarz and E. Witten, “Superstring Theory”, 2 vols. (Cambridge Univ. Press, Cambridge, UK, 1987); J. Polchinski, “String theory”, 2 vols. (Cambridge Univ. Press, Cambridge, UK, 1998); B. Zwiebach, “A first course in string theory”, (Cambridge Univ. Press, Cambridge, UK, 2004); K. Becker, M. Becker and J. H. Schwarz, “String theory and M-theory: A modern introduction”, (Cambridge Univ. Press, Cambridge, UK, 2007); E. Kiritsis, “String theory in a nutshell”, (Princeton Univ. Press, Princeton, NJ, USA, 2007). “Higher-Spin Gauge Theories”, Proceedings of the First Solvay Workshop, held in Brussels on May 12-14, 2004, eds. R. Argurio, G. Barnich, G. Bonelli and M. Grigoriev (Int. Solvay Institutes, 2006). This collection contains some contributions closely related, in spirit, to the present work, including: M. Bianchi and V. Didenko, arXiv:hep-th/0502220; D. Francia and C. M. Hull, arXiv:hep-th/0501236; N. Bouatta, G. Compere and A. Sagnotti, arXiv:hep-th/0409068; X. Bekaert, S. Cnockaert, C. Iazeolla and M. A. Vasiliev, arXiv:hep-th/0503128; A. Sagnotti, E. Sezgin and P. Sundell, arXiv:hep-th/0501156. Other recent reviews include: D. Sorokin, AIP Conf. Proc. [**767**]{} (2005) 172 \[hep-th/0405069\]; D. Francia and A. Sagnotti, J. Phys. Conf. Ser.  [**33**]{} (2006) 57 \[arXiv:hep-th/0601199\]; A. Fotopoulos and M. Tsulaia, arXiv:hep-th/0805.1346; A. Campoleoni, Riv. Nuovo Cim.  [**033** ]{} (2010) 123-253. \[arXiv:0910.3155 \[hep-th\]\]; A. Sagnotti, arXiv:hep-th/1002.3388; X. Bekaert, N. Boulanger and P. Sundell, arXiv:1007.0435 \[hep-th\]. W. Siegel, Nucl. Phys. [**B263**]{} (1986) 93; W. Siegel and B. Zwiebach, Nucl. Phys. [**B263**]{} (1986) 105; T. Banks and M. E. Peskin, Nucl. Phys. [**B264**]{} (1986) 513; M. Kato and K. Ogawa, Nucl. Phys. [**B212**]{} (1983) 443; N. Ohta, Phys. Rev. [**D33**]{} (1986) 1681, Phys. Lett. [**B179**]{} (1986) 347, Phys. Rev. Lett.  [**56**]{} (1986) 440 \[Erratum-ibid.  [**56**]{} (1986) 1316\]; A. Neveu, H. Nicolai and P. C. West, Nucl. Phys. [**B264**]{} (1986) 573; A. Neveu and P. C. West, Nucl. Phys. [**B268**]{} (1986) 125; E. Witten, Nucl. Phys. [**B268**]{} (1986) 253. M. Bianchi, J. F. Morales and H. Samtleben, JHEP [**0307**]{} (2003) 062 \[arXiv:hep-th/0305052 \[hep-th\]\]; N. Beisert, M. Bianchi, J. F. Morales and H. Samtleben, JHEP [**0402**]{} (2004) 001 \[hep-th/0310292\], JHEP [**0407**]{} (2004) 058 \[hep-th/0405057\]. J. M. Maldacena, Adv. Theor. Math. Phys.  [**2** ]{} (1998) 231-252. \[hep-th/9711200\]; E. Witten, Adv. Theor. Math. Phys.  [**2** ]{} (1998) 253-291. \[hep-th/9802150\]; S. S. Gubser, I. R. Klebanov, A. M. Polyakov, Phys. Lett.  [**B428** ]{} (1998) 105-114. \[hep-th/9802109\]. For a review see: O. Aharony, S. S. Gubser, J. M. Maldacena, H. Ooguri, Y. Oz, Phys. Rept.  [**323** ]{} (2000) 183-386. \[hep-th/9905111\]. E. Majorana, Nuovo Cim.  [**9**]{} (1932) 335. P. A. M. Dirac, Proc. Roy. Soc. Lond.  [**155A**]{} (1936) 447; M. Fierz, Helv. Phys. Acta [**12** ]{} (1939) 3-37; M. Fierz and W. Pauli, Proc. Roy. Soc. Lond.  [**A173**]{} (1939) 211. M. A. Vasiliev, Annals Phys.  [**190**]{} (1989) 59, Phys. Lett.  [**B238**]{} (1990) 305, Phys. Lett.  [**B243**]{} (1990) 378, Class. Quant. Grav.  [**8**]{} (1991) 1387, Phys. Lett.  [**B257**]{} (1991) 111, Phys. Lett.  [**B285**]{} (1992) 225, Fortsch. Phys.  [**52**]{} (2004) 702 \[arXiv:hep-th/0401177\];\ For a review see also: M. A. Vasiliev, Int. J. Mod. Phys.  [**D5**]{} (1996) 763 \[arXiv:hep-th/9611024\], arXiv:hep-th/9910096. D. Francia, J. Mourad and A. Sagnotti, Nucl. Phys.  [**B773**]{} (2007) 203 \[arXiv:hep-th/0701163\]; Nucl. Phys.  [**B804** ]{} (2008) 383-420. \[arXiv:0803.3832 \[hep-th\]\]. T. Curtright, Phys. Lett.  [**B165**]{} (1985) 304; C. S. Aulakh, I. G. Koh and S. Ouvry, Phys. Lett.  [**B173**]{} (1986) 284; C. S. Aulakh, I. G. Koh and S. Ouvry, Phys. Lett.  [**B173**]{} (1986) 284; S. Ouvry and J. Stern, Phys. Lett.  [**B177**]{} (1986) 335; W. Siegel and B. Zwiebach, Nucl. Phys.  [**B282**]{} (1987) 125; W. Siegel, Nucl. Phys.  [**B284**]{} (1987) 632; J. M. F. Labastida and T. R. Morris, Phys. Lett.  [**B180**]{} (1986) 101; J. M. F. Labastida, Phys. Rev. Lett.  [**58**]{} (1987) 531; J. M. F. Labastida, Nucl. Phys.  [**B322**]{} (1989) 185. X. Bekaert and N. Boulanger, Commun. Math. Phys.  [**245**]{} (2004) 27 \[arXiv:hep-th/0208058\], Phys. Lett.  [**B561**]{} (2003) 183 \[arXiv:hep-th/0301243\]; P. de Medeiros and C. Hull, JHEP [**0305**]{} (2003) 019 \[arXiv:hep-th/0303036\]. A. Campoleoni, D. Francia, J. Mourad and A. Sagnotti, Nucl. Phys.  [**B815**]{} (2009) 289 \[arXiv:hep-th/0810.4350\], Nucl. Phys.  [**B828**]{} (2010) 405 \[arXiv:hep-th/0904.4447\]. C. Fronsdal, Phys. Rev.  [**D18**]{} (1978) 3624; J. Fang and C. Fronsdal, Phys. Rev.  [**D18**]{} (1978) 3630. D. Francia and A. Sagnotti, Phys. Lett.  [**B543**]{} (2002) 303 \[arXiv:hep-th/0207002\], Class. Quant. Grav.  [**20**]{} (2003) S473 \[arXiv:hep-th/0212185\]; Phys. Lett.  [**B624**]{} (2005) 93 \[arXiv:hep-th/0507144\]. A. K. H. Bengtsson, Phys. Lett.  [**B182**]{} (1986) 321; M. Henneaux and C. Teitelboim, in “Quantum Mechanics of Fundamental Systems, 2”, eds. C. Teitelboim and J. Zanelli (Plenum Press, New York, 1988), p. 113; G. Bonelli, Nucl. Phys.  [**B669** ]{} (2003) 159-172. \[hep-th/0305155\]; A. K. H. Bengtsson, Class. Quant. Grav.  [**5**]{} (1988) 437; A. Fotopoulos and M. Tsulaia, Int. J. Mod. Phys.  [**A24**]{} (2009) 1 \[arXiv:0805.1346 \[hep-th\]\]. E. P. Wigner, Annals Math.  [**40** ]{} (1939) 149-204. T. Regge, Nuovo Cim.  [**14** ]{} (1959) 951. For a review see: M. Gunaydin, Lect. Notes Phys.  [**180** ]{} (1983) 192-213. P. A. M. Dirac, J. Math. Phys.  [**4** ]{} (1963) 901-909. C. Fronsdal, Rev. Mod. Phys.  [**37** ]{} (1965) 221-224; M. Flato, C. Fronsdal, Lett. Math. Phys.  [**2** ]{} (1978) 421-426; E. Angelopoulos, M. Flato, C. Fronsdal, D. Sternheimer, Phys. Rev.  [**D23** ]{} (1981) 1278. R. Casalbuoni, PoS [**EMC2006** ]{} (2006) 004. \[arXiv:hep-th/0610252 \[hep-th\]\]; X. Bekaert, M. R. de Traubenberg, M. Valenzuela, JHEP [**0905** ]{} (2009) 118. \[arXiv:0904.2533 \[hep-th\]\]; S. Esposito, \[arXiv:1110.6878 \[physics.hist-ph\]\]. W. Rarita, J. Schwinger, Phys. Rev.  [**60** ]{} (1941) 61. D. Z. Freedman, P. van Nieuwenhuizen and S. Ferrara, Phys. Rev. [**D13**]{} (1976) 3214; S. Deser and B. Zumino, Phys. Lett. [**B62**]{} (1976) 335. S. Weinberg, Phys. Rev.  [**135** ]{} (1964) B1049-B1056. S. R. Coleman, J. Mandula, Phys. Rev.  [**159** ]{} (1967) 1251-1256. G. Velo and D. Zwanziger, Phys. Rev.  [**186**]{} (1969) 1337, Phys. Rev.  [**188**]{} (1969) 2218; G. Velo, Nucl. Phys.  [**B43**]{} (1972) 389. C. Aragone and S. Deser, Phys. Lett.  [**B86**]{} (1979) 161, Nuovo Cim.  [**B57**]{} (1980) 33. S. Weinberg, E. Witten, Phys. Lett.  [**B96** ]{} (1980) 5. M. Porrati, Phys. Rev.  [**D78** ]{} (2008) 065016 \[arXiv:0804.4672 \[hep-th\]\]. L. P. S. Singh and C. R. Hagen, Phys. Rev.  [**D9**]{} (1974) 910, Phys. Rev.  [**D9**]{} (1974) 898. A. Pashnev and M. M. Tsulaia, Mod. Phys. Lett.  [**A12**]{} (1997) 861 \[arXiv:hep-th/9703010\]; Mod. Phys. Lett.  [**A13**]{} (1998) 1853 \[arXiv:hep-th/9803207\], C. Burdik, A. Pashnev and M. Tsulaia, Nucl. Phys. Proc. Suppl.  [**102**]{} (2001) 285 \[arXiv:hep-th/0103143\]; I. L. Buchbinder, A. Pashnev and M. Tsulaia, Phys. Lett.  [**B523**]{} (2001) 338 \[arXiv:hep-th/0109067\] arXiv:hep-th/0206026; X. Bekaert, I. L. Buchbinder, A. Pashnev and M. Tsulaia, Class. Quant. Grav.  [**21**]{} (2004) S1457 \[arXiv:hep-th/0312252\]; I. L. Buchbinder, V. A. Krykhtin and A. Pashnev, Nucl. Phys.  [**B711**]{} (2005) 367 \[arXiv:hep-th/0410215\]. C. Becchi, A. Rouet and R. Stora, Commun. Math. Phys.  [**42**]{} (1975) 127, Annals Phys.  [**98**]{} (1976) 287; I.V. Tyutin, preprint FIAN n. 39 (1975); E. S. Fradkin and G. A. Vilkovisky, Phys. Lett. [**B55**]{} (1975) 224; I. A. Batalin and G. A. Vilkovisky, Phys. Lett. [**B69**]{} (1977) 309; M. Henneaux, Phys. Rept.  [**126**]{} (1985) 1. J. S. Schwinger, Reading, USA: Addison-Wesley (1989) 306 p. (Advanced book classics series). D. Francia, Nucl. Phys. [**B796** ]{} (2008) 77-122. \[arXiv:0710.5378 \[hep-th\]\], Fortsch. Phys.  [**56** ]{} (2008) 800-808. \[arXiv:0804.2857 \[hep-th\]\]. I. L. Buchbinder, A. V. Galajinsky, V. A. Krykhtin, Nucl. Phys.  [**B779** ]{} (2007) 155-177. \[hep-th/0702161\]; I. L. Buchbinder, V. A. Krykhtin, A. A. Reshetnyak, Nucl. Phys.  [**B787** ]{} (2007) 211-240. \[hep-th/0703049\]; I. L. Buchbinder, A. V. Galajinsky, JHEP [**0811** ]{} (2008) 081. \[arXiv:0810.2852 \[hep-th\]\]. B. de Wit and D. Z. Freedman, Phys. Rev.  [**D21**]{} (1980) 358. A. Sagnotti and M. Tsulaia, Nucl. Phys.  [**B682**]{} (2004) 83 \[arXiv:hep-th/0311257\]. D. Francia, J. Phys. Conf. Ser.  [**222** ]{} (2010) 012002. \[arXiv:1001.3854 \[hep-th\]\]; Phys. Lett.  [**B690**]{} (2010) 90 \[arXiv:1001.5003 \[hep-th\]\]. M. Taronna, “Higher-Spin Interactions: four-point functions and beyond,” \[arXiv:1107.5843 \[hep-th\]\]. M. Taronna, “Higher Spins and String Interactions”, Master Thesis, arXiv:1005.3061 \[hep-th\]; A. Sagnotti, M. Taronna, Nucl. Phys.  [**B842** ]{} (2011) 299-361. \[arXiv:1006.5242 \[hep-th\]\]. R. Manvelyan, K. Mkrtchyan, W. Ruhl, Nucl. Phys.  [**B836** ]{} (2010) 204-221. \[arXiv:1003.2877 \[hep-th\]\]; Phys. Lett. [**B696**]{} (2011) 410 \[arXiv:1009.1054 \[hep-th\]\]; R. Manvelyan, K. Mkrtchyan, W. Ruhl, M. Tovmasyan, Phys. Lett.  [**B699** ]{} (2011) 187-191. \[arXiv:1102.0306 \[hep-th\]\]. For a review, see: K. Mkrtchyan, arXiv:1101.5643 \[hep-th\]. P. C. Argyres, C. R. Nappi, Phys. Lett.  [**B224** ]{} (1989) 89. M. Porrati, R. Rahman, A. Sagnotti, Nucl. Phys.  [**B846** ]{} (2011) 250-282. \[arXiv:1011.6411 \[hep-th\]\]. E. S. Fradkin and M. A. Vasiliev, Nucl. Phys. [**B291**]{} (1987) 141, Phys. Lett. [**B189**]{} (1987) 89. J. E. Paton and H. M. Chan, Nucl. Phys.  [**B10**]{} (1969) 516; J. H. Schwarz, “Gauge Groups For Type I Superstrings”, CALT-68-906-REV *Presented at 6th Johns Hopkins Workshop on Current Problems in High-Energy Particle Theory, Florence, Italy, Jun 2-4, 1982*; N. Marcus and A. Sagnotti, Phys. Lett.  [**B119**]{} (1982) 97, Phys. Lett.  [**B188**]{} (1987) 58. For reviews see: J. H. Schwarz, Phys. Rept.  [**89** ]{} (1982) 223-322; C. Angelantonj, A. Sagnotti, Phys. Rept.  [**371** ]{} (2002) 1-150. \[hep-th/0204089\]. R. R. Metsaev, arXiv:hep-th/0712.3526, Nucl. Phys.  [**B759**]{} (2006) 147 \[arXiv:hep-th/0512342\]. A. K. H. Bengtsson, I. Bengtsson and L. Brink, Nucl. Phys. [**B227**]{} (1983) 31, 41. N. Boulanger, S. Leclercq and P. Sundell, JHEP [**0808**]{} (2008) 056 \[arXiv:hep-th/0805.2764\]. A. Sagnotti, talk at Strings 09, http://strings2009.roma2.infn.it/program.html. H. van Dam and M. J. G. Veltman, Nucl. Phys. [**B22**]{}, 397 (1970); V. I. Zakharov, JETP Lett.  [**12**]{} (1970) 312 \[Pisma Zh. Eksp. Teor. Fiz.  [**12**]{} (1970) 447\]. X. Bekaert, E. Joung and J. Mourad, JHEP [**0905**]{} (2009) 126 \[arXiv:hep-th/0903.3338\]. JHEP [**1102**]{} (2011) 048 \[arXiv:1012.2103 \[hep-th\]\]. A. Higuchi, Nucl. Phys. [**B282**]{} (1987) 397, J. Math. Phys.  [**28**]{} (1987) 1553 \[Erratum-ibid.  [**43**]{} (2002) 6385\]; M. Porrati, Phys. Lett.  [**B498**]{}, 92 (2001) \[arXiv:hep-th/0011152\]; I. I. Kogan, S. Mouslopoulos and A. Papazoglou, Phys. Lett. [**B503**]{}, 173 (2001) \[arXiv:hep-th/0011138\]. J. Dixmier, “Représentations intégrables du groupe de De Sitter”, Bulletin de la Société Mathématique de France, 89 (1961), p. 9-41. Earlier works on the representations of SO(1,4) include: L. Thomas, “On unitary representations of the group of the De Sitter group," Ann. of Math. [**42**]{} (1941) 113; T. Newton, “A note on the representation of the De Sitter group," Ann. of Math. [**52**]{} (1950) 730; A. W. Knapp and E. M. Stein, “Interwining operators for semisimple groups," Ann. of Math. (2) [**93**]{} (1971) 489; E. Thieleker, “The unitary representations of the generalized Lorentz groups," Transactions of the American mathematical Society [**199**]{} (1974) 327; N. Ja. Vilenkin and A. U. Kilmyk, “Representation of Lie groups and special fucntions," Kluwer Academic Publishers (1993). The two-dimensional case was considered in V. Bargmann, Annals Math.  [**48**]{} (1947) 568. S. Deser and R. I. Nepomechie, Annals Phys.  [**154**]{} (1984) 396; S. Deser and A. Waldron, Phys. Rev. Lett.  [**87**]{} (2001) 031601 \[arXiv:hep-th/0102166\], Nucl. Phys. [**B607**]{} (2001) 577 \[arXiv:hep-th/0103198\], Phys. Lett. [**B501**]{}, 134 (2001) \[arXiv:hep-th/0012014\]; C. Fronsdal, Phys. Rev. [**D20**]{} (1979) 848. J. Fang and C. Fronsdal, Phys. Rev. [**D22**]{} (1980) 1361; T. Biswas and W. Siegel, JHEP [**0207**]{} (2002) 005 \[arXiv:hep-th/0203115\]; K. Hallowell and A. Waldron, Nucl. Phys. [**B724**]{} (2005) 453 \[arXiv:hep-th/0505255\]; X. Bekaert, I. L. Buchbinder, A. Pashnev and M. Tsulaia, Class. Quant. Grav.  [**21**]{} (2004) S1457 \[arXiv:hep-th/0312252\]. See, for instance, E.S. Whittaker and G.N. Watson, *A Course of Modern Analysis* (Cambridge Univ. Press, Cambridge, 1973). F. A. Berends, G. J. H. Burgers and H. van Dam, Nucl. Phys. [**B271**]{} (1986) 429, Z. Phys. [**C24**]{} (1984) 247, Nucl. Phys. [**B260**]{} (1985) 295. D. J. Gross and P. F. Mende, Phys. Lett. [**B197**]{} (1987) 129, Nucl. Phys. [**B303**]{} (1988) 407. E. Joung and M. Taronna, arXiv:1110.5918 \[hep-th\]. D. Polyakov, Phys. Rev. [**D82**]{} (2010) 066005 \[arXiv:0910.5338 \[hep-th\]\]; Int. J. Mod. Phys. [**A25**]{} (2010) 4623 \[arXiv:1005.5512 \[hep-th\]\]. [^1]: The arguments of [@bbms], inspired by the AdS/CFT correspondence [@adscft], are a notable example of this body of evidence. [^2]: This mass deformation should not be confused with the special mass parameters of Section \[sec:adsdeformations\], which guarantee the massless gauge symmetry in (A)dS backgrounds. [^3]: Let me stress that retaining all these gauge–invariant terms is a weaker condition than simply letting $\a^\prime\to \infty$, which would only leave the leading terms with $s_1+s_2+s_3$ derivatives that I shall describe shortly. The leading terms, incidentally, were first identified long ago by Gross and Mende [@grossmende]. [^4]: I cannot fail to mention that a very recent (A)dS extension [@adsembedding] shows in a concise and elegant fashion how the covariant derivatives generate Fradkin–Vasiliev minimal–like dressings, the counterparts for cubic vertices of the mass–like terms of Section \[sec:adsdeformations\], although this goes well beyond the Erice School. [^5]: Appendix A of [@st] explains how to pass to the compensator formulation of Section \[sec:freesymmetric\].
{ "pile_set_name": "ArXiv" }
--- abstract: 'We investigate the problem of a single ion in a radio-frequency trap and immersed in an ultracold Bose gas either in a condensed or a non-condensed phase. We develop master equation formalism describing the sympathetic cooling and we determine the cooling rates of ions. We show that cold atomic reservoir modifies the stability diagram of the ion in the Paul trap creating the regions where the ion is either cooled or heated due to the energy quanta exchanged with the time-dependent potential.' author: - 'Micha[ł]{} Krych' - Zbigniew Idziaszek bibliography: - 'master.bib' title: Description of ion motion in a Paul trap immersed in a cold atomic gas --- I. Introduction =============== Hybrid systems combining cold atomic gases with single ions or ionic crystals attract an increasing attention [@Grier2009; @Zipkes2010; @Schmid2010; @Rellergert2011; @Ravi2011; @Hall2011; @Hall2012; @Sivarajah2012; @Denschlag2013; @Denschlag2013a]. They have been proposed for implementation of quantum gates [@Idziaszek2007; @Doerk2010; @Nguyen2012], realization of new mesoscopic quantum states [@Cote2002; @Massignan2005], probing quantum gases [@Sherkunov2009], studying controlled chemical reactions at low temperatures [@Rellergert2011; @Hall2011; @Hall2012; @Hall2013] or emulating some well-known condensed-matter physics phenomena [@Gerritsma2012; @Bissbort2013]. The theoretical framework to describe atom-ion collisions in the quantum regime has been developed [@Cote2000; @Bodo2008; @Idziaszek2009; @Gao2010; @Idziaszek2011; @Gao2011; @Gao2013], however the ab-initio potentials [[@Makarov2003; @Knecht2010; @Krych2011; @Hall2013; @Hall2013a; @Tomza2014]]{} are not known with accuracy sufficient for precise determination of scattering lengths. Their values can be measured in experiments, for instance by applying technique of Feshbach resonances [@Idziaszek2009], provided the ions immersed in cold atomic gas are cooled down to the quantum regime where scattering takes place only in the lowest partial waves. Such low temperatures can be reached, for instance, via sympathetic cooling of ions in contact with cold atomic reservoir [@Goodman2012; @Hudson2009; @Ravi2011]. This method, however, suffers both due to some technical issues (e.g. excess micromotion [@Denschlag2013]), or due to some fundamental limitations resulting from the ion dynamics in the time-dependent Paul trap [@Dehmelt1968; @Ravi2011; @Vuletic2012; @Goodman2012]. Apart from the Paul traps there are first successful experimental attempts on optical ion trapping [[@Huber2014; @Schneider2010]]{}. So far the problem of atom-ion sympathetic cooling has been studied only in the classical regime [@DeVoe2010; @Vuletic2012; @Chen2013]. The purpose of this paper is to provide a consistent framework to describe the process of sympathetic cooling in the quantum regime and to study its limitations for experimental systems where a single ion is immersed in a Bose-Einstein condensate [@Zipkes2010; @Schmid2010]. [The paper is organized as follows.]{} [In Sec. II. we [introduce]{} the master equation formalism and [discuss]{} two reservoirs: an non-condensed ultracold gas and a [Bose-Einstein]{} condensate. A regularization of an atom-ion interaction potential is presented in Sec. III. Next, in Sec. IV. we derive the master equation. After that, in Sec. V. we discuss the evolution of the position operator in the time-dependent Paul trap. In Sec. VI. we present the equations of motion of the ion in contact with the cold reservoir. Finally, in Sec. VII. we discuss the ion cooling rates [and stability regimes]{} for different experimental parameters.]{} II. Master equation formalism ============================= We describe the system treating the atomic gas as a reservoir, and deriving an effective equation for the dynamics of the ion. The total Hamiltonian consists of the following parts: $\hat{H}=\hat{H}_S+\hat{H}_R+\hat{H}_{RS}$, where $\hat{H}_S$ is the Hamiltonian of the ion [@leibfried] $$\begin{split} \label{Ham1} \hat{H}_S & =\frac{\hat{{\mathbf}{p}}^2}{2M} + \frac{M}2 \sum_j \omega_j^2(t) \hat{r}_j^2, \end{split}$$ where the time-dependent trapping frequency consists of the static and dynamic parts: $\omega_j^2(t) = \Omega^2\frac14\left( a_j+ 2 q_j \cos \left( \Omega t \right)\right)$ and [$j=x,y,z$ is a spatial direction.]{} Here, $\hat{{\mathbf}{p}} = (\hat{p}_x,\hat{p}_y,\hat{p}_z)$ and $\hat{{\mathbf}{r}} = (\hat{r}_x,\hat{r}_y,\hat{r}_z)$ denote the momentum and position operators, respectively, $M$ is the ion mass and $\Omega$ is the radio frequency of the dynamic part. A homogeneous gas of atoms is described by [$$\label{HamR} \begin{split} \hat{H}_R&= \int \!\! d^3 r_a \hat{\Psi}^\dagger({\mathbf}{r}_a) \left(\frac{\hat{{\mathbf}{p}}_a^2}{2 m} + \frac{g}{2} \hat{\Psi}^\dagger({\mathbf}{r}_a) \hat{\Psi}({\mathbf}{r}_a) \right) \hat{\Psi}({\mathbf}{r}_a)\\ &=\sum_{{\mathbf}{k}}\hbar \omega_{{\mathbf}{k}} \hat{a}^\dagger_{\mathbf}{k} \hat{a}_{\mathbf}{k}, \end{split}$$]{} [where $\hat{\Psi}({\mathbf}{r}_{\hat{a}})=\sum_{{\mathbf}{k}}e^{i {\mathbf}{k} {\mathbf}{r}} {\hat{a}}_{\mathbf}{k}/\sqrt{L^3}$ is the field operator, $L$ is the size of the quantization box, $a_{\mathbf}{k}$ is the annihilation operator for mode ${\mathbf}{k}$, $\hbar\omega_k$ is energy of this mode, $\hat{{\mathbf}{p}}_a$ is the [atomic]{} momentum operator, $m$ is the atomic mass, and $g = 4 \pi \hbar^2 a /m $ is the interaction constant with $a$ denoting the $s$-wave scattering length. The ion-atom interaction is given by]{} $$\label{HamRS} \hat{H}_{RS} = \int d^3 r_a \hat{\Psi}^\dagger({\mathbf}{r}_a) V({\mathbf}{r}_a - \hat{{\mathbf}{r}}) \hat{\Psi}({\mathbf}{r}_a),$$ where $V({\mathbf}{r})$ is the atom-ion interaction potential. In our approach we treat the atomic gas in the second quantization formalism, while the ion is described by position and momentum operators. For an ion immersed in a Bose-Einstein condensate (BEC) we describe the reservoir in a Bogoliubov approximation $$\label{Bogoliubov} \hat{H}_R=E_0+\sum_{{\mathbf}{q}}\varepsilon({\mathbf}{q})\hat{b}^\dagger_{\mathbf}{q}\hat{b}_{\mathbf}{q}$$ where $\hat{b}^\dagger_{\mathbf}{q}$ and $\hat{b}_{\mathbf}{q}$ are the creation and annihilation operators for Bogoliubov excitations with momentum $\hbar {\mathbf}{q}$ and energy $\varepsilon({\mathbf}{q})$ and $E_0$ is the ground state energy of the superfluid. [Even in the ground state of the Paul trap the speed of motion of the ion is typically much larger than the speed of sound in the condensate, and ion couples only to the particle part of the Bogoliubov spectrum $\varepsilon({\mathbf}{q})\approx \hbar^2 q^2/(2m)$.]{} The ion and the superfluid are coupled by the density-density interaction [@Pitaevskii; @DaleyZoller2004] $$\label{oddzialywanieBEC} \begin{split} \hat{H}_{RS}&=\int d^3 r d^3 r' V({\mathbf}{r}-{\mathbf}{r}')\delta\hat{\rho}({\mathbf}{r})\delta\hat{\rho}_{ion}({\mathbf}{r}')\\ &=\int d^3 r V({\mathbf}{r}-{\mathbf}{\hat{r}})\delta\hat{\rho}({\mathbf}{r}) \end{split}$$ where $\delta\hat{\rho}({\mathbf}{r})=\hat{{\mathbf}{\Psi}}^\dagger \hat{{\mathbf}{\Psi}}-\rho_0$ and $\hat{{\mathbf}{\Psi}}=\sqrt{\rho_0}+\delta\hat{{\mathbf}{\Psi}}$ is the quantized field operator for the superfluid and $\rho_0$ is the condensate density, $\delta\hat{\rho}_{ion}({\mathbf}{r}')=\delta^3 ({\mathbf}{r}'-\hat{{\mathbf}{r}})$ is the density operator of the position of the ion. The field operator for the excitations of the superfluid is given by [@DaleyZoller2004] $\delta\hat{\Psi}=L^{-3/2}\sum_{\mathbf}{q}\left(u_{\mathbf}{q} \hat{b}_{\mathbf}{q} exp(i{\mathbf}{q r})+v_{\mathbf}{q} \hat{b}_{\mathbf}{q}^\dagger exp(-i{\mathbf}{q r})\right)$, where $L$ is the size of a box, and we assume the periodic boundary conditions. III. Regularized potential {#App:RegulPot} ========================== Long range interaction between ion and atom is described by the polarization potential $-C_4/r^4$, but at short distances this potential is singular and it needs to be regularized. We introduce a regularized version of the polarization potential $$\label{potential} V(r)=-C_4\frac{r^2-c^2}{r^2+c^2}\frac{1}{(b^2+r^2)^2}$$ that mimics at large distances the behavior of $-C_4/r^4$ tail. For small separations between particles it supports a single minimum [(cf. Fig. \[Fig:potencjal\])]{}. The short-range repulsive part is finite, which simplifies numerical calculations. We choose $b$ and $c$ parameters in such a way, that scattering amplitude calculated in the first-order Born approximation $$\begin{split} \label{fq} f(q)&= \frac{\pi (R^{\star})^2}{4b(b^2-c^2)^2 q}\left(-4b c^2 e^{-cq}\right.\\ &\left.+e^{-bq}(4b c^2+(b^2-c^2)(b^2+c^2)q)\right) \end{split}$$ is equal to the exact scattering amplitude of the potential at zero energy. Here, $q=|{\mathbf}{k}-{\mathbf}{k}^\prime|$ is the magnitude of momentum transfer. In this way our description within the master equation formalism [@Carmichael], treating system-reservoir interactions in the Born approximation will be accurate for a single collision in the ultracold regime. ![\[Fig:potencjal\] (Color online). A regularized potential (solid line) and a pure $-C_4/r^4$ (dashed line). ](potencjal.eps){width="0.6\linewidth"} With the polarization potential one can associate a characteristic length $R^\ast = \left( 2 \mu C_4/\hbar^2\right)^{1/2}$ and a characteristic energy $E^\ast = \hbar^2/\left[2 \mu (R^\ast)^2\right]$. We choose $b$, $c$ parameters in such a way, that a scattering amplitude calculated in the first order Born approximation is equal to the exact scattering amplitude of the potential at zero energy. For example for $b=0.0781R^{\star}$ and $c=0.2239R^{\star}$ potential supports a single bound state, its scattering length is equal $a_{sc}=R^{\star}$ and the zero-energy scattering amplitudes calculated exactly and from Eq.  are equal to $-R^{\star}$. IV. Master equation {#App:Srednie} =================== The interaction Hamiltonian of the ion with the ultracold gas (Eq. ) can be rewritten explicitly as $$\label{oddzialywaniegazrozwiniete} \begin{split} \hat{H}_{RS}&=\sum_{{\mathbf}{k},{\mathbf}{k}'}\hat{a}_{\mathbf}{k'}^\dagger \hat{a}_{{\mathbf}{k}} L^{-3}\int d^3 r_a e^{-i({\mathbf}{k}'-{\mathbf}{k}){\mathbf}{r}_a} V({\mathbf}{r}_a-\hat{{\mathbf}{r}}_j)\\ &=\sum_{{\mathbf}{k},{\mathbf}{k}'}\hat{a}_{\mathbf}{k'}^\dagger \hat{a}_{{\mathbf}{k}} e^{-i({\mathbf}{k}'-{\mathbf}{k})\hat{{\mathbf}{r}}_j} L^{-3}\int d^3 r e^{-i({\mathbf}{k}'-{\mathbf}{k}){\mathbf}{r}} V({\mathbf}{r}), \end{split}$$ where $V({\mathbf}{r})$ denotes the interaction potential. It is easy to [separate ion and reservoir operators in a general form used in derivation of the master equation [@Carmichael]]{} $$\hat{H}_{RS}=\hbar \sum_{{\mathbf}{k},{\mathbf}{k}'} \hat{s}_{{\mathbf}{k}{\mathbf}{k}'} \hat{\Gamma}_{{\mathbf}{k}{\mathbf}{k}'},$$ where the ion part reads $$\hat{s}_{{\mathbf}{k}{\mathbf}{k}'}=e^{-i({\mathbf}{k}'-{\mathbf}{k})\hat{{\mathbf}{r}}} c_{{\mathbf}{k},{\mathbf}{k}'}/\hbar,$$ and $c_{kk'}$ is a Fourier transform of the interaction potential $$c_{{\mathbf}{k},{\mathbf}{k}'}= L^{-3} \int d^3 r e^{i({\mathbf}{k}-{\mathbf}{k}')\hat{{\mathbf}{r}}} V({\mathbf}{r}).$$ Gas operators can be written down as $$\hat{\Gamma}_{{\mathbf}{k}{\mathbf}{k}'}=\hat{a}_{\mathbf}{k}^\dagger \hat{a}_{{\mathbf}{k}'}.$$ General textbook form of a master equation [for a reduced density matrix]{} (page 8 in [@Carmichael]) after the Markov approximation reads $$\label{masterksiazka} \begin{split} \dot{\hat{\tilde{\rho}}}=\!\!\!\!\sum_{{\mathbf}{k},{\mathbf}{k}',{\mathbf}{l},{\mathbf}{l}'}\!\int_0^t \!\!\!\! & dt' \!\left( [\hat{\tilde{s}}_{{\mathbf}{l}{\mathbf}{l}'}(t')\hat{\tilde{\rho}}(t),\hat{\tilde{s}}_{{\mathbf}{k}{\mathbf}{k}'}(t)]\langle\hat{\tilde{\Gamma}}_{{\mathbf}{k}{\mathbf}{k}'}(t)\hat{\tilde{\Gamma}}_{{\mathbf}{l}{\mathbf}{l}'}(t')\rangle_{R_0} \right.\\ &\left.+ [\hat{\tilde{s}}_{{\mathbf}{k}{\mathbf}{k}'}(t),\hat{\tilde{\rho}}(t)\hat{\tilde{s}}_{{\mathbf}{l}{\mathbf}{l}'}(t')]\langle\hat{\tilde{\Gamma}}_{{\mathbf}{l}{\mathbf}{l}'}(t')\hat{\tilde{\Gamma}}_{{\mathbf}{k}{\mathbf}{k}'}(t)\rangle_{R_0}\right). \end{split}$$ Here $\langle\dots\rangle_{R_0}$ is a trace over the reservoir density matrix $R_0$, tilded operators denote an interaction picture with respect to the noninteracting system (ion in a Paul trap Eq. ) and a reservoir (noninteracting gas Eq. ) $$\hat{\tilde{X}}(t)=U^\dagger (0,t) e^{(i/\hbar)\hat{H}_R t} \hat{X} e^{-(i/\hbar)\hat{H}_R t}U (0,t)$$ where $U(t_1,t_2)={\cal{T}}\exp(-\frac{i}{\hbar}\int_{t_1}^{t_2}d\tau \hat{H}_S(\tau))$ is the time-ordered evolution operator of an ion immersed in a Paul trap. Let us consider the free evolution of the gas operators $$\hat{\tilde{\Gamma}}_{{\mathbf}{k}{\mathbf}{k}'}(t)=\hat{\tilde{a}}_{\mathbf}{k}^{\dagger}(t)\hat{\tilde{a}}_{{\mathbf}{k}'}(t)=\hat{a}_{\mathbf}{k}^{\dagger}\hat{a}_{{\mathbf}{k}'} e^{i(\omega_{\mathbf}{k}-\omega_{{\mathbf}{k}'})t}.$$ We denote the mean occupation number of the atomic modes as $$\bar{n}_{\mathbf}{k}\equiv \langle \hat{a}_{\mathbf}{k}^{\dagger}\hat{a}_{{\mathbf}{k}}\rangle_{R_0}$$ where $\bar{n}_{\mathbf}{k}=1/(\exp(\hbar^2 {\mathbf}{k}^2/(2m k_B T)-\mu)-1)$ and $\mu$ denotes the chemical potential, $k_B$ is the Boltzmann constant nad $T$ is a temperature. Using the commutation relations for creation and annihilation operators we are able to eliminate the reservoir degrees of freedom and rewrite the master equation with an explicit form of gas correlation functions $$\label{masterkolejne} \begin{split} \dot{\hat{\tilde{\rho}}}=&\sum_{{\mathbf}{k},{\mathbf}{k}'}\int_0^t dt'\\ &\left([\hat{\tilde{s}}_{{\mathbf}{k}' {\mathbf}{k}}(t')\hat{\tilde{\rho}}(t),\hat{\tilde{s}}_{{\mathbf}{k}{\mathbf}{k}'}(t)] e^{i(t-t')(\omega_{\mathbf}{k}-\omega_{{\mathbf}{k}'})} \bar{n}_{{\mathbf}{k}}(\bar{n}_{\mathbf}{k'}+1)\right. \\ &+\left.[\hat{\tilde{s}}_{{\mathbf}{k}{\mathbf}{k}'}(t),\hat{\tilde{\rho}}(t)\hat{\tilde{s}}_{{\mathbf}{k}' {\mathbf}{k}}(t')] e^{i(t-t')(\omega_{\mathbf}{k}-\omega_{{\mathbf}{k}'})} \bar{n}_{{\mathbf}{k'}}(\bar{n}_{\mathbf}{k}+1)\right). \end{split}$$ In the next step we [transform the master equation back from the interaction picture]{} $$\dot{\hat{\rho}}=\frac{1}{i\hbar}[\hat{H}_S,\hat{\rho}]+U(0,t) \dot{\hat{\tilde{\rho}}} U^\dagger(0,t).$$ [ After changing the order of terms in commutators and of the summation indices in the last line of and substituting $t-t'\equiv \tau$, the master equation reads $$\label{masterkolejnekolejne} \begin{split} \dot{\hat{\rho}}&=\frac{1}{i\hbar}[\hat{H}_S,\hat{\rho}]-\sum_{{\mathbf}{k},{\mathbf}{k}'} \bar{n}_{\mathbf}{k}(\bar{n}_{{\mathbf}{k}'}+1)\\ &\times \int_0^t d\tau \left( e^{i\tau(\omega_{\mathbf}{k}-\omega_{{\mathbf}{k}'})} \left[{\hat{s}}_{{\mathbf}{k}{\mathbf}{k}'},{\hat{s}}_{{\mathbf}{k}' {\mathbf}{k}}(t,-\tau){\hat{\rho}}\right]\right.\\ &+\left.e^{-i\tau(\omega_{\mathbf}{k}-\omega_{{\mathbf}{k}'})} \left[{\hat{\rho}} {\hat{s}}_{{\mathbf}{k} {\mathbf}{k}'}(t,-\tau),{\hat{s}}_{{\mathbf}{k}' {\mathbf}{k}}\right]\right). \end{split}$$ For any operator $\hat{x}$ we define $$\begin{split} \hat{x}(t,-\tau)&\equiv U(0,t)U^\dagger(0,t-\tau)\hat{x}U(0,t-\tau)U^\dagger(0,t)\\ &\equiv U(0,t)\hat{x}(t-\tau)U^\dagger(0,t) \end{split}$$ In case of a time independent Hamiltonian this would reduce to $\hat{x}(t,-\tau)=U^\dagger(0,-\tau)\hat{x}U(0,-\tau)$. However, we note that in a Paul trap the evolution cannot be reduced to the single evolution operator with the time difference as in energy-conserving system, since $U(0,t)U^\dagger(0,t-\tau) \neq U^\dagger(0,-\tau)$. The integration over $\tau$ in the master equation is dominated by short timescales, because correlation functions in a large reservoir vanish quickly in time.]{} In this way we can extend the integration limit up to the infinity, as is usually done in the derivation of the master equation [@Carmichael]. [Substituting an explicit form of operators $\hat{s}_{{\mathbf}{k}{\mathbf}{k}'}$ into yields]{} $$\label{masterexp} \begin{split} \dot{\hat{\rho}}&=\frac{1}{i\hbar}[\hat{H}_S,\hat{\rho}]-\sum_{{\mathbf}{k},{\mathbf}{k}'}\bar{n}_{{\mathbf}{k}}(\bar{n}_{{\mathbf}{k}'}+1) c_{{\mathbf}{k},{\mathbf}{k}'} c_{{\mathbf}{k}',{\mathbf}{k}}/\hbar^2 \\ &\times \int_0^{\infty}\!\!\! d\tau\left( e^{i\tau (\omega_{\mathbf}{k}-\omega_{{\mathbf}{k}'})} \left[e^{i({\mathbf}{k}-{\mathbf}{k}')\hat{{\mathbf}{r}}}, e^{-i({\mathbf}{k}-{\mathbf}{k}')\hat{{\mathbf}{r}}(t,-\tau)}{\hat{\rho}}\right]\right.\\ &+\left. e^{-i\tau (\omega_{\mathbf}{k}-\omega_{{\mathbf}{k}'})} \left[{\rho}e^{i({\mathbf}{k}-{\mathbf}{k}')\hat{{\mathbf}{r}}(t,-\tau)}, e^{-i({\mathbf}{k}-{\mathbf}{k}')\hat{{\mathbf}{r}}}\right]\right). \end{split}$$ [In case of a non-condensed buffer gas the dynamics of the reduced density operator $\hat{\rho}$ obtained by tracing over the reservoir modes, derived within Born and Markov approximations [@Gardiner; @Carmichael] can be straightforwardly rewritten in a compact form]{} $$\label{MasterGas} \dot{\hat{\rho}}(t)= \frac{1}{i\hbar}\left[\hat{H}_S(t),\hat{\rho}\right]-\sum_{{\mathbf}{k},{\mathbf}{k}'} \Omega_{{\mathbf}{k},{\mathbf}{k}'}^2 \left\{ \left[ \hat{Z}_{{\mathbf}{k},{\mathbf}{k}'},\hat{W}_{{\mathbf}{k},{\mathbf}{k}'}(t) \hat{\rho} \right] + H. c. \right\}$$ [where ${\mathbf}{k}$ and ${\mathbf}{k}^\prime$ are the quantized wave vectors of atoms in a box [of size $L^3$]{}, $\Omega_{{\mathbf}{k},{\mathbf}{k}'}^2 = \bar{n}_{\mathbf}{k} (\bar{n}_{{\mathbf}{k}'}+1) |c_{{\mathbf}{k},{\mathbf}{k}'}|^2/\hbar^2$. Furthermore, $\hat{Z}_{{\mathbf}{k},{\mathbf}{k}'} = e^{i({\mathbf}{k}-{\mathbf}{k}')\hat{{\mathbf}{r}}}$, and $\hat{W}_{{\mathbf}{k},{\mathbf}{k}'}(t) = \int_{0}^{\infty} d\tau e^{i\tau (\omega_{{\mathbf}{k}}-\omega_{\mathbf}{k}')} e^{i({\mathbf}{k}'-{\mathbf}{k})\hat{{\mathbf}{r}}(t,-\tau)}$.]{} For the Bose condensed reservoir in order to derive the master equation we use the interaction Hamiltonian given by Eq.  instead of Hamiltonian and we describe the reservoir in the Bogoliubov approximation (Eq. ). All the following steps of the derivation are analogous to the non-condensed case: (i) [we start with]{} the general textbook form of the master equation for a reduced density matrix after the Markov approximation (Eq. ), (ii) perform the trace operation with respect to the BEC degrees of freedom, (iii) [transform back]{} from the interaction picture, (iv) extend the time integration up to the infinity. [Therefore in case of the BEC reservoir the master equation reads]{} $$\label{MasterBose} \begin{split} \dot{\hat{\rho}}(t)&= \frac{1}{i\hbar}\left[\hat{H}_S(t),\hat{\rho}\right]-\sum_{{\mathbf}{q}} \Omega_{{\mathbf}{q}}^2 \left\{\bar{n}_{\mathbf}{q}\left[\hat{Z}_{0,{\mathbf}{q}},\hat{W}_{0,{\mathbf}{q}}(t) \hat{\rho} \right]\right.\\ &\left.+(1+\bar{n}_{\mathbf}{q})\left[ \hat{\rho}\hat{W}_{0,{\mathbf}{q}}(t),\hat{Z}_{0,{\mathbf}{q}} \right] + H. c. \right\} \end{split}$$ where $\Omega_{{\mathbf}{q}}^2=\rho_0 L^3|c_{0,{\mathbf}{q}}|^2/\hbar^2$ [and $\bar{n}_{\mathbf}{q}=1/(\exp(\hbar^2 {\mathbf}{q}^2/(2m k_B T))-1)$ denotes the mean occupation number of the Bogoliubov quasiparticles]{}. [We note that the master equation for a BEC (Eq. ) is equivalent to the low temperature limit of the master equation for the non-condensed gas (Eq. ). This is because of the fact that for typical parameters the ion couples only to the particle region of the Bogoliubov excitations.]{} V. Time dependence of the position operator =========================================== In order to determine the coefficients of master equations and we have to find the [evolution of the position operator $\hat{{\mathbf}{r}}(t,-\tau)$ of the ion in a Paul trap in the absence of the ultracold gas.]{} We start from Heisenberg equations of motion for position $\dot{\hat{r}}_j=\hat{p}_j/M$ and momentum $\dot{\hat{p}}_j=-M \omega_j^2(t) \hat{r}_j$ derived with help of Hamiltonian from Eq. , where $j=x,y,x$ denotes the direction in space. Their combination leads to the second order differential equation $$\label{heisenberg} \ddot{\hat{r}}_j+\omega_j^2(t) \hat{r}_j=0,$$ where the time-dependent trapping frequency consists of the static and dynamic parts: $\omega_j^2(t) = \Omega^2\frac14\left( a_j+ 2 q_j \cos \left( \Omega t \right)\right)$, as was defined before. Above equation has two linearly independent $\cal{C}$-number solutions $u_j(t)$ and $u_j(-t)$, where $u_j(t)=e^{i(\beta_j/2) \Omega t} \sum_{n=-\infty}^{\infty}C^j_{n} e^{in\Omega t}$, $u_j(0)=1$ and $C^j_n$ are expansion coefficients of Mathieu functions describing the time evolution of a single ion in a Paul trap and $(\beta_j/2) \Omega$ is an effective secular frequency [@leibfried]. Eq.  does not have the first order derivative, so [the Wronskian of $u_j(t)$ and $u_j(-t)$ is constant in time]{} $$\label{wr0} \begin{split} W\left(u_j(t),u_j(-t)\right)&=u_j(-t)\dot{u}_j(t)-u_j(t)\dot{u}_j(-t)\\ &=2 i \nu_j={\mathrm}{const}, \end{split}$$ where $\nu_j=\dot{u}_j(0)/i=\Omega \sum_{n=-\infty}^{\infty} C^j_{n} (\beta_j/2+n)$ [is called a reference harmonic oscillator frequency. One can also define two other Wronskians between position operator and $u_j(t)$, which is proportional to]{} $$\label{wr1} \begin{split} \hat{c}_{j,1}(t)&=i\sqrt{M/(2\hbar \nu_j)}\times W\left({\hat{r}}_j(t),u_j(t)\right)\\ &=i\sqrt{M/(2\hbar \nu_j)}\left(u_j(t)\dot{\hat{r}}_j(t)-\dot{u}_j(t)\hat{r}_j(t)\right)\\ &=\hat{c}_{j,1}(0)=1/\sqrt{2 M \hbar \nu_j}(M \nu_j \hat{r}_j(0)+i \hat{p}_j(0)) \end{split}$$ [or between position operator and $u_j(-t)$, which is proportional to]{} $$\label{wr2} \begin{split} \hat{c}_{j,2}(t)&=-i\sqrt{M/(2\hbar \nu_j)}\times W\left({\hat{r}}_j(t),u_j(-t)\right)\\ &=-i\sqrt{M/(2\hbar \nu_j)}\left(u_j(-t)\dot{\hat{r}}_j(t)-\dot{u}_j(-t)\hat{r}_j(t)\right)\\ &=\hat{c}_{j,2}(0)=1/\sqrt{2 M \hbar \nu_j}(M \nu_j \hat{r}_j(0)-i \hat{p}_j(0)) \end{split}$$ [Since Wronskians of Eq.  must be constant in time, operators $\hat{c}_{j,1}(t)$ and $\hat{c}_{j,2}(t)$ are constant.]{} For $\beta_j$ real one can show that they are equivalent to creation and annihilation operator (respectively) of the reference harmonic oscillator of frequency $\nu_j$ [@leibfried]. Both of the above equations connect $\hat{r}_j(t)=U^\dagger(0,t)\hat{r}_j(0)U(0,t)$ and $\hat{p}_j(t)=U^\dagger(0,t)\hat{p}_j(0)U(0,t)$ with their values for $t=0$ ($\hat{r}_j(0)\equiv\hat{r}_j$ and $\hat{p}_j(0)\equiv\hat{p}_j$). Multiplying Eqs. , by $u_j(-t)$ and $u_j(t)$, respectively, adding the first one to the second one and using Eq.  we can express $\hat{r}_j(t)$ as a function of $\hat{r}_j(0)$ and $\hat{p}_j(0)$. $$\label{wr3} \begin{split} \hat{r}_j(t)&=\frac{\hat{r}_j(0)}{2}(u_j(t)+u_j(-t))+\frac{\hat{p}_j(0)}{2 i M \nu_j}(u_j(t)-u_j(-t)) \end{split}$$ Basing on Heisenberg equations of motion $\hat{p}_j=M \dot{\hat{r}}_j$ [we can express]{} $\hat{p}_j(t)$ also as a function of $\hat{r}_j(0)$ and $\hat{p}_j(0)$. $$\label{wr4} \begin{split} \hat{p}_j(t)&=\frac{M\hat{r}_j(0)}{2}(\dot{u}_j(t)+\dot{u}_j(-t)+\frac{\hat{p}_j(0)}{2 i \nu_j}(\dot{u}_j(t)-\dot{u}_j(-t))) \end{split}$$ [Starting from Eqs.  and it]{} is straightforward to derive the following result $$\begin{split}\label{r0} \hat{r}_j(0)&=\frac{\hat{r}_j(t)}{2i\nu_j}(\dot{u}_j(t)-\dot{u}_j(-t))-\frac{\hat{p}_j(t)}{2 i M \nu_j}(u_j(t)-u_j(-t)) \end{split}$$ and $$\label{p0} \begin{split} \hat{p}_j(0)&=-\frac{M\hat{r}_j(t)}{2}(\dot{u}_j(t)+\dot{u}_j(-t))+\frac{\hat{p}_j(t)}{2}(u_j(t)+u_j(-t)). \end{split}$$ Now we are able to calculate $$\begin{split} \hat{r}_j(t,-\tau)&\equiv U(0,t) U^\dagger(0,t-\tau) \hat{r}_j U(0,t-\tau) U^\dagger(0,t) = U(0,t)\hat{r}_j(t-\tau) U^\dagger(0,t)\\ &=U^\dagger (t,0)\left(\frac{\hat{r}_j(0)}{2}(u_j(t-\tau)+u_j(-t+\tau))+\frac{\hat{p}_j(0)}{2 i M \nu_j}(u_j(t-\tau)-u_j(-t+\tau))\right)U(t,0) \end{split}$$ where we have used the identity $U^\dagger(0,t)\equiv U(t,0)$. [With help of]{} Eqs.  and we perform the time evolution from $t$ to $0$. With an explicit form of $u_j(\pm t)$ this can be arranged in a compact form $$\label{rodtau} \hat{r}(t,-\tau)=\sum_{n,m}\!C_n C_m\!\!\left[\hat{r} \left(\frac{\!\beta}{2}\!+\!m\!\right)\!\frac{\Omega}{\nu} \cos I_{nm}^\tau\!-\!\frac{\hat{p}}{\nu M} \sin I_{nm}^\tau\right]$$ where we have omitted $j$ index for simplicity, $I_{nm}^\tau\equiv \Omega\left(\left(\frac{\beta}{2}+n\right)\tau-(n-m)t\right)$. VI. Equations of motion {#App:Srednie} ======================= We expand exponential terms of master equation (Eq. ) in the small Lamb-Dicke parameter $\zeta=a_{i}/\lambda_T$ up to the second-order terms, where $a_{i}=\sqrt{\hbar/(M\nu)}$ is a length scale of the secular potential with a reference oscillator frequency (which is [comparable to the size of the ion wavefunction]{}) and $\lambda_T=\sqrt{2\pi \hbar^2/(m k_B T)}$ is de Broglie wavelength, with $T$ denoting the temperature of the reservoir (a typical change of the atomic momenta $({\mathbf}{k}-{\mathbf}{k'})$ during a single atom-ion collision is of the order of $\lambda_T^{-1}$). For example $$\begin{split} &\left[e^{i({\mathbf}{k}-{\mathbf}{k'})\hat{{\mathbf}{r}}}, e^{-i({\mathbf}{k}-{\mathbf}{k'})\hat{{\mathbf}{r}}(t,-\tau)}{\rho}\right]=[i({\mathbf}{k}-{\mathbf}{k'})\hat{{\mathbf}{r}},\rho]+\\ &+\sum_j (k_j-k'_j)^2(\frac12[\rho, \hat{r}_j^2]+[\hat{r}_j, \hat{r}_j(t,-\tau)\rho])+\dots \end{split}$$ where $j=x,y,z$. We [note that the terms with odd powers of $(k_j-k_j')$ vanish due to the symmetric]{} summation in the master equation. Every odd term of the expansion is antisymmetric in $(k_j-k_j')$, so the first neglected term is of the fourth order in a small Lamb-Dicke parameter $\zeta$. The master equation exact up to the third order in $\zeta$ is given by $$\begin{split} \dot{\hat{\rho}}=&\frac{1}{i\hbar}[\hat{H}_S,\hat{\rho}]-\sum_j\sum_{k_j,k_j'}\frac{c_{{\mathbf}{k},{\mathbf}{k}'}c_{{\mathbf}{k}', {\mathbf}{k}}}{2\hbar^2} (k_j-k_j')^2 \bar{n}_{{\mathbf}{k}}(\bar{n}_{{\mathbf}{k}'}+1)\\&\times\int_0^{\infty} d\tau \left(e^{i\tau (\omega_{{\mathbf}{k}}-\omega_{{\mathbf}{k}'})} \left([\hat{\rho},\hat{r}_j^2]+2[\hat{r}_j, \hat{r}_j(t,-\tau)\hat{\rho}]\right)\right.\\ &-\left. e^{-i\tau (\omega_{{\mathbf}{k}}-\omega_{{\mathbf}{k}'})} \left([\hat{r}_j^2,\hat{\rho}]+2[\hat{r}_j,\hat{\rho} \hat{r}_j(t,-\tau)]\right)\right). \end{split} \label{2rzad}$$ [We note that in this approximation spatial directions are not coupled.]{} In order to derive equations of motion [for expectation values of position and momentum operators]{} we multiply master equation for ultracold gas or BEC, respectively, by $\hat{r}$ or $\hat{p}$ operators and perform tracing over ion degrees of freedom. In this way we obtain $$\label{liniowej} \begin{split} \dot{\overline{r}}_j&=\overline{p}_j/M\\ \dot{\overline{p}}_j&=- M\omega_j^2(t) \overline{r}_j +K^j(t)\overline{r}_j- 2 G^j_{\eta,\delta}(t)\overline{p}_j, \end{split}$$ [where in general, expectation values are defined as follows $${\overline{x}}\equiv{\mathrm}{tr}\left\{\hat{x} {\rho}\right\}$$ $$\dot{\overline{x}}\equiv{\mathrm}{tr}\left\{\hat{x} \dot{\rho}\right\}$$]{} Similar procedure may be applied to equations involving expectation values of operators quadratic in position and momentum $$\begin{aligned} \label{kwadratowej} (\dot{\overline{r_j p_j}}+\dot{\overline{p_j r_j}})=&-2M\omega_j^2(t) \overline{r_j^2}+2\overline{p_j^2}/M+2K_j(t)\overline{r_j^2} \nonumber\\ &-2G^j_{\eta,\delta}(t)(\overline{r_j p_j}+\overline{p_j r_j})+2\hbar G^j_{\gamma,-\mu}(t),\nonumber\\ \dot{\overline{r_j^2}}=&\frac{1}{M} (\overline{r_j p_j}+\overline{p_j r_j}),\\ \dot{\overline{p_j^2}}=&-M\omega_j^2(t) (\overline{r_j p_j}+\overline{p_j r_j})-4 G^j_{\eta,\delta}(t)\overline{p_j^2}\nonumber\\ &+K^j(t) (\overline{r_j p_j}+\overline{p_j r_j})+2\hbar D^j_{\mu,\gamma}(t).\nonumber\end{aligned}$$ Coefficients in Eqs.  and have the following form $$\begin{aligned} G^j_{\eta,\delta}(t) = & \sum_{n,m}\frac{C_n^j C_m^j}{\nu_j}\frac{\hbar}{M} \\ &\times\big[\eta_{jn}\cos\left((n\!-\!m)\Omega t\right)+\delta_{jn}\sin\left((n\!-\!m)\Omega t\right)\big], \nonumber \\ G^j_{\gamma,-\mu}(t) = & \sum_{n,m}\frac{C_n^j C_m^j}{\nu_j}\frac{\hbar}{M} \\ &\times\big[\gamma_{jn}\cos\left((n\!-\!m)\Omega t\right)-\mu_{jn}\sin\left((n\!-\!m)\Omega t\right)\big], \nonumber \\ K^j(t) = & 4\hbar\kappa_j- 2 D^j_{\delta_,-\eta}(t), \\ D^j_{\mu,\gamma}(t) = & \sum_{n,m}\frac{C_n^j C_m^j}{\nu_j}\left(\frac{\beta_j}{2}+m\right)\hbar\Omega \\ &\times\big[\mu_{jn}\cos\left((n\!-\!m)\Omega t\right)+\gamma_{jn}\sin\left((n\!-\!m)\Omega t\right)\big], \nonumber \\ D^j_{\delta,-\eta}(t) = & \sum_{n,m}\frac{C_n^j C_m^j}{\nu_j}\left(\frac{\beta_j}{2}+m\right)\hbar\Omega \\ &\times\big[\delta_{jn}\cos\left((n\!-\!m)\Omega t\right)-\eta_{jn}\sin\left((n\!-\!m)\Omega t\right)\big]. \nonumber\end{aligned}$$ Here, $G^j_{\eta,\delta}(t)$ plays the role of the time-dependent friction force, where sequences of constants $\eta_{jn}$ and $\delta_{jn}$ can be calculated for a given atom-ion potential, $\nu_j$ is the frequency of a reference oscillator[, $\beta_j/2$ denotes a characteristic exponent]{}, $C_n^j$, $C_m^j$ are the coefficients in a solution of Mathieu equation of an ion in a Paul trap without the buffer gas [@leibfried], [and $\kappa_j$ is some coefficient that will be defined later separately for a condensed or a non-condensed reservoir.]{} One can check that free terms $G^j_{\gamma,-\mu}(t)$ and $D^j_{\mu,\gamma}(t)$ assure that the ion energy cannot drop below the ground state energy of the secular trap even if the temperature of the atomic gas is lower. $\mu_{jn}$ and $\gamma_{jn}$ are sequences of constants depending on interaction potential. The form of Eq.  and Eq.  is general regardless of the interaction potential. They are valid both for an ultracold gas and a Bose-Einstein condensate, but the sequences of constants $\eta_{jn}$, $\delta_{jn}$, $\mu_{jn}$, $\gamma_{jn}$ and $\kappa_j$ have different functional forms [that will be introduced soon]{}. [In order to derive their explicit form we use the following identity to calculate the integrals with respect to time variable]{} $$\begin{split} &\int_0^{\infty} d\tau e^{s_1 i \tau(\omega_{{\mathbf}{k}}-\omega_{{\mathbf}{k}'})+s_2 i \Omega((\beta_j/2+n)\tau-(n-m)t) }\\ &=e^{-i s_2 \Omega (n-m)t}\left(\pi \delta(s_1 (\omega_{{\mathbf}{k}}-\omega_{{\mathbf}{k}'})+s_2\Omega(\beta_j/2+n))\right.\\ &\left.+i\frac{{\cal{ P}}}{s_1 (\omega_{{\mathbf}{k}}-\omega_{{\mathbf}{k}'})+s_2\Omega(\beta_j/2+n)}\right), \end{split}$$ where $s_1$ and $s_2$ can be equal to $+$, $-$ or $0$ and $\delta(\dots)$ and $\cal{P}\dots$ denote Dirac delta and Principal value distributions. With help of above relation and Eq.  in case of a reservoir consisting of a non-condensed Bose gas [we obtain]{} $$\begin{aligned} \label{KappaGas} \kappa_j = & \frac{1}{2\hbar^2}\sum_{{\mathbf}{k},{\mathbf}{k'}} |c_{{\mathbf}{k},{\mathbf}{k'}}|^2 (k_j-k'_j)^2 (\bar{n}_{{\mathbf}{k'}}\bar{n}_{{\mathbf}{k}}+\bar{n}_{{\mathbf}{k}}) \frac{\cal{P}}{\omega_{{\mathbf}{kk'}}}, \\ \eta_{jn} = & \frac{\pi}{2\hbar^2}\sum_{{\mathbf}{k},{\mathbf}{k'}} |c_{{\mathbf}{k},{\mathbf}{k'}}|^2 (k_j-k'_j)^2 (\bar{n}_{{\mathbf}{k'}}\bar{n}_{{\mathbf}{k}}+\bar{n}_{{\mathbf}{k}})\\ & \times\big[\delta(\omega_{{\mathbf}{kk'}}+\Omega(\beta_j/2+n))-\delta(\omega_{{\mathbf}{kk'}}-\Omega(\beta_j/2+n))\big], \nonumber \\ \mu_{jn} = & \frac{\pi}{2\hbar^2}\sum_{{\mathbf}{k},{\mathbf}{k'}} |c_{{\mathbf}{k},{\mathbf}{k'}}|^2 (k_j-k'_j)^2 (\bar{n}_{{\mathbf}{k'}}\bar{n}_{{\mathbf}{k}}+\bar{n}_{{\mathbf}{k}})\\ & \times\big[\delta(\omega_{{\mathbf}{kk'}}+\Omega(\beta_j/2+n))+\delta(\omega_{{\mathbf}{kk'}}-\Omega(\beta_j/2+n))\big], \nonumber \\ \delta_{jn} = & \frac{1}{2\hbar^2}\sum_{{\mathbf}{k},{\mathbf}{k'}} |c_{{\mathbf}{k},{\mathbf}{k'}}|^2 (k_j-k'_j)^2 (\bar{n}_{{\mathbf}{k'}}\bar{n}_{{\mathbf}{k}}+\bar{n}_{{\mathbf}{k}})\\ & \times \left( \frac{\cal{P}}{\omega_{{\mathbf}{kk'}}+\Omega(\beta_j/2+n)} +\frac{\cal{P}}{\omega_{{\mathbf}{kk'}}-\Omega(\beta_j/2+n)}\right), \nonumber \\ \label{GammaGas} \gamma_{jn} = & \frac{1}{2\hbar^2}\sum_{{\mathbf}{k},{\mathbf}{k'}} |c_{{\mathbf}{k},{\mathbf}{k'}}|^2 (k_j-k'_j)^2 (\bar{n}_{{\mathbf}{k'}}\bar{n}_{{\mathbf}{k}}+\bar{n}_{{\mathbf}{k}})\\ & \times \left( \frac{\cal{P}}{\omega_{{\mathbf}{kk'}}+\Omega(\beta_j/2+n)} -\frac{\cal{P}}{\omega_{{\mathbf}{kk'}}-\Omega(\beta_j/2+n)}\right), \nonumber\end{aligned}$$ where $\omega_{{\mathbf}{kk'}}\equiv \omega_{{\mathbf}{k}}-\omega_{{\mathbf}{k'}}$. For a Bose-condensed reservoir $$\begin{aligned} \label{KappaBEC} \kappa_j = & -\frac{\rho_0 L^3}{2\hbar^2}\sum_{{\mathbf}{q}}(u_{\mathbf}{q}+v_{\mathbf}{q})^2 |c_{0,{\mathbf}{q}}|^2 q_j^2 \frac{\cal{P}}{\omega_q}, \\ \eta_{jn} = & -\frac{\pi \rho_0 L^3}{2\hbar^2}\sum_{{\mathbf}{q}}(u_{\mathbf}{q}+v_{\mathbf}{q})^2 |c_{0,{\mathbf}{q}}|^2 q_j^2 \\ &\times\big[\delta(\omega_{{\mathbf}{q}}+\Omega(\beta_j/2+n))-\delta(\omega_{{\mathbf}{q}}-\Omega(\beta_j/2+n))\big], \nonumber \\ \mu_{jn} = & \frac{\pi \rho_0 L^3}{2\hbar^2}\sum_{{\mathbf}{q}}(u_{\mathbf}{q}+v_{\mathbf}{q})^2 |c_{0,{\mathbf}{q}}|^2 q_j^2 2\left(\bar{n}_{{\mathbf}{q}}+\frac12\right)\\ &\times\big[\delta(\omega_{{\mathbf}{q}}+\Omega(\beta_j/2+n))+\delta(\omega_{{\mathbf}{q}}-\Omega(\beta_j/2+n))\big], \nonumber \\ \delta_{jn} = & -\frac{\rho_0 L^3}{2\hbar^2}\sum_{{\mathbf}{q}}(u_{\mathbf}{q}+v_{\mathbf}{q})^2 |c_{0,{\mathbf}{q}}|^2 q_j^2\\ &\times \left( \frac{\cal{P}}{\omega_{{\mathbf}{q}}+\Omega(\beta_j/2+n)} +\frac{\cal{P}}{\omega_{{\mathbf}{q}}-\Omega(\beta_j/2+n)}\right), \nonumber \\ \label{GammaBEC} \gamma_{jn} = & \frac{\rho_0 L^3}{2 \hbar^2}\sum_{{\mathbf}{q}}(u_{\mathbf}{q}+v_{\mathbf}{q})^2 |c_{0,{\mathbf}{q}}|^2 q_j^2 2\left(\bar{n}_{{\mathbf}{q}}+\frac12\right)\\ &\times \left( \frac{\cal{P}}{\omega_{{\mathbf}{q}}+\Omega(\beta_j/2+n)} -\frac{\cal{P}}{\omega_{{\mathbf}{q}}-\Omega(\beta_j/2+n)}\right). \nonumber\end{aligned}$$ [In equations of motion and in the following text we omit]{} contributions from principal values present in $\kappa_j$, $\gamma_{jn}$ and $\delta_{jn}$ coefficients. One can check that their main role is renormalization of the trap $a_j$ and $q_j$ parameters due to the interactions with surrounding atomic gas. In typical experimental realizations, such effects are negligibly small. Moreover, we have verified that omitting terms containing principal values [do]{} not change the cooling rate and the final energies of the ion for small $a_j$ and $q_j$ parameters, relevant for current experiments. [ With this simplifications equations of motion read $$\label{liniowe} \begin{split} \dot{\overline{r}}_j&=\overline{p}_j/M\\ \dot{\overline{p}}_j&=- M\tilde{\omega}^2_j(t) \overline{r}_j - \overline{p}_j \sum_{n,m}C^j_n C^j_m \tilde{\eta}_{jn}\cos\left((n\!-\!m)\Omega t\right), \end{split}$$ [where $\tilde{\eta}_{jn}=\frac{2 \hbar}{\nu_j M} \eta_{jn}$, $\tilde{\omega}_j^2(t) = \omega_j^2(t)-\sum_{n,m}C^j_n C^j_m \Omega (\beta_j/2+m)\tilde{\eta}_{jn}\sin\left((n\!-\!m)\Omega t\right)$.]{} [In case of a Bose-Einstein condensate $\tilde{\eta}_{jn}$ can be calculated analytically $$\begin{split} \tilde{\eta}_{jn}=&\frac{4\sqrt{2}\pi}{3} \frac{m^{1/2}(m+M)^2}{M^{5/2}} \rho {a^j_{i}}^3\left(\frac{f(\tilde{q}_{jn})}{a^j_{i}}\right)^2\\ &\times \frac{\Omega^{3/2}}{\nu_j^{1/2}}|\beta_j/2+n|^{3/2}{\mathrm}{sign}(\beta_j/2+n) \end{split}$$ where $a^j_{i}=\sqrt{\hbar/(M\nu_j)}$, $\tilde{q}_{jn}=\sqrt{2 m \Omega |\beta_j/2+n|/\hbar}$.]{} The term $\sum_{n,m}C^j_n C^j_m \tilde{\eta}_{jn}\cos\left((n\!-\!m)\Omega t\right)$ in the above equations plays the role of the time-dependent friction force,]{} $f(\tilde{q}_{jn})$ denotes the scattering amplitude for potential $V(r)$ calculated in the first-order Born approximation. Similar procedure may be applied to equations involving expectation values of operators quadratic in position and momentum, which yield $$\begin{aligned} \label{kwadratowe} (\dot{\overline{r_j p_j}}+\dot{\overline{p_j r_j}})=&-2M\tilde{\omega}_j^2(t) \overline{r_j^2}+2\overline{p_j^2}/M \nonumber \\ -\sum_{n,m}&\left[ (\overline{r_j p_j}+\overline{p_j r_j}) \tilde{\eta}_{jn}\chi^j_{nm}(t)+2 \hbar \tilde{\mu}_{jn}\sigma^j_{nm}(t)\right],\nonumber\\ \dot{\overline{r_j^2}}=&\frac{1}{M} (\overline{r_j p_j}+\overline{p_j r_j}),\\ \dot{\overline{p_j^2}}=&-M\tilde{\omega}^2(t) (\overline{r_j p_j}+\overline{p_j r_j}) \nonumber \\ -2& \sum_{n,m} \left[ \overline{p_j^2}\tilde{\eta}_{jn}-\hbar \Omega M\!\! \left(\!\frac{\beta_j}{2}\!+\!m\!\right)\!\tilde{\mu}_{jn} \right] \chi^j_{nm}(t),\nonumber\end{aligned}$$ [where $\tilde{\mu}_{jn}=\frac{2 \hbar}{\nu_j M} \mu_{jn}$,]{} $\sigma^j_{nm}(t) = C^j_n C^j_m \sin((n-m)\Omega t)$ and $\chi^j_{nm}(t) = C^j_n C^j_m \cos((n-m)\Omega t)$. [Free term $\tilde{\mu}_{jn}$ assures that the ion energy cannot drop below the ground state energy of the secular trap even if the temperature of the atomic gas is lower. In case of a BEC it can be calculated analytically]{} $$\begin{split} \tilde{\mu}_{jn}=&\frac{2\sqrt{2}\pi}{3} \frac{m^{1/2}(m+M)^2}{M^{5/2}} \rho {a^j_{i}}^3\left(\frac{f(\tilde{q}_{jn})}{a^j_{i}}\right)^2\\ &\times \frac{\Omega^{3/2}}{\nu^{1/2}}|\beta/2+n|^{3/2}(2 \bar{n}_{\tilde{q}_n}+1). \end{split}$$ [ The form of the equations of motion is similar to classical equations describing particle in harmonic potential with time-dependent frequency and time-dependent damping. For small $a_j,q_j^2\ll 1$ (typical in experimental realizations) and sufficiently small gas density only limited number of terms in sums containing gas-dependent coefficients would be important from the point of view of the dynamics.]{} In order to solve Eq.  we use the following ansatz ${{\mathbf}v}(t) = \sum_{n=-\infty}^{\infty} {{\mathbf}w}_n e^{i \lambda \Omega t} e^{i n \Omega t}$ (similar to used in [@IdziaszekZoller2011; @Nguyen2012; @Zoller]), while for Eq.  we use ${{\mathbf}v}(t) = \sum_{n=-\infty}^{\infty} {{\mathbf}w}_n e^{i \lambda \Omega t} e^{i n \Omega t}+\sum_{n=-\infty}^\infty {{\mathbf}u}_n e^{i n \Omega t}$ where ${{\mathbf}v}(t)$ represents $(\overline{r}(t),\overline{p}(t))$ for Eq.  and $(\overline{rp}(t)+\overline{pr}(t),\overline{r^2}(t),\overline{p^2}(t))$ for Eq. , and $\lambda$ is a complex-valued characteristic exponent. Real part of $\lambda$ describes the secular frequency of oscillations of the ion in an effective trap and the imaginary part describes the cooling rate [(see Figs. \[Fig:bifurkacja\] and \[Fig:urojone\])]{}. In case of cooling ($\Im(\lambda)>0$ - discussed below) the first part of the ansatz for Eq.  goes to zero for large times and the second one represents the asymptotic solution with ${\mathbf}{u}_0$ being the average value of ${\mathbf}{v}(t)$ with respect to [the time scale given by the RF frequency $\Omega$.]{} \ \ \ VII. Cooling rates ================== Let us analyze the ion dynamics for some typical experimental parameters. Fig. \[Fig:bifurkacja\] shows the cooling rates of a Ba$^{+}$ ion immersed in an ultracold gas of Rb atoms for parameters similar to used in the experiment [@Schmid2010] (middle panel) and for a weaker trap and a higher gas density (top panel). [The bottom panel of Fig. \[Fig:bifurkacja\] presents the cooling rates of Yb$^{+}$ ion immersed in an ultracold gas of Li atoms.]{} The white color represents the unstable regions of the Paul trap without the buffer gas. In general there are three different regimes of ion dynamics immersed in an ultracold gas: (i) under-damped, (ii) over-damped harmonic motion, (iii) heating. They are marked as points $\alpha$, $\beta$ and $\gamma$, respectively in the top panel of Fig. \[Fig:bifurkacja\]. In regime (ii) the ion motion is similar to over-damped harmonic oscillator. The real part of the characteristic exponent $\lambda$ drops to zero and the energy asymptotically reaches its final value. However, there still exists the micromotion, that adds the periodic modulation of the energy. For parameters of Fig. \[Fig:bifurkacja\](a) the region of over-damped motion dominates, but this is only due to relatively weak trapping and high density of the atomic gas. In typical experimental realizations, however, the stability region corresponds mainly to the cooling behavior of the under-damped motion, see Fig. \[Fig:bifurkacja\](b). [ The time dependence of the energy for three described regimes is shown in Fig. \[Fig:energie\], for parameters corresponding to points $\alpha$, $\beta$ and $\gamma$, respectively (cf. Fig. \[Fig:bifurkacja\] (a)). In the regime represented by point $\alpha$ (upper panel) the ion energy decreases exponentially in time while the secular frequency of the trap is renormalized in the presence of cold reservoir. The inset shows the asymptotic behaviour at large $t$. Since the energy is not conserved in the presence of time-dependent potential, one can observe remaining oscillations around the final value. For point $\gamma$ (bottom panel) the motion of the ion is unstable - it gains energy exponentially from the time-dependent trap. The pink line depicts the growing amplitude of the oscillatory motion. For parameters of point $\beta$ (middle panel) one can observes the net cooling effect (green line), after averaging out over fast micromotion. However, the time-scale of cooling in this regime is much longer than for cooling in region represented by point $\alpha$.]{} ![\[Fig:energie\] [(color online). Energy of the ion (expectation value of $H_S$, blue line), average energy with respect to the period of the radio frequency modulation (green), asymptotic average energy (red), and energy amplitude (pink) for parameters of point $\alpha$ (top panel), $\beta$ (middle panel), and $\gamma$ (bottom panel) marked in Fig. \[Fig:bifurkacja\].]{} ](4.eps){width="\linewidth"} ![\[Fig:urojone\] (Color online). Imaginary part of the characteristic exponent for $q=0.3$ without (left) and with (right) the buffer gas (parameters as in the top panel of Fig. \[Fig:bifurkacja\]). [Grey line corresponds to stable solution for the Paul trap ($\Im (\lambda)=0$), dark blue line - the region of under-damped harmonic oscillator ($\Im (\lambda)>0$, $\Re(\lambda)\neq 0$), light blue line - over-damped harmonic oscillator ($\Im (\lambda)>0$, $\Re(\lambda)=0$), red line - nonstable harmonic oscillator ($\Im (\lambda)<0$).]{} ](Urojone.eps){width="\linewidth"} Fig. \[Fig:urojone\] shows the imaginary part of the characteristic exponent $\lambda$ for an isolated ion (left panel) and for the ion immersed in cold reservoir (right panel) for parameters of Fig. \[Fig:bifurkacja\](a). For values larger than zero the ion motion is being damped (regime (i) or (ii)), while for negative values the motion is unstable (regime (iii)). This figure explains the presence of three different regions of ion dynamics. In the absence of the buffer gas there are only two regions - stable and unstable and there is no cooling inside the stable region, because the imaginary part of the characteristic exponent is zero there. Introduction of the buffer gas increases $\Im(\lambda)$ and in the comparison to [the isolated ion case a new region appears (all solutions marked in blue) where the ion motion is damped, and its energy decreases.]{} There is a possibility of creating molecular states of an atom and an ion in the course of collisions. Such states have been predicted to emerge in a classical simulation of ion-atom collisions [@Vuletic2012] in the presence of the time-dependent RF potential. Ref. [@Vuletic2012] shows that during the collision the particles can be bound for relatively long time, and the work performed by the time-dependent electric field constitutes a significant source of heating, shifting the final temperatures of ions in the sympathetic cooling up to mK regime. Our formalism neglects the effect of bound states association in the atom-ion collisions, but our analysis shows that this process should not be significant in the low-energy quantum regime, because the probability of creation of a molecular complex will be significant only when the resonance condition is fulfilled. In order to verify this assumption we estimated the probability of transition to molecular complex during a single collision using time-dependent perturbation theory. For Rb-Yb$^+$ collisions and $a_{sc} = R^\ast = 307$nm the probability of association of a molecular complex in a single collision is $P = 0.007$, while for $a_{sc} = - R^\ast = - 307$nm is larger: $P = 0.127$. In the association process one or more energy quanta are transferred from the collision complex to RF field, and the probability of association strongly depends on the fulfillment of the resonance condition between the initial and final molecular states. Therefore, by appropriate selection of the trap parameters it should be possible to detune from the resonance, and finally reduce the probability of association. VIII. Concluding Remarks ======================== [ [Based on the theory of quantum stochastic processes we have developed a master equation for the system in the time-dependent external potential, in which a single trapped ion is brought into a contact with an]{} ultracold gas in a condensed or an non-condensed phase. We have investigated three different stability regimes of the ion motion. Furthermore, we have studied experimentally relevant sets of parameters and we have [calculated]{} cooling rates for Ba$^+$ ion immersed in a [Rb]{} atoms and Yb$^+$ ion immersed in a Li reservoir. [In typical experimental realizations also so called excess micromotion constitutes an additional source of heating. We plan to investigate this issue in future research.]{} ]{} Acknowledgments =============== We thank K. Jachymski, R. Gerritsma and E. Hudson for stimulating discussions and C. Sias, A. Härter and A. Krükow for providing us with experimental parameters. This work was supported by the Foundation for Polish Science International PhD Projects Programme co-financed by the EU European Regional Development Fund and National Science Centre [(ZI) (Grants No. DEC-2011/01/B/ST2/02030 (ZI) and DEC-2012/07/N/ST2/02879 (MK)).]{} Appendix: Probability of associating bound states during the collision process {#App:BoundStates} ============================================================================== In order to estimate the probability of transition to molecular states during a single atom-ion collision, we have developed a one-dimensional model assuming that atom approaches the ion along one-dimensional trajectory, which should be a good approximation to real three-dimensional scattering process in a time-dependent field [@Vuletic2012]. First, we have transformed the total Hamiltonian with the help of the Cook, Shankland, Wells transformation [@Nguyen2012], separating it into static and time-dependent parts: $$\begin{aligned} H(t) = & H_{0} + \tilde{H}(t) \\ H_0 = & \frac{\hat{p}_i^2}{2 M} + \frac{M}{2} \nu^2 x^2+ \frac{\hat{p}_a^2}{2 m} + V(|x-x_a|) \\ \tilde{H}(t) = & - M (\gamma \nu)^2 x^2 \cos \left(2 \Omega t\right) \\ & + 2 i \hbar \gamma \nu \left( x \frac{\partial}{\partial x} + \frac12 \right) \sin \left(\Omega t \right)\end{aligned}$$ The first part $H_{0}$ contains kinetic energies of the atom and ion, the static part of the Paul trap with $\nu$ denoting the frequency of the reference harmonic oscillator [@leibfried], and the atom-ion interaction is given by $V(|x|)$. The second part $\tilde{H}(t)$ contains two time-dependent terms oscillating with frequency of the RF field $\Omega$ and $2 \Omega$, respectively. The coefficient $\gamma$ $$\gamma=\frac{1}{\sqrt{2(1+\frac{2a}{q^2})}}.$$ depends on the ratio of the static and dynamic amplitudes of the RF field. We calculate the transition probability from the scattering state to the bound state of $H_0$ in the first-order time-dependent perturbation theory, treating $\tilde{H}(t)$ as perturbation [@Nguyen2012]. The initial and final states of $H_0$ can be represented as $$\Psi(x,x_a) = \sum_{n=0}^{\infty} \phi_n(x) \psi_n(x_a)$$ where $\phi_n(x)$ are the wave functions of the ion in the static trap $$\left(\frac{\hat{p}^2}{2 M} + \frac12 M \nu^2 x^2 \right) \phi_n(x) = E_n \phi_n(x)$$ and $\psi_n(x_a)$ are corresponding wave functions of the atom. We assume that initially the ion is in the ground-state of the Paul trap and the asymptotic kinetic energy of the free atom $E_a= (\hbar^2 k^2)/(2 m) < \hbar \nu$. In this case wave function of the atom for the channel $n=0$ takes the following asymptotic form at large distances: $$\psi_0(x_a) \stackrel{|x_a| \to \infty}{\longrightarrow} e^{i k x_a} + f_{+} e^{i k |x_a|} + f_{-} \textrm{sgn}(x_a) e^{i k |x_a|}$$ Here, $f_{+}$ and $f_{-}$ denote the scattering amplitudes corresponding to even and odd scattered waves, respectively. The transition probability in the first-order perturbation theory is given by Fermi’s golden rule. We sum independent contributions due to the transitions induced by the first and the second term in $\tilde{H}(t)$. Moreover, we include only the processes in which the energy quanta are transferred from the collision complex to the RF field, which lead to creation of molecular states. Hence the total probability of bound-state association in the single collision can be approximated by $$P \approx P_{1} + P_{2},$$ where $$\begin{aligned} P_{1} & = \frac{2 \pi}{\hbar} \left|\langle\Psi_i|\tilde{H}_1|\Psi_f\rangle\right|^2 \rho(E-2 \hbar \Omega) j^{-1} \\ P_{2} & = \frac{2 \pi}{\hbar} \left|\langle\Psi_i|\tilde{H}_2|\Psi_f\rangle\right|^2 \rho(E-\hbar \Omega) j^{-1} ,\end{aligned}$$ with $$\begin{aligned} \tilde{H}_1 & = - \frac12 M (\gamma \nu)^2 x^2\\ \tilde{H}_2 & = \hbar \gamma \nu \left(x \frac{\partial}{\partial x} + \frac12 \right)\end{aligned}$$ Here, $E = \hbar^2 k^2/(2 m) + \frac12 \hbar \nu$ denotes the energy of the initial scattering state $\Psi_i$, while the energy of the final state $\Psi_f$ is $E-2\hbar \nu$ and $E-\hbar \nu$ for transitions induced by $\tilde{H}_1$ and $\tilde{H}_2$ terms, respectively, $j = \hbar k/m$ is the probability flux of the atoms scattering on the ion, and $\rho(E)$ is the density of states. The density of states for energy corresponding to the resonance was determined from the numerical energy spectrum calculated by the diagonalization of the Hamiltonian $H_0$. The initial and final states were determined by numerically diagonalizing Hamiltonian $H_0$ in a box of size much larger than $R^\ast$. The secular frequency was $\nu = 2 \pi \times 1$MHz. We calculated the probability of transition to bound state for Rb-Yb$^{+}$ system, for two different scattering lengths $a_{sc} = R^\ast = 307$nm and $a_{sc} = - R^\ast = - 307$nm. The RF frequency in both cases was about $2 \pi \times 10$MHz, and its exact value was chosen in order to be resonant with a molecular level resonant with the transition induced by $\tilde{H}_2(t)$ (dominating perturbation). This should overestimate the probability of creation of a molecular state and in real experimental conditions one can always try to modify the RF frequency in order to detune from the resonance.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We propose classical equations of motion for a charged particle with magnetic moment, taking radiation reaction into account. This generalizes the Landau-Lifshitz equations for the spinless case. In the special case of spin-polarized motion in a constant magnetic field (synchrotron motion) we verify that the particle does lose energy. Previous proposals did not predict dissipation of energy and also suffered from runaway solutions analogous to those of the Lorentz-Dirac equations of motion.' author: - 'Arnab Kar[^1]  and S. G. Rajeev[^2]' title: On The Relativistic Classical Motion of a Radiating Spinning Particle in a Magnetic Field --- Department of Physics and Astronomy\ University of Rochester\ Rochester NY 14627\ USA Introduction ============ What is the classical motion of an electron in a constant magnetic field? Surprisingly, this question does not have a definitive theoretical answer even now. The complications arise when the radiation from the particle is considered. Without it, the answer is quite elementary: a helical path around the magnetic field[@EM] obtained by solving the Lorentz equation of motion.[^3] =F\_u\^ But we know that matters are not so simple in reality: an accelerated charge radiates, and the radiation emitted by it exerts a force back on it. If the energy is not replenished the path should be a spiral, the radius of the circle decreasing with the energy. Determining the force exerted on a point particle by its own field is fraught with conceptual difficulties: the force is infinite. Dirac[@Dirac38] made a major step forward by showing that this divergence can be removed by a renormalization of the mass. The resulting equation (the Lorentz-Dirac equation of motion) [^4] =F\_u\^+,= is still not the correct answer. It predicts runaway solutions: even in the absence of any external fields, a particle can accelerate away to infinity with exponentially **growing** energy. Part of the problem is that the radiative terms persist even in the absence of external forces. Also, the equation involves the third derivative of position; so it needs a new initial condition. Dirac suggested that this be chosen such that the energy remains finite forever. However, such solutions suffer from another problem: a violation of causality; the particle would accelerate **before** the external field is turned on. The textbook of Landau and Lifshitz[@LandauLifshitz] proposed (without proof) a way out of this dilemna. Replace the derivative of acceleration by its leading approximation in an expansion in powers of $\epsilon$: =F\_u\^+,= The radiative term now vanishes when the extrenal force is zero; and the equation now involves only the second order derivative of position. The free particle does not run away. Further analysis shows that the orbits in a constant magnetic field [@Spohn] and in a Coulomb field [@Rajeev] decay as expected. But why should we stop at first order in the iteration in powers of $\epsilon$? Although it works, the ad hoc prescription of Landau and Lifshitz is unsatisfactory. To get sensible solutions to the Lorentz-Dirac equation, its new degree of freedom must be continuously fine-tuned so that the trajectory stays on a submanifold of the phase space (the critical submanifold). Spohn used modern ideas in the theory of dynamical systems to determine the equation of motion projected to this critical submanifold: it is precisely the above Landau-Lifshitz equation! Thus we consider the century-old problem of determining the equation of motion of a radiating point charge to be settled finally. However, the electron has a magnetic moment as well as electric charge; radiation arises not only from the acceleration of the charge but also the precession of the magnetic moment. In the relativistic case, these are of comparable magnitude. Therefore, it is necessary to extend the Lorentz-Dirac and Landau-Lifshitz equations to the case of a spinning charged particle. Bargman, Michel and Telegdi[@BMT] proposed, more than a half century ago, an equation for a (non-radiating) classical particle with charge and magnetic moment. However, it does not seem to be the classical limit of the Dirac wave equation of the electron. The correct equation must have a canonical form whose quantization leads to the Dirac wave equation. The BMT equation does not appear to satisfy this condition. A better approach was proposed more recently by Barut and Zanghi[@BZ]. We will review their ideas, which show that a canonical quantization does lead to the Dirac equation. We will also find solutions of the Barut-Zanghi equation for a non-radiating charged particle in a constant magnetic field: itself a new result. Barut and Unal[@BU] also derived the analogue of the Lorentz-Dirac equation, after a mass renormalization. However, it suffers from the same runaway solutions of the free particle, as the spinless case. In this paper, we propose an equation of motion for a radiating spinning charged particle with gyromagnetic ratio $2$ (e.g., the electron or the muon) by applying a prescription analogous to that of Landau and Lifshitz to the Barut-Zanghi equation. We will show that, at least for spin-polarized orbits in a constant magnetic field, the orbits decay as expected physically. It should be possible to test our equations using a small (few MeV) electron synchrotron or a muon storage ring, such as the one used for the determination of $g-2$. Also, it should not be difficult to revise the predictions for an electron accelerated by a laser field[@LaserLL]. Current predictions do not include the effect of the magnetic moment. We do not attempt, in this paper, to derive our prescription from first principles as Spohn did. Non-Radiating Spinning Charged Particle ======================================= Can we talk of the motion of a particle with spin in the classical limit? At first this would appear to be impossible. But we do it all the time. The motion of electrons in synchrotron of a radius of several meters (if not kilometers) is surely classical: even if the spin were polarized, the differences in the quantum energy levels of the electron are much too small to be of significance. If we can find a lagrangian whose canonical quantization is the Dirac wave equation, it can be the basis for such a classical treatment. The lagrangian proposed by Barut and Zanghi[^5] is, L=+p\_\^-H,H=|[z]{}\^z,|[z]{}=z\^\_[0]{} Here $z$ is a Dirac spinor with four complex (not Grassmann) numbers as components. The dimensions of $z$ are that of the square root of action (=mass[\*]{}length) and that of $\tau$ is the inverse of mass, in units where $c=1$ but $\hbar\neq1.$ Then $qF_{\mu\nu}$ has dimensions of mass over length. The equations of motion are, with $\pi_{\mu}=p_{\mu}-qA_{\mu}$ $$\begin{aligned} \frac{dx^{\mu}}{d\tau}&=&\bar{z}\gamma^{\mu}z \nonumber \\ -i \frac{dz}{d\tau}&=&\gamma^{\mu}\pi_{\mu}z \nonumber \\ \frac{d\pi_{\mu}}{d\tau}&=&qF_{\mu\nu}\bar{z}\gamma^{\nu}z \nonumber \\\end{aligned}$$ Note that, $\tau$ is not proper time as $\dot{x}^{\mu}$ is not necessarily of constant length. But we have the orthogonality condition \^\_=0. These equations follow from the Hamiltonian H=|[z]{}\^z\_ and Poisson Brackets (P.B.) are obtained using {f,g}=i{-}+g\_{-} { \_,\_} =-qF\_,{ \_,x\^} =\_\^ { z,|[z]{}} =i,{ \_,z} =0={ z,x} All other pairs of P.B. are zero. We can introduce a constraint |[z]{}z=a consistent with the P.B. and the equations of motion. The quantity $a$, with the dimensions of action, sets the scale for the intrinsic angular momentum of the particle. Upon quantization, the wave function can be thought of as a function of $x$, $p_{\mu}=\frac{\hbar}{i}\frac{\partial}{\partial x^{\mu}}$; it is also a polynomial in $\bar{z}$, with $z=-\hbar\frac{\partial}{\partial\bar{z}}$ satisfying the constraint -|[z]{}=a. Thus the parameter $a$ is quantized in multiples of $\hbar$. The smallest value is $a=\hbar$ for which $\psi(x,\bar{z})=\bar{z}\psi(x)$ is a linear function. The Dirac wave equation is the eigenvalue equation for $H$, with eigenvalue $am$: i\^=m Other quantizations, for which$\psi(\bar{z})$ is a higher degree polynomial, correspond to the wave equations of Bhabha and Harish-Chandra. They could be interesting as wave equations for composite particles like hadrons[@SkyrmeWaveEqn]. It is convenient to introduce the velocity and spin variables v\^=|[z]{}\^z,S\^=|[z]{}\[\^,\^\]z so that $$\begin{aligned} \frac{dx^{\mu}}{d\tau}&=&v^{\mu} \nonumber \\ \frac{dv^{\mu}}{d\tau}&=&4S^{\mu\nu}\pi_{\nu} \nonumber\\ \frac{d\pi_{\mu}}{d\tau}&=&qF_{\mu\nu}v^{\nu} \nonumber \\ \frac{dS^{\mu\nu}}{d\tau}&=&v^{\nu}\pi^{\mu}-v^{\mu}\pi^{\nu} \nonumber\\\end{aligned}$$ These equations describe a relativistic analogue of the isotropic precessing top of Poinsot [@LandauLifshitzMech], as we explain in the Appendix. The hamiltonian H=v\^\_ describes a coupling to the external field which causes a precession of the velocity and spin. For a constant electromagnetic field, $K=\pi\cdot\pi-qS^{\mu\nu}F_{\mu\nu}$ is conserved as well: =2qF\_\^v\^-2qF\_v\^\^=0. This is the classical analogue of the square of the Dirac operator, corresponding to the gyromagnetic ratio $2.$ We will see later that it is proportional to the angular momentum in the plane of the electromagnetic field. A Non-Radiating Spinning Particle in a Constant Magnetic Field ============================================================== Now let us turn to solving the equations of motion in the case of a magnetic field pointed along the third direction: $F_{12}=-F_{21}>0$; and all other components are zero. Before we solve the problem including radiation damping, we must recover the solution without radiation. Even this appears to be new: Barut and collaborators only solved the case of a free particle. We consider only orbits that lie in the $12$ plane. That is, we restrict to the case where all the tensor indices $\mu,\nu,\rho\cdots$ are in the $012$ subspace. It is useful to introduce the ‘dual’ variables $$\begin{aligned} q F_{\mu\nu}&=&\epsilon_{\mu\nu\rho}F^{\rho} \nonumber\\ S^{\mu\nu}&=&\epsilon^{\mu\nu\rho}S_{\rho} \nonumber \\ S_{\rho}&=&\frac{1}{2}\epsilon^{\mu\nu\rho}S_{\mu\nu} \nonumber\\\end{aligned}$$ For a constant magnetic field, only the time component $F^{0}$ is non-zero. We can assume it to be positive without loss of generality. Then $$\begin{aligned} \frac{d\pi_{\mu}}{d\tau}&=&\epsilon_{\mu\nu\rho}v^{\nu}F^{\rho} \nonumber\\ \frac{dv^{\mu}}{d\tau}&=&-4\epsilon^{\mu\nu\rho}S_{\nu}\pi_{\rho} \nonumber\\ \frac{dS_{\mu}}{d\tau}&=&-\epsilon_{\mu\nu\rho}v^{\nu}\pi^{\rho} \nonumber\\\end{aligned}$$ If we define $$\lambda_{\mu}=\frac{1}{2}(S_{\mu}+\frac{1}{2}v_{\mu}),\quad\rho_{\mu}=\frac{1}{2}(S_{\mu}-\frac{1}{2}v_{\mu})$$ they reduce to $$\begin{aligned} \frac{d\pi_{\mu}}{d\tau}&=&2\epsilon_{\mu\nu\rho}\left[\lambda^{\nu}-\rho^{\nu}\right]F^{\rho} \nonumber \\ \frac{d\lambda^{\mu}}{d\tau}&=&-2\epsilon^{\mu\nu\rho}\lambda_{\nu}\pi_{\rho} \nonumber \\ \frac{d\rho_{\mu}}{d\tau}&=&2\epsilon_{\mu\nu\rho}\rho^{\nu}\pi^{\rho} \nonumber \\\end{aligned}$$ Spin-Polarized Orbits --------------------- There is a special class of ‘left-handed’ orbits for which $\rho_{\mu}=0$. (Also, the opposite case with $\lambda_{\mu}=0$.) For these, the spin and velocity are synchronized in the following way: S\_=\_v\^ In the non-relativistic limit, this means the spin is pointed along the magnetic field (times $q$). Since these solutions are periodic (not chaotic), orbits of this kind in a synchrotron should be of special interest. If the gyromagnetic ratio is slightly different from two, the spin would depart from this orientation, which is the principle behind a famous experiment for the measurement of $g-2$ for the muon. The equations reduce to $$\begin{aligned} \frac{d\pi_{\mu}}{d\tau}&=&\epsilon_{\mu\nu\rho}v^{\nu}F^{\rho} \nonumber \\ \frac{dv^{\mu}}{d\tau}&=&-2\epsilon^{\mu\nu\rho}v_{\nu}\pi_{\rho} \nonumber \\\end{aligned}$$ The tensor $\epsilon_{\mu\nu}^{\sigma}=\epsilon_{\mu\nu\rho}\eta^{\rho\sigma}$ form the structure constants of the three dimensional Lorentz Lie algebra $SO(1,2)$. This is analogous to the way that the Levi-Civita tensor gives the stucture constants of the Lie algebra of rotations in Euclidean three dimensional space. Denote the Lie product of $SO(1,2)$ as a relativistic cross product (uw)\_=\_\^u\_w\_ This differs from the familiar cross product in Euclidean geometry by some signs: (uw)\_[0]{}=u\_[1]{}w\_[2]{}-u\_[2]{}w\_[1]{},(uw)\_[1]{}=-u\_[2]{}w\_[0]{}+u\_[0]{}w\_[2]{},(uw)\_[2]{}=-u\_[0]{}w\_[1]{}+u\_[1]{}w\_[0]{} Then our equations are a relativistic version of the equation for a precessing top: =vF,=-2v It follows that v=0 so that $v\cdot v=a^{2}$ is conserved; $a$ is some number (like Planck’s constant) with the dimensions of action; the spin is of magnitude $\frac{a}{2}.$ Also, the parameter $\tau$ is proportional to proper time in this special case. The energy $\pi_{0}=E$ and K=\_\^-v\^F\_ are conserved[^6]. Now, \_[0]{}=2(v\_[2]{}\_[1]{}-v\_[1]{}\_[2]{}) and \_[0]{}\^[2]{}=4(-\^[2]{}) Re-expressing the r.h.s. in terms of conserved quantities, \_[0]{}=2 Weierstrass elliptic function satisfy this differential equation. The remaining components of $\pi,v$ can be obtained by solving the other constraint equations. It is also possible to understand the solution as an integrable hamiltonian system, a relativistic analogue of the precessing top. (See the Appendix) The Equation of Motion for Radiating Spinning Particles ======================================================= Barut and Unal followed the method of Dirac to derive the analogue of the Lorentz-Dirac equation of motion including the self-force. Only the equation for $\pi_{\mu}$ is affected: =qF\_v\^+q\^2{ -} The last term is the contribution from self-interactions, after a renormalization. It is a singular perturbation i.e. $q^{2}$ is small and it changes the order of the equations. It has runaway solutions, just like the Lorentz-Dirac equation. Let us apply to it a prescription like that of Landau and Lifshitz, to get a second order equation of motion. Naively, the Landau-Lifshitz prescription amounts to using \^=4\^\_+4S\^ and replacing $\dot{\pi}$ by its zeroth order contribution: \^4\^\_+4S\^qF\_v\^=4\_+4S\^qF\_v\^ But this can’t be right: for a free particle ($F_{\mu\nu}=0$), the radiation reaction force should be zero. Dropping the terms not proportional to $F_{\mu\nu}$, =qF\_v\^+S\^F\_v\^ Along with the previous equations for $v$ and $S,$ this system of ordinary differential equations is our proposal for the equation of motion of a spinning radiating charged particle. We must verify that the solutions are physically sensible: there should not be runaway solutions. We will study the case of polarized orbits in a constant magnetic field and show that the energy decreases with time. The ultimate test of our proposal must be experimental, especially since we do not yet have a derivation from first principles. Radiating Spin-Polarized Particles in a Magnetic Field ====================================================== Now let us turn to our proposed equations for a radiating spinning charged particle. Again, we will look for solutions in the special case where the spin is polarized S\_=v\_ Since the radiative terms do not change the equations for $v^{\mu},\, S^{\mu\nu}$ this condition is still preserved by time evolution i.e. $\dot{S}^{\mu}=\frac{1}{2}\dot{v}^{\mu}$. Unpolarized orbits are chaotic and the radiation damping is likely to be even larger. (See Appendix) The equations then reduce to $$\begin{aligned} \frac{d\pi_{\mu}}{d\tau}&=&qF_{\mu\nu}v^{\nu}+\frac{4q^{3}}{3v\cdot v}\left[\eta_{\mu\nu}-\frac{v_{\mu}v_{\nu}}{v\cdot v}\right]\epsilon^{\nu\rho\alpha}v_{\alpha}F_{\rho\sigma}v^{\sigma} \nonumber \\ \frac{dv^{\mu}}{d\tau}&=&-2\epsilon^{\mu\nu\rho}v_{\nu}\pi_{\rho} \nonumber \\\end{aligned}$$ We know that, $v\cdot\dot{v}=0$ and $v\cdot v=v_{0}^{2}-(v_{1}^{2}+v_{2}^{2})=a^{2}.$ =\_v\^F\^+F\^ where = In terms of relativistic cross products =v,=-2v Thus, even the radiation damping terms have a simple meaning in terms of the relativistic cross product of the Lie algebra $SO(1,2).$ Note that H=v is still conserved and $m=\frac{H}{a}$ continues to have the physical meaning of mass. But energy $E=\pi_{0}$ and $K$ are no longer conserved. The evolution of $v_{0}$ continues to be determined by the equation we derived earlier. Thus we get the system $$\begin{aligned} \frac{dE}{d\tau}&=&-\gamma F_{0}(v_{0}^{2}-a^{2}) \nonumber \\ \frac{dK}{d\tau}&=&2\gamma F_{0}\left[a^{2}E-Hv_{0}\right] \nonumber \\ \dot{v}_{0}&=&2\sqrt{\left[v_{0}^{2}-a^{2}\right]\left[E^{2}-K+v_{0}F_{0}\right]-\left[H-Ev_{0}\right]^{2}} \nonumber \\\end{aligned}$$ Since $E$ and $K$ are no longer constants, the solution is no longer an elliptic function. Also, noting that $F_{0}>0$ in our convention and $v_{0}^{2}-a^{2}=(v_{1}^{2}+v_{2}^{2})>0$, we see that energy is monotonically decreasing. Acknowledgement {#acknowledgement .unnumbered} =============== We thank Gregory Bentsen, V. Sreedhar and Tamar Friedman for discussions. This work was supported in part by a grant from the US Department of Energy under contract DE-FG02-91ER40685. Appendix: Hamiltonian Formalism {#appendix-hamiltonian-formalism .unnumbered} =============================== The Barut-Zanghi equations for a spinning charges particle follow from the P. B. forming the Lie algebra $SO(2,3)$: $$\begin{aligned} \left\{ v^{\mu},v^{\nu}\right\} &=& 4S^{\mu\nu} \nonumber \\ \left\{ S^{\mu\nu},v^{\rho}\right\} &=& \eta^{\mu\rho}v^{\nu}-\eta^{\nu\rho}v^{\mu} \nonumber \\ \left\{ S^{\mu\nu},S^{\rho\sigma}\right\} &=& \eta^{\mu\rho}S^{\nu\sigma}+\eta^{\nu\sigma}S^{\mu\rho}-\eta^{\nu\rho}S^{\mu\sigma}-\eta^{\mu\sigma}S^{\nu\rho} \nonumber \\\end{aligned}$$ Using previous relations we can also find the following P.B. : $$\begin{aligned} \left\{ v_{\mu},v_{\nu}\right\} &=& 4S_{\mu\nu} \nonumber \\ \left\{ S_{\mu},S_{\nu}\right\} &=& S_{\mu\nu} \nonumber \\ \left\{ S_{\mu},v_{\nu}\right\} &=& \epsilon_{\mu\nu\sigma}v^{\sigma} \nonumber \\\end{aligned}$$ which generalizes the rotation algebra. When the third axis is ignored, this reduces to $SO(2,2)\approx SO(1,2)\oplus SO(1,2)$. This splitting can be seen by in terms of the linear combinations $\lambda_{\mu},\rho_{\mu}:$ $$\begin{aligned} \left\{ \lambda_{\mu},\lambda_{\nu}\right\} &=& \epsilon_{\mu\nu\sigma}\lambda^{\sigma} \nonumber \\ \left\{ \rho_{\mu},\rho_{\nu}\right\} &=& \epsilon_{\mu\nu\sigma}\rho^{\sigma} \nonumber \\ \left\{ \lambda_{\mu},\rho_{\nu}\right\} &=& 0 \nonumber \\\end{aligned}$$ We have also the P.B., $$\begin{aligned} \left\{ \lambda_{\mu},\pi_{\nu}\right\} &=& 0 \nonumber \\ \left\{ \rho_{\mu},\pi_{\nu}\right\} &=& 0 \nonumber\\ \left\{ \pi_{\mu},\pi_{\nu}\right\} &=& -\epsilon_{\mu\nu\rho}F^{\rho} \nonumber \\\end{aligned}$$ The hamiltonian is then H=2\[-\] Non-Radiating Spin-Polarized Orbits in Constant Magnetic Field {#non-radiating-spin-polarized-orbits-in-constant-magnetic-field .unnumbered} -------------------------------------------------------------- Since v\^[0]{}= and $\pi_{0}=E$ is conserved we only have four co-ordinates in the phase space: $\pi_{a},v_{a}$. The P.B. are { \_[1]{},\_[2]{}} =-F\_[0]{} and { v\_[1]{},v\_[2]{}} =2 The latter corresponds to the standard symplectic form on hyperbolic space. It is not hard to find canonical co-ordinates $$\{\frac{\pi_{1}}{\sqrt{F_{0}}},\frac{\pi_{2}}{\sqrt{F_{0}}}\}\rightarrow\{\sqrt{2p}\cos\theta,\sqrt{2p}\sin\theta\},\\ \{v_{1},v_{2}\}\rightarrow\{a\,\sinh\theta\,\cos\phi,a\,\sinh\theta\,\sin\phi\}$$ so that- { p,} =1,p=(\_[1]{}\^[2]{}+\_[2]{}\^[2]{}),= { v\_[0]{},} =1,= Note that $J=p+\frac{1}{2}v_{0}=\frac{K}{2F_{0}}$ is the generator of rotations; it is a conserved quantity corresponding to angular momentum. The hamiltonian is then H=Ev\_[0]{}-(-) Since pd+v\_[0]{}d=Jd+v\_[0]{}d,=(-) we can use the pairs $J,\theta$ and $v_{0},\alpha$ as canonical variables: H=Ev\_[0]{}-2 The equation of motion is =-=-22 Thus \^[2]{}=4F\_[0]{}(2J-v\_[0]{})-4(H-Ev\_[0]{})\^[2]{} which is the Weierstrass equation. Because of its polarization, this also determines the spin. Also ==-2 Thus = which determines $\theta$ also in terms of elliptic integrals. To determine the position variable, we note the that =0 This leads to the conserved quantity (corresponding to translation invariance) X\_=\_-\_x\^F\^ Thus (in a co-ordinate system in which $F_{0}>0,F_{a}=0$ and $X_{c}=0$) we have x\_[b]{}=-\_[bc]{}\_[c]{} In other words, knowing $\pi_{1},\pi_{2}$ is the same knowing the spatial orbit, except for a rotation by $90^{\circ}$ and a scaling by $F_{0}.$ In polar co-ordinates this amounts to knowing $p=J-\frac{1}{2}v_{0}$ and $\theta$ as functions of $\tau$, a kind of elliptic curve: \^[2]{}=2F\_[0]{}p-(H-2EJ+2Ep)\^[2]{},=+ It is not correct to interpret the $x$ as the position of the electron: due to the phenomenon of zitterbewegung. Even for a free particle, $x^{\mu}(\tau)$ is a spiral and not a straightline. Barut et. al. have shown how to recover a center of mass variable for the electron from this complicated solution. A Symplectic Reduction of the General Planar Orbit {#a-symplectic-reduction-of-the-general-planar-orbit .unnumbered} -------------------------------------------------- The general case of planar but not spin-polarized orbits does not appear to be integrable. It corresponds to chaotic motion of an electron in a synchrotron. We can use the conserved quantities \_[0]{}=E, a\^[2]{}, b\^[2]{} to reduce it to a conservative system with three degrees of freedom: the phase space is a plane ($\pi_{a}$) times a pair of hyperboloids parametrized by ($\lambda_{\mu},\rho_{\mu}$). It is easy to check that { p,} =1 with p=\_[a]{}\_[a]{}, = Also, $$\begin{aligned} \left\{ \lambda_{0},\phi\right\} =1,\quad\phi=\arctan\frac{\lambda_{2}}{\lambda_{1}} \nonumber \\ \left\{ \rho_{0},\chi\right\} =1,\quad\chi=\arctan\frac{\rho_{2}}{\rho_{1}} \nonumber \\\end{aligned}$$ In terms of these $$\begin{aligned} H&=&2E(\lambda_{0}-\rho_{0})-2(\lambda_{a}-\rho_{a})\pi_{a} \nonumber \\ &=&E(\lambda_{0}-\rho_{0})-\sqrt{2F_{0}\left[\lambda_{0}^{2}-a^{2}\right]p}\cos[\phi-\theta]+\sqrt{2F_{0}\left[\rho_{0}^{2}-b^{2}\right]p}\cos[\chi-\theta] \nonumber \\\end{aligned}$$ Another conserved quantity is the generator of rotations[^7] J=p+\_[0]{}+\_[0]{} The reduced system has three degrees of freedom and two conserved quantities $H,J$ : just one conserved quantity short of an integrable system. When $\lambda_{0}=a$ (or $\rho_{0}=b$ ) one degree of freedom decouples and we get an integrable system. Circular orbits correspond to the case $\phi=\chi,\lambda_{0}+\rho_{0}=\mathrm{constant}.$ It may be more convenient to use $\alpha=\phi-\theta,\beta=\chi-\theta,\theta$ as the independent position variables. Then p\_d+p\_d+p\_d=pd+\_[0]{}d+\_[0]{}d=pd+\_[0]{}d\[+\]+\_[0]{}d\[+\] $$\begin{aligned} p_{\alpha}&=&\lambda_{0} \nonumber \\ p_{\beta} &=& \rho_{0} \nonumber \\ p_{\theta} &=& p+\lambda_{0}+\rho_{0}=J \nonumber \\ \end{aligned}$$ Then we get a system with two degrees of freedom with hamiltonian H=E(\_[0]{}-\_[0]{})-+ This appears to be a chaotic system. In general, the information about the polarization of spin in the initial condition will be lost rapidly and the particle beam will look unpolarized. For the special case $\rho_{0}=0=b$ the spin remains polarized and solutions can be obtained in terms of elliptic functions as we saw above. [11]{} J. D. Jackson, *Classical Electrodynamics,* Wiley(1998); F. Rohrlich, *Classical Charged Particles,* Westview Press (1994); P. A. M. Dirac, Proc. Roy. Soc. Lond A167 (1938) 148 L. D. Landau and E. M. Lifshitz, *The Classical Theory of Fields* Butterworth-Heinemann (1982); especially Section 76 H. Spohn, Europhys. Lett. 50 287 (2000); H. Spohn, *Dynamics of Charged Particles And Their Radiation Field* , Cambridge University Press (2004) S. G. Rajeev Ann. Phys. 323, 2654 (2008) V. Bargmann, L. Michel, and V. L. Telegdi, Phys. Rev. Lett. 2, 435 (1959) A. O. Barut and N. Zanghi, Phys. Rev. Lett. 52, 20092012 (1984) A. O. Barut and N. Unal, Phys. Rev. A 40, 54045406 (1989) Y. Hadad, L. Labun, J. Rafelski, N. Elkina, C. Klier and H. Ruhl, arxiv/1005.3980 S. G. Rajeev, Ann. Phys. 323, Pages 2873 (2008) L. D. Landau and E. M. Lifshitz, *Mechanics*, Butterworth-Heinemann (1982) [^1]: [email protected] [^2]: Also at the Department of Mathematics [^3]: In units with $c=1$. $q$ and $m$ are the charge and mass respectively. Also, $u^{\mu}=\frac{dx^{\mu}}{d\tau}$ is the four-velocity and $\tau$ is proper time. Note the constraint $\eta_{\mu\nu}u^{\mu}u^{\nu}=1$ and its derivative $\eta_{\mu\nu}u^{\mu}\frac{du^{\nu}}{d\tau}=0.$ [^4]: The last term includes a projection operator that ensures that the constraint $\eta_{\mu\nu}u^{\mu}\frac{du^{\nu}}{d\tau}=0$ is satisfied. [^5]: They use a slightly different system of units. [^6]: $K$ is proportional to angular momentum. [^7]: This is linearly related to the quantity $K$ we introduced earlier: $$K=2F_{0}J$$
{ "pile_set_name": "ArXiv" }
--- author: - '\' bibliography: - 'references.bib' title: 'Cross-Layer Design of Wireless Multihop Networks over Stochastic Channels with Time-Varying Statistics' --- Introduction {#sec:intro} ============ Network Utility Maximization (NUM) is a very popular tool in the communications research community, for cross-layer design and optimization of wireless networks. Typically, a utility function is assigned to each network flow (source-destination pair), and the sum of all utilities over the network is maximized, subject to network stability constraints. Most approaches in literature applying NUM, consider ideal or stationary and ergodic wireless channels. However, under realistic conditions, fading occurring in wireless channels hampers the performance in wireless communications, leading to stochastic, i.e. time and space varying, random (thus unknown), wireless link capacities possibly bearing time-varying statistics. Therefore, it becomes necessary to reformulate/redesign and solve the basic NUM problem for incorporating realistic stochastic wireless channel conditions by addressing non-stationarity issues, and also transient phenomena occurring when the network operates for a finite time interval. This paper aims at treating this problem by proposing a novel optimization framework for joint congestion control, routing, scheduling and power control based on NUM, under stochastic possibly non-stationary Long Term Fading (LTF) or Short Term Fading (STF) wireless channels. Congestion control determines the optimal sources’ data production rates, routing determines the optimal routes of the flows within the network, while the set of transmitting links and their corresponding transmission powers are chosen based on scheduling and power control. The proposed optimization framework refers to a finite network operation and it is developed for two cases, where the first one deals with non-orthogonal access to the wireless medium and the second deals with orthogonal access to the wireless medium. In the first case, scheduling is performed via power control, while in the latter case, scheduling and power control are separated. On the one hand, non-orthogonal access leads to a very hard to solve cross terminal power control/scheduling problem in the physical layer. On the other hand, orthogonal access to the medium allows for more efficient and distributed power control in the physical layer, although introducing the need of the NP-hard computation of all the independent sets of the network graph for scheduling. Importantly, in the case of orthogonal access, the role of power control is to further boost link capacity and consequently source rates by exploiting good channel states and to save energy when the channel state is destructive for the transmitted signal. The structure of the rest of the paper is as follows. Section \[sec:relatedwork\] presents the related literature while summarizing the basic contributions of this paper. Then, Section \[sec:channelmodels\] describes the considered STF and LTF channel models derived with the use of stochastic differential equations, while Section \[sec:sysmodel\] presents the considered system model. Sections \[sec:nonortho\] and \[sec:orthogonal\] focus on the analysis and solution of the proposed optimization framework in the cases of non-orthogonal and orthogonal access to the wireless medium respectively. Finally, in Section \[sec:simulation\] numerical results are presented to evaluate the proposed approach and Section \[sec:conclusions\] concludes the paper. Related Work & Contributions {#sec:relatedwork} ============================ Several works exist in the literature targeting at incorporating the stochastic wireless channels in the NUM problem’s formulation and solution. In [@chiang05], the channel quality is expressed via the SIR (Signal-to-Interference-Ratio), since the latter is affected by interference due to parallel transmissions and LTF. It is assumed that the LTF parameters are deterministic and slowly varying, allowing for the algorithms which perform joint congestion and power control to converge in the meantime of their change. In [@papandriopoulos08], [@fischione11], the case of composite fading (LTF and STF) is examined, considering channel conditions that vary faster than the algorithm’s convergence, via the use of outage-probabilities in the NUM problem formulation. In [@papandriopoulos08], [@fischione11], STF follows Rayleigh and Nakagami distributions respectively, where however the statistics of the distributions remain invariant. A different approach is followed in [@neill09; @firouzabadi10], where NUM is extended to Wireless NUM (WNUM), with random channel conditions. WNUM leads to policies for controlling the network by responding optimally to the change of the channel state, based on random samples without a priori knowledge of the wireless channels’ statistics. Although the statistics of the channel may be unknown, the latter is considered as stationary and ergodic. In the sequel, in [@marques], NUM is employed to perform, joint congestion control, power control, routing and scheduling assuming that the channel fading process is stationary and ergodic. Similarly in [@chen06; @huang13], joint congestion control, routing and scheduling is performed in the framework of NUM while assuming that there is a finite number of channel states. In these works, the network functions in time slots, where during each time slot the channel state remains stable and changes randomly and independently on the boundary of time slots. Finally, in [@chen11] the convergence of primal-dual algorithms for solving NUM is studied under wireless fading channels with time-varying parameters (and thus statistics). Time-varying statistics of wireless channels lead to time-varying optimal solutions of the NUM problem necessitating the study of how well the solution algorithms track the changes in the optimal values. However, it is assumed that the channel fading parameters vary following a Finite State Markov Chain. In a nutshell, in the existing body of research work in literature, the wireless channel modeling in the framework of NUM is characterized by one or combinations of the following assumptions: (a) The channel process is stationary and ergodic. (b) The statistics of the wireless channel are fixed in time (and known), or vary at slower rate than the one of the network control algorithms’ convergence. (c) The statistics of the channel change according to a Finite State Markov Chain. All previous approaches are not capable of capturing and tracking complex time and space variations in the propagation environment of realistic systems [@olama10]. In [@olama10], wireless channel models for both LTF and STF are introduced based on Stochastic Differential Equations (SDEs) [@oksendal03] in order to capture higher order dynamics of the wireless channel. In this case the wireless channel is modeled via stochastic processes which may have time-varying statistics. By means of SDEs, it is possible to express an LTF or an STF channel capturing both space and time variations [@olama10], [@menemenlis99], as it will be described in detail in Section \[sec:channelmodels\]. In our paper, the NUM problem is reformulated and solved using the SDE model to capture the wireless channel state. Emphasis is placed on LTF especially for demonstration purposes. Due to the possible non-stationarity of the wireless channel we cannot formulate the NUM problem based on the stationary mean values of the involved optimization variables (e.g. [@marques], [@chen06], [@stai14]). On the contrary we will adopt the stochastic optimal control problem’s formulation [@fleming06] based on expected values over time integrals, thus also allowing for the consideration of a finite time duration of the network’s operation. In [@stai15], we proposed a preliminary version of this approach focused on congestion and power control in the case of orthogonal access to the wireless medium. The basic contributions of this paper can be summarized as: - We develop a framework for the cross-layer design and control of the operation of wireless multihop networks, i.e. for congestion and power control, routing and scheduling, over wireless channels (LTF or STF) that are stochastic but not necessarily stationary. - The proposed problem formulation adopts a (more realistic) finite duration of the network’s operation (e.g. corresponding to the case of finite battery levels of the wireless nodes or of finite flows) where the wireless channel may still operate in a transient state, even though there may exist a limiting stationary distribution. The extension to an infinite duration of the network’s operation is discussed. - Utilities are not necessarily convex functions but adopt a more general form (more specifically a continuous differentiable one), contrary to the related papers in literature (e.g. [@marques], [@chen06], [@stai14]) that assume convex utility forms. This fact is important since it allows for addressing the case of real time traffic modeled, for example, by sigmoidal utilities [@tychogiorgos13]. Zero duality gap is analytically proven in this case of general utilities, following a technique that leverages from the wireless channels’ continuous stochastic modeling. - The proposed problem formulation (expected values over time integrals) allows for the adoption of time-varying utility functions. This serves the purpose of evolving users’ preferences/needs and is also aligned with the finite network duration, i.e. nodes may desire to produce significantly less data close to the end of the network operation. To the best of our knowledge, this fact has not been addressed in the literature. - Power control is shown to further boost link capacity and thus source rates by exploiting good channel conditions while saving energy in case of destructive channel conditions. - Finally, we prove an interesting theorem in the case of LTF, elucidating a basic advantage with respect to the optimal users’ utilities when exploiting the random channel fluctuations, compared to the conventional NUM problem formulation (e.g. [@chen06], [@stai14]). Specifically, it is proven that, contrary to what is possibly expected, a higher value of the diffusion coefficient of the wireless channels’ power loss leads to higher optimal sum of users’ utilities, fact that cannot be captured by the conventional NUM modeling approach. We also show via numerical evaluations that a higher diffusion coefficient achieves simultaneously a reduced power consumption leading to energy efficiency. This result emphasizes the importance of utilizing a more realistic power loss model such as the one of an SDE, as opposed to the mean power loss model used in the conventional NUM problem formulation. Background on Wireless Channel Modeling via SDEs {#sec:channelmodels} ================================================ The objective of this section is to briefly describe the SDE-based LTF and STF channel models, developed in the literature, and their assumptions, upon which, the optimization problems of the following sections will be formulated and solved. Long Term Fading (LTF) {#sec:ltf} ---------------------- LTF consists of path loss and shadowing [@goldsmith05]. Path loss is due to the dissipation of the transmitted power and the effects of the propagation channel, while shadowing is caused by obstacles between the transmitter and the receiver. LTF depends on the geographical area and occurs in sparsely populated or suburban areas. Before describing the dynamic in time LTF model for the wireless channels [@olama10], we recall the conventional LTF model, where the power loss $PL$ along a given link $(i,j)$ between nodes $i$ and $j$ in Euclidean distance $d_{ij}$, is given [@goldsmith05]: $$PL(d_{ij})[dB]=\overline{PL}(d_{0})[dB]+ 10\gamma \log \big(\frac{d_{ij}}{d_{0}} \big)+\tilde{Z}, ~d_{ij}\geq d_0, \label{eq:pathloss}$$ where $\gamma$ is the power loss exponent and depends on the wireless propagation medium, $d_{0}$ is the reference distance, $\overline{PL}(d_{0})$ is the expected power loss on the reference distance, and $\tilde{Z}\sim \mathcal{N}(0;\sigma^2)$, is a gaussian random variable with zero mean and variance $\sigma^2$, used to model any uncertainty in the propagation environment. Note that the statistics (mean (denoted as $\overline{PL}(d_{ij})$) and variance) of the conventional LTF model are invariant in time. In the following, we describe the extension of the LTF model to dynamically changing conditions in time, as it is developed in [@olama10]. Specifically, the random variable, $PL(d_{ij})[dB]$, of Eq. (\[eq:pathloss\]), becomes a stochastic process denoted as $\{X_{ij}(t)\}_{t\geq t_0}$ (\[dB\]), where $t$ represents time. Time dependence is used to capture time variations of the propagation environment due to e.g. movement of objects and scatterers in the area surrounding the network. In a similar spirit with Eq. (\[eq:pathloss\]), $\{X_{ij}(t)\}_{t\geq t_0}$, represents the power lost by the signal during a transmission from $i$ to $j$ at a particular distance $d_{ij}$. Although, $\{X_{ij}(t)\}_{t\geq t_0}$ depends on the distance $d_{ij}$, we do not explicitly model this dependence as the network considered is static. In [@olama10], $\{X_{ij}(t)\}_{t\geq t_0}$, $\forall (i,j)$, are modeled as solutions of mean reverting linear SDEs, given as: $$\begin{aligned} dX_{ij}(t)=\beta_{ij}(t)\big(\gamma_{ij}(t)-X_{ij}(t)\big)dt+\delta_{ij}(t)dW_{ij}(t),~X_{ij}(t_0)\sim \mathcal{N}(\overline{PL}(d_{ij})[dB];\sigma^2), \label{eq:sdepathloss}\end{aligned}$$where $\{W_{ij}(t)\}_{t\geq t_0}$, $\forall (i,j)$, are independent standard Brownian motions defined over a filtered probability space $(\Omega,\mathcal{F},\{\mathcal{F}_t\}_{t\geq t_0},\mathbf{P})$ and each one being independent of the corresponding $X_{ij}(t_0)$. $\{\mathcal{F}_t\}_{t\geq t_0}$ is the filtration produced by $X_{ij}(t_0)$, $\forall (i,j)$, and the Brownian motions themselves. For each $(i,j)$, $\gamma_{ij}(t)$ is the power loss level $X_{ij}(t)$ is attracted to, $\beta_{ij}(t)$ is the positive speed of this adjustment and finally, $\delta_{ij}(t)$ is the diffusion coefficient of the SDE, determining the “noise" of the channel. The parameters $\beta_{ij}(t),\gamma_{ij}(t),\delta_{ij}(t)$, $\forall (i,j)$, are assumed to be deterministic and can be estimated directly from signal measurements following the approaches in [@olama06], [@olama09], [@charalambous08], which can be implemented online, i.e. while receiving the signal measurements. The existence of a strong solution to the SDE (\[eq:sdepathloss\]) is satisfied if the relation $\int\limits_s^T \left\{\beta_{ij}(t)\left|\gamma_{ij}(t)\right|+\delta _{ij}^2(t)\right\}dt < \infty , \forall (i,j)$ holds [@oksendal03]. The time dependent attenuation coefficient (in squared magnitude) equals to: $a_{ij}(t)=e^{-\frac{\ln10}{10}X_{ij}(t)}=e^{K X_{ij}(t)},~\forall (i,j),~K=-\frac{\ln10}{10}$. In [@olama10], it is shown that when all the parameters of the SDE (\[eq:sdepathloss\]) are time independent, its solution tends to the conventional LTF model (Eq. (\[eq:pathloss\])) as $t\rightarrow \infty$ (which is stationary). In general when the parameters of the SDE (\[eq:sdepathloss\]) change with time, $X_{ij}(t)$ is gaussian with time-varying statistics and a stationary distribution may not exist. Short Term Fading (STF) {#sec:stf} ----------------------- In a similar spirit as LTF, in [@menemenlis99], [@olama09], [@olama10] a stochastic model for STF wireless channels has been developed, alleviating the assumption of stationarity. This kind of signal fading is due to the constructive and destructive addition of multipath components [@goldsmith05] created from reflections, diffractions and scattering and usually occurring in densely built-up areas. The statistics of the STF models usually applied in the literature (e.g. Rayleigh, Nakagami, Ricean, etc. [@papandriopoulos08], [@goldsmith05]), are assumed constant over local areas (i.e. at a microscopic level) [@menemenlis99]. However, STF wireless channels are of stochastic nature with time varying statistics mainly due to the continuous and arbitrary change of the propagation environment if the transmitter, the receiver or objects between them move. The latter is the main reason why in this paper, we adopt a stochastic process with time varying statistics for modeling STF channels. For the models developed in [@menemenlis99], [@olama09], the inphase and quadrature components of the wireless fading channels are assumed conditionally uncorrelated gaussian random variables (thus conditionally independent). In the case of flat fading, the multipath components are not resolvable and can be considered as a single path. Then, the inphase, $I$, and quadrature, $Q$, components over one link (e.g. $(i,j)$) can be realized as: $$\begin{aligned} dX_{I}(t)=A_I(t) X_{I}(t)dt+B_I(t) dW_I(t),~X_{I}(t_0),~ I(t)=C_I X_{I}(t),\nonumber\\ dX_{Q}(t)=A_Q(t) X_{Q}(t)dt+B_Q(t) dW_Q(t),~X_{Q}(t_0),~ Q(t)=C_Q X_{Q}(t),\nonumber \label{eq:sdestf}\end{aligned}$$where $X_{I}(t),~X_{Q}(t)$ are the state vectors of the inphase and quadrature components and $\{W_I(t)\}_{t\geq t_0}$, $\{W_Q(t)\}_{t\geq t_0}$ are independent standard Brownian motions corresponding to the inphase and the quadrature components respectively, defined over a filtered probability space $(\Omega,\mathcal{F},\{\mathcal{F}_t\}_{t\geq t_0},\mathbf{P})$. The same model describes every link with different parameter values $A_I(t),~A_Q(t)$, $B_I(t)$, $B_Q(t)$, $C_I,~C_Q$ and (independent) Brownian motions $W_I(t),~W_Q(t)$, but this fact is not explicitly modeled for ease of presentation. The attenuation coefficient (in squared magnitude) is [@charalambous08] $a_{ij}(t)=I(t)^2+Q(t)^2.$ As in the case of LTF, the coefficients $A_I(t),~A_Q(t),~B_I(t),~B_Q(t),~C_I,~C_Q$ can be obtained directly via signal measurements following the methodology proposed in [@charalambous08], [@olama09] using the EM algorithm together with Kalman filtering. This model leads to time-varying mean and variance for the inphase and the quadrature components and thus for the STF wireless channel and includes the Ricean, Rayleigh and Nakagami distributions as special cases [@menemenlis99]. In the rest of the paper, the vectors $X(t),~W(t)$ denote collectively (for all links) the channel states and the Brownian motions respectively at time $t$. Note that we assume that the wireless channels are uncorrelated. This assumption is also made in [@neely10], where it is argued that inter-link correlations do not impact the network capacity region and therefore the maximum utility. System Model & Assumptions {#sec:sysmodel} ========================== We consider a static wireless multihop network with $N$ nodes and $E$ directed links forming the set $\mathcal{E}$. The network serves $F$ overlaying flows (source-destination pairs) over a finite duration (lifetime) $[s,T]$. At time $t \in [s,T]$, $\lambda_i^d(t)$ data (e.g. packets) are produced from the source node $i$ for its destination node $d$. Let $S_r(d)$ be the set of sources for node $d$. Then, $\lambda_i^d(t)=0,~\forall i \notin S_r(d),~\forall t \in [s,T]$. We denote with $r_{ij}^d(t)$ the communication traffic on the link $(i,j)$ for destination $d$ at time $t\in [s,T]$. Then, $R(t),~\Lambda(t),$ denote collectively the variables $\{r_{ij}^d(t)\}_{\forall d,(ij)},~\{\lambda_i^d(t)\}_{\forall i,d},$ respectively, at time $t$. The set $\mathcal{R}(i,d)$ consists of the one-hop (out-)neighbors of node $i$ which are allowed to serve as next-hop nodes towards $d$ according to the routing protocol under consideration. If there are no routing constraints, we consider $\mathcal{R}(i,d)=\mathcal{N}^{out}(i)$, where $\mathcal{N}^{out}(i)=\{j|(i,j)\in \mathcal{E}\}$. Also, $r_{ij}^d(t)=0$, if $j \notin \mathcal{R}(i,d)$ and $r_{ii}^d(t)=0,~r_{di}^d(t)=0,~\forall i,d,t\in [s,T]$. Furthermore, we assume that the transmitter of the link $(i,j)$, i.e. the node $i$, transmits with power $P_{ij}(t)$ at time $t \in [s,T]$. $P(t)$ expresses collectively the transmission powers of all links at time $t \in [s,T]$. Each source node associates its satisfaction for its produced data for destination $d$, $\lambda_i^d$, at time $t \in [s,T]$, with a time-varying continuous differentiable utility function $U_i^d(\lambda_i^d, t)$. Several utility functions used in literature belong in this category, such as the strictly convex and increasing $a$-fair utility, including the logarithmic one [@eryilmaz06]. It is further assumed that $U_i^d(\lambda_i^d, t)$ is increasing with $\lambda_i^d$ and uniformly bounded as $t\rightarrow \infty$. Also, a (continuous differentiable) cost function, $J_{ij}(P_{ij})$, is assigned to each directed link $(i,j)$ with respect to its transmission power $P_{ij}(t)$, $t \in [s,T]$. In literature, $J_{ij}(P_{ij})$ is often assumed to be a strictly convex function [@marques]. The proposed cross-layer framework includes routing, scheduling, power and congestion control. Routing (network layer) determines the amount of traffic for each destination that will be served by every link, by optimizing $R(t)$, $\forall t \in [s,T]$. Scheduling and power control (MAC and physical layers) determine which links are going to transmit and their transmission power by optimizing $P(t)$, $\forall t \in [s,T]$. Finally, congestion control (transport layer) optimizes $\Lambda(t)$, $\forall t \in [s,T]$. Therefore, the proposed cross-layer scheme aims at determining the optimal values of the control variables $R(t),~\Lambda(t),~P(t)$, $\forall t \in [s,T]$, according to an optimality criterion designed with the aid of the utility and cost functions defined above, considering the channel models of Section \[sec:channelmodels\]. Due to the considered underlying channel processes, the control variables $R,~\Lambda,~P$ should be in general defined as stochastic processes. Let us define the value range for each $\lambda_i^d$, $U_{\lambda}=[0,\lambda_{\max}]$, and the corresponding feasible set $\mathcal{U}_{\lambda}=\left\{\lambda:[s,T]\times \Omega\rightarrow U_{\lambda}: \lambda \mathrm{~is~} \{\mathcal{F}_t\}_{t\geq s} \mathrm{~adapted} \right\}$. Then, $\Lambda \in \mathcal{U}_{\lambda}^F$. Similarly, we define the value range for each $r_{ij}^d$, $U_{r}=[0,R_{\max}]$, and the corresponding feasible set $\mathcal{U}_{r}=\left\{r:[s,T]\times \Omega\rightarrow U_{r}: r \mathrm{~is~} \{\mathcal{F}_t\}_{t\geq s} \mathrm{~adapted} \right\}$. Then, $R \in \mathcal{U}_{r}^{E\times (N-1)}$. Finally, we define the value range for each $P_{ij}$, $U_{P}=[0,P_{\max}]$, and the corresponding feasible set $\mathcal{U}_{P}=\left\{P:[s,T]\times \Omega\rightarrow U_{P}: P \mathrm{~is~} \{\mathcal{F}_t\}_{t\geq s} \mathrm{~adapted} \right\}$. Then, $P \in \mathcal{U}_{P}^{E}$. We will use ${\mathbb{E}}_{s,x}$ to denote expectations given the initial condition $X(s)=x$. At this point, we distinguish two cases with respect to the access to the wireless medium. Based on the two types of access, two cross-layer problems are developed. The first case concerns non-orthogonal access to the wireless medium, in which the transmitters are allowed to access the wireless medium simultaneously (one frequency carrier is assumed), while the interfering transmissions are considered as noise. For this case we define the Signal-to-Interference-plus-Noise-Ratio $(SINR)$ for the link $(i,j)$ as follows: $SINR_{ij}(t)=\frac{a_{ij}(t)P_{ij}(t)}{N_0+\sum_{(k,l)\in \mathcal{I}_{ij}}a_{kj} P_{kl}(t)}$, where $\mathcal{I}_{ij}$ denotes the subset of $\mathcal{E}$ containing the links that interfere with the link $(i,j)$. $N_0$ (Watts) stands for the average background noise at the receiver’s ($j$) side and $a_{ij}(t)$ is defined in Section \[sec:channelmodels\] depending on the fading type. The capacity of the link $(i,j)$ is given by the Shannon’s formula in bits/sec as $C_{ij}(P(t))=B_{ij}\log_2(1+SINR_{ij}(t))$, where $B_{ij}$ ($Hertz$) is the wireless channel’s bandwidth at link $(i,j)$. The second case refers to the orthogonal access to the wireless medium, where only non-interfering links can access simultaneously the wireless medium. Orthogonal access decreases the complexity of the proposed framework’s operation in the physical layer (power control), as it will be shown in later sections, while it is nearly optimal when interference is strong [@marques]. In this case, the connectivity graph of the wireless multihop network is important in identifying the feasible schedules. Based on the latter, the finite set of all possible independent sets of links (i.e. links that do not interfere with each other) is constructed. Only links belonging to the same independent set can access the wireless medium simultaneously. In this case, the capacity of the link $(i,j)$ is a concave function of $P_{ij}$, given in bits/sec as $C_{ij}(P_{ij}(t))=B_{ij}\log_2\left(1+\frac{a_{ij}(t)P_{ij}(t)}{N_0}\right)$. Non-orthogonal Access to the Medium {#sec:nonortho} =================================== Problem Formulation & Analysis {#sec:Problform} ------------------------------ According to the discussion in Section \[sec:sysmodel\], the optimization framework, denoted as $\mathbf{P_1}$, is formulated as follows. $$\begin{aligned} P_1:=\max_{\Lambda \in \mathcal{U}_{\lambda}^F, ~R \in \mathcal{U}_{r}^{E\times (N-1)},~ P \in \mathcal{U}_{P}^{E}} {\mathbb{E}}_{s,x}\left[\int_s^T \left(\sum_{i,d:i\in S_r(d)} U_i^d(\lambda_i^d(t),t)- \sum_{(i,j)}J_{ij}(P_{ij}(t))\right)dt\right]\nonumber \\ s.t.\nonumber\\ {\mathbb{E}}_{s,x}\left[\int_s^T \lambda_i^d(t)dt +\int_s^T \sum_{j:i \in \mathcal{R}(j,d)} r_{ji}^d(t)dt\right]\leq {\mathbb{E}}_{s,x}\left[\int_s^T \sum_{j\in \mathcal{R}(i,d)} r_{ij}^d(t)dt\right],~ \forall i,d \label{const:1}\\ {\mathbb{E}}_{s,x}\left[\int_s^T \sum_{d} r_{ij}^d(t)dt\right]\leq {\mathbb{E}}_{s,x}\left[\int_s^T C_{ij}(P(t))dt\right], ~\forall (i,j)\in \mathcal{E}\label{const:2}\\ {\mathbb{E}}_{s,x}\left[\int_s^T\sum_{j\in \mathcal{N}^{out}_i}P_{ij}(t)dt\right]\leq P_{i,\max}, ~\forall i\label{const:3}\end{aligned}$$ The objective function expresses the trade-off between the accumulated for all sources utilities/satisfaction of producing data and the accumulated cost due to power consumption for link transmissions. Thus, its maximization targets at improving energy efficiency by maximizing source rates while penalizing the cost of power consumption for achieving them when $J_{ij}\neq 0,~\forall (i,j)$, i.e. when power control is applied. The first constraint relates to the flow conservation at each node and for each destination, used to ensure high throughput for the examined time interval $[s,T]$. The second constraint relates to the capacity restriction (right side) due to power, channel and interference limitations for each link. Finally, the third constraint, relates to a limitation on the total power consumption of each node for the examined time interval, $[s,T]$, (left side) according to its energy storage, denoted as $P_{i,\max}>0$ (right side). It is important to note that we consider continuous time network operation (thus, using time integrals), for ease of presentation, due to the continuous time evolution of the channel state (Section \[sec:channelmodels\]). The case of discrete time network operation can be obtained trivially by replacing the integrals $\int_s^T$ by sums $\sum_{t=0}^{t=N_L}$ where $N_L$ is the number of time slots in $[s,T]$ considered for the network operation and each denoted by $t$. In this case the channel state will be sampled at each time slot as described in the subsequent sections. Furthermore, for considering an infinite $T$, we should also divide every integral by $T$, thus considering time averages. This problem is non-convex due to the forms of the capacity, utility and cost functions. Even if the utility and cost functions were concave and convex respectively as commonly assumed in literature, the problem would still be non-convex due to the capacity function forms. However, we will prove that its duality gap is zero, which is an important fact as it renders the Lagrange (dual)-based optimization method optimal. The latter allows for devising efficient algorithmic solutions as it leads to a separable optimization problem with respect to the variables of each layer while using the Lagrange multipliers for the communication between adjacent layers for achieving a cross-layer optimal solution. Let $\mu_i^d\geq 0, ~\forall i,d, ~l_{ij}\geq 0,~\forall (i,j)\in \mathcal{E},~\nu_i\geq 0, ~\forall_i$ be the Lagrange multipliers associated with the constraints (\[const:1\]), (\[const:2\]), (\[const:3\]) correspondingly. Denote with $L$ the whole set of the Lagrange multipliers. Then, the dual function is formulated as follows: $$\begin{aligned} \label{eq:dual} L_A(L)=\max_{\Lambda \in \mathcal{U}_{\lambda}^F, ~R \in \mathcal{U}_{r}^{E\times (N-1)},~ P \in \mathcal{U}_{P}^{E}} ~{\mathbb{E}}_{s,x}\left[\int_s^T \left(\sum_{i,d:i\in S_r(d)} U_i^d(\lambda_i^d(t),t)- \sum_{(i,j)}J_{ij}(P_{ij}(t))\right)dt\right]\nonumber \\ -\sum_{i,d} \mu_i^d {\mathbb{E}}_{s,x}\left[\int_s^T \lambda_i^d(t)dt +\int_s^T \sum_{j:i \in \mathcal{R}(j,d)} r_{ji}^d(t)dt-\int_s^T \sum_{j\in \mathcal{R}(i,d)} r_{ij}^d(t)dt\right] \nonumber\\ -\sum_{(i,j)\in\mathcal{E}}l_{ij}{\mathbb{E}}_{s,x}\left[\int_s^T \sum_{d} r_{ij}^d(t)dt-\int_s^T C_{ij}(P(t))dt\right]- \sum_{i} \nu_i {\mathbb{E}}_{s,x}\left[\int_s^T\sum_{j\in \mathcal{N}^{out}_i}P_{ij}(t)dt- P_{i,\max}\right].\end{aligned}$$Consequently, the dual problem of $P_1$ is defined as $D_1:=\inf_{L}(L_A(L))$. The problem $P_1$ has zero dual gap, i.e. if $P_1^*$ its optimal value and $D_1^*$ the optimal value of the dual problem, then $P_1^*=D_1^*$.\[thm:1\] To prove this theorem we proceed in analogy with Theorem $1$ in [@ribeiro10]. The proof relies on the fact that the channel’s cumulative distribution function (cdf) is continuous and thus no channel realization has strictly positive probability. It uses the definition of nonatomic measures and the Lyapunov’s convexity theorem [@ribeiro10]. To prove the zero duality gap, we consider a perturbed version of the problem $P_1$, obtained by perturbing the constraints used to define the Lagrangian. Let $P_1(\Delta)$ be the function that assigns to each perturbation set $\Delta=(\{\Delta_{i,d}^1\}_{\forall i,d}, \{\Delta_{i,j}^2\}_{\forall (i,j)}, \{\Delta_{i}^3\}_{\forall i})$, the solution of the following perturbed optimization problem. $$\begin{aligned} P_1(\Delta)=\max_{\Lambda \in \mathcal{U}_{\lambda}^F, ~R \in \mathcal{U}_{r}^{E\times (N-1)},~ P \in \mathcal{U}_{P}^{E}} {\mathbb{E}}_{s,x}\left[\int_s^T \left(\sum_{i,d:i\in S_r(d)} U_i^d(\lambda_i^d(t),t)- \sum_{(i,j)}J_{ij}(P_{ij}(t))\right)dt\right]\nonumber \\ s.t.\nonumber\\ {\mathbb{E}}_{s,x}\left[\int_s^T \lambda_i^d(t)dt +\int_s^T \sum_{j:i \in \mathcal{R}(j,d)} r_{ji}^d(t)dt-\int_s^T \sum_{j\in \mathcal{R}(i,d)} r_{ij}^d(t)dt\right]\leq \Delta_{i,d}^1,~ \forall i,d \label{eq:constr1per}\\ {\mathbb{E}}_{s,x}\left[\int_s^T \left(\sum_{d} r_{ij}^d(t)- C_{ij}(P(t))\right)dt\right]\leq \Delta_{i,j}^2, ~\forall (i,j)\in \mathcal{E}, ~ ~{\mathbb{E}}_{s,x}\left[\int_s^T\sum_{j\in \mathcal{N}^{out}_i}P_{ij}(t)dt\right]- P_{i,\max}\leq \Delta_{i}^3, ~\forall i \label{eq:constr3per}\end{aligned}$$i.e. the constraints can be violated by $\Delta$ amounts. In order to prove zero duality gap, we should show that the function $P_1(\Delta)$ is a concave function of $\Delta$ [@ribeiro10]. Let $\underline{\Delta}=(\{\underline{\Delta}_{i,d}^1\}_{\forall i,d}, \{\underline{\Delta}_{i,j}^2\}_{\forall (i,j)}, \{\underline{\Delta}_{i}^3\}_{\forall i})$, $\overline{\Delta}=(\{\overline{\Delta}_{i,d}^1\}_{\forall i,d}, \{\overline{\Delta}_{i,j}^2\}_{\forall (i,j)}, \{\overline{\Delta}_{i}^3\}_{\forall i})$ be two arbitrary sets of perturbations with respective optimal values $\underline{P_1}=P_1(\underline{\Delta})$, $\overline{P_1}=P_1(\overline{\Delta})$ and respective solutions $(\underline{\Lambda}, \underline{R}, \underline{P})$ and $(\overline{\Lambda}, \overline{R}, \overline{P})$. Then, for an arbitrary $a\in [0,1]$, we define the perturbation $\hat{\Delta}=a \underline{\Delta}+(1-a) \overline{\Delta}$ and for feasible solutions $(\hat{\Lambda}, \hat{R}, \hat{P})$, i.e. satisfying the constraints (\[eq:constr1per\]), (\[eq:constr3per\]), we need to show $$P_1(\hat{\Delta})= P_1(a \underline{\Delta}+(1-a) \overline{\Delta})\geq a P_1(\underline{\Delta})+(1-a) P_1(\overline{\Delta}).\label{eq:convex}$$ Consider the set of all possible state ($X$) realizations $\mathcal{H}$ and the Borel field, $\mathbb{B}$, on $\mathcal{H}$. For $A \in \mathbb{B}$, let ${\mathbb{E}}_{s,x}^A$ be the expected value restricted on channel realizations included in $A$. We define the following measures. $$\begin{aligned} \theta_{id}(A)=\left[{\mathbb{E}}_{s,x}^A\left[\int_s^T U_i^d(\underline{\lambda}_i^d(t),t)dt\right], {\mathbb{E}}_{s,x}^A\left[\int_s^T U_i^d(\overline{\lambda}_i^d(t),t)dt \right] \right],~ \forall i,d:i \in S_r(d),\\ \phi_{id}(A)=\left[{\mathbb{E}}_{s,x}^A\left[\int_s^T \underline{\lambda}_i^d(t)dt\right], {\mathbb{E}}_{s,x}^A\left[\int_s^T \overline{\lambda}_i^d(t)dt \right] \right],~ \forall i,d,\\ w_{ij}(A)=\left[{\mathbb{E}}_{s,x}^A\left[\int_s^T C_{ij}(\underline{P}(t))dt\right], {\mathbb{E}}_{s,x}^A\left[\int_s^T C_{ij}(\overline{P}(t))dt \right] \right],~ \forall (i,j)\in \mathcal{E},\\ v_{ij}(A)=\left[{\mathbb{E}}_{s,x}^A\left[\int_s^T J_{ij}(\underline{P}_{ij}(t))dt\right], {\mathbb{E}}_{s,x}^A\left[\int_s^T J_{ij}(\overline{P}_{ij}(t))dt \right] \right],~ \forall (i,j) \in \mathcal{E},\\ \xi_{ij}(A)=\left[{\mathbb{E}}_{s,x}^A\left[\int_s^T \underline{P}_{ij}(t)dt\right], {\mathbb{E}}_{s,x}^A\left[\int_s^T \overline{P}_{ij}(t)dt \right] \right],~ \forall (i,j) \in \mathcal{E},\end{aligned}$$while we also define $\theta_{id}(\varnothing)=\phi_{id}(\varnothing)=w_{ij}(\varnothing)=v_{ij}(\varnothing)=\xi_{ij}(\varnothing)=0$. These measures are nonatomic [@ribeiro10] since the channel cdf is continuous and all the control variables are bounded. Thus, there are no channel realizations with positive measure, i.e., $\theta_{id}(A)=\phi_{id}(A)=w_{ij}(A)=v_{ij}(A)=\xi_{ij}(A)=0$ for all singleton sets $A \in \mathcal{H}$. Let $W(A)$ be the vector measure expressing collectively all measures $\theta_{id},\phi_{id},w_{ij},v_{ij},\xi_{ij}$, which is also obviously nonatomic. Then from Lyapunov’s convexity theorem [@ribeiro10], the range of $W$ is convex. Therefore, the value $w_0=a W(\mathcal{H})+(1-a) W(\varnothing)=a W(\mathcal{H})$ belongs to the range of values of $W$. As a result there exists $A_0 \in \mathbb{B}$ such that $W(A_0)=a W(\mathcal{H})$, i.e., $\theta_{id}(A_0)=a \theta_{id}(\mathcal{H}),~\phi_{id}(A_0)=a \phi_{id}(\mathcal{H})$, etc. Moreover, due to the additivity of measures, for the complement of $A_0$, $A_0^c$, we have $W(A_0^c)= W(\mathcal{H})-W(A_0)=(1-a)W(\mathcal{H})$. Then, we define the following controls for the new perturbation $\hat{\Delta}$. $$\begin{aligned} \hat{r}_{ij}^d(t)=a \underline{r}_{ij}^d(t)+ (1-a) \overline{r}_{ij}^d(t), ~\forall d,(i,j)\in \mathcal{E}, t \in [s,T], ~\mathbf{P}-a.s.\end{aligned}$$ $$\begin{aligned} \hat{\lambda}_{i}^d(t)=\begin{cases}\underline{\lambda}_{i}^d(t) ~\text{within}~ A_0 \nonumber\\ \overline{\lambda}_{i}^d(t)~ \text{within}~ A_0^c \end{cases}\end{aligned}$$ $$\begin{aligned} \hat{P}_{ij}(t)=\begin{cases}\underline{P}_{ij}(t) ~\text{within}~ A_0\nonumber \\ \overline{P}_{ij}(t)~ \text{within}~ A_0^c \end{cases} \forall i,d:i \in S_r(d), \forall (i,j)\in \mathcal{E}, t \in [s,T],~ \mathbf{P}-a.s.\end{aligned}$$ Obviously since $\underline{\Lambda},~ \overline{\Lambda}~ \in \mathcal{U}_{\lambda}^F, ~\underline{R}, ~\overline{R} ~\in \mathcal{U}_{r}^{E\times (N-1)},~ \underline{P}, ~\overline{P}~ \in \mathcal{U}_{P}^{E}$, it also holds that $\hat{\Lambda} \in \mathcal{U}_{\lambda}^F, ~\hat{R} \in \mathcal{U}_{r}^{E\times (N-1)},~ \hat{P} \in \mathcal{U}_{P}^{E}$. Now, based on the above, we check if the controls defined above for the perturbation $\hat{\Delta}$ satisfy Ineqs. (\[eq:constr1per\]), (\[eq:constr3per\]) and if Ineq. (\[eq:convex\]) holds. For the constraint (\[eq:constr1per\]), we have: $$\begin{aligned} {\mathbb{E}}_{s,x}\left[\int_s^T \left(\hat{\lambda}_i^d(t)+\sum_{j:i \in \mathcal{R}(j,d)} \hat{r}_{ji}^d(t)- \sum_{j\in \mathcal{R}(i,d)} \hat{r}_{ij}^d(t)\right)dt\right]= {\mathbb{E}}_{s,x}^{A_0}\left[\int_s^T \underline{\lambda}_i^d(t)dt\right] + {\mathbb{E}}_{s,x}^{A_0^c}\left[\int_s^T \overline{\lambda}_i^d(t)dt\right]+\nonumber\\ +{\mathbb{E}}_{s,x}\left[\int_s^T \sum_{j:i \in \mathcal{R}(j,d)}\left( a \underline{r}_{ji}^d(t)+ (1-a) \overline{r}_{ji}^d(t)\right)dt-\int_s^T \sum_{j\in \mathcal{R}(i,d)} \left(a \underline{r}_{ij}^d(t)+ (1-a) \overline{r}_{ij}^d(t)\right)dt\right]=\nonumber\\ a{\mathbb{E}}_{s,x}\left[\int_s^T \underline{\lambda}_i^d(t)dt +\int_s^T \sum_{j:i \in \mathcal{R}(j,d)} \underline{r}_{ji}^d(t)dt -\int_s^T \sum_{j\in \mathcal{R}(i,d)} \underline{r}_{ij}^d(t)dt\right]\nonumber\\+(1-a){\mathbb{E}}_{s,x}\left[\int_s^T \overline{\lambda}_i^d(t)dt +\int_s^T \sum_{j:i \in \mathcal{R}(j,d)} \overline{r}_{ji}^d(t)dt -\int_s^T \sum_{j\in \mathcal{R}(i,d)} \overline{r}_{ij}^d(t)dt\right]\leq a\underline{\Delta}_{i,d}^1 +(1-a) \overline{\Delta}_{i,d}^1= \hat{\Delta}_{i,d}^1.\end{aligned}$$ With respect to the first constraint in (\[eq:constr3per\]), we have: $$\begin{aligned} {\mathbb{E}}_{s,x}\left[\int_s^T \sum_{d} \hat{r}_{ij}^d(t)dt\right]- {\mathbb{E}}_{s,x}\left[\int_s^T C_{ij}(\hat{P}(t))dt\right]=a{\mathbb{E}}_{s,x}\left[\int_s^T \sum_{d} \underline{r}_{ij}^d(t)dt-\int_s^T C_{ij}(\underline{P}(t))dt\right]\nonumber \\+ (1-a){\mathbb{E}}_{s,x}\left[\int_s^T \sum_{d} \overline{r}_{ij}^d(t)dt-\int_s^T C_{ij}(\overline{P}(t))dt\right]\leq a\underline{\Delta}_{i,j}^2 +(1-a) \overline{\Delta}_{i,j}^2= \hat{\Delta}_{i,j}^2.\end{aligned}$$ Similarly, with respect to the second constraint in (\[eq:constr3per\]), we have: $$\begin{aligned} {\mathbb{E}}_{s,x}\left[\int_s^T\sum_{j\in \mathcal{N}^{out}_i}\hat{P}_{ij}(t)dt\right]- P_{i,\max}= {\mathbb{E}}_{s,x}^{A_0}\left[\int_s^T\sum_{j\in \mathcal{N}^{out}_i} \underline{P}_{ij}(t)dt\right]+{\mathbb{E}}_{s,x}^{A_0^c}\left[\int_s^T\sum_{j\in \mathcal{N}^{out}_i} \overline{P}_{ij}(t)dt\right]- P_{i,\max}=\nonumber\\ a{\mathbb{E}}_{s,x}\left[\int_s^T\sum_{j\in \mathcal{N}^{out}_i}\underline{P}_{ij}(t)dt\right]+(1-a){\mathbb{E}}_{s,x}\left[\int_s^T\sum_{j\in \mathcal{N}^{out}_i}\overline{P}_{ij}(t)dt\right]- P_{i,\max}\leq a\underline{\Delta}_{i}^3 +(1-a) \overline{\Delta}_{i}^3= \hat{\Delta}_{i}^3.\end{aligned}$$ Finally, $P_1(\hat{\Delta})\geq {\mathbb{E}}_{s,x}\left[\int_s^T \left(\sum_{i,d:i\in S_r(d)} U_i^d(\hat{\lambda}_i^d(t),t)- \sum_{(i,j)}J_{ij}(\hat{P}_{ij}(t))\right)dt\right]$, i.e., $$\begin{aligned} P_1(\hat{\Delta})\geq= {\mathbb{E}}_{s,x}^{A_0}\left[\int_s^T \left(\sum_{i,d:i\in S_r(d)} U_i^d(\underline{\lambda}_i^d(t),t)-\sum_{(i,j)}J_{ij}(\underline{P}_{ij}(t))\right)dt\right]\nonumber\\+ {\mathbb{E}}_{s,x}^{A_0^c}\left[\int_s^T \left( \sum_{i,d:i\in S_r(d)} U_i^d(\overline{\lambda}_i^d(t),t)-\sum_{(i,j)}J_{ij}(\overline{P}_{ij}(t))\right)dt\right]= aP_1(\underline{\Delta})+(1-a)P_1(\overline{\Delta}),\end{aligned}$$which concludes the proof. Problem Solution ---------------- Since the dual gap corresponding to the problem $P_1$ is null, we can obtain its optimal value by solving its dual problem via a subgradient methodology [@bertsekas09]. The subgradient algorithm, repeatedly renews the Lagrange multipliers until converging to their optimal solutions. We use the symbol $\eta=\{0,1,..\}$ for the repetitions of the subgradient algorithm. Then, the renewal equations of the Lagrange multipliers take the form: $$\begin{aligned} \label{eq:lagrangea} \mu_i^d(\eta+1)=\left\{\mu_i^d(\eta)+\kappa(\eta)\cdot {\mathbb{E}}_{s,x}\left[\int_s^T \lambda_i^{d*}(t)dt +\int_s^T \sum_{j:i \in \mathcal{R}(j,d)} r_{ji}^{d*}(t)dt-\int_s^T \sum_{j\in \mathcal{R}(i,d)} r_{ij}^{d*}(t)dt\right]\right\}^+,~ \forall i,d,\\ l_{ij}(\eta+1)=\left\{l_{ij}(\eta)+\kappa(\eta)\cdot {\mathbb{E}}_{s,x}\left[\int_s^T \sum_{d} r_{ij}^{d*}(t)dt-\int_s^T C_{ij}(P^*(t))dt\right]\right\}^+, ~\forall (i,j)\in\mathcal{E},\label{eq:lagrangeb} \\ \nu_i(\eta+1)=\left\{\nu_i(\eta)+\kappa(\eta)\cdot {\mathbb{E}}_{s,x}\left[\int_s^T\sum_{j\in \mathcal{N}^{out}_i}P_{ij}^*(t)dt- P_{i,\max}\right]\right\}^+,~\forall i.\label{eq:lagrangec}\end{aligned}$$where $\left\{\right\}^+$ denotes projection to $[0,\infty)$ and the values $\mu_i^d(0)\geq 0, ~\forall i,d, ~l_{ij}(0)\geq 0,~\forall (i,j)\in \mathcal{E},~\nu_i(0)\geq 0, ~\forall i$, are considered given. The subgradient method is known to converge to a close neighborhood of the optimal values for the Lagrange multipliers if constant step-size, $\kappa(\eta)$, is used, while diminishing, non summable but square summable step size allows for convergence to the optimal values [@bertsekas09], [@gatsis11]. The values of the stochastic processes $\lambda_i^{d*},~\forall i,d,~r_{ij}^{d*},~\forall (i,j),d,~P_{ij}^*,~\forall (i,j),$ are computed while obtaining the dual function (Eq. (\[eq:dual\])) with the current solution for the Lagrange multipliers, i.e. for the iteration $\eta$. For performing this maximization, we can observe that one cannot achieve a better objective value than the one achieved by choosing at each time $t\in[s,T]$, each control variable optimally as a function of $X(t)$ and the current values of the Lagrange multipliers (iteration $\eta$). Therefore, at the iteration $\eta$, given $\mu_i^d(\eta), ~\forall i,d, ~l_{ij}(\eta)~\forall (i,j)\in \mathcal{E},~\nu_i(\eta), ~\forall_i$, the controls are computed as follows: - The optimal $\lambda_i^{d*},~\forall i,d: i \in S_r(d)$, at the transport layer, are computed source-wise by $$\label{eq:lamda} \frac{\vartheta U_i^d(\lambda_i^d,t)}{\vartheta \lambda_i^d}-\mu_i^d(\eta)=0, \forall~t\in [s,T],$$while taking into account that $\lambda_i^{d*} \in U_{\lambda}$. From Eq. (\[eq:lamda\]), it is observed that $\lambda_i^{d*}$ is time-varying but not random as it depends only on the deterministic Lagrange multiplier $\mu_i^d(\eta)$ (which is constant in time $t\in [s,T]$) and the deterministic time-varying utility function $U_i^d$ but not on the channel state $X(t)$. In the case that $U_i^d(\lambda_i^d,t),~\forall i,d:i\in S_r(d)$ are invariant with $t$ then $\lambda_i^{d*},~\forall i,d:i\in S_r(d)$ are constants over $[s,T]$, too. - The optimal routing variables $r_{ij}^{d*},~\forall (i,j),d$ (network layer) are computed by $$\label{eq:r} \max_{R}\sum_d \sum_{(i,j)|j\in R(i,d)}r_{ij}^d(\mu_i^d(\eta)-\mu_j^d(\eta)-l_{ij}(\eta)),~ \forall~t\in [s,T],$$i.e. $r_{ij}^{d*}(t)=R_{\max},~\forall~t\in [s,T]$, if $(\mu_i^d(\eta)-\mu_j^d(\eta)-l_{ij}(\eta))>0$, while it is observed (from Eq. (\[eq:r\])) that each $r_{ij}^{d*}$ is constant in time. - The optimal transmission power values at the physical layer, $P^*$, are computed via solving: $$\label{eq:p} \max_{P} \sum_{(i,j)}\left(-J_{ij}(P_{ij})+l_{ij}(\eta)C_{ij}(P)-\nu_i(\eta) P_{ij}\right), \forall~t\in [s,T],$$while considering the $U_{P}$ constraints (Section \[sec:sysmodel\]). From Eq. (\[eq:p\]), it is observed that $P$ is a stochastic process since the link capacity $C_{ij},~\forall (i,j)$ depends on the stochastic process $X(t)$ representing the wireless channels (Section \[sec:sysmodel\]). The computation of the expected values involved in the Lagrange multipliers’ renewal equations requires Monte Carlo simulations. However, since the controls $\Lambda^*,~R^*$, have been shown to be deterministic, the expected value involved in the renewal equation of the Lagrange multipliers $\mu_i^d, ~\forall i,d$ (Eq. (\[eq:lagrangea\])) is superfluous. For the rest of the Lagrange multipliers, the expected values are obtained via the following algebraic computations. Firstly, the processes $\{X_{ij}(t):\ t\ge 0\}$ are discretized [@higham01]. We compute $\delta t=\frac{T-s}{n}$, where $n$ is a design parameter representing the number of samples of the channel over the time interval $[s,T]$ of the network’s operation, and we sample on the time instants $\{\tau_b=s+b \cdot \delta t\}_{b=1:n}$. In view of (\[eq:sdepathloss\]) it is not hard to see that for every $i,j$ and $b$ we have $$X_{ij}(\tau_b)=\rho_{ij}(b)X_{ij}(\tau_{b-1})+\zeta_{ij}(b)+Z_{ij}(b)$$ where $$\rho_{ij}(b)=\exp(-\int_{\tau_{b-1}}^{\tau_b}\beta_{ij}(s)ds)$$ $$\zeta_{ij}(b)=\int_{\tau_{b-1}}^{\tau_b}\beta_{ij}(s)\gamma_{ij}(s)\exp\big(-\int_{s}^{\tau_b}\beta_{ij}(q)dq\big)ds$$ and $$Z_{ij}(b)=\int_{\tau_{b-1}}^{\tau_b}\delta_{ij}(s)\exp\big(-\int_{s}^{\tau_b}\beta_{ij}(q)dq\big)dW_{ij}(s).$$ Note that $\{Z_{ij}(b)\}_{i,j,b}$ are independent random variables with distributions ${\cal N}(0,\sigma_{ij}^2(b))$, where $$\sigma_{ij}^2(b)=\int_{\tau_{b-1}}^{\tau_b}\delta_{ij}^2(s)\exp(-2\int_{s}^{\tau_b}\beta_{ij}(q)dq)ds.$$ The discretized scheme then becomes for every $(i,j)$: $$\label{eq:sdenumerical} X_{ij}(\tau_b)=\rho_{ij}(b)X_{ij}(\tau_{b-1})+\zeta_{ij}(b)+\sigma_{ij}(b) \xi_{ij}(b),\ b=1,2,\ldots,n,\quad X_{ij}(s)=x_{ij0}$$ where $\{\xi_{ij}(b)\}_{i,j,b}$ are independent samples from a standard normal random variable. After numerically computing the solution of the SDE (\[eq:sdepathloss\]), we compute $P^*(\tau_b)$ for each $b$ from Eq. (\[eq:p\]) and we use a Riemann sum approximation for a sample of the integral $\int_{s}^{T} C_{ij}(P^*(t))dt$, that is $\int_{s}^{T} C_{ij}(P^*(t))dt\simeq\sum_{b=0}^{n-1} C_{ij}(P^*(\tau_b))\delta t$, where ${C_{ij}}({P^*}({\tau _b})) = {B_{ij}}{\log _2}\left( {1 + \frac{{{e^{K{X_{ij}}({\tau _b})}}P_{ij}^*({\tau _b})}}{{{N_0} + \sum\limits_{(k,l) \in {{\cal I}_{ij}}} {{e^{K{X_{kj}}({\tau _b})}}P_{kl}^*({\tau _b})} }}} \right)$ (Sections \[sec:channelmodels\], \[sec:sysmodel\]). Finally, we repeat the above procedure $M$ times to obtain $M$ independent samples of the preceding integral, and we average these samples to estimate the expected capacity of link $(i,j)$ over the time interval $[s,T]$, i.e. ${\mathbb{E}}_{s,x}\left[\int_{s}^{T} C_{ij}(P^*(t))dt\right]$, appearing in the Eq. (\[eq:lagrangeb\]). Similarly we obtain ${\mathbb{E}}_{s,x}\left[\int_{s}^{T} P_{ij}^*(t)dt\right]$. Since the computation of $P^*$ from Eq. (\[eq:p\]) involves the Lagrange multipliers’ values, the Monte Carlo computations of the expected values should be performed at each iteration of the subgradient algorithm. Notably, $\rho_{ij}(b),\zeta_{ij}(b),\sigma_{ij}(b)$ are deterministic so we only need to compute them once. The optimal power allocation problem at the physical layer, i.e., the solution of Eq. (\[eq:p\]) determines the complexity of the whole problem since everything else is simple algebraic computations. Indeed the computations of Eqs. (\[eq:lamda\]), (\[eq:r\]) can be distributed to the sources and links correspondingly. This cross-terminal optimization problem at the physical layer constitutes an important challenge in wireless networking [@ribeiro10] which is treated in other works in literature [@ribeiro10], [@gatsis11] and is out of the scope of this paper. In the next section, we formulate and solve the same problem in the case of orthogonal access to the wireless medium where the capacity functions take much simpler concave forms leading to tractable analytic solutions. We summarize below the steps of the algorithm proposed in this section for obtaining $D_1^*$. 1. Initialize the Lagrange multipliers, $\eta=0,~\mu_i^d(0), ~\forall i,d, ~l_{ij}(0),~\forall (i,j)\in \mathcal{E},~\nu_i(0), ~\forall_i$. 2. Compute $\lambda_i^{d*},~\forall i,d,~r_{ij}^{d*},~\forall (i,j),d$ using Eqs. (\[eq:lamda\]), (\[eq:r\]) respectively. 3. Compute the expected values ${\mathbb{E}}_{s,x}\left[\int_{s}^{T} C_{ij}(P^*(t))dt\right],~{\mathbb{E}}_{s,x}\left[\int_{s}^{T} P_{ij}^*(t)dt\right],~\forall (i,j)$, as described. 4. Compute $\mu_i^d(\eta+1), ~\forall i,d, ~l_{ij}(\eta+1),~\forall (i,j)\in \mathcal{E},~\nu_i(\eta+1), ~\forall_i$ from Eqs. (\[eq:lagrangea\]), (\[eq:lagrangeb\]), (\[eq:lagrangec\]) and set $\eta \leftarrow \eta+1$. 5. Repeat steps $2,~3,~4$ until convergence. Discussion ---------- It is important to note that the time scale of the renewal of the Lagrange multipliers should be distinguished from the time interval $[s,T]$ of the network’s operation. In principle, the above algorithm should run off-line, i.e. prior to the network operation to determine the optimal source rates, routing variables and Lagrange multipliers and afterwards, the online network operation will be designed based on these optimal values and the solution of the cross-terminal power allocation problem of Eq. (\[eq:p\]). Note that convergence of the dual variables close to their optimal values does not imply convergence of the primal variables except if the primal variables change continuously with respect to the optimal Lagrange multipliers (e.g. source rates). Following the approach of [@gatsis11], we can compute optimal routing variables while performing the subgradient iterations. Specifically let as assume that $N_o$ is the total number of subgradient iterations, while the index $\eta \in {0,...,N_o-1}$ is used to distinguish each one iteration. Then, if applying $\bar{r}_{ij}^d(t,N_o)=\frac{\sum_{\eta=0}^{N_o-1}r_{ij}^{d*}(t,\eta)}{N_o},~\forall (i,j),d$ as optimal routing variables for each $t\in [s,T]$, we can achieve a close to the optimal value of $P_1$, using diminishing step size. This can be proven as in [@gatsis11], if we first reformulate $P_1$ in an equivalent form replacing the objective function by the optimization variable $P'$ and adding the constraint ${\mathbb{E}}_{s,x}\left[\int_s^T \left(\sum_{i,d:i\in S_r(d)} U_i^d(\lambda_i^d(t),t)- \sum_{(i,j)}J_{ij}(P_{ij}(t))\right)dt\right]\geq P'$. Then, obviously, Theorem \[thm:1\] holds. More discussion on a possible (suboptimal) online implementation of the proposed algorithm is made in Section \[sec:simulation\]. We also note that instantaneous values of the controls (online approach - as functions of $X(t)$) during the network operation for such a problem may be obtained via a dynamic programming solution methodology (Hamilton-Jacobi-Bellman partial differential equation (HJB pde)) [@fleming06] which adds dramatically to complexity for a wireless multihop network (specifically the solution of the HJB pde is completely inefficient [@cui14], [@olama06]). Orthogonal Access to the Medium {#sec:orthogonal} =============================== In this section, we redesign the problem $P_1$ allowing only orthogonal access to the wireless medium, and thus leading to convex link capacity forms (Section \[sec:sysmodel\]) since the noise from interference will become negligible. In order to achieve this, we introduce new optimization variables for each independent set $\iota$, denoted as $\pi_{\iota}$, expressing the activation percentage of the corresponding independent set at time $t\in [s,T]$, and further satisfying the relations: $ \sum_{\iota} \pi_{\iota}(t) \leq 1, ~0\leq \pi_{\iota}(t) \leq 1,~\forall \iota, t \in [s,T].$ In the following, $\Pi$ stands for the collection of all $\pi_{\iota},~\forall \iota$ and $I_n$ is the number of the independent sets of the network’s connectivity graph. Since the channel state is a stochastic process, similarly to the definition of the rest of the control variables (Section \[sec:sysmodel\]), we define the value range for each $\pi_{\iota}$, $U_{\pi}=[0,1]$, and the corresponding feasible set $\mathcal{U}_{\pi}=\left\{\pi:[s,T]\times \Omega\rightarrow U_{\pi}: \pi \mathrm{~is~} \{\mathcal{F}_t\}_{t\geq s} \mathrm{~adapted} \right\}$. Then, $\Pi \in \mathcal{U}_{\Pi}=\{\mathcal{U}_{\pi}^{I_n}: \sum_{\iota=1}^{I_n} \pi_{\iota}(t) \leq 1, ~\forall t \in [s,T]\}$. The new optimization variables impose time-sharing among the independent sets, thus they render the interference levels negligible and the capacity of each link $(i,j)$ is given by the concave function in Section \[sec:sysmodel\]. The time share corresponding to the link $(i,j)$ at $t\in [s,T]$, is given by $\sum_{\iota: (i,j) \in \iota} \pi_{\iota}(t)$. The new optimization problem $\mathbf{P_2}$ is formulated as: $$\begin{aligned} P_2:=\max_{\Lambda \in \mathcal{U}_{\lambda}^F, ~R \in \mathcal{U}_{r}^{E\times (N-1)},~ P \in \mathcal{U}_{P}^{E}, ~\Pi \in \mathcal{U}_{\Pi}} {\mathbb{E}}_{s,x}\left[\int_s^T \left(\sum_{i,d:i\in S_r(d)} U_i^d(\lambda_i^d(t),t)- \sum_{(i,j)}\sum_{\iota: (i,j) \in \iota} \pi_{\iota}(t)J_{ij}(P_{ij}(t))\right)dt\right]\nonumber\\ s.t.\nonumber\end{aligned}$$ $$\begin{aligned} {\mathbb{E}}_{s,x}\left[\int_s^T \lambda_i^d(t)dt +\int_s^T \sum_{j:i \in \mathcal{R}(j,d)} r_{ji}^d(t)dt\right]\leq {\mathbb{E}}_{s,x}\left[\int_s^T \sum_{j\in \mathcal{R}(i,d)} r_{ij}^d(t)dt\right],~ \forall i,d \label{const:1b}\\ {\mathbb{E}}_{s,x}\left[\int_s^T \sum_{d} r_{ij}^d(t)dt\right]\leq {\mathbb{E}}_{s,x}\left[\int_s^T \sum_{\iota: (i,j) \in \iota} \pi_{\iota}(t)C_{ij}(P_{ij}(t))dt\right], ~\forall (i,j)\in \mathcal{E}\label{const:2b}\\ {\mathbb{E}}_{s,x}\left[\int_s^T\sum_{j\in \mathcal{N}^{out}_i}\sum_{\iota: (i,j) \in \iota} \pi_{\iota}(t)P_{ij}(t)dt\right]\leq P_{i,\max}, ~\forall i\label{const:3b}\end{aligned}$$ The formulation of $P_2$ is similar to the one of $P_1$ (Section \[sec:nonortho\]), with the difference that in $P_2$, we have introduced the new optimization variables $\Pi \in \mathcal{U}_{\Pi}$, the link capacities are concave and the link transmission powers, $P_{ij}(t)$, the link costs with respect to the transmission powers, $J_{ij}(P_{ij}(t))$, and the link capacities, $C_{ij}(P_{ij}(t))$, are replaced by their effective values $\sum_{\iota: (i,j) \in \iota} \pi_{\iota}(t)P_{ij}(t)$, $\sum_{\iota: (i,j) \in \iota} \pi_{\iota}(t)J_{ij}(P_{ij}(t))$, $\sum_{\iota: (i,j) \in \iota} \pi_{\iota}(t)C_{ij}(P_{ij}(t))$, correspondingly [@marques]. $P_2$ is non-convex due to the appearance of the control variables in multiplicative form in the objective function and the constraints (Eqs. (\[const:2b\]), (\[const:3b\])) in addition to the general forms of the utility and cost functions. However, in a similar way as for $P_1$, it can be shown that $P_2$ has a zero duality gap. Let $\mu_i^d\geq 0, ~\forall i,d, ~l_{ij}\geq 0,~\forall (i,j)\in \mathcal{E},~\nu_i\geq 0, ~\forall i$, be the Lagrange multipliers associated with the constraints (\[const:1b\]), (\[const:2b\]), (\[const:3b\]), respectively. Denote with $L$ the whole set of the Lagrange multipliers. Then, the dual function is formulated as follows: $$\begin{aligned} \label{eq:dual2} L_A(L)=\max_{\Lambda \in \mathcal{U}_{\lambda}^F, ~R \in \mathcal{U}_{r}^{E\times (N-1)},~ P \in \mathcal{U}_{P}^{E},~\Pi \in \mathcal{U}_{\Pi}} ~{\mathbb{E}}_{s,x}\left[\int_s^T \left(\sum_{i,d:i\in S_r(d)} U_i^d(\lambda_i^d(t),t)- \sum_{(i,j)}\sum_{\iota: (i,j) \in \iota} \pi_{\iota}(t)J_{ij}(P_{ij}(t))\right)dt\right]\nonumber \\ -\sum_{i,d} \mu_i^d {\mathbb{E}}_{s,x}\left[\int_s^T \lambda_i^d(t)dt +\int_s^T \sum_{j:i \in \mathcal{R}(j,d)} r_{ji}^d(t)dt-\int_s^T \sum_{j\in \mathcal{R}(i,d)} r_{ij}^d(t)dt\right] -\nonumber\\\sum_{(i,j)\in\mathcal{E}}l_{ij}{\mathbb{E}}_{s,x}\left[\int_s^T \left(\sum_{d} r_{ij}^d(t)-\sum_{\iota: (i,j) \in \iota} \pi_{\iota}(t) C_{ij}(P_{ij}(t))\right)dt\right]- \sum_{i} \nu_i {\mathbb{E}}_{s,x}\left[\int_s^T\sum_{j\in \mathcal{N}^{out}_i}\sum_{\iota: (i,j) \in \iota} \pi_{\iota}(t)P_{ij}(t)dt- P_{i,\max}\right].\end{aligned}$$Then, the dual problem is defined as $D_2:=\inf_{L}(L_A(L))$. The problem $P_2$ has zero dual gap.\[thm:2\] The proof is briefly described in Appendix B, provided as a supplementary file, as it is very similar to the proof of Theorem \[thm:1\]. Now, we obtain the optimal value of the problem $P_2$ via solving its dual. The renewal equations of the Lagrange multipliers are given by ($\eta=\{0,1,2..\}$): $$\begin{aligned} \label{eq:lagrange2} \mu_i^d(\eta+1)=\left\{\mu_i^d(\eta)+\kappa(\eta)\cdot {\mathbb{E}}_{s,x}\left[\int_s^T \lambda_i^{d*}(t)dt +\int_s^T \sum_{j:i \in \mathcal{R}(j,d)} r_{ji}^{d*}(t)dt-\int_s^T \sum_{j\in \mathcal{R}(i,d)} r_{ij}^{d*}(t)dt\right]\right\}^+,~ \forall i,d,\end{aligned}$$ $$\begin{aligned} l_{ij}(\eta+1)=\left\{l_{ij}(\eta)+\kappa(\eta)\cdot {\mathbb{E}}_{s,x}\left[\int_s^T \sum_{d} r_{ij}^{d*}(t)dt-\int_s^T \sum_{\iota: (i,j) \in \iota} \pi_{\iota}^*(t)C_{ij}(P_{ij}^*(t))dt\right]\right\}^+, ~\forall (i,j)\in\mathcal{E}, \label{eq:lagrange2b}\\ \nu_i(\eta+1)=\left\{\nu_i(\eta)+\kappa(\eta)\cdot {\mathbb{E}}_{s,x}\left[\int_s^T\sum_{j\in \mathcal{N}^{out}_i}\sum_{\iota: (i,j) \in \iota} \pi_{\iota}^*(t)P_{ij}^*(t)dt- P_{i,\max}\right]\right\}^+,~\forall i.\label{eq:lagrange2c}\end{aligned}$$with given values $\mu_i^d(0)\geq 0, ~\forall i,d, ~l_{ij}(0)\geq 0,~\forall (i,j)\in \mathcal{E},~\nu_i(0)\geq 0, ~\forall_i$. The optimal values of the stochastic processes $\lambda_i^{d*},~\forall i,d,~r_{ij}^{d*},~\forall (i,j),d$, for each $\eta$ are computed by Eqs. (\[eq:lamda\]), (\[eq:r\]) correspondingly and the same observations hold. Regarding the optimal values $P_{ij}^*,~\forall (i,j)\in \mathcal{E},~\pi_{\iota}^*,~\forall \iota$, for each $\eta$, they are obtained by solving $\forall~t\in [s,T]$: $$\begin{aligned} \label{eq:p2} \max_{P,\Pi} \sum_{(i,j)}\sum_{\iota: (i,j) \in \iota} \pi_{\iota} \left(l_{ij}(\eta)C_{ij}(P_{ij})-J_{ij}(P_{ij})-\nu_i(\eta) P_{ij}\right)= \max_{P,\Pi} \sum_{\iota}\pi_{\iota} \sum_{(i,j) \in \iota} \left(l_{ij}(\eta)C_{ij}(P_{ij})-J_{ij}(P_{ij})-\nu_i(\eta) P_{ij}\right),\end{aligned}$$which constitutes a maximum weight matching problem over the independent sets. Specifically, for its solution, each link $(i,j)$ computes the stochastic process $P_{ij}^*(t)$, $t\in[s,T]$ (depending on the state’s, $X$, path) by maximizing $-J_{ij}(P_{ij})+l_{ij}(\eta)C_{ij}(P_{ij})-\nu_i(\eta) P_{ij}$, i.e. solving $$\label{eq:p3} -\frac{\vartheta J_{ij}(P_{ij})}{\vartheta P_{ij}}+l_{ij}(\eta)\frac{\vartheta C_{ij}(P_{ij})}{\vartheta P_{ij}}-\nu_i(\eta) =0,~ \forall~t\in [s,T],$$while taking into account the $U_P$ constraints (Section \[sec:sysmodel\]). Then, each link $(i,j)\in \mathcal{E}$ is assigned a weight equal to $W_e(i,j,t)=\left(-J_{ij}(P_{ij}^*(t))+l_{ij}(\eta)C_{ij}(P_{ij}^*(t))-\nu_i(\eta) P_{ij}^*(t)\right)$ and finally the independent set $\iota^*$ maximizing the sum $\sum_{(i,j) \in \iota}W_e(i,j,t)$ receives $\pi_{\iota^*}(t)=1$, while $\pi_{\iota}(t)=0$ for $\iota\neq \iota^*$ at $t\in [s,T]$. In this paper, we assume that ties break arbitrarily, however, a study on how to break ties can be found in [@marques]. The computation of the expected values involved in the Lagrange multipliers follows the lines of the Monte Carlo simulations described in Section \[sec:nonortho\] and the algorithm for solving $D_2$ is similar to the one in Section \[sec:nonortho\]. Note that Eq. (\[eq:p3\]) can be solved link-wise in a very efficient manner, contrary to Eq. (\[eq:p\]) which involves the complex cross-terminal problem. The observations regarding the convergence to the optimal Lagrange multipliers and primal values are of similar nature to the ones of Section \[sec:nonortho\]. At this point we study the solution of Eq. (\[eq:p3\]) in more detail in order to gain more insight regarding the optimal power control. Let us assume LTF and convex link costs of the form $J_{ij}(P_{ij})=V P_{ij}^2$, where $V>0$ is a constant for all $(i,j)\in \mathcal{E}$. Then, for a given $\eta$, the solution of Eq. (\[eq:p3\]) is explicitly given by: $$\begin{aligned} \label{eq:p3solution} P_{ij}^*(X_{ij}(t))= \max\big\{0,\min\{\tilde{P}_{ij}^*(X_{ij}(t)),P_{\max}\}\big\},~\mathrm{where} \nonumber \end{aligned}$$ $$\begin{aligned} \tilde{P}_{ij}^*(X_{ij}(t))=\frac{1}{2}\left[-\left(N_0 e^{-KX_{ij}(t)}+\frac{\nu_i(\eta)}{2V} \right)+\sqrt{\left(N_0 e^{-KX_{ij}(t)}+\frac{\nu_i(\eta)}{2V} \right)^2-4\left( \frac{\nu_i(\eta) N_0e^{-KX_{ij}(t)}}{2V}-\frac{l_{ij}(\eta) B_{ij}}{2V \log(2)}\right)}\right]\end{aligned}$$ Note that the optimally controlled power at $\eta$ never exceeds the value $$\begin{aligned} \label{eq:optPowermax}P_{ij}^{* \max}=\frac{1}{2}\left(\sqrt{\left( \frac{\nu_i(\eta)}{2V}\right)^2+\frac{4l_{ij}(\eta) B_{ij}}{2V \log(2)}} -\frac{\nu_i(\eta)}{2V} \right),\end{aligned}$$ irrespectively of the $P_{\max}$ value, while this value is achieved asymptotically when $X_{ij}(t)\rightarrow -\infty$. Note also that $P_{ij}^*(X_{ij}(t))\rightarrow 0$ when $X_{ij}(t)\rightarrow \infty$. Therefore, the optimal power control exploits low power loss values created by random fluctuations to increase the link capacity as much as possible, thus positively affecting the flows’ throughput (rates). On the other hand, transmission power is not wasted when the power loss is high. Note that in case of orthogonal access to the medium the aim of power control is not scheduling (which is defined via the $\Pi$ variables) but to take advantage of the channel when it is favorable and to avoid depleting energy when the channel is destructive. In the following, we assume no power control and scheduling and we prove an interesting counter-intuitive theorem regarding the relation between the achieved utility by the network and the channel’s diffusion coefficient in the case of LTF. Absence of power control means that in $P_2$, $J_{ij}(t)=0,~\forall (i,j), t \in [s,T]$, the constraint of Eq. (\[const:3b\]) is dropped and finally, $P_{ij}(t),~\forall (i,j), t \in [s,T]$ are constant and predefined. Absence of scheduling means that the time share of each link is constant and predefined, e.g. $\sum_{\iota: (i,j) \in \iota} \pi_{\iota}(t)=\zeta_{ij}\geq 0,~\forall (i,j), t \in [s,T]$. Utilities as Increasing Functions of the Channel’s Diffusion Coefficient in the case of LTF {#sec:utilities} ------------------------------------------------------------------------------------------- We examine the effect of the channel’s diffusion coefficient, on the users’ optimal utility. Contrary to what is perhaps expected, we prove that an increase of a channel’s diffusion coefficient (SDE (\[eq:sdepathloss\])) leads to increased achieved users’ utility. This result emphasizes the importance of a more realistic model for the power loss such as the one of SDE (\[eq:sdepathloss\]), as opposed to the mean power loss model used in the conventional NUM problem formulation. Let us consider two networks $1$, $2$, one noisier than the other, but otherwise identical. Precisely, for $k\in\{1,2\},~\forall\,(i,j),~\forall\,t\in [t_0,T],$ $$\begin{aligned} dX_{ij}^k(t)=\beta_{ij}(t)\big(\gamma_{ij}(t)-X_{ij}^k(t)\big)\, dt+\delta_{ij}^k(t)\, dW_{ij}^k(t),~ X_{ij}^k(t_0)=x_{ij,0},~\mathrm{and}~\delta_{ij}^1(t)\leq \delta_{ij}^2(t). \label{eq:sdepathlossi}\end{aligned}$$ If $J^*_k, ~k\in\{1,2\},$ is the optimal objective value achieved for each network, then $J^*_1 \leq J^*_2$. \[thm:volatilityincrease\] The solutions of the SDEs in Eq. (\[eq:sdepathlossi\]) can be written, $\forall (i,j)$, as $$\begin{aligned} X_{ij}^k(t)=X_{ij}^{det}(t)+\int_{t_0}^t \delta_{ij}^k(s)e^{-\int_s^t \beta_{ij}(r)dr}dW_{ij}^k(s),~k\in\{1,2\}, \label{eq:solutionpowerlossi}\end{aligned}$$ where $X_{ij}^{det}(t)=e^{-\int_{t_0}^t \beta_{ij}(s)ds}x_{ij0}+ \int_{t_0}^t \beta_{ij}(s) \gamma_{ij}(s)e^{-\int_s^t \beta_{ij}(r)dr}ds$ is the solution for a noiseless channel. Note that ${\mathbb{E}}_{t_0,x_0}\big[X_{ij}^k(t)\big]=X_{ij}^{det}(t),~k\in\{1,2\}.$ By Proposition $3.1$ in [@hirsch14] we have that $X_{ij}^1(t)\stackrel{(c)}{\leq} X_{ij}^2(t), ~\forall t \in [t_0,T]$, where $\stackrel{(c)}{\leq}$ stands for partial ordering in the convex order. That is, for every convex function $\psi$ we have ${\mathbb{E}}_{t_0,x_0}\left[ \psi\big(X_{ij}^1(t)\big) \right] \leq {\mathbb{E}}_{t_0,x_0}\left[ \psi\big(X_{ij}^2(t)\big) \right], ~\forall t \in [t_0,T]$. One such convex function is Shannon’s formula for the channel’s capacity $C_{ij}$, i.e. $x\rightarrow B_{ij}\log_2\left(1+\frac{e^{Kx}P_{ij}}{N_{0}}\right)$. Hence, $$\begin{aligned} {\mathbb{E}}_{t_0,x_0}\left[ C_{ij}^1(t) \right] \leq {\mathbb{E}}_{t_0,x_0}\left[ C_{ij}^2(t) \right], ~\forall t \in [t_0,T]. \label{eq:result3}\end{aligned}$$ In particular, let us denote by $\mathfrak{L}^k,~k\in\{1,2\}$, the set of the deterministic $\Lambda \in \mathcal{U}_{\lambda}^F,~R \in \mathcal{U}_{r}^{E\times (N-1)}$ that satisfy the constraints of $P_2$ imposed on each network: $$\begin{aligned} {\mathbb{E}}_{t_0,x_0}\left[\int_{t_0}^T \lambda_i^d(t)dt +\int_{t_0}^T \sum_{j:i \in \mathcal{R}(j,d)} r_{ji}^d(t)dt\right]\leq {\mathbb{E}}_{t_0,x_0}\left[\int_{t_0}^T \sum_{j\in \mathcal{R}(i,d)} r_{ij}^d(t)dt\right],~ \forall i,d, \\ {\mathbb{E}}_{t_0,x_0}\left[\int_{t_0}^T \sum_{d} r_{ij}^d(t)dt\right]\leq {\mathbb{E}}_{t_0,x_0}\left[\int_{t_0}^T \zeta_{ij} C_{ij}^k(t)dt\right], ~\forall (i,j)\in \mathcal{E},~k=\{1,2\}. $$ In view of Rel. (\[eq:result3\]) we have that $\mathfrak{L}^1\subseteq \mathfrak{L}^2$ and therefore, as asserted, $$\begin{aligned} J_1^*=\max_{\Lambda,~R \in \mathfrak{L}^1} {\mathbb{E}}_{t_0,x_0}\left[\int_{t_0}^T \sum_{i,d} U_i^d(\lambda_i^d(t),t) dt\right]\leq \max_{\Lambda,~R \in \mathfrak{L}^2} {\mathbb{E}}_{t_0,x_0}\left[\int_{t_0}^T \sum_{i,d} U_i^d(\lambda_i^d(t),t) dt\right]=J_2^*.\end{aligned}$$ Note that $C_{ij}\big(\delta_{ij}(.)\big)\ge B_{ij}{\mathbb{E}}^{A_0}\Big[\log_2\big(1+\frac{P_{ij}e^{KX_{ij}(t)}}{N_0}\big)\Big] \ge \mathbf{P}(A_0) B_{ij}\log_2\frac{P_{ij}}{N_0}+\frac{B_{ij}K}{\log 2}{\mathbb{E}}^{A_0}\big[X_{ij}(t)\big]$, where $A_0=\big\{X_{ij}(t)\ge {\mathbb{E}}[X_{ij}(t)]\big\}$. Since $X_{ij}(t)$ are gaussian, $\mathbf{P}(A_0)=1/2$ by symmetry, and ${\mathbb{E}}^{A_0}[X_{ij}(t)]= \frac{1}{2}{\mathbb{E}}[X_{ij}(t)]+\sqrt{\frac{V(X_{ij}(t))}{2\pi}}$ where $V(X_{ij}(t))$ is the variance of $X_{ij}(t)$ given by $\int_{s}^{t}\delta_{ij}^2(r)\exp(-2\int_{r}^{t}\beta_{ij}(q)dq)dr$. Therefore, the link capacity may take arbitrarily large values if $\int_s^T \delta_{ij}^2(t)\,dt$ is sufficiently large. Numerical Results {#sec:simulation} ================= In this section, we present and discuss indicative numerical results evaluating the proposed schemes focusing on the case of orthogonal access to the medium and LTF. After describing the evaluation setting and some general observations, we illustrate numerically Theorem \[thm:volatilityincrease\] and the behavior of the proposed framework in case of congestion control and routing (i.e. optimization in transport and network layers). The latter is then compared with the behavior of the joint routing, scheduling, congestion and power control scheme (i.e. cross-layer optimization). In addition, we examine the behavior of the proposed framework in the case of time varying utilities and the possibility of applying the solution procedure of the dual problem $D_2$, during the network operation (online deployment). We consider a wireless multihop network of $N=16$ nodes, forming a $4\times 4$ grid topology. For ease of presentation, we consider that the parameters of the SDE (\[eq:sdepathloss\]), $\gamma_{ij}(t),\beta_{ij}(t),\delta_{ij}(t)$, as well as the initial states $X_{ij}(s)=x$ are identical for every link $(i,j)$. Here, we consider $M=200$ paths, $N_{0}=0.1W$, $s=0$. Furthermore, $\forall\, (i,j),\tau_b$, $B_{ij}=10^6 Hz$, $\gamma_{ij}(\tau_ b)=\gamma \bigg(1+0.15 e^{(-2 \frac{\tau_b}{n})} \sin(10 \pi \frac{\tau_b}{n})\bigg)$ [@olama10], $X_{ij}(0)=\gamma$, $n=500$, and $\beta_{ij}(\tau_b)=100$, $\gamma=70 dB$ unless differently mentioned. Moreover, $\delta_{ij}(\tau_b)=\delta,\forall\, (i,j),\tau_b,$ where $\delta$ will be tuned in each numerical experiment. Note that the value $n=500$ determines the sampling rate for computing the Riemann sums that approximate the integrals e.g. in Eqs. (\[eq:lagrange2b\]), (\[eq:lagrange2c\]) and it is chosen so that the Riemann sum is close to the corresponding integral value, while simultaneously being small enough for trading-off the cost of sampling. Low sampling rate impacts performance since the Riemann sum does not converge to the actual value of the corresponding integral. On the contrary high sampling rate may induce extra cost without offering significant improvement in the Riemann sum’s accuracy in approximating the corresponding integral. The periodic behavior of the LTF wireless channel parameters may be due to an absorbing obstacle intervening periodically between the transmitter and the receiver (as in an example of [@goldsmith05]). In the absence of scheduling optimization, for each link $(i,j)$ the value $\zeta_{ij}$ is computed based on scheduling all the maximal independent sets of the network topology for equal percentage of time. Regarding the traffic model, each node chooses a random destination among its non-physically connected nodes, and becomes the source for this destination. Initially, we consider logarithmic utilities, i.e. $U_i^d(\lambda_i^d)=\log (\lambda_i^d)$, a common choice in the literature to model elastic traffic [@georgiadis06]. Each numerical experiment for the determination of the optimal control variables and the optimal Lagrange multipliers runs until convergence is ensured and specifically until the sum of the changes between consecutive values of the Lagrange multipliers over the preceding eight iterations is less than $0.001$. The learning rate is chosen as $\kappa(\eta)=\frac{A'}{\eta},~\forall\,\eta$ [@bertsekas09], where $A'=0.1$ so that convergence is allowed in a reasonable time duration with respect to the chosen convergence criterion. Specifically, by decreasing $A'$ by one or more orders of magnitude our criterion of convergence is satisfied too soon, impeding the subgradient algorithm approach to the global optimal values. On the other hand, increasing $A'$ or using constant values of $\kappa(\eta),~\forall\,\eta$ do not lead to convergence within a reasonable time interval. In order to obtain an intuition regarding the wireless channel’s stochastic behavior, Figs. \[fig:capacitypaths\], \[fig:capacitypaths1\] show typical sample paths of the solution to the SDE (\[eq:sdepathloss\]) for $\delta=25$ and $\delta=50$, respectively. As expected, we observe larger deviations of the power loss from $\gamma_{ij}(\cdot)$ for the higher choice of $\delta$. Consequently, the capacity may achieve higher values due to random fluctuations in this case. Note that although the curve of $\gamma_{ij}(\cdot)$ tends to weaken to a line as time increases, it fluctuates considerably for the duration of the network’s operation (transient state). Fig. \[fig:capacitypaths3\] shows typical sample paths of the solution to the SDE (\[eq:sdepathloss\]) for time varying speed of adjustment $\beta_{ij}(\tau_b)$. When $\beta_{ij}(\tau_b)$ attains low values (i.e. $\tau_b\leq166$) the power loss diverges more from its attraction curve $\gamma_{ij}(\tau_b)$, while the attraction to $\gamma_{ij}(\tau_b)$ becomes faster when $\beta_{ij}(\tau_b)=500$ (i.e. $\tau_b>333$). Congestion Control & Routing ---------------------------- In the sequel, we examine the case of applying only congestion control and routing, assuming the transmitter of each link $(i,j)$ has a constant power of $P_{ij}(t)=2 W$, $\forall~t\in [0,T]$. Fig. \[fig:routingcongestiondelta\] shows the optimal source rates (i.e. after convergence since they change continuously with respect to the optimal Lagrange multipliers) for different choices of the diffusion coefficient $\delta$. It can be observed that as $\delta$ increases, random fluctuations to higher capacity values are exploited to offer increased optimal source rates, hence verifying numerically the statement of Theorem \[thm:volatilityincrease\]. It is important to note here that the noise level $\delta$ does not affect the mean power loss, as indicated in the proof of Theorem \[thm:volatilityincrease\]. In other words, if the mean power loss is used to determine capacity, higher capacity values due to random fluctuations cannot be tracked and exploited for increasing the source rates. Time-varying $\delta$ (TV) is also applied, and specifically $\delta=15 sin(10 \pi d/n)+35$ i.e. taking values between $\delta=20$ and $\delta=50$, thus leading to optimal source rates in between the ones corresponding to these two values of $\delta$. For benchmarking purposes, in Fig. \[fig:routingcongestiondelta\] we have added the case of $\delta=0$, which corresponds to time-varying but deterministic channels. We observe that the optimal source rates achieved under deterministic wireless channels are the lowest (approximately the same as in the case of a low value of noise, i.e. the case that $\delta=5$), indicating the improvement in system’s performance when accounting for randomness. Figs. \[fig:convM\], \[fig:convL\] depict the behavior of the proposed scheme considering convergence to the optimal Lagrange multipliers. We also study the impact of $T$ on the time to convergence and the achieved source rates. The value of $n$ is adapted for each $T$ as it is shown in Fig. \[fig:Limpact\]. We observe that as the duration of the network’s operation, $T$, increases, the time to convergence also increases (Fig. \[fig:Limpact\]), while the achieved arrival source rates decrease for all flows (Fig. \[fig:Timpact\]). Finally, the behavior of the proposed algorithm in case of time varying $\beta_{ij}(t)$ with respect to the optimal source rates is shown in Fig. \[fig:betaimpact\]. We observe that when $\beta_{ij}(t)$ is high at the beginning (close to time $s=0$) the optimal rates are higher than when $\beta_{ij}(t)$ is initially low and increases later in time. Joint Scheduling, Routing, Congestion & Power Control Scheme ------------------------------------------------------------ In order to evaluate the joint scheduling, routing, congestion and power control scheme, we assume that each link can vary its transmission power between $1W$ and $3W=P_{i,\max}, ~\forall i$. The network topology and traffic along with the rest of the parameters remain the same as in the previous experiments. Fig. \[fig:capacitypower\] depicts the optimal transmission power at each repetition of channel state’s sampling derived from Eq. (\[eq:p3solution\]) and the corresponding path of the link capacity. It is observed that the optimal power increases when power loss decreases (thus capacity increases) and attains low values for high values of power loss, as expected from the analysis of Section \[sec:orthogonal\]. Therefore, the transmission power increases only when there is an opportunity for an important capacity improvement due to random dips of power loss. On the contrary, transmission power is not wasted when the stochastic power loss does not support capacity increase. Fig. \[fig:powerroutingcongestiondelta\] shows the optimal source rates (i.e. after convergence) for different choices of the diffusion coefficient $\delta$. As in the previous case (Fig. \[fig:routingcongestiondelta\]), it can be observed that as $\delta$ increases the optimal source rates increase for all flows. This is beyond the scope of Theorem \[thm:volatilityincrease\], that does not account for power control and scheduling in the optimization problem. Fig. \[fig:powerroutingcongestiondelta\] also includes the optimal source rates when $\delta$ is time-varying (TV) and when channels are deterministic ($\delta=0$), leading to similar conclusions as in the case of congestion control and routing (Fig. \[fig:routingcongestiondelta\]). Figs. \[fig:convMP\], \[fig:convLP\] depict the convergence to the optimal Lagrange multipliers. Fig. \[fig:ratescomparison\] shows that the joint scheduling, routing, congestion and power control scheme improves the optimal source rates for all flows, compared to the congestion control and routing scheme. As also stated in [@goldsmith05], applying power control at the transmitter is like having channel state information at both the transmitter and the receiver which improves the network capacity compared with the case of constant power which is equivalent to having channel state information at the receiver’s side only. Furthermore, we study the expected per link transmission power over the entire time interval $[0,T]$ for the numerical experiments of Fig. \[fig:powercongroutsch\]. It can be easily computed that it is equal to $1.0399W$ for $\delta=50$, $1.2658W$ for $\delta=20$, $1.5634W$ for $\delta=5$, $1.6249W$ for $\delta=0$. On the one hand these values are much smaller than the constant power of $2W$ used to evaluate the routing and congestion control scheme (Fig. \[fig:routingcongestiondelta\]), thus achieving a more “green" network operation in addition to throughput improvement (Figs. \[fig:powerroutingcongestiondelta\], \[fig:ratescomparison\]) leading to energy efficiency. On the other hand, we observe that as $\delta$ increases, the expected per link transmission power decreases, indicating the importance of taking randomness into account in operating wireless channels with energy efficiency. Time Varying Utilities & Online Deployment ------------------------------------------ Finally, we study the case of time-varying utilities when applying routing and congestion control. The utilities take the form $U_i^d(\lambda_i^d)=\frac{\log (\lambda_i^d)}{t},~t\in[s,T],~s>0~\forall i,d:i \in S_r(d)$, i.e., they decrease with time modeling the decreasing willingness of users to produce high data amounts when approaching the end of the network’s operation. The optimal function to which the source rates converge over $[s,T]$ is depicted in Fig. \[fig:RoutingCongestionTimeVarying\]. We observe that as time increases, it dominates over the Lagrange multiplier for the determination of the source rates (Eq. (\[eq:lamda\])). At this point we will make another interesting observation regarding the online application of the proposed approach (Section \[sec:orthogonal\]) during the network’s operation which is initially discussed in Section \[sec:nonortho\]. In order to obtain an online algorithm for the network control, i.e. during the network operation, we may consider that the network decisions are taken at $\tau_b$ times when the channel is sampled, while also considering that $\tau_b\equiv \eta$. Then, we should consider the optimal control values in the interval $(\tau_b,T]$ as “predicted" and the ones in the interval $[s,\tau_b)$ as “corrections". However, we should study under what conditions convergence is achieved (in practice) early enough (for small $\tau_b$) so that optimality with respect to the achieved value of $P_2$ is not affected. In Figs. \[fig:ntimeconv1\], \[fig:ntimeconv2\], it is shown that when increasing $n$, the time for convergence (according to the imposed criterion) of the proposed scheme is not significantly affected for time invariant utilities while it is affected in a concave manner for time varying utilities. As a result, if $n$ is large enough, the decisions taken at times $\tau_b$ will converge fast enough compared to the whole duration $T$. Therefore, the online application of the proposed approach during the network operation will be suboptimal only at the beginning barely affecting the optimal objective value of $P_2$. Conclusions {#sec:conclusions} =========== In this paper we presented, analyzed and evaluated a framework of NUM for performing routing, scheduling, congestion and power control under stochastic possibly non-stationary LTF or STF wireless channels modeled by SDEs. The continuous stochastic non-stationary wireless channels along with the consideration of transient phenomena lead to a problem formulation that can also tackle non-convex and time-varying objective functions in an optimal way. Power control aims at increasing users’ experience allowing for higher source rates while also improving the energy efficiency. In the case of LTF, we prove that higher values of the diffusion coefficient of the power loss lead to higher optimal users’ utilities, a fact that cannot be captured by the conventional NUM problem’s formulation. Numerical results evaluate the latter along with the convergence properties of our proposed algorithms and the effect of diverse parameters on it. The efficiency of power control and the conditions under which an online implementation of the proposed approach is possible are also investigated. Finally, our proposed NUM-based framework may constitute a core for devising efficient cross-layer algorithms for the network operation that incorporate transient or non-stationary phenomena. Acknowledgment ============== This research is co-financed by the European Union (European Social Fund) and Hellenic national funds through the Operational Program ’Education and Lifelong Learning’ (NSRF 2007-2013). (under “ARISTEIA" 1260). M.L. acknowledges support from the NSRF Research Funding Program Thales: Optimal Management of Dynamical Systems of the Economy and the Environment MIS375586. [10]{} M. Chiang, “Balancing Transport and Physical Layers in Wireless Multihop Networks: Jointly Optimal Congestion Control and Power Control", *IEEE Journal on Selected Areas in Communications*, Vol. 23, No. 1, pp. 104-116, Jan. 2005. J. Papandriopoulos, S. Dey, J. S. Evans, “Optimal and Distributed Protocols for Cross-Layer Design of Physical and Transport Layers in MANETs", *IEEE/ACM Trans. on Networking*, Vol. 16, No. 6, pp. 1392-1405, Dec. 2008. C. Fischione, M. D’Angelo, M. Butussi, “Utility Maximization via Power and Rate Allocation with Outage Constraints in Nakagami-Lognormal Channels", *IEEE Trans. on Wireless Communications*, Vol. 10, No. 4, pp. 1108-1120, April 2011. D. O’Neill, B. Sim Thian, A. Goldsmith, S. Boyd, “Wireless NUM: Rate and Reliability Tradeoffs in Random Environment", *in Proc. of the IEEE Wireless Communications and Networking Conference (WCNC)*, pp. 1-6, April 2009. S. Firouzabadi, D. O’Neill, A. Goldsmith, “Distributed Wireless Network Utility Maximization", *In Proc. of the 44th Annual Conference on Information Sciences and Systems (CISS)*, pp.1-6, March 2010. A. G. Marques, N. Gatsis, G. B. Giannakis, “Optimal Cross-Layer Design of Wireless Fading Multi-hop Networks", *Cross Layer Designs in WLAN Systems, Leicester, UK:Troubador Publishing*, 2011. L. Chen, S. H. Low, M. Chiang, J. C. Doyle, “Cross-Layer Congestion Control, Routing and Scheduling Design in Ad Hoc Wireless Networks", *in Proc. of IEEE INFOCOM*, pp. 1-13, April 2006. L. Huang, S. Moeller, M. J. Neely, B. Krishnamachari, “LIFO-Backpressure Achieves Near-Optimal Utility-Delay Tradeoff", *IEEE/ACM Trans. on Networking*, Vol. 21, No. 3, pp. 831-844, June 2013. J. Chen, V. K. N. Lau, Y. Chen, “Distributive Network Utlity Maximization Over Time-Varying Fading Channels”, *IEEE Transactions on Signal Processing*, Vol. 59, No. 5, pp. 2395-2404, May 2011. M. Olama, S. Djouadi, C. Charalambous, “Wireless Fading Channel Models: from Classical to Stochastic Differential Equations", *Stochastic Control, Chris Myers (Ed.), ISBN: 978-953-307-121-3, InTech, DOI: 10.5772/9738*, Aug. 2010. B. Oksendal, “Stochastic Differential Equations", *6th Edition Springer-Verlag*, 2003. C. D. Charalambous, N. Menemenlis, “Stochastic Models for Short-term Multipath Fading Channels: Chi-square and Ornstein-Uhlenbeck Processes", *In Proc. of the 38th IEEE Conf. on Decision and Control*, Vol. 5, pp. 4959-4964, 1999. E. Stai, S. Papavassiliou, “User Optimal Throughput-Delay Trade-off in Multihop Networks under NUM Framework”, *IEEE Communications Letters*, Vol. 18, No. 11, pp. 1999-2002, Nov. 2014. W. H. Fleming, H. M. Sooner, “Controlled Markov Processes and Viscosity Solutions", *2nd Edition, Springer-Verlag*, 2006. E. Stai, M. Loulakis, S. Papavassiliou, “Congestion & Power Control of Wireless Multihop Networks over Stochastic LTF Channels", *in Proc. of IEEE Wireless Communications and Networking Conference (WCNC)*, pp. 1769-1774, March 2015. G. Tychogiorgos, A. Gkelias, K. K. Leung, “A Non-Convex Distributed Optimization Framework and its Application to Wireless Ad-hoc Networks", *IEEE Trans. on Wireless Communications*, Vol. 12, No. 9, pp. 4286-4296, September 2013. A. Goldsmith, “Wireless Communications", *Cambridge University Press*, 2005. M. M. Olama, S. M. Djouadi, C. D. Charalambous, “Stochastic Differential Equations for Modeling, Estimation and Identification of Mobile-to-Mobile Communication Channels", *IEEE Transactions on Wireless Communications*, Vol. 8, No. 4, pp. 1754-1763, April 2009. C. D. Charalambous, R. J. C. Bultitude, X. Li, J. Zhan, “Modeling Wireless Fading Channels via Stochastic Differential Equations: Identification and Estimation Based on Noisy Measurements", *IEEE Transactions on Wireless Communications*, Vol. 7, No. 2, pp. 434-439, February 2008. M. J. Neely, “Delay-Based Network Utility Maximization”, *in Proc. of IEEE INFOCOM*, pp. 1-9, March 2010. A. Eryilmaz, R. Srikant, “Joint Congestion Control, Routing and MAC for Stability and Fairness in Wireless Networks”, *IEEE Journal on Selected Areas in Communications*, Vol. 24, No. 8, pp. 1514-1524, August 2006. A. Ribeiro, G. B. Giannakis, “Separation Principles in Wireless Networking", *IEEE Transactions on Information Theory*, Vol. 56, No. 9, pp. 4488-4505, Sept. 2010. D. P. Bertsekas, “Nonlinear Programming: 2nd Edition", *Athena Scientific*, 2009. N. Gatsis, A. Ribeiro, G. B. Giannakis, “A Class of Convergent Algorithms for Resource Allocation in Wireless Fading Networks", *IEEE Trans. on Wireless Communications*, Vol. 9, No. 5, pp. 1808-1823, May 2010. D. J. Higham, “An Algorithmic Introduction to Numerical Simulation of Stochastic Differential Equations", *SIAM REVIEW*, Vol. 43, No. 3, pp. 525-546, 2001. M. Olama, S. Djouadi, C. Charalambous, “Stochastic Power Control for Time-Varying Long-Term Fading Wireless Networks", *EURASIP Journal on Advances in Signal Processing*, pp. 1-13, June 2006. Y. Cui, E. M. Yeh, “Delay Optimal Control and its Connection to the Dynamic Backpressure Algorithm", *in Proc. of the IEEE International Symposium on Information Theory (ISIT)*, pp. 451-455, June-July 2014. F. Hirsch, M. Yor, “Comparing Brownian Stochastic Integrals for the Convex Order", *Modern Stochastics and Apps., Springer Optimization and Its Apps., Springer Int’l Publishing*, Vol. 90, pp. 3-19, 2014. L. Georgiadis, M. J. Neely, L. Tassiulas, “Resource Allocation and Cross-Layer Control in Wireless Networks”, *Foundations and Trends in Networking*, Vol. 1, No. 1, pp. 1-144, 2006.
{ "pile_set_name": "ArXiv" }
--- abstract: | Networks are powerful data structures, but are challenging to work with for conventional machine learning methods. Network Embedding (NE) methods attempt to resolve this by learning vector representations for the nodes, for subsequent use in downstream machine learning tasks. Link Prediction (LP) is one such downstream machine learning task that is an important use case and popular benchmark for NE methods. Unfortunately, while NE methods perform exceedingly well at this task, they are lacking in transparency as compared to simpler LP approaches. We introduce ExplaiNE, an approach to offer counterfactual explanations for NE-based LP methods, by identifying existing links in the network that explain the predicted links. ExplaiNE is applicable to a broad class of NE algorithms. An extensive empirical evaluation for the NE method ‘Conditional Network Embedding’ in particular demonstrates its accuracy and scalability. bibliography: - 'paper.bib' --- Introduction\[sec:introduction\] ================================ Network embeddings (NEs) have exploded in popularity in both the machine learning and data mining communities. By mapping a network’s nodes into a vector space, NEs enable the application of a variety of machine learning methods on networks for important tasks such as link prediction (LP): the task to predict whether nodes are likely to be(come) connected in incomplete or evolving networks. LP has wide-ranging applications, for friendship recommendations, recommender systems, knowledge graph completion, etc. While there are numerous conventional LP methods that predict links based on heuristic statistics computed over networks (e.g., based on the number of common neighbors) [see, e.g., @martinez2017survey], recently proposed NE-based methods typically outperform those heuristic approaches [e.g., @grover2016node2vec; @kang2018conditional]. While the superior performance of NE-based LP methods is an advantage, a major disadvantage is that they do not easily allow for human-intelligible explanations of the predicted links. Yet, the ability to understand link predictions is important and useful for several reasons: (a) recommender systems that provide explanations are more easily trusted and more effective, (b) it allows data analysts to have a better understanding of the network characteristics such as node features and network dynamics, (c) transparency of automated processing systems is required in a growing number of regulations, and explanations can increase transparency. We present ExplaiNE, a mathematically principled counterfactual reasoning approach for explaining NE-based link predictions. In its simplest form, ExplaiNE quantifies how the probability of a predicted link [$\{i,j\}$]{} would be affected by weakening an existing link [$\{i,k\}$]{}. Links [$\{i,k\}$]{} that after weakening most strongly reduce the probability of the predicted link [$\{i,j\}$]{} then serve as counterfactual explanations. ![In Zachary’s karate club network, we explain the predicted link between $i=33$ and $j=24$. The colored nodes with dashed or dotted edges are the neighbors of node $33$. According to ExplaiNE, the links to the orange nodes with dotted edge have a positive effect on the probability of link $\{i,j\}$ to exist, while the effect of the links to blue nodes with dashed edge is negative. \[fig:intro\_illustr\_karate\]](figures/intro_illustr_karate "fig:"){width="50.00000%"} -0.7cm [[**Example.**]{}]{} We show the idea of ExplaiNE on Zachary’s karate club network [@zachary1977information]. The network consists of $34$ karate club members (nodes), with $78$ friendship links. An NE-based LP method[^1] to predict a link for node $i=33$ (green pentagon) indicates a high probability link to node $j=24$ (green square). Figure \[fig:intro\_illustr\_karate\] visualizes the embedding and highlights which existing links incident to $i$ ExplaiNE deems explanatory for this prediction in a positive (orange circle, dotted edge) or negative sense (blue circle, dashed edge). It concludes this because weakening links to the orange nodes would reduce the link probability [$\{i,j\}$]{}, whereas weakening links to the blue nodes would increase it. Note that these effects are quite intuitive given the geometry of the embedding: the orange nodes ‘pull’ node $33$ closer to $24$, while the blue nodes pull node $33$ away from $24$. ExplaiNE is first derived as generically as possible, allowing for explanations not only in terms of links incident to the predicted link, but also in terms of other links as well as non-links. We then reduce its scope to explanations of the type used in the example above (i.e., only incident links), and make an approximation (which we justify empirically), in order to obtain a still generic but highly scalable approach. Next we apply ExplaiNE to Conditional Network Embedding [CNE; @kang2018conditional], a recent state-of-the-art NE method. The application of ExplaiNE to CNE is particularly transparent, thanks to the mathematical elegance of the CNE model and its straightforward use in LP, requiring no training once the embedding is found. We also outline how ExplaiNE can be applied to NE methods based on skip gram with negative sampling-based such as LINE [@tang2015line], DeepWalk [@perozzi2014deepwalk], PTE [@tang2015pte], and node2vec [@grover2016node2vec]. [[**Contributions.**]{}]{} The main contributions are: - ExplaiNE, a mathematically principled counterfactual reasoning approach for explaining link predictions based on network embeddings (Sec. \[sec:method\_explaine\]). - A scalable tight approximation of ExplaiNE (Sec. \[sec:appr\]). - A detailed application of ExplaiNE to CNE (Sec. \[sec:explaine\_cne\]) - An outline of how to apply ExplaiNE to NE methods based on skip gram with negative sampling (Sec. \[sec:explaine\_other\]) - Quantitative and run time analyses showing the stability and scalability of the approximation. (Sec. \[sec:exp\_approx\],\[sec:exp\_runtime\]) - Qualitative and quantitative realistic case studies confirming the usefulness of ExplaiNE. (Sec. \[sec:exp\_qualitative\],\[sec:exp\_quantitative\]) Methods\[sec:method\] ===================== To introduce ExplaiNE in full generality, we first provide a simple but generic description of NE-based link prediction methods in Section \[sec:method\_ne\_lp\]. We then formalize ExplaiNE in a generic manner in Section \[sec:method\_explaine\], before describing a scalable approximation in Section \[sec:appr\]. In Section \[sec:explaine\_cne\] we develop ExplaiNE in detail for CNE. In Section \[sec:explaine\_other\] we outline how ExplaiNE can also be applied to other popular NE methods. But before all that, we first introduce some notation. An undirected network is denoted $\gG = (V, E)$ where $V$ is a set of $n=|V|$ nodes and $E\subseteq \binom{V}{2}$ is the set of links (also known as edges). A link is denoted by an unordered node pair $\{i,j\} \in E$. Let $\mA$ denote the adjacency matrix, with element $a_{ij} = 1$ for $\{i,j\} \in E$ and $a_{ij} = 0$ otherwise. The symbol ${\hat{\mA}}$ will be used to denote the adjacency matrix of a particular observed network. NE methods find a mapping $f: V \to \sR^d$ from nodes to $d$-dimensional real vectors. An embedding is denoted as $\mX = (\vx_1, \vx_2,\ldots,\vx_n)' \in \sR^{n \times d}$, with $\mX^*$ denoting an optimal embedding for adjacency matrix $\mA$ (suppressing the dependency of $\mX$ on $\mA$ for conciseness—see below), and similarly ${{\hat{\mX}}^*}$ optimal for ${\hat{\mA}}$. Network Embedding-based Link Predictions\[sec:method\_ne\_lp\] -------------------------------------------------------------- All well-known NE methods aim to find an embedding $\mX^*$ for given graph $\gG$ (with adjacency matrix $\mA$) that maximizes a continuously differentiable[^2] objective function ${\mathcal{L}}(\mA, \mX)$ for the given adjacency matrix $\mA$. Thus $\mX^*$ must satisfy the following necessary condition for optimality: $$\begin{aligned} \label{eq:opt_condt} \nabla_{\mX} {\mathcal{L}}(\mA, \mX^*) = \mathbf{0}.\end{aligned}$$ Defining $\mF(\mA,\mX)\triangleq\nabla_{\mX}{\mathcal{L}}(\mA, \mX)$, the optimal embedding $\mX^*$ is thus a solution to $\mF\left(\mA, \mX^*\right)=\mathbf{0}$. Based on an embedding $\mX$, it is common to predict the existence of a link between any pair of nodes $i$ and $j$ by computing a link probability (or other score) $g_{ij}(\mX)$, using a differentiable function $g_{ij}:\sR^{nd}\rightarrow\sR$. In practice, $g_{ij}$ often only depends on the embeddings $\vx_i$ and $\vx_j$ of $i$ and $j$, and often it can be written as $g_{ij}(\mX)=g(\vx_i,\vx_j)$ for some function $g:\sR^d\times\sR^d\rightarrow\sR$. It is often found by training a classifier (e.g., logistic regression) on a set of known linked and unlinked node pairs (see Sec. \[sec:explaine\_other\]), but sometimes it follows directly from the NE model (e.g., for CNE). We also introduce the function $g^*_{ij}:\sR^{n\times n}\rightarrow\sR$ defined as $g^*_{ij}(\mA)\triangleq g_{ij}(\mX^*)$ where $\mX^*$ is optimal w.r.t. $\mA$. I.e., $g^*_{ij}$ directly computes the link probability w.r.t. an optimal embedding for a specified adjacency matrix. ExplaiNE as a generic approach\[sec:method\_explaine\] ------------------------------------------------------ ExplaiNE uses a counterfactual reasoning approach to explain link predictions based on a NE. Namely, it quantifies the change of the link probability (or other score) of a node pair $\{i,j\}$ if the presence of a link between a given pair of nodes $\{k,l\}$ were to be altered. Consider first the situation where $\{k,l\}\in E$. Then, if removing the link between them strongly decreases the probability of a link between $i$ and $j$, the link $\{k,l\}$ is a good counterfactual explanation of this predicted link. Conversely, consider the situation where $\{k,l\}\not\in E$. Then, if adding a link between them strongly decreases the probability of a link between $i$ and $j$, it is the absence of a link between $k$ and $l$ that is a good counterfactual explanation of this predicted link. Intuitively, adding or removing an existing link will alter the probability of a link between $i$ and $j$ because it will alter the optimal embedding, which in turn will change the link probability of the target pair. For the ExplaiNE strategy to be effective, we must be able to compute and combine these two effects in an efficient manner. A naive approach would be to recompute the embedding with a link added or removed, and to quantify how much this changes the probability of a link between $i$ and $j$. However, recomputing the embedding is computationally demanding, and is practically impossible to do even for a moderate number of pairs $\{k,l\}$. Moreover, even adding or removing a single link can dramatically change the optimization landscape. As there are potentially many local optima, this can change the optimal embedding entirely (even if initialized with the original embedding), making a change in link probability erratic and hard to interpret. Instead, ExplaiNE investigates the effect of an *infinitesimal* change to $a_{kl}$ around its observed value ${\hat{a}}_{kl}$, on the link probability as computed by $g^*_{ij}$. Specifically, ExplaiNE seeks explanations as node-pairs $\{k,l\}$ ($k\neq l$ and $\{k, l\} \neq \{i, j\}$) for which $\frac{\partial g_{ij}^*}{\partial a_{kl}}({\hat{\mA}})$ is large in absolute value, with a positive sign if ${\hat{a}}_{kl}=1$ (as then decreasing $a_{kl}$ down from ${\hat{a}}_{kl}=1$ by a small amount would maximally decrease $g_{ij}^*$), and with a negative sign if ${\hat{a}}_{kl}=0$ (as then increasing $a_{kl}$ up from ${\hat{a}}_{kl}=0$ by a small amount would maximally decrease $g_{ij}^*$). This can be done analytically. Indeed, applying the chain rule: $$\begin{aligned} \label{eq:chainrule} \frac{\partial g_{ij}^*}{\partial a_{kl}}({\hat{\mA}}) = \nabla_{{\mX}}g_{ij}\left({{\hat{\mX}}^*}\right)^T\cdot\frac{\partial \mX^*}{\partial a_{kl}}({\hat{\mA}}).\end{aligned}$$ For many NE methods the first factor can be computed analytically from the expression for $g_{ij}$, as we will see in the next subsections. The second factor can be computed using the *implicit function theorem* [see, e.g., @chiang1984fundamental]. Rephrased for our specific setting, this theorem states (note that we are overloading the symbol $\mX^*$ here to also signify a function): Let $\mF:\sR^{n\times n}\times\sR^{n\times d}\rightarrow \sR^{n\times d}$ be a continuously differentiable function with arguments denoted $\mA\in\sR^{n\times n}$ and $\mX\in\sR^{n\times d}$. Moreover, let ${\hat{\mA}}$ and ${{\hat{\mX}}^*}$ be such that $\mF({\hat{\mA}},{{\hat{\mX}}^*})=\mathbf{0}$. If the Jacobian matrix $\nabla_{\mX}\mF({\hat{\mA}}, {{\hat{\mX}}^*})$ is invertible, then there exists an open set $S\subset\sR^{n\times n}$ with ${\hat{\mA}}\in S$ such that there exists a continuously differentiable function $\mX^*:S\rightarrow \sR^{n\times d}$ with: $$\begin{aligned} \mX^*({\hat{\mA}})&={{\hat{\mX}}^*},\ \mbox{and}\\ \mF(\mA,\mX^*(\mA))&=\mathbf{0}\ \mbox{for all}\ \mA\in S,\end{aligned}$$ and: $$\begin{aligned} \frac{\partial \mX^*}{\partial a_{kl}}(\mA) &= -(\nabla_{\mX}\mF(\mA, \mX^*(\mA)))^{-1}\cdot\frac{\partial\mF}{\partial a_{kl}}(\mA, \mX^*(\mA)).\end{aligned}$$ It is the latter expression, evaluated at ${\hat{\mA}}$, that we need in in order to evaluate Eq. (\[eq:chainrule\]). Note that the Jacobian $\nabla_{\mX}\mF$ is in fact the Hessian of ${\mathcal{L}}$ with respect to $\mX$. This means that $\nabla_{\mX}\mF({\hat{\mA}},{{\hat{\mX}}^*})$ is negative definite (as ${{\hat{\mX}}^*}$ is optimal for ${\hat{\mA}}$). While for some NE-methods it may not be *strictly* negative definite and thus not invertible as required by the theorem (because, e.g., any translation of ${{\hat{\mX}}^*}$ may be equally optimal according to ${\mathcal{L}}$), this situation can be avoided by adding a regularizer to ${\mathcal{L}}$ on, e.g., the Frobenius norm of ${{\hat{\mX}}^*}$ with very small weight. Without going into detail, we note that as this regularization constant approaches zero, this becomes equivalent with using the pseudo-inverse of the Hessian, instead of its inverse. This is the approach we have taken whenever this situation arose. Denoting this Hessian evaluated at ${\hat{\mA}}$ and ${{\hat{\mX}}^*}$ as $\mH$, can thus write: $$\begin{aligned} \label{eq:ift} \frac{\partial \mX^*}{\partial a_{kl}}({\hat{\mA}}) &=-\mH^{-1}\cdot\frac{\partial\mF}{\partial a_{kl}}({\hat{\mA}}, {{\hat{\mX}}^*}).\end{aligned}$$ Putting Eqs. (\[eq:chainrule\]) and (\[eq:ift\]) together, we now can compute the derivative of $g^*_{ij}$ with respect to $a_{kl}$ as follows: $$\begin{aligned} \label{eq:exact_der} \frac{\partial g_{ij}^*}{\partial a_{kl}}({\hat{\mA}}) &=-\nabla_{{\mX}}g_{ij}\left({{\hat{\mX}}^*}\right)^T\cdot\mH^{-1}\cdot\frac{\partial\mF}{\partial a_{kl}}({\hat{\mA}}, {{\hat{\mX}}^*}).\end{aligned}$$ For efficiency, one can compute the partial derivatives for a given predicted link $\{i,j\}$ and for all pairs $\{k,l\}$ by pre-computing the vector $\nabla_{{\mX}}g_{ij}\left({{\hat{\mX}}^*}\right)^T\cdot\mH^{-1}$ by solving a linear system with $nd$ variables and equations, and right multiplying it with the vectors $\frac{\partial\mF}{\partial a_{kl}}({\hat{\mA}}, {{\hat{\mX}}^*})$ which depend on $k$ and $l$. Unfortunately, the computational cost of solving this linear system is $\gO((nd)^3)$ in practice, limiting scalability both in network size and dimensionality. Thus, while this is a clear improvement over the naive approach, it is still not sufficient for realistic network sizes. The next subsection describes how to make ExplaiNE tractable also for large networks and dimensionalities. Making ExplaiNE scalable {#sec:appr} ------------------------ First, we choose to focus on explanations in terms of linked pairs $\{k,l\}$, rather than in terms of unlinked pairs. Such positive explanations are arguably more insightful than negative ones, and especially in sparse networks. Second, experiments (see supplement Sec. 4.1) show that the best explanation for a predicted link $\{i, j\}$ for a node $i$, tends to be a link $\{k, l\}$ that is incident to node $i$, i.e., for which $l=i$. This is arguably because links adjacent to node $i$ affect the link probability $g_{ij}^*({\hat{\mA}})$ by directly affecting the embedding $\vx^*_i$, whereas links not incident to $i$ are likely to have a secondary effect only. Besides this, we also believe that nodes incident to $i$ are likely to be more meaningful from node $i$’s perspective than other links, in practical applications. Thus, we can restrict ourselves to seeking an explanation for a predicted link from node $i$ to node $j$ in terms of an existing link $\{i,k\}$ for which $\frac{\partial g_{ij}^*}{\partial{\hat{a}}_{ik}}({\hat{\mA}})$ is large and positive. Third, we consider only NE methods where $g_{ij}(\mX^*)$ only depends on $\vx_i^*$ and $\vx_j^*$.[^3] Thus, Eq. (\[eq:chainrule\]) can be written as: $$\begin{aligned} \label{eq:chainrule2} \frac{\partial g_{ij}^*}{\partial a_{ik}}({\hat{\mA}}) &= \nabla_{{\vx_i}}g_{ij}\left({{\hat{\mX}}^*}\right)^T\cdot\frac{\partial \vx^*_i}{\partial a_{ik}}({\hat{\mA}}) \\ &+ \nabla_{{\vx_j}}g_{ij}\left({{\hat{\mX}}^*}\right)^T\cdot\frac{\partial \vx^*_j}{\partial a_{ik}}({\hat{\mA}}).\nonumber\end{aligned}$$ Finally, we make an approximation inspired by the fact that changing $a_{ik}$ will have a direct effect on the optimal embeddings $\vx^*_i$ and $\vx^*_k$, but only indirectly (and thus typically less so) on the embedding of the other nodes—including on $\vx^*_j$. This means that the second term in Eq. (\[eq:chainrule2\]) can be neglected. What remains to be computed is thus $\frac{\partial \vx^*_i}{\partial a_{ik}}({\hat{\mA}})$. To do so, we consider the optimality condition of the embedding w.r.t. $\vx_i^*$ alone, considering all other node embeddings fixed to their optimal embeddings in ${{\hat{\mX}}^*}$ for the observed ${\hat{\mA}}$. Letting ${{\hat{\mX}}^*}_{(i)}$ denote the set of $\hat\vx_l$ with $l\neq i$, this optimality condition can be written as: $$\begin{aligned} \nabla_{\vx_i}{\mathcal{L}}({\hat{\mA}},\vx_i,{{\hat{\mX}}^*}_{(i)})=\mathbf{0}.\end{aligned}$$ For conciseness, let us define $\hat\mF_i(\mA, \vx_i)\triangleq \nabla_{\vx_i}{\mathcal{L}}(\mA,\vx_i,{{\hat{\mX}}^*}_{(i)})$. Optimality of $\hat\vx_i^*$ given the observed network ${\hat{\mA}}$ then requires that $\hat\mF_i({\hat{\mA}}, \hat\vx_i^*)=\mathbf{0}$. We can now use the implicit function theorem on this optimality condition to approximate $\frac{\partial \vx^*_i}{\partial a_{ik}}$ as: $$\begin{aligned} \label{eq:ift2} \frac{\partial \vx^*_i}{\partial a_{ik}}(\hat{a}_{ik}) &= -\mH_i^{-1}\cdot\frac{\partial \hat\mF_i}{\partial a_{ik}}.\end{aligned}$$ Here, $\mH_i = \nabla_{\vx_i} \hat\mF_i({\hat{\mA}}, \hat\vx_i^*)$ is the Jacobian of $\hat\mF_i$ or equivalently the Hessian of ${\mathcal{L}}$ w.r.t. $\vx_i$, evaluated at $({\hat{\mA}},{{\hat{\mX}}^*})$. Putting Eqs. (\[eq:ift2\]) and (\[eq:chainrule2\]) (neglecting the second term as discussed) together, this yields: $$\begin{aligned} \label{eq:approx_der} \frac{\partial g_{ij}^*}{\partial a_{ik}}({\hat{\mA}}) &= -\nabla_{{\vx_i}}g_{ij}\left({{\hat{\mX}}^*}\right)^T\cdot \mH_i^{-1} \cdot\frac{\partial \hat\mF_i}{\partial a_{ik}}({\hat{\mA}},\hat\vx_i^*).\end{aligned}$$ Comparing Eq. (\[eq:exact\_der\]) with Eq. (\[eq:approx\_der\]) reveals the dramatic complexity reduction achieved: Inverting $\mH_i\in\sR^{d\times d}$ has a practical complexity of only $\gO(d^3)$, which is entirely feasible given common dimensionalities used in the literature (often 128). The experiments will validate that the approximations made are entirely justified in practice. ExplaiNE for Conditional Network Embedding\[sec:explaine\_cne\] --------------------------------------------------------------- We now apply the generic ExplaiNE approach to Conditional Network Embedding (CNE), a specific NE method. Detailed derivations are deferred to the supplement Sec. 1. CNE proposes a probability distribution for the network conditional on the embedding, and finds the optimal embedding by maximum likelihood estimation. Specifically, the objective function ${\mathcal{L}}$ in CNE is the log-probability of the network conditioned on the embedding: $$\begin{aligned} {\mathcal{L}}({\hat{\mA}},\mX)=\log(P({\hat{\mA}}|\mX)) &= \sum_{\{i,j\}:{\hat{a}}_{ij}=1} \log{P_{ij}(a_{ij}=1|\mX)} \\ &+ \sum_{\{i,j\}:{\hat{a}}_{ij}=0} \log{P_{ij}(a_{ij}=0|\mX)}.\end{aligned}$$ Here, the link probabilities $P_{ij}$ conditioned on the embedding are defined as follows: $$\begin{aligned} \label{eq:posterior} & P_{ij}(a_{ij}=1|\mX) = 1-P_{ij}(a_{ij}=0|\mX)=\\ & \frac{P_{{\hat{\mA}},ij}{\mathcal{N}}_{+,\sigma_1}(\|\vx_i-\vx_j\|)}{P_{{\hat{\mA}},ij}{\mathcal{N}}_{+,\sigma_1}(\|\vx_i-\vx_j\|) + (1-P_{{\hat{\mA}},ij}){\mathcal{N}}_{+,\sigma_2}(\|\vx_i-\vx_j\|)},\nonumber\end{aligned}$$ where ${\mathcal{N}}_{+,\sigma}$ denotes a half-Normal distribution [@leone1961folded] with spread parameter $\sigma$, $\sigma_2>\sigma_1=1$, and where $P_{{\hat{\mA}},ij}$ is a prior probability for a link to exist between nodes $i$ and $j$ as inferred from the degrees of the nodes (or based on other information about the structure of the network)—see, e.g., @adriaens2017subjectively [@van2016subjective]. CNE, being based on a probabilistic model for the graph conditioned on the embedding, naturally allows for LP using the probabilities $P_{ij}({\hat{a}}_{ij}=1|\mX)$. In other words, $g_{ij}(\mX)=P_{ij}(a_{ij}=1|\mX)$ as shown in Eq. (\[eq:posterior\]). Note that it depends on $\vx_i$ and $\vx_j$ alone, as required for the approximate version of ExplaiNE to be applicable (third assumption). Next we show how to apply approximated ExplaiNE to CNE.[^4] First, we derive the optimality condition: $$\begin{aligned} \hat\mF_i({\hat{\mA}}, \hat\vx_i^*) &= \nabla_{\vx_i^*}\log(P({\hat{\mA}}|{{\hat{\mX}}^*})) \\ &= \gamma\sum_{j\neq i} (\hat\vx_i^*-\hat\vx_j^*)\left(P\left(a_{ij}=1|{{\hat{\mX}}^*}\right)-{\hat{a}}_{ij}\right) \\ &= \mathbf{0}.\end{aligned}$$ Denoting $\gamma = \frac{1}{\sigma_1^2} - \frac{1}{\sigma_2^2}$, and $\hat{P}^*_{ij}\triangleq g_{ij}^*({\hat{\mA}})=P_{ij}(a_{ij} = 1 | {{\hat{\mX}}^*})$ (the probability of a link between $i$ and $j$ given the optimal embedding ${{\hat{\mX}}^*}$ for ${\hat{\mA}}$), we can now derive the three factors in Eq. (\[eq:approx\_der\]):[^5] $$\begin{aligned} \nabla_{\vx_i}g_{ij}\left({{\hat{\mX}}^*}\right) &= -\gamma(\vx_i^*-\vx_j^*)\hat{P}^*_{ij}(1-\hat{P}^*_{ij}).\\ \mH_i &= \nabla_{\vx_i}\hat\mF_i({\hat{\mA}}, \hat\vx_i^*)\\ &= \gamma \mI\sum_{l\neq i} \left(P_{il}^*-{\hat{a}}_{il}\right) \\ & -\gamma^2\sum_{l\neq i} (\vx_i^*-\vx_l^*)(\vx_i^*-\vx_l^*)'\hat{P}^*_{il}(1-\hat{P}^*_{il}).\\ \frac{\partial \hat\mF_i}{\partial a_{ik}}({\hat{\mA}},\hat\vx_i^*) &= \gamma(\vx_k^* - \vx_i^*).\end{aligned}$$ This means: $$\begin{aligned} \frac{\partial g_{ij}^*}{\partial a_{ik}}({\hat{\mA}}) &= (\vx_i^*-\vx_j^*)^T\left(\frac{-\mH_i}{\gamma^2\hat{P}^*_{ij}(1-\hat{P}^*_{ij})}\right)^{-1}(\vx_i^* - \vx_k^*).\end{aligned}$$ Note that the Hessian should be invertible and negative definite, if $\hat\vx_i^*$ is indeed a local maximum. Interestingly, this expression has an intuitive interpretation: without the inverted Hessian, it would be an inner product between the distance of $\vx_i^*$ to the embeddings of both nodes $\vx_j^*$ and $\vx_k^*$, indicating that the best explanation is located as far as possible in the direction of $\vx_j^*$ as seen from $\vx_i^*$. Yet, the Hessian modulates the metric and reduces the explanatory power in directions where there are lots of embedded nodes $l$ for which $\hat{P}^*_{il}(1-\hat{P}^*_{il})$ is large, i.e., for which the model is undecided whether there should be a link. ExplaiNE for other NE methods\[sec:explaine\_other\] ---------------------------------------------------- Here we illustrate the generic applicability of ExplaiNE by outlining the steps of applying it to NE methods based on skip gram with negative sampling (SGNS) (e.g., LINE, PTE, DeepWalk, node2vec). In Sec. 3 of the supplement, we derive a concrete example for LINE [@tang2015line]. In those methods, $g_{i,j}(\mX)=g(\vx_i,\vx_j)$, where $g\triangleq \sigma\circ h$ with $\sigma:\sR^d\rightarrow \sR$ a linear classifier (often logistic regression) applied to edge embeddings, whereby the embedding $h(\vx_i,\vx_j)$ of an edge $\{i,j\}$ is computed by applying an edge embedding operator $h: \sR^d\times \sR^d \to \sR^d$ (e.g., element-wise product) to the embeddings of the nodes at its end-points. @levy2014neural and @qiu2018network found that SGNS-based NE methods all share the same objective: $$\begin{aligned} {\mathcal{L}}= \sum_{i=1}^{|V|}\sum_{j=1}^{|V|}\log\sigma(\vx_i\cdot\vy_j) + b\sum_{i=1}^{|V|}\E_{j'\sim P_N}\left[\log \sigma(-\vx_i\cdot\vy_{j'})\right],\end{aligned}$$ where $\vx_i$ is the target embedding of node $i$, $\vy_i$ is the embedding of node $j$ as context (usually discarded, node2vec does not differentiate target and context), $\sigma(\cdot)$ is a sigmoid function, $P_N$ is known as the noise that generates negative samples, and $b$ is the number of negative samples. Moreover, @qiu2018network showed that ${\mathcal{L}}$ often has a closed form representation (or converges to one in probability). This makes it possible to obtain an analytical expression of the NE optimality condition, and thus of the function $\mF(\mA,\mX)$. Given this, both exact and approximated ExplaiNE can be derived. Experiments\[sec:experiments\] ============================== We investigated the following questions: **Q1** How does the approximation compare to the exact version? **Q2** Does ExplaiNE give sensible explanations? **Q3** Does the proposed method scale? All experiments are based on CNE with parameters $\sigma_1=1$, $\sigma_2=2$. Any weights associated to the links in the networks are ignored. We used the following networks. [[**Game of Thrones’ (GoT) network.**]{}]{}[^6] Consisting of $796$ characters (nodes) and $2823$ links between characters that are mentioned within $15$ words of one another in books 1-5. We used a $2$-dimensional embedding of this network to assess the quality of the approximated ExplaiNE approach. [[**DBLP co-authorship network [@tang2008arnetminer].**]{}]{}[^7] Containing papers published up to year 2017, from which we selected all papers published at ICML, NeurIPS, ICLR, JMLR, MLJ, KDD, ECML-PKDD, and DMKD. This results in 23,359 authors (nodes) and 20,545 papers, converted into 66,597 links between authors who co-authored at least one paper. We conducted both qualitative and quantitative evaluations on a $32$-dimensional embedding of this network. [[**MovieLens dataset [@harper2016movielens].**]{}]{}[^8] Containing 100,000 ratings by 943 users on 1,682 movies. The network is thus bipartite and consists of 943+1,682 nodes and 100,000 edges. The dataset also contains metadata such as title and genre, which we have used as external validation sources. We conducted qualitative and quantitative experiments on a $16$-dimensional embedding of this network. In Sec. \[sec:exp\_approx\] we analyze the quality of the approximation. In Sec. \[sec:exp\_qualitative\], we conduct a qualitative analysis of explanations on the DBLP and MovieLens networks. In Sec. \[sec:exp\_quantitative\] we quantitatively analyze the quality of the explanations. Finally, in Sec. \[sec:exp\_runtime\] we consider the scalability of ExplaiNE. Quality of the ExplaiNE approximation\[sec:exp\_approx\] -------------------------------------------------------- Before applying approximated ExplaiNE to real world dataset, we first evaluate the quality of the approximation (**Q1**). We will assess the extent to which the top $K$ explanations for a predicted link $\{i,j\}$ incident to a given node $i$, as given by approximated ExplaiNE, overlap with the top-$K$ explanations given by exact ExplaiNE. Relevant parameters here are (1) the value of $K$ and (2) the number of neighbors. As we consider only links to neighbors as candidate explanations, $K$ must be smaller than the number of neighbors of $i$. Moreover, if the number of neighbors is not much larger than $K$, a substantial overlap in the top-K explanations of the exact and approximate method is not surprising. Indeed, if $i$ has $m$ neighbors, two random subset of $K$ neighbors would share $l$ elements with probability $\binom{K}{l}\binom{m-K}{K-l}/\binom{m}{K}$, which is large for large $l$ if $m$ is not much larger than $K$. Thus, we performed a stratified analysis, computing the size of the overlap of the top-$K$ explanations, aggregated in a histogram over nodes with a specific degree. We did this on the GoT dataset for $K$ from $1$ to $5$ This experiment revealed that the top-1 is always identical between the approximated and exact versions, while the elements further in the ranked list very rarely swapped positions ($2$ to $3$ differences out of $796$ on ranks $2$,$3$,$4$, and $7$ differences out of $796$ for rank $5$, see supplement Sec. 4.2). In the supplement Sec. 4.3 we also compared the complete ranking of the neighbors between the approximated and exact ExplaiNE versions, and this for the most probable link for every node (i.e., seeking explanations for links that are actually present in the network). We computed the normalized Kendall tau distance[^9] between the ranked explanations given by approximated and exact ExplaiNE. The average normalized Kendall tau distance is $0.05\pm 0.08$. For comparison, the average Kendall tau distance between a random ranking and exact ExplaiNE is $0.51 \pm 0.15$. Now confident in its accuracy, we can now evaluate the behavior of approximated ExplaiNE on two realistic networks. Qualitative evaluation\[sec:exp\_qualitative\] ---------------------------------------------- Here we apply ExplaiNE to explain the predicted links in two real world networks (the DBLP co-authorship and the MovieLens rating networks) to assess whether ExplaiNE gives sensible explanations to the predicted links (**Q2**). [[**DBLP network.**]{}]{} In the co-authorship network, a predicted link between authors $i$ and $j$ suggests a collaboration between them. While ExplaiNE uses no external information to provide its explanations for such suggested collaborations, our experiments indicate that such explanations tend to be existing collaborators working on a topic on which the suggested collaborator is active as well. As an example, we predict links for ICML’19 general chair Eric P. Xing (node $i$), and compute the explanations for his top recommendation (node $j$): Adams Wei Yu. It turns out that the existing co-authors of Eric P. Xing identified by ExplaiNE as top-$5$ explanations for this recommendation (see Table \[tb:dblp\_explanation\]) are either colleagues or co-authors of Adams Wei Yu, with a shared interest in large scale optimization and deep learning. --- -------------------- -- 1 Adams Wei Yu 2 Jure Leskovec 3 Sunita Sarawagi 4 Tong Zhang 5 Soumen Chakrabarti --- -------------------- -- : Predicted/recommended collaborations for Eric P. Xing. The top link (author: Adams Wei Yu) predicted by CNE are explained through co-authors of Eric P. Xing that are also colleagues or co-authors of Adams Wei Yu. The most relevant five co-authors of Eric P. Xing also cover major parts of Adams Wei Yu’s research interests: large scale optimization and deep learning.[]{data-label="tb:dblp_explanation"} [[**MovieLens network.**]{}]{} In the rating network, a predicted link between a user $i$ and movie $j$ amounts to a recommendation of movie $j$ to user $i$. In making this recommendation CNE did not have access to any meta-data of the users or movies, and neither does ExplaiNE to identify explanations. Yet, we can make use of this meta-data to qualitatively assess whether the explanations make sense. As an example, we computed the recommendation for the first user (uid=$0$) in the user list (See Table. \[tb:imdb\_explanation\]). The top recommended movie is ‘Batman’ with genre tags ‘Action’, ‘Adventure’, ‘Crime’, and ‘Drama’. The genres of the top explanations given by ExplainNE arguably have strongly overlapping genre tags (e.g., all top-$5$ are tagged with ‘Action’). Moreover, the second-highest ranked explanation is ‘Batman Forever’. ----- ---------------------------- ------------------------------------ 1 Batman Action, Adventure, Crime, Drama 2 E.T. the Extra-Terrestrial Children’s, Drama, Fantasy, Sci-Fi 3 The Secret of Roan Inish Adventure $k$ 1 Supercop Action, Thriller 2 Batman Forever Action, Adventure, Comedy, Crime 3 The Crow Action, Romance, Thriller 4 Full Metal Jacket Action, Drama, War 5 Young Guns Action, Comedy, Western ----- ---------------------------- ------------------------------------ : Recommended movie to user uid=$0$. The top movie recommended by CNE (Batman) is explained through movies already seen by user uid=$0$. The top-ranked explanations have genres that overlap with the recommended movie. \[tb:imdb\_explanation\] -0.6cm More case studies are given in the supplement Sec. 5. These results suggest that ExplaiNE gives sensible explanations. The next subsection aims to quantify these findings. Quantitative evaluation\[sec:exp\_quantitative\] ------------------------------------------------ Objectively evaluating the quality of an explanation is conceptually non-trivial, due to a lack of datasets with ground-truth explanations for LP. Yet, as we show in this section, it is possible to use metadata to derive reasonable ground truth explanations, and compare with those. [[**DBLP network.**]{}]{} Here, we can construct ground truth explanations for *existing* links (as opposed to *predicted* ones). While this is not the intended use case of ExplaiNE, it is perfectly legitimate and justified here given our intention to objectively validate the quality of the explanations. Our approach is based on the intuition that a one-time co-author $j$ of a given author $i$ could have been introduced to that author $i$ by another co-author $k$ on the same paper, thus explaining the link $\{i,j\}$. While this will of course not always be true, we postulate that it is sufficiently common for ExplaiNE—providing it works well—to highlight the other co-authors as explanations for the observed link $\{i,j\}$. Given an author $i$ and a one-time co-author $j$ of $i$, we used ExplaiNE to rank the other co-authors of $i$, from more to less explanatory (according to Eq. \[eq:approx\_der\]). We then took the top-$r$ of this ranked list as predicted co-authors on the paper $i$ co-authored with $j$. Based on this, we created a confusion matrix. Clearly, the hardness of this prediction task is different for papers with different numbers of authors. Thus, in order to get a more aggregate assessment, we summed the top-$r$ confusion matrices for all one-time co-authors of node $i$ on papers with a given number of co-authors $L$, and this for different $L$ between 3 and 5. For a given author-list length, the confusion matrices with different $r$ were then used to create precision-recall curves or ROC curves. Figure \[fig:exp\_quant\_dblp\] shows the ROC curves for Eric P. Xing as node $i$ and three author-list lengths. For comparison, also ROC curves computed based on a randomly ranked list is shown (as the size of the data is rather small, these are not always close to the diagonal). ![image](figures/exp_quant_dblp){width="\textwidth"} -0.5cm ROC curves for other nodes $i$ as well as Precision-Recall curves can be found in the supplement. All results indicate that the explanations are remarkably effective at this task, indicating that ExplaiNE performs well. [[**MovieLens network.**]{}]{} A good explanation $k$ of a predicted link between a movie-user pair $\{i,j\}$ should arguably have a similar list of genres as $j$. To test this, we computed the top-$5$ explanations for user $i$ and her top recommended movie $j$. Then we averaged the Jaccard similarity between the set of genres for movie $j$ and the set of genres of each of the $5$ explanations. To assess the significance of this average, we computed an empirical $p$-value for it by randomly sampling $50$ sets of $5$ ‘explanations’ drawn from the watched movies of $i$, resulting in $50$ random average Jaccard similarities to compare with the one obtained by ExplaiNE. Thus we obtained an empirical $p$-value for each user $i$, indicating the significance of the overlap between the set of genres of the recommended movie $j$ and the top-$5$ explanations. A histogram of these $p$-values is shown in Fig. \[fig:exp\_quant\_imdb\]. While $p$-values are uniformly distributed under the null hypothesis that the explanations have genres unrelated to those of $j$, here this is not the case—indicating the null hypothesis is false. A Kolmogorov-Smirnoff test indeed shows an extremely high significance ($p$-value numerically $0$). ![$P$-values that indicates the significance of the correlation between the genre recommended and the genres in the explanation. Each $p$-value is computed against $50$ random explanations. Those explanations are drawn from user’s watched movies. The empirical distribution has Kolmogorov-Smirnov test statistic $0.32$ and a $p$-value that is numerically $0.0$ against uniform distribution. This shows the significance of positive correlation between the recommended movies and the explanations made by ExplaiNE. \[fig:exp\_quant\_imdb\]](figures/exp_quant_imdb "fig:"){width="50.00000%"} -0.4cm Scalability and runtime\[sec:exp\_runtime\] ------------------------------------------- To address **Q3**, we measured the runtime of exact and approximated ExplaiNE when computing $\frac{\partial g^*_{ij}}{\partial a_{ik}}({\hat{\mA}})$ for all $k\not\in\{i,j\}$, as per Eqs. (\[eq:exact\_der\]) and (\[eq:approx\_der\]), on average over random pairs of nodes $\{i,j\}$. The runtime was measured on a PC with quad-core 2.7GHz Intel Core i5 and 16GB 1600MHz DDR3 RAM. Table \[tb:exp\_runtime\], shows that approximated ExplaiNE is efficient and applicable to large networks with higher dimensionality, while exact ExplaiNE is not. Network $\#$nodes dim time exact time approx ----------- ----------- ------ ------------------------- --------------------- Karate $34$ $2$ $0.03$ $1.8\mathrm{e}{-4}$ GoT $796$ $2$ $64.1$ $4.1\mathrm{e}{-4}$ GoT $796$ $8$ $1490.2$ $9.8\mathrm{e}{-4}$ MovieLens $2625$ $16$ $\sim1.63\mathrm{e}{6}$ $6.8\mathrm{e}{-3}$ DBLP $23359$ $32$ — $0.02$ : Average runtime (in sec., $10$ trials) of exact and approximated ExplaiNE in computing the explanations for a random pair of nodes $\{i,j\}$. Note that the exact method also has substantial memory cost: $13.1$ Gb for MovieLens and on DBLP we went out of memory. On MovieLens, the time was computed only for one $k$, and multiplied by $n-2$ to get an estimated total time for all $k$. \[tb:exp\_runtime\] -0.8cm Related Work\[sec:relatedwork\] =============================== LP, as an important network analysis task, has recently been extensively studied in the NE literature [@hamilton2017representation; @cui2018survey]. By embedding the nodes in a vector space, the link prediction task can be addressed using traditional machine learning (ML) methods. This has led to new and accurate approaches for LP [@grover2016node2vec; @kang2018conditional], but at the expense of explainability. In parallel, the importance of accountability of AI has sparked growing research interest in interpretable ML. Approaches to interpretable machine learning research can be categorized into model-based and post-hoc approaches [@du2018techniques; @murdoch2019interpretable]. The first category focuses on incorporating interpretability (e.g., sparsity) while constructing the ML model. ExplaiNE belongs to the second category of interpretable ML methods: it is a post-hoc method that focuses on interpreting the local structure of ML models (here, NE models). The most strongly related work (although not for LP) are @ribeiro2016should and @lundberg2017unified, who provide a model-agnostic explanation via local approximation of the target model. More closely related, @simonyan2013deep and @koh2017understanding propose to compute the gradient of the loss function of a (black-box) model with respect to the input to gauge the relevance of the input features. The first of these computes the gradient using back-propagation, while the second approximates the gradient using a Taylor series expansion. ExplaiNE is the first generic approach (and, as far as we know, the first approach at all) for explaining link predictions based on a NE. Moreover, to the best of our knowledge, ExplaiNE is the first method that uses the implicit function theorem for explainability. This proved to be a crucial element for computing the gradient of the link probability w.r.t. the network structure, as it allowed us to rigorously track the optimal embedding given an infinitesimal change in the input network. We believe this theorem can prove valuable also for other tasks, particularly those where an intermediate representation is obtained by optimizing an unsupervised objective function (e.g., an autoencoder), to be fed into a subsequent model that is trained in a supervised manner. Conclusions\[sec:conclusions\] ============================== Link Prediction (LP) in networks is an important task, with applications to social networks, recommenders, and knowledge graphs. State-of-the-art approaches are based on first embedding the nodes in a vector space, followed by a LP step. Unfortunately, while accurate, these approaches offer no insight in their predictions. To remedy this, we introduced ExplaiNE, a generic approach to explain LPs based on Network Embeddings (NEs) in terms of existing links in the network. ExplaiNE is applicable for a wide range of NE methods. We applied it to CNE, a state-of-the-art NE method, and outlined how it can be applied for a wide range of other NE methods. Extensive qualitative and quantitative evaluations show the usefulness of ExplaiNE, and its ability to scale to large networks. In the future we aim to develop ExplaiNE for other NE methods, apply it to recommender systems, and extend it to offer explanations in terms of the presence of dense communities or other larger substructures in the network. ### Acknowledgments {#acknowledgments .unnumbered} The research leading to these results has received funding from the European Research Council under the European Union’s Seventh Framework Programme (FP7/2007-2013) / ERC Grant Agreement no. 615517, from the FWO (project no. G091017N, G0F9816N), and from the European Union’s Horizon 2020 research and innovation programme and the FWO under the Marie Sklodowska-Curie Grant Agreement no. 665501. [^1]: The embedding used is 2-dimensional and derived using Conditional Network Embedding [@kang2018conditional]. [^2]: Note that, although NE methods are often described for unweighted networks (i.e., a binary adjacency matrix), the objective ${\mathcal{L}}(\mA, \mX)$ is often continuously differentiable also w.r.t. the adjacency matrix $\mA$. This is required for ExplaiNE to be applicable, but as we will see this requirement is often satisfied. [^3]: This is true for all NE methods we are aware of, and thus hardly a limitation at all. [^4]: Due to the limited space, here we only show how to apply approximated ExplaiNE to CNE, as the exact version is not used in the experiments except for the experiment validating the approximated version. For the application of exact ExplainNE to CNE, we refer the reader to the supplement Sec. 2. From now on, we drop the modifier ‘approximated’ when the context is clear. [^5]: Detailed derivations are provided in the supplementary material Sec. 1 due to space constraints. [^6]: <https://github.com/mathbeveridge/asoiaf> [^7]: DBLP dataset V10: <https://aminer.org/citation> [^8]: <https://grouplens.org/datasets/movielens/100k/> [^9]: <https://en.wikipedia.org/wiki/Kendall_tau_distance>
{ "pile_set_name": "ArXiv" }
--- abstract: 'We propose a universal language to assess macroscopic quantumness in terms of coherence, with a set of conditions that should be satisfied by any measure of macroscopic coherence. We link the framework to the resource theory of asymmetry. We show that the quantum Fisher information gives a good measure of macroscopic coherence, enabling a rigorous justification of a previously proposed measure of macroscopicity. This picture lets us draw connections between different measures of macroscopicity and evaluate them; we show that another widely studied measure fails one of our criteria.' author: - Benjamin Yadin - Vlatko Vedral bibliography: - 'macro\_coherence.bib' title: A general framework for quantum macroscopicity in terms of coherence --- [*[Introduction.—]{}*]{} One of the most difficult challenges in quantum theory is to explain how it can be compatible with the classical world observed at the macroscopic scale. We are accustomed to the fact that microscopic systems can exist in superposition states, but quantum theory also allows this behaviour in principle at the macroscopic scale. Various attempts have been made to address this; one common approach may be broadly termed ‘decoherence’, in which the quantumness of a large system is highly susceptible to being destroyed by interactions with a noisy environment [@zurek2003decoherence]; similarly, it has been argued that classical behaviour emerges from limitations on the precision of measurements [@kofler2007classical; @sekatski2014difficult; @raeisi2011coarse]. Others have suggested fundamental modifications to quantum theory that are negligible at microscopic scales but cause macroscopic superpositions to quickly appear classical [@bassi2013models]. Ultimately, it is up to experiments to probe the boundary between the quantum and classical worlds, and much progress has been made in this direction in recent decades [@leggett2002testing; @arndt2014testing; @farrow2015classification]. A wide variety of systems have been explored, including interference of large molecules [@nairz2003quantum; @eibenberger2013matter], superpositions of coherent states in photonic systems [@gao2010experimental], superconducting circuits behaving as large-scale qubits [@vanderwal2000quantum], and the control of low-lying vibrational states of micromechanical oscillators [@oconnell2010quantum]. Given this diversity, we need a general way to quantify how well each experiment has achieved its aim of creating large-scale quantum coherence. Beginning with ideas by Leggett [@leggett1980macroscopic; @leggett2002testing], a variety of measures of ‘quantum macroscopicity’ have been proposed, each motivated differently and aiming to capture a different potentially macroscopic quantum property. Some assume that the given state is a superposition of the form $\ket{\psi_0}+\ket{\psi_1}$ and quantify to what extent the two branches are macroscopically distinct [@marquardt2008measuring; @korsbakken2007measurement; @sekatski2014size] or useful for interferometry [@bjork2004size]. Others apply to more general states of systems of many qubits [@frowis2012measures], continuous-variable modes [@lee2011quantification] and mechanical objects [@nimmrichter2013macroscopicity]. These various approaches do not share any precise unifying principles. Inspired by recent work on quantifying coherence [@baumgratz2014quantifying; @aberg2006quantifying], we propose a mathematical framework of very general applicability for defining macroscopic coherence and give criteria to be satisfied by measures. This aims to provide a rigorous meaning to the notion of a macroscopic superposition. We show that this is mathematically equivalent to an instance of the resource theory of asymmetry, which describes the available states and operations under symmetry constraints. Via this approach, we give a solid motivation to a previously proposed measure of macroscopicity based on the quantum Fisher information, and show that a particular measure based on phase space structure fails one of the criteria. [*[Defining the problem.—]{}*]{} Let us suppose that, for a given experiment of the kind mentioned above, there is some particular ‘macroscopic observable’ $A$ of interest. This is always a quantity which makes sense to use for the classical macroscopic scale description of the system: the centre of mass for molecular interference or micromechanical oscillators; a field quadrature for a photonic cat state; the current for a superconducting circuit. We can say that such an experiment has created macroscopic coherence when there is a superposition of states with macroscopically different values of $A$. Thus we must first be able to quantify coherence associated with a superposition of states differing from one another in $A$-value by a fixed amount $\delta$; we will later call this $\delta$-coherence. There has recently been progress in rigorously quantifying coherence (with respect to some fixed basis) using the framework of resource theories [@baumgratz2014quantifying]. In general resource theories, ‘free states’ are states containing no resource and ‘free operations’ are those which are unable to create the resource. When the resource is coherence, the free (i.e., incoherent) states are all diagonal states in the preferred basis; the free (i.e., incoherent) operations are those which take any incoherent state to another incoherent state. A measure of a general resource must vanish for free states and never increase under free operations – thus measures of coherence can be defined. One might attempt to define macroscopic coherence in this way, with reference to the eigenbasis $\ket{i}$ of the observable $A = \sum_i a_i \ketbra{i}{i}$. However, this formalism is not appropriate: the incoherent operations permit arbitrary permutations of the basis states, allowing for instance a ‘microscopic superposition’ $(\ket{i}+\ket{j})/\sqrt{2}$ with a small difference $\delta=a_i-a_j$ to be turned into one with a macroscopic value of $\delta$. In short, the resource theory of coherence is ignorant of the values $a_i$ of the observable. To rectify this, we introduce the notion of $\delta$-coherence to be coherence associated with a superposition of $\ket{i}$ and $\ket{j}$ such that $a_i-a_j = \delta$. Formally, by expanding a state $\rho = \sum_{i,j} \rho_{i,j} \ketbra{i}{j}$, we can identify the part containing all the $\delta$-coherence as $$\label{eqn:mode_decomp} \rho^{(\delta)} := \sum_{i,j:\, a_i-a_j=\delta} \rho_{i,j} \ketbra{i}{j}.$$ The entire state can be expressed as $\rho = \sum_{\delta \in \Delta} \rho^{(\delta)}$, where $\Delta := \{a_i - a_j\}_{i,j}$ is the set of all eigenvalue differences of $A$. We consider the amounts of $\delta$-coherence for each $\delta$ to all be resources of independent interest. Hence we define the free operations to be such that $\delta$-coherence cannot be created from a state with none initially. To do this, we must impose that $\delta$-coherence cannot be turned into $\delta'$-coherence for $\delta \neq \delta'$ – without this constraint, a state with only $\delta$-coherence could be turned into one with nonzero $\delta'$-coherence. Thus a free operation ${\mathcal{E}}$ is any quantum operation (it need not be trace-preserving, as we allow nondeterministic operations) such that $$\label{eqn:free_ops} {\mathcal{E}}(\rho)^{(\delta)} = {\mathcal{E}}(\rho^{(\delta)}) \; \forall \delta \in \Delta.$$ As we have seen, in accordance with resource theories, we demand several properties of a measure ${\mathcal{M}}$. Firstly, it must be nonnegative and vanish for states with no coherence: (M1) : ${\mathcal{M}}(\rho) \geq 0$ and ${\mathcal{M}}(\rho) = 0 \Leftrightarrow \rho = \rho^{(0)}$. Furthermore, ${\mathcal{M}}$ should not increase under a free operation – and if there are various possible outcomes, it should not increase on average: (M2a) : For a deterministic free operation ${\mathcal{E}}$ (such that $\tr {\mathcal{E}}(\rho) = 1$), ${\mathcal{M}}(\rho) \geq {\mathcal{M}}({\mathcal{E}}(\rho))$ (M2b) : For ${\mathcal{E}} = \sum_\alpha {\mathcal{E}}_\alpha$ such that each ${\mathcal{E}}_\alpha$ satisfies (\[eqn:free\_ops\]), ${\mathcal{M}}(\rho) \geq \sum_\alpha p_\alpha {\mathcal{M}}(\sigma_\alpha)$, where the outcome $\sigma_\alpha = {\mathcal{E}}_\alpha(\rho)/p_\alpha$ occurs with probability $p_\alpha = \tr {\mathcal{E}}_\alpha(\rho)$. Measures satisfying these conditions are said to be monotones. Another condition which one often demands is convexity, meaning that a measure does not increase under mixing: (M3) : ${\mathcal{M}}(\sum_i p_i \rho_i) \leq \sum_i p_i {\mathcal{M}}(\rho_i)$. These conditions should be familiar from properties of entanglement measures, where the free states are separable and the free operations are LOCC (local operations and classical communication). They are not all independent – (M2b) and (M3) together imply (M2a). The theory we have built up is actually a special case of the resource theory of quantum reference frames, also known as asymmetry [@bartlett2007reference]. Our free operations are symmetric (or covariant) with respect to the transformations $T_x(\rho) := e^{-ixA} \rho e^{ixA},\, x \in \mathbb{R}$, meaning that ${\mathcal{E}}(T_x(\rho)) = T_x({\mathcal{E}}(\rho))$. Equation (\[eqn:mode\_decomp\]) is called a decomposition into modes of asymmetry in [@marvian2014modes]. As noted there, the trace norm ${\| X \|_{\tr}} := \tr(\sqrt{X^\dagger X})$ provides a set of monotones ${\| \rho^{(\delta)} \|_{\tr}}$ satisfying (M2). This in fact works for any norm $\| \cdot \|$ which is contractive under trace-preserving operations, meaning $\|{\mathcal{E}}(X)\| \leq \|X\|$, and convex. Then $\|\rho^{(\delta)}\|$ is a measure of the $\delta$-coherence in $\rho$. (A weaker form of (M1) is satisfied, as this vanishes when the state has no $\delta$-coherence.) The concept of $\delta$-coherence answers the question ‘how much coherence does a state have at a scale $\delta$?’ To quantify macroscopic coherence, we would like a measure taking all scales into account, giving higher weight to larger $\delta$. For this we suggest an additional condition imposing a strict order on the simplest superpositions: (M4) : Let $\ket{\psi} = \frac{1}{\sqrt{2}} (\ket{i} + \ket{j})$ and $\ket{\phi} = \frac{1}{\sqrt{2}} (\ket{k} + \ket{l})$. If $\abs{a_i - a_j} > \abs{a_k - a_l}$ then ${\mathcal{M}}(\ketbra{\psi}{\psi}) > {\mathcal{M}}(\ketbra{\phi}{\phi})$. [*[Constructing measures.—]{}*]{} We now use some results derived in asymmetry to construct measures satisfying (M1-4). The first result applies to the variance $V(\ket{\psi},A) := \braXket{\psi}{A^2}{\psi} - \braXket{\psi}{A}{\psi}^2$: \[lem:variance\_monotone\] $V(\ket{\psi},A)$ satisfies (M1), (M2a), (M2b) and (M4) for pure states. Lemma 8 in [@gour2008resource] proves (M1), (M2a) and (M2b) for the case of $U(1)$ reference frames. Simple modifications suffice to prove the slightly more general case considered here (see Appendix \[app:variance\]). (M4) is clear from $V((\ket{i}+\ket{j})/\sqrt{2},A) = \frac{1}{4}(a_i - a_j)^2$. An extension to mixed states is achieved by the quantum Fisher information [@braunstein1994statistical]. In its most general form the quantum Fisher information ${\mathcal{F}}(\rho_x)$ is defined for a family of states $\rho_x$ depending on a continuous real parameter $x$ (or possibly multiple parameters), and evaluated at some $x_0$. Then $\left. {\mathcal{F}}(\rho_x) \right|_{x=x_0} := 4 \left. \partial_x^2 D_B^2(\rho_{x_0},\rho_x) \right|_{x=x_0}$, in terms of the Bures distance $D_B(\rho,\sigma) := 2(1 - F(\rho,\sigma))$, where $F(\rho,\sigma) = \tr \sqrt{\sqrt{\rho}\sigma\sqrt{\rho}}$ is the fidelity. Thus ${\mathcal{F}}(\rho_x)$ measures the rate of change of the state with respect to $x$. We need the Fisher information measuring the sensitivity of the state under the transformations $T_x$, ${\mathcal{F}}(\rho, A) := \left. {\mathcal{F}}( e^{-ixA} \rho e^{ixA}) \right|_{x=0}$. This can be expressed as $$\label{eqn:fisher_expansion} {\mathcal{F}}(\rho, A) = 2 \sum_{a,b} \frac{(\lambda_a - \lambda_b)^2}{\lambda_a + \lambda_b} \abs{\braXket{\psi_a}{A}{\psi_b}}^2,$$ where $\rho = \sum_a \lambda_a \ketbra{\psi_a}{\psi_a}$ is a spectral decomposition and the sum is over all $a,\,b$ such that $\lambda_a + \lambda_b >0$. For pure states, ${\mathcal{F}}(\ketbra{\psi}{\psi},A) = 4V(\ket{\psi},A)$. \[thm:fisher\_monotone\] The quantum Fisher information ${\mathcal{F}}(\rho, A)$ satisfies (M1-4) for all states. Any state $\rho$ can be decomposed in many ways as a mixture of pure states: $\rho = \sum_\mu p_\mu \ketbra{\psi_\mu}{\psi_\mu}$, where the $p_\mu$ form a probability distribution (the $\ket{\psi_\mu}$ need not be orthogonal). This lets us do a convex roof construction to extend any real-valued function $f$ of pure states to mixed states: $f_{CR}(\rho) := \inf_{ \{ p_\mu, \ket{\psi_\mu} \} }\, \sum_\mu p_\mu f(\ket{\psi_\mu})$, where the optimisation is over all possible such decompositions. By construction, this is convex and reduces to $f$ for pure states. It was shown in [@toloui2011constructing] in the context of asymmetry that the convex roof of any pure state measure satisfying (M1-2) is a monotone on all states. Given Lemma \[lem:variance\_monotone\], the convex roof of the variance is therefore a monotone. This quantity was called the ‘frameness of formation’ in [@toloui2011constructing], but it has actually been shown more recently that $\frac{1}{4}{\mathcal{F}}(\rho,A)$ is the convex roof of $V(\ket{\psi},A)$ [@toth2013extremal; @yu2013quantum]. Note that other asymmetry monotones satisfying (M1-3) exist, such as the relative entropy measure $\min_{\sigma \in {\mathcal{I}}}S(\rho || \sigma)=\min_{\sigma \in {\mathcal{I}}} \tr(\rho \log \rho - \rho \log \sigma)$, where ${\mathcal{I}}$ is the set of incoherent states. However, this does not depend on the eigenvalues of $A$, so fails (M4). The Wigner-Yanase-Dyson skew information [@wigner1963information] $I_{sk}(\rho,A):=-\frac{1}{2}\tr([\sqrt{\rho},A]^2)$ is also a monotone [@marvian2014extending; @girolami2014observable] and satisfies (M4), but is essentially equivalent to ${\mathcal{F}}$ in that $I_{sk}(\rho,A) \leq {\mathcal{F}}(\rho,A) \leq 2I_{sk}(\rho,A)$ [@petz2011introduction]. The theorem below, following [@gour2008resource; @toloui2011constructing], demonstrates the meaning of ${\mathcal{F}}$ in terms of the equivalence of many copies of a given state with many copies of a chosen reference state. \[thm:asymptotic\] Let $\ket{\phi}$ be a reference state with $V(\ket{\phi},A)=A_0$. Coherence for $n$ copies is defined via the observable $\sum_{i=1}^n A_i$. Then in the limit of large $n$, $\ket{\psi}^{\ox n}$ and $\ket{\phi}^{\ox m}$ have the same $\delta$-coherence for all $\delta$, such that $m/n = V(\ket{\psi},A)/A_0$. For mixed states, the minimal average ratio $m/n$ over all pure state decompositions $\rho = \sum_\mu p_\mu \ketbra{\psi_\mu}{\psi_\mu}$ is ${\mathcal{F}}(\rho,A)/(4A_0)$. See Appendix \[app:asymptotic\] for details. The idea is that the central limit theorem guarantees that the distribution of $\sum_{i=1}^n A_i$ tends to a normal distribution which is characterised fully by its mean and variance. The $\delta$-coherence depends only on this distribution; it is independent of the mean but proportional to the number of copies. The final statement follows from the convex roof property. [*[Evaluating existing measures.—]{}*]{} *i) The measure by Fröwis and Dür*: In [@frowis2012measures], the authors compared various measures of macroscopicity for systems of $N$ qubits [@bjork2004size; @korsbakken2007measurement; @marquardt2008measuring; @shimizu2005detection]. They proposed a measure based on the quantum Fisher information: $$\label{eqn:nf} {N_{\mathcal{F}}}(\rho) := \max_{A \in {\mathcal{A}}}\, \frac{1}{4N} {\mathcal{F}}(\rho, A),$$ where ${\mathcal{A}}$ is the set of all observables that can be written as $A = \sum_{i=1}^N A_i$ over local $A_i$, each acting on a single qubit $i$ and with bounded spectrum ($\| A_i \| = 1$ for some norm) [^1]. Thought of as a total spin, these could reasonably be described as macroscopic observables. This measure was motivated by its ability to capture some supposedly macroscopic quantum properties. For instance, ${N_{\mathcal{F}}}> k-1$ indicates genuine $k$-body entanglement [@toth2012multipartite; @hyllus2012fisher]. Fisher information is also a central quantity in quantum metrology [@paris2009quantum]. ${\mathcal{F}}(\rho,A) = O(N^2)$ enables an unknown parameter $x$ encoded in a state via $T_x$ to be estimated with precision $\Delta x \propto 1/N$ compared with the classical ‘shot-noise’ scaling $\Delta x \propto 1/\sqrt{N}$ for ${\mathcal{F}}(\rho,A) = O(N)$. So ${N_{\mathcal{F}}}$ measures the highest quantum improvement of a given state over the classical bound for estimating transformations $T_x$ generated by any observable in ${\mathcal{A}}$. ${N_{\mathcal{F}}}$ has been called an ‘effective size’; it is supposed to measure the size up to which quantum properties are significant. Our framework helps justify ${N_{\mathcal{F}}}$ more rigorously as a measure of macroscopicity: The quantity ${N_{\mathcal{F}}}$ measures the maximum macroscopic coherence, according to the quantum Fisher information, over all observables in ${\mathcal{A}}$. Note that if we take the reference state in Theorem \[thm:asymptotic\] to be a product of single-qubit states $\bigotimes_{i=1}^N \ket{\phi_i}$ with $V(\ket{\phi_i},A_i)=1$, then the average ratio of copies is exactly ${\mathcal{F}}(\rho,A)/(4N)$. This measure could in principle be applied to any physical system, as long as one can convincingly choose a set ${\mathcal{A}}$ of macroscopic observables. To illustrate this, we shall look at an example. For $N$ continuous-variable bosonic modes with annihilation and creation operators $a_i$ and $a^\dagger_i$, it has been suggested to maximize the Fisher information over sums of quadratures [@frowis2014linking; @oudot2014two]. Then the set ${\mathcal{A}}$ contains $x^{{\boldsymbol{\theta}}} := \sum_{i=1}^N x_i^{\theta_i}$, where ${\boldsymbol{\theta}}:=\{\theta_1,\dots,\theta_N\}$ and $x_i^{\theta_i} := \cos(\theta_i)x_i + \sin(\theta_i)p_i$ is a rotated quadrature of the $i$th mode with associated conjugate quadratures satisfying the canonical commutation relation $[x_i,p_i]=i$. For a single mode and a fixed quadrature $x$, $V(\ket{\psi},x)$ is the (squared) dispersion in $x$ of the phase space distribution of $\ket{\psi}$. The free operations also have a simple interpretation using Lemma 2 of [@gour2008resource]: they can be constructed from phase space displacements, $\rho \to e^{\alpha a^\dagger - \alpha^* a} \rho e^{-\alpha a^\dagger + \alpha^* a}$, and generalized measurements of $x$, $\rho \to M \rho M^\dagger$ with $M = \int dx \, f(x) \ketbra{x}{x}$. A reasonable choice of reference state for Theorem \[thm:asymptotic\] is any coherent state $\ket{\alpha}$ (defined by $a \ket{\alpha} = \alpha \ket{\alpha}$), which has many classical properties. Since $V(\ket{\alpha},x)=1/2$, we find that $n$ copies of $\ket{\psi}$ are equivalent to $n V(\ket{\psi},x)/2$ copies of $\ket{\alpha}$. Maximizing over quadratures, ${N_{\mathcal{F}}}(\ketbra{\psi}{\psi})$ is the largest dispersion over all directions in phase space. For a number eigenstate $\ket{n}$, a superposition of coherent states $\ket{\alpha}+\ket{-\alpha}$ and a quadrature-squeezed coherent state $e^{(\xi^* a^2 - \xi {a^\dagger}^2)/2}\ket{\alpha}$ – all typically described as nonclassical states – it is simple to show that ${N_{\mathcal{F}}}$ scales approximately with the expected number of particles. *ii) The measure by Lee and Jeong* [@lee2011quantification]: This measure for systems of $N$ modes aims to quantify nonclassicality associated with oscillations in the Wigner function in phase space [@gerry2005introductory]. We define a displacement operator $D({\boldsymbol{\alpha}}) := \prod_i e^{\alpha_i a_i^\dagger - \alpha_i^* a_i}$, where ${\boldsymbol{\alpha}} = (\alpha_1,\alpha_2,\dots,\alpha_N)$. We write their measure as $$\label{eqn:I_LJ} {N_{LJ}}(\rho) := \frac{1}{2\pi^N} \int d^{2N} {\boldsymbol{\alpha}} \; \abs{{\boldsymbol{\alpha}}}^2 \abs{\chi_\rho({\boldsymbol{\alpha}})}^2,$$ where $\abs{{\boldsymbol{\alpha}}}^2 := \sum_{i=1}^N \abs{\alpha_i}^2$ and $\chi_\rho({\boldsymbol{\alpha}}) := \tr[\rho D({\boldsymbol{\alpha}})]$ is the characteristic function [^2]. \[thm:I\_LJ\] The measure ${N_{LJ}}$ has the following properties: 1. It measures the rate of decrease of purity under a certain decoherence model: ${N_{LJ}}(\rho) = -\frac{1}{2} \partial_t \tr(\rho^2)$ with $\partial_t \rho = {\mathcal{L}}(\rho) := -\frac{1}{4} \sum_{i=1}^N [x_i,[x_i,\rho]] + [p_i,[p_i,\rho]]$. 2. Define the quantity $I_L(\rho,A) := -\frac{1}{2} \tr([\rho,A]^2)$, which was proposed as an observable lower bound to coherence in [@girolami2014observable]. Then ${N_{LJ}}(\rho) = \frac{1}{2} \sum_{i=i}^N I_L(\rho, x_i) + I_L(\rho, p_i)$. 3. $\frac{1}{N} {N_{LJ}}(\rho) \leq {N_{\mathcal{F}}}(\rho) := \max_{{\boldsymbol{\theta}}} \, \frac{1}{4N} {\mathcal{F}}(\rho, x^{{\boldsymbol{\theta}}})$. Part (a) follows from a direct calculation that shows $\partial_t \chi({\boldsymbol{\alpha}}) = -\frac{1}{2} \abs{{\boldsymbol{\alpha}}}^2 \chi({\boldsymbol{\alpha}})$ (see Appendix \[app:decoherence\]), while (b) is a simple matter of evaluating ${N_{LJ}}(\rho) = -\tr[\rho {\mathcal{L}}(\rho)]$ and using $\tr(\rho [A,[A,\rho]]) = \tr([\rho, A]^2)$. For (c), it suffices to write $I_L(\rho,A) = \frac{1}{2} \sum_{a,b} (\lambda_a - \lambda_b)^2 \abs{\braXket{\psi_a}{A}{\psi_b}}^2$ and compare with (\[eqn:fisher\_expansion\]) to see that $I_L(\rho,A)\leq \frac{1}{4}{\mathcal{F}}(\rho,A)$. It is easy to check that ${\mathcal{L}}$ is isotropic in the sense that $x_i,\,p_i$ can be replaced with $x_i^{\theta_i},\,x_i^{\theta_i+\pi/2}$ for any $\theta_i$. Then ${N_{LJ}}(\rho) \leq \max_{{\boldsymbol{\theta}}}\, I_L(\rho,x^{{\boldsymbol{\theta}}}) \leq \max_{{\boldsymbol{\theta}}}\, \frac{1}{4}{\mathcal{F}}(\rho, x^{{\boldsymbol{\theta}}})$. The original paper related the measure to a decoherence model slightly different from that in (a). We note in passing that this provides a connection with a measure based on the rate of decoherence under an intrinsic collapse model in mechanical phase space [@nimmrichter2013macroscopicity]. As shown in Appendix \[app:nh\_model\], in a natural limit their decoherence model for a single massive object is $\partial_t \rho \approx - c_x [x,[x,\rho]] - c_p [p,[p,\rho]]$, where $c_x,\, c_p$ are coefficients. However, the measure in [@nimmrichter2013macroscopicity] is not a straightforward function of the state, so we cannot make precise statements about (M1-4). As in [@frowis2014linking], we see from (b) that ${N_{LJ}}$ is additive with respect to modes and hence cannot capture correlations – this explains the $1/N$ normalisation in (c). To remedy this problem, while putting the measure on an equal footing with ${N_{\mathcal{F}}}$ by using a maximization over observables, we define a refined version by $$\label{eqn:IPrime_LJ} {\tilde{N}_{LJ}}(\rho) := \max_{{\boldsymbol{\theta}}}\, \frac{1}{N} I_L(\rho, x^{{\boldsymbol{\theta}}}),$$ satisfying ${\tilde{N}_{LJ}}(\rho) \leq {N_{\mathcal{F}}}(\rho)$. Moreover, ${\tilde{N}_{LJ}}(\ketbra{\psi}{\psi}) = {N_{\mathcal{F}}}(\ketbra{\psi}{\psi})$ for any pure state $\ket{\psi}$ because $I_L$, like ${\mathcal{F}}/4$, then equals $V$. To determine whether ${\tilde{N}_{LJ}}$ is a good measure of macroscopicity in the same way as ${N_{\mathcal{F}}}$, we must ask if $I_L$ satisfies (M1-4). The following answers this in the negative: $I_L$ can increase under the free operations defined by (\[eqn:free\_ops\]). Suppose we have the system and an ancilla in a state $\rho \ox \sigma$ with $\tr(\sigma^2)<1$. Then $I_L(\rho \ox \sigma, A \ox I) = I_L(\rho,A) \tr(\sigma^2)$. Thus the measure increases by replacing the ancilla with some pure state, which is a free operation with respect to $A \ox I$. This problem stems from the relation of $I_L$ to the Hilbert-Schmidt distance $D_{HS}(\rho, \sigma) := \tr[(\rho-\sigma)^2]$. We have $I_L(\rho,A) = \frac{1}{2} \left. \partial_x^2 D_{HS}(\rho, e^{-ixA} \rho e^{ixA}) \right|_{x=0}$, as for ${\mathcal{F}}$ and $D_B$. The distance $D_{HS}$ is problematic when used to define various quantum information-theoretic quantities, notably measures of entanglement and discord, because it is not contractive under trace-preserving operations [@ozawa2000entanglement; @piani2012problem]. This cannot be fixed by simply dividing by the purity, as it leads to a poor characterisation of mixed state superpositions [@jeong2011reply]. We finally address whether $I_L$ is a faithful lower bound to ${\mathcal{F}}$. This is true in the sense that $I_L(\rho,A)=0 \Leftrightarrow {\mathcal{F}}(\rho,A)=0$. However, for macroscopicity we are concerned with the scaling of the measure. We provide an $N$-qubit example where the two measures exhibit entirely different scaling. It is possible to have $I_L = O(N)$ but ${\mathcal{F}} = O(N^2)$. Let $N$ be even and define $\rho_N := \sum_{k=0}^{N/2-1} \frac{2}{N} \ketbra{\psi^N_k}{\psi^N_k}$, where $\ket{\psi^N_k} := \frac{1}{\sqrt{2}} \left( \ket{0^{N-k}1^k} + \ket{1^{N-k}0^k} \right)$. Then $I_L(\rho_N, \sum_i \sigma^z_i) = O(N)$, ${\mathcal{F}}(\rho_N, \sum_i \sigma^z_i) = O(N^2)$. The calculations proceed from the expansion (\[eqn:fisher\_expansion\]) and the equivalent for $I_L$. See Appendix \[app\_scaling\] for details. [*[Conclusions.—]{}*]{} In this Letter, we have introduced a formal framework for quantifying the performance of experiments that aim to create macroscopic quantum coherence. This may be seen as a more restricted version of the recent work on coherence monotones [@baumgratz2014quantifying]. We have demonstrated an equivalence with a type of asymmetry resource theory, enabling us to use existing results about monotones from that field. We have proved the quantum Fisher information to be a monotone and demonstrated its meaning in terms of the coherence of many copies of a state. This justifies an existing measure of quantum macroscopicity [@frowis2012measures] that uses the Fisher information, but was not previously motivated rigorously. We have compared this with another measure based on phase space structure [@lee2011quantification], which can be seen as an attempt to quantify quadrature coherence. However, we have shown it is related to a measure which is not a monotone, and can significantly underestimate the Fisher information. We expect our result to be applicable to a wide range of experiments that test the macroscopic limits of quantum mechanics, since it has a very general justification and the proposed measure is relatively simple to calculate. It should be possible to assess and compare a variety of existing systems, as in [@nimmrichter2013macroscopicity], by taking comparable mechanical observables such as total momenta or currents. There have been recent insights about ways to feasibly measure the quantum Fisher information [@girolami2014observable; @frowis2015detecting; @hauke2015measuring], demonstrating that macroscopic coherence could be witnessed in the laboratory.\ The authors acknowledge funding from the National Research Foundation (Singapore), the Ministry of Education (Singapore), the EPSRC (UK), the Templeton Foundation, the Leverhulme Trust, the Oxford Martin School and Wolfson College, University of Oxford. The authors also thank Davide Girolami, Oscar Dahlsten, Raam Uzdin and Chiara Marletto for helpful discussions and comments. Characterisation of the free operations {#app:free_ops} ======================================= Let ${\mathcal{E}}$ be a completely positive trace-non-increasing operation with associated Kraus operators $K_\alpha$, such that ${\mathcal{E}}(\rho) = \sum_\alpha K_\alpha \rho K_\alpha^\dagger$ and $\sum_\alpha K_\alpha^\dagger K_\alpha \leq I$. Applying Lemma 1 of [@gour2008resource] to the case where $T_x$ forms a representation of either $U(1)$ or the real line under addition $\mathbb{R}_+$, we can always find a set of Kraus operators for ${\mathcal{E}}$ satisfying $$\label{eqn:kraus_condition} T_x(K_\alpha) = e^{-ixA} K_\alpha e^{ixA} = e^{-ix\delta_\alpha} K_\alpha \quad \forall x \in \mathbb{R},$$ for some set of $\delta_\alpha \in \mathbb{R}$. (The functions $e^{-ix\delta_\alpha}$ are irreducible representations of the relevant group.) Writing an arbitrary operator in the eigenbasis of $A$, we have $$\begin{aligned} K_\alpha & = \sum_{\delta \in \Delta} \sum_{(i,j)\sim\delta} c_{\alpha,i,j} \ketbra{i}{j}, \\ T_x(K_\alpha) & = \sum_{\delta \in \Delta} e^{-ix\delta} \sum_{(i,j)\sim\delta} c_{\alpha,i,j} \ketbra{i}{j},\end{aligned}$$ where the notation $(i,j)\sim\delta$ means that $a_i-a_j=\delta$. Thus condition (\[eqn:kraus\_condition\]) can only be satisfied when $\delta_\alpha \in \Delta$ and then is equivalent to $K_\alpha = K_\alpha^{(\delta_\alpha)}$. Monotonicity of variance {#app:variance} ======================== We will not repeat the proof of Lemma 8 in [@gour2008resource], but instead note that the only point where a proof of our case differs slightly is in the evaluation of the commutator $[A,K_\alpha]$. The original assumes that $A = \hat{N}$ has a harmonic spectrum (its eigenvalues are all the nonnegative integers), and uses the fact that $[\hat{N},K_\alpha] = k_\alpha K_\alpha$ for some $k_\alpha \in \mathbb{Z}$. In our case, as shown above, the Kraus operators can always be written in the form $K_\alpha = \sum_{(i,j)\sim\delta_\alpha} K_{\alpha,i,j} \ketbra{i}{j}$ for some $\delta_\alpha \in \Delta$. Then $$\begin{aligned} [A, K_\alpha] & = \sum_{(i,j)\sim\delta_\alpha} c_{\alpha,i,j} [A, \ketbra{i}{j}] \nonumber \\ & = \sum_{(i,j)\sim\delta_\alpha} c_{\alpha,i,j} (a_i-a_j) \ketbra{i}{j} \nonumber \\ & = \delta_\alpha K_\alpha,\end{aligned}$$ and the remainder of the proof continues unaffected by the replacement $\hat{N} \to A$, $k_\alpha \to \delta_\alpha$. Asymptotic meaning of Fisher information {#app:asymptotic} ======================================== Our proof follows the ideas in [@schuch2004quantum; @gour2008resource]. We take the observable of interest, $A$, to have a continuous spectrum (although the idea of the proof is the same in the discrete case). For any pure state we write $\ket{\psi} = \int \! dx \, \psi(x) \ket{\psi_x}$, such that each $\ket{\psi_x}$ is normalized by $\braket{\psi_x}{\psi_y} = \delta(x-y)$ and $A\ket{\psi_x} = x\ket{\psi_x}$. Then, for $n$ copies, $$\begin{aligned} \ket{\psi}^{\ox n} & = \int \! dX \, \sqrt{p_n(X)} \ket{\Psi_X} , \\ \sqrt{p_n(X)} \ket{\Psi_X} & := \int_{\sum_{x=1}^n x_i =X} \! d\mathbf{x} \, \bigotimes_{i=1}^n \psi(x_i) \ket{\psi_{x_i}} , \\ p_n(X) & = \int_{\sum_{x=1}^n x_i =X} \! d\mathbf{x} \, \prod_{i=1}^n \abs{\psi(x_i)}^2 ,\end{aligned}$$ where $\mathbf{x} = (x_1,\dots,x_n)$. Now $p_n(X)$ is just the probability density that the values of the observables $A_i$ all sum to $X$, given that they individually have the distribution $\abs{\psi(x_i)}^2$. Under a set of reasonable assumptions, the central limit theorem says that as $n \to \infty$, $p_n(X)$ tends to a normal distribution with mean $n \braXket{\psi}{A}{\psi}$ and variance $\sigma^2 = n V(\ket{\psi},A)$. For the $\delta$-coherence, we have $$\left( \ketbra{\psi}{\psi}^{\ox n} \right)^{(\delta)} = \int \! dX \, \sqrt{p_n(X+\delta) p_n(X)} \ketbra{\Psi_{X+\delta}}{\Psi_{X}}.$$ If we now do the same for $m$ copies of the reference state $\ket{\phi}$, replacing the distribution $p_n(X)$ with $q_m(X)$, it is clear that for any norm, $\| ( \ketbra{\psi}{\psi}^{\ox n} )^{(\delta)} \| = \| ( \ketbra{\phi}{\phi}^{\ox m} )^{(\delta)} \|$ as long as $p_n(X) = q_m(X-X_0)$ with some constant shift $X_0$. This is guaranteed by the central limit theorem when $m/n = V(\ket{\psi},A) / V(\ket{\phi},A)$. As in [@schuch2004quantum; @gour2008resource], we must note that if $\psi(x)$ or $\phi(x)$ vanish anywhere, then the application of the central limit theorem does not necessarily work. However, negligible amounts of extra resources suffice to remove these gaps. Decoherence model for $I_{LJ}$ {#app:decoherence} ============================== We first assemble some useful facts for a single mode [@gerry2005introductory]: $$\begin{aligned} \braket{\beta}{\alpha} & = e^{-\frac{1}{2}\abs{\beta-\alpha}^2 + \frac{1}{2}(\alpha \beta^* - \alpha^* \beta)}, \\ D(\alpha)D(\beta) & = e^{\frac{1}{2}(\alpha \beta^* - \alpha^* \beta)} D(\alpha + \beta), \\ \frac{1}{\pi} \int \! d^2\alpha \, \ketbra{\alpha}{\alpha} & = I.\end{aligned}$$ Using these, we have $$\begin{aligned} \tr D(\alpha) & = \frac{1}{\pi} \int \! d^2\beta \, \braXket{\beta}{D(\alpha)}{\beta} \nonumber \\ & = \frac{1}{\pi} e^{-\abs{\alpha}^2} \int \! d^2\beta \, e^{\alpha \beta^* - \alpha^* \beta} \nonumber \\ & = \pi\, \delta^2(\alpha),\end{aligned}$$ from which it follows that we can write (in the $N$-mode case) $$\begin{aligned} \rho = \frac{1}{\pi^N} \int \! d^{2N}{\boldsymbol{\alpha}} \; \chi({\boldsymbol{\alpha}}) D(-{\boldsymbol{\alpha}}).\end{aligned}$$ Under the model in Theorem 4a, using $[x_i,D({\boldsymbol{\alpha}})] = \sqrt{2} \Re(\alpha_i) D({\boldsymbol{\alpha}})$ and $[p_i,D({\boldsymbol{\alpha}})] = \sqrt{2} \Im(\alpha_i) D({\boldsymbol{\alpha}})$, we have $$\begin{aligned} \partial_t \rho & = -\frac{1}{4} \sum_{i=1}^N \frac{1}{\pi^N} \int \! d^{2N}{\boldsymbol{\alpha}} \; (2 \Re(\alpha_i)^2 + 2\Im(\alpha_i)^2) \chi({\boldsymbol{\alpha}}) D(-{\boldsymbol{\alpha}}) \nonumber \\ & = -\frac{1}{2\pi^N} \int \! d^{2N}{\boldsymbol{\alpha}} \; \abs{{\boldsymbol{\alpha}}}^2 \chi({\boldsymbol{\alpha}}) D(-{\boldsymbol{\alpha}}),\end{aligned}$$ and so $\partial_t \chi({\boldsymbol{\alpha}}) = -\frac{1}{2}\abs{{\boldsymbol{\alpha}}}^2 \chi({\boldsymbol{\alpha}})$. Finally, we calculate the rate of loss of purity: $$\begin{aligned} -\frac{1}{2} \partial_t \tr(\rho^2) & = - \tr(\rho\, \partial_t \rho) \nonumber \\ & = -\frac{1}{\pi^{2N}} \int \! d^{2N}{\boldsymbol{\alpha}}\, d^{2N}{\boldsymbol{\beta}}\; \chi({\boldsymbol{\alpha}}) \partial_t \chi({\boldsymbol{\beta}}) \, \tr[D(-{\boldsymbol{\alpha}})D(-{\boldsymbol{\beta}})] \nonumber \\ & = -\frac{1}{\pi^{2N}} \int \! d^{2N}{\boldsymbol{\alpha}}\, d^{2N}{\boldsymbol{\beta}}\; \chi({\boldsymbol{\alpha}})^* \partial_t \chi({\boldsymbol{\beta}}) \, \tr[D({\boldsymbol{\alpha}})D(-{\boldsymbol{\beta}})] \nonumber \\ & = -\frac{1}{\pi^{2N}} \int \! d^{2N}{\boldsymbol{\alpha}}\, d^{2N}{\boldsymbol{\beta}}\; \chi({\boldsymbol{\alpha}})^* \partial_t \chi({\boldsymbol{\beta}}) \, e^{\frac{1}{2}(-{\boldsymbol{\alpha}}.{\boldsymbol{\beta}}^* + {\boldsymbol{\alpha}}^*.{\boldsymbol{\beta}})} \tr[D({\boldsymbol{\alpha}}-{\boldsymbol{\beta}})] \nonumber \\ & = -\frac{1}{\pi^{2N}} \int \! d^{2N}{\boldsymbol{\alpha}}\, d^{2N}{\boldsymbol{\beta}}\; \chi({\boldsymbol{\alpha}})^* \partial_t \chi({\boldsymbol{\beta}}) \,e^{\frac{1}{2}(-{\boldsymbol{\alpha}}.{\boldsymbol{\beta}}^* + {\boldsymbol{\alpha}}^*.{\boldsymbol{\beta}})} \pi^N \, \delta^{2N}({\boldsymbol{\alpha}}-{\boldsymbol{\beta}}) \nonumber \\ & = -\frac{1}{\pi^N} \int \! d^{2N}{\boldsymbol{\alpha}} \; \chi({\boldsymbol{\alpha}})^* \partial_t \chi({\boldsymbol{\alpha}}) \nonumber \\ & = \frac{1}{2\pi^N} \int \! d^{2N}{\boldsymbol{\alpha}} \; \abs{{\boldsymbol{\alpha}}}^2 \abs{\chi({\boldsymbol{\alpha}})}^2 \nonumber \\ & = I_{LJ}(\rho)\end{aligned}$$ as required. Decoherence model of Nimmrichter and Hornberger {#app:nh_model} =============================================== The general form of their modification is rather complex; for simplicity, while highlighting the important features, we take the case where the system may be well approximated by a point-like object of mass $M$ moving in one dimension. Then we can write the modification in a single-particle form $$\label{eqn:NH_model} {\mathcal{L}}_1(\rho) = \frac{1}{\tau} \left( \int\! ds\,dq \; g(s,q) W(s,q) \rho W(s,q)^\dagger - \rho \right),$$ where $W(s,q) := e^{i(s\hat{p}-q\hat{x})/\hbar}$ is a displacement of the position and momentum of the object. $\tau$ and $g(s,q)$ both depend on $M$, as well as some chosen reference constants that specify the scale of the problem. The exact form of $g(s,q)$ is not supposed to be essential – the authors take it to be a normalized Gaussian. Its standard deviations in $s$ and $q$ are $\frac{m_e}{M} \sigma_s$ and $\sigma_q$, where $m_e$, $\sigma_s$ and $\sigma_q$ are some of the reference constants. $m_e$ is chosen to be the electron mass, while the restrictions $\sigma_s \lesssim 20\; \mathrm{pm}$ and $\hbar/\sigma_q \gtrsim 10\; \mathrm{fm}$ are imposed in order to stay within the nonrelativistic regime. It is simple to show from (\[eqn:NH\_model\]) that when the all the dynamics come from ${\mathcal{L}}_1$, then $\partial_t \chi_\rho(r,p) = -\frac{1}{\tau}[1 - \tilde{g}(r,p)] \chi_\rho(r,p)$, where $\tilde{g}(r,p) := \int\! ds\,dq\, g(s,q) e^{-i(ps-rq)/\hbar}$ [@nimmrichter2014macroscopic]. Note that the bounded nature of $\tilde{g}(r,p)$ means that the decoherence rate always saturates at some maximum [@nimmrichter2013macroscopicity]. Whenever we are far from saturation, we have $\partial_t\chi_\rho \approx -\frac{1}{\tau}\left[ (\frac{m_e \sigma_s r}{M \hbar})^2 + (\frac{\sigma_q p}{\hbar})^2 \right] \chi_\rho(r,p)$. Given that $\frac{1}{\tau} \propto M^2$, it follows that the decoherence model may be well approximated by $$\partial_t \rho \approx - \text{constant} \times \left[ (m_e \sigma_s)^2 [\hat{x},[\hat{x},\rho]] + (M \sigma_q)^2 [\hat{p},[\hat{p},\rho]] \right].$$ Scaling of $I_L$ versus ${\mathcal{F}}$ {#app_scaling} ======================================= Here we calculate ${\mathcal{F}}(\rho,Z)$, where $Z := \sum_{i=1}^N \sigma^z_i$, for any $N$-qubit state $\rho$ of the form $$\begin{aligned} \rho & = \sum_{{\mathbf{x}},s} p^s_{\mathbf{x}} \ketbra{\psi^s_{\mathbf{x}}}{\psi^s_{\mathbf{x}}}, \\ \ket{\psi^\pm_{\mathbf{x}}} & = \frac{1}{\sqrt{2}} \left( \ket{0} \ket{{\mathbf{x}}} \pm \ket{1} \ket{{\mathbf{\bar{x}}}} \right),\end{aligned}$$ where ${\mathbf{x}} = (x_2,x_3,\dots,x_N) \in \{0,1\}^{N-1}$, $\bar{x}:=1-x$ and $s \in \{+,-\}$. Note that the form given above is a spectral decomposition of $\rho$. One can also easily verify the matrix elements $$\begin{aligned} \braXket{\psi^+_{\mathbf{x}}}{Z}{\psi^+_{\mathbf{y}}} & = \braXket{\psi^-_{\mathbf{x}}}{Z}{\psi^-_{\mathbf{y}}} = 0, \nonumber \\ \braXket{\psi^+_{\mathbf{x}}}{Z}{\psi^-_{\mathbf{y}}} & = (N - 2n({\mathbf{x}}))\, \delta_{{\mathbf{x}},{\mathbf{y}}},\end{aligned}$$ where $n({\mathbf{x}}) := \sum_{i=2}^N x_i$. Putting this into equation (4) in the main text, we obtain $${\mathcal{F}}(\rho,Z) = 4\sum_{\mathbf{x}} \frac{(p^+_{\mathbf{x}} - p^-_{\mathbf{x}})^2}{p^+_{\mathbf{x}}+p^-_{\mathbf{x}}} (N-2n({\mathbf{x}}))^2.$$ In particular, when all $p^-_{\mathbf{x}} = 0$, we have $${\mathcal{F}}(\rho,Z) = 4 \sum_{\mathbf{x}} p^+_{\mathbf{x}} (N-2n({\mathbf{x}}))^2.$$ The example state $\rho_N$ in Observation 3 is a special case of the above, so that $$\begin{aligned} {\mathcal{F}}(\rho_N,Z) & = 4 \sum_{k=0}^{N/2-1} \frac{2}{N} (N-2k)^2 \nonumber \\ & = \frac{4}{3}(N+1)(N+2) = O(N^2).\end{aligned}$$ On the other hand, one finds that $$I_L(\rho_N,Z) = \sum_{k=0}^{N/2} \left( \frac{2}{N} \right)^2 (N-2k)^2 = O(N).$$ [^1]: ${\mathcal{A}}$ may be extended to include ‘$k$-local’ $A_i$ acting on groups of $k$ qubits, with $k$ bounded independent of $N$, which case the denominator of (\[eqn:mode\_decomp\]) contains the number of groups in place of $N$. This modification lets the measure capture correlations between groups of sites. [^2]: Note that their original definition instead used $|{\boldsymbol{\alpha}}|^2 - N$, but removing the constant shift ensures that the measure is nonnegative [@jeong2011reply]
{ "pile_set_name": "ArXiv" }
--- abstract: 'A collectivised fund is a proposed form of pension investment, in which all investors agree that any funds associated with deceased members should be split among survivors. For this to be a viable financial product, it is necessary to know how to manage the fund even when it is heterogeneous: that is when different investors have different preferences, wealth and mortality. There is no obvious way to define a single objective for a heterogeneous fund, so this is not an optimal control problem. In lieu of an objective function, we take an axiomatic approach. Subject to our axioms on the management of the fund, we find an upper bound on the utility that can be achieved for each investor, assuming a complete markets and the absence of systematic longevity risk. We give a strategy for the management of such heterogeneous funds which achieves this bound asymptotically as the number of investors tends to infinity.' author: - 'John Armstrong, Cristin Buescu' bibliography: - 'collectivization.bib' title: Asymptotically Optimal Management of Heterogeneous Collectivised Investment Funds --- Introduction {#introduction .unnumbered} ============ We study the problem of maximizing the benefit one can obtain from one’s pension if one is willing to pursue a collective strategy. This is a strategy in which a group of individuals agree that all assets left by an individual who dies are shared among the survivors. By collectivising investments one counters idiosyncratic longevity risk. Unlike an annuity which guarantees a fixed real-terms income until retirement, a collectivised fund may pursue a risky strategy to take advantage of the equity risk-premium and so yield a higher utility for investors. Additionally, a fund may also exploit intertemporal substitution of consumption, by delaying consumption until investments have grown. These considerations imply that a collectivised fund should out-perform both an annuity and a individually managed pension funder. For a more detailed discussion of the benefits of collectivised pension investment, see [@ab-main]. We will model a collective fund of $n$ individuals who have already retired and invested their capital in the fund. We will then consider the optimal pattern of investment and consumption. For simplicity we assume that the only stochastic risk factors are given by the market and idiosyncratic mortality risk. Our modelling for the market, mortality and individual preferences is described in detail in \[sec:marketAndMortality\]. We will also give a number of remarks on how systematic longevity risk could be included in the analysis. In Section \[sec:homogeneousModel\] we give a mathematical description of the optimal investment problem for a homogeneous fund: that is to say, a collective fund where all the individuals are isomorphic. We also show that an overall objective function for the fund can be derived using the ideas of robust optimization or the notion of “The Veil of Ignorance” introduced in [@rawls]. Both approaches yield the same objective function. The management of the fund may then be written as an optimal control problem. We state the optimal control problems for both the cases of continuous and discrete time consumption. We do not solve these problem in this paper. The discrete time problem is solved analytically in the case of homogeneous Epstein–Zin preferences in the Black–Scholes model in [@ab-ez] building on the techniques of [@merton1969lifetime; @campbellViceira]. A numerical algorithm for exponential Kihlstrom–Mirman preferences is given in [@ab-exponential]. In Section \[sec:basicProperties\] we prove some basic properties of collectivised pension investment. Specifically we prove that, under very mild assumptions, collectivisation is always beneficial. We also give a sufficient condition for constant consumption to be the optimal strategy. These conditions are extremely stringent and indicate that an annuity will typically be a suboptimal investment. The mathematical novelty in the paper appears in Section \[sec:heterogeneous\] where we consider how to manage heterogeneous funds. These are funds where preferences, capital and mortality distributions vary between investors. This problem is challenging because it is not obvious how to combine each individual’s preferences to obtain a single objective for the fund as a whole. Rather than attempt this, we take an axiomatic approach describing the properties that any management strategy for the fund should possess. From these axioms we are able to deduce an upper bound on the utility of each investor in a heterogeneous fund, on the assumption that the market is complete. Given the times-scale of pension investments this assumption is a very reasonable approximation to reality. We are then able to describe a strategy that asymptotically achieves this bound as the number of investors tends to infinity. Market, preference and mortality models {#sec:marketAndMortality} ======================================= Market model ------------ We will assume that a fund may invest in a market determined by a filtered probability space $(\Omega^M, {\cal F}, {\cal F}^M_{t\in \R^+}, \P^M)$. In particular we will only consider continuous time markets. We assume there are $k$ available assets and that the price of asset $j$ at time $t$ in real terms is given by $S^j_t$. Arbitrary quantities of assets can be bought or sold at these prices. We assume that the asset $1$ is risk free, so that $\ed S^1_t=r_t S^1_t \ed t$ where $r_t$ is the short-rate. We assume as our no-arbitrage condition that there there is an equivalent measure $\Q^M$ such that $(S^0_t)^{-1}S^i_t$ is a $\Q$-Martingale for each $i$. We will call such a market an [*infinitely-liquid*]{} market. Mortality model --------------- We will later assume that consumption only occurs at times in a set ${\cal T}$ which may be either $[0,T)$ or the evenly spaced time grid $\{ 0,\, \delta t,\, 2 \delta t,\, 3 \delta t, \ldots, T-\delta t \}$ where $T$ is an upper bound on an individual’s possible age which may be infinite. As a result we may assume that mortality events are also restricted to ${\cal T}$. We write $\DT$ for the measure determined by the index set: this will be the Lebesgue measure on $[0,\infty)$ in the continuous case, or the sum of Dirac masses of mass $\delta t$ at each point in ${\cal T}$ for the discrete case. We assume that the random variables $\tau_i$ representing the time of death of individual $i$ are independent and absolutely continuous with respect to $\DT$, with distribution given by $p^i_t \, \DT$. We will write $(\Omega^{L}, {\cal F}^L, {\cal F}^{L}_t, \P^L)$ for the filtered probability space generated by all the $\tau_i$, the filtration is obtained by requiring that each $\tau_i$ is a stopping time. We will write $F_\tau(t)$ for the distribution function of $\tau$. The assumption of independence means that we are only considering idiosyncratic longevity risk, but we will give some remarks on how systematic longevity risk will affect our findings later in the paper. We will write $n_t$ for the number of survivors at time $t$, that is the number of individuals whose time of death is greater than or equal to $t$. This convention ensures that $n_0=n$ and works well with our convention that cashflows received at the time of death are still consumed. Note, however, that $n_{t+\delta t}$ will be ${\cal F}^L_t$ measurable. Preference model ---------------- We assume that there are $n$ individuals. We model a “pension outcome” for individual $i$ as a pair $(\gamma^i,\tau_i)$ consisting of a stochastic process $\gamma^i_t$, representing the rate of pension payments to individual $i$ at time $t$, and the random variable $\tau_i$ representing the time of death. The underlying filtered probability space will be denoted by $(\Omega, {\cal F}, {\cal F}_t, \P)$ and is assumed to satisfy the usual conditions. The units of $\gamma_t$ should be taken to be in real terms which ensures that our models for inflation and preferences are separate. We will later assume that consumption only occurs at times in a set ${\cal T}$. It will occasionally be convenient to allow the cashflow $\gamma_t$ to be non-zero when $t>\tau$, but this cash will not be consumed. In the discrete case we assume that cashflow at the moment of death $\gamma_\tau$ is still consumed. So the total consumption over the lifetime of an individual is $$\int_0^\tau \gamma_t \, \DT.$$ We will assume that the preferences of our individuals are given by a $\R \cup \{ \pm \infty \}$ valued [*gain function*]{} ${\cal J}_i(\gamma,\tau_i)$ which acts on pension outcomes. The individual prefers pension outcomes which yield higher values of the gain function. We will assume that the gain function takes the value $-\infty$ whenever $P(\gamma_t \id_{t\leq \tau}(t) < 0)>0$. Following [@ab-main] we make the following definitions. The preferences are said to be [*invariant*]{} if they are invariant under filtration preserving automorphisms of the probability space. We will say that ${\cal J}_i$ is [*concave*]{} if it is concave as a function of $\gamma$ for all $\tau_i$. The preferences are [*monotonic*]{} if ${\cal J}_i(\gamma, \tau) \leq {\cal J}_i(\gamma^\prime, \tau)$ if $\gamma_t \leq \gamma^\prime_t$ for all $t \in {\cal T}$. A gain function [*does not saturate*]{} if whenever ${\cal J}(\gamma, \tau)$ is finite we have $J(\gamma+\epsilon,\tau)> {\cal J}(\gamma, \tau)$ for all positive $\epsilon$. \[def:vnm\] [*Von Neumann–Morgernstern preferences with mortality*]{} are determined by a choice of concave, increasing utility function $u:\R_{\geq 0} \to \R$ and a discount rate $b$. The gain function is given by $${\cal J}(\gamma,\tau) = \E\left( \int_0^\tau e^{-bt} u( \gamma_t )\, \DT \right).$$ [*Exponential Kihlstrom–Mirman preferences with mortality*]{} are determined by a choice of concave, increasing utility function $u:\R_{\geq 0} \to \R$. The gain function is given by $${\cal J}(\gamma,\tau) = \E\left( -\exp\left( -\int_0^\tau u( \gamma_t )\, \DT \right) \right).$$ Both of these gain functions give rise to concave, invariant, monotonic preferences. As is explained in detail in [@ab-main], these gain functions are also [*stationary*]{} (meaning that the preferences do not depend upon the time period being considered) and [*law-invariant*]{} (meaning that the preferences depend only on the probability law of the consumption and not the time at which information is received). A larger class of stationary preferences over consumption may be obtained if one drops the requirement that preferences are law-invariant. Such preferences were studied in discrete time by Kreps and Porteus, [@krepsPorteus], and the particular case of Epstein–Zin preferences, introduced in [@epsteinZin1], has proved popular in applications as they allow for separate risk-aversion and intertemporal substitution parameters while maintaining mathematical tractability. Such preferences have been used to resolve a number of asset pricing puzzles, see for example [@bansalYaron; @bansal; @benzoniEtAl; @bhamraEtAl]. In [@xing], a continuous time analogue of Epstein–Zin preferences is defined using BSDEs on probability spaces equipped with a filtration generated by Brownian motion, building on the continuous time analogues of Kreps–Porteus preferences introduced in [@duffieEpstein]. A detailed literature review of the development of such preferences is given in [@xing]. Motivated by [@xing] we may define continuous-time Epstein–Zin preferences with mortality as follows. Suppose we are working with continuous time mortality. Suppose that $\Omega=\Omega^M \times \Omega^L$ and that the filtered probability space $\Omega^M$ is generated by d-dimensional Brownian motion. Let $N^i_t={\1}_{\tau^i \leq t}$ be the jump process associated with the death of individual $i$. Let $\Lambda^i_t$ be the predictable compensator of $N^i_t$ given by $$\Lambda^i_t = \int_0^{t \wedge \tau^i} \lambda_s \, \ed s.$$ where $\lambda_t$ is the force of mortality. Let $M^i_t$ be the compensated martingale process defined by $M^i_t=N^i_t-\Lambda^i_t$. Let ${\cal S}_2$ denote the set of $\R$ valued progressively measurable processes $\{Y_t\}$ $(0\leq t\leq T)$ such that $$\|Y\|^2_{{\cal S}_2}:=\E[ \sup_{t\geq 0} |Y_{t \wedge T}|^2]<\infty.$$ Let $L^2(W)$ denote the set of $\R^d$ valued progressively measurable processes $Z_t$ $(0\leq t\leq T)$ such that $$\|Z\|^2_{L^2(W)}:=\E[ \int_0^\infty \|Z_t\|^2 \ed t ]<\infty.$$ Let $L^2(\lambda)$ denote the set of $\R^n$ valued progressively measurable processes $\zeta_t$ $(0\leq t\leq T)$ such that $$\|\zeta\|^2_{L^2(\lambda)}:=\E[ \sum_{i=1}^n \int_0^{t \wedge \tau^i} |\zeta_s^i|^2 \lambda_s \,\ed s ] < \infty.$$ Let $b>0$ be a discount rate. Let $0\neq\rho<1$ and $0\neq\alpha<1$ be parameters determining the individual’s satiation and risk preferences. The Epstein–Zin aggregator $f:[0,\infty) \times (-\infty,0] \to \R$ is defined by $$f(\gamma,v):=b \frac{ \alpha v }{\rho} \left( \left( \frac{\gamma}{(\alpha v)^\frac{1}{\alpha}} \right)^\rho - 1 \right).$$ We will assume further that $\alpha<0$ and $0<\rho<1$. These parameter restrictions are justified in [@xing] [^1] on the grounds of empirical relevance. We require one additional parameter $a>0$ which we call the [*adequacy level*]{}. A stopping time $\tau$ is admissible if $\tau < T$ almost surely. Given an admissible stopping time $\tau$, the set of admissible consumption streams is defined by $${\cal C}:=\{ \gamma \in {\cal R}_+ \mid \E[ \int_0^\tau e^{-b s} \gamma_s \,\ed s]<\infty \}$$ where ${\cal R}_+$ is the set of all non-negative progressively measurable processes. Given an admissible consumption and mortality $(\gamma_t,\tau)$ we define $(V_t,Z_t,\zeta_t)$ to be the solution of the BSDE $$\ed V_t = f(\gamma_t, V_t) {\1}_{t \leq \tau} \, \ed t - Z_t \, \ed W_t - \sum_{i=1}^n \zeta_t^i \, \ed M^i_t, \quad 0 \leq t \leq T; \quad V_T = U \label{eq:bsde1}$$ where $(V,Z,\zeta) \in {\cal S}_2 \times L^2(W) \times L^2(\tilde{\lambda})$ and where we set $U=\frac{a^\alpha}{\alpha}.$ The existence and uniqueness of solutions of the BSDE may be established using the technique of the proof of Proposition 2.2 of [@xing] combined with the properties of BSDEs with default jumps established in [@dumitrescu]. Since we assume that $\tau$ is bounded above by $T$, the techniques of [@aurandHuang] are not required. We may then define the Epstein–Zin utility with mortality for adequacy level $a$ associated with $(\gamma,\tau)$ by $${\cal J}(\gamma,\tau):=V_0.$$ In the classical Epstein–Zin utility without mortality, the term $U$ in is a ${\cal F}_\tau$-measurable random variable given by a $U=\frac{c_\tau^\alpha}{\alpha}$ where $c_\tau$ denotes a bequest at time $\tau$. Such a formulation is attractive mathematically as it ensures the preferences are positively homogeneous. This symmetry then allows one to reduce the dimension of control problems involving such preferences. However, we note that when $\alpha<0$, this modelling choice would associate an infinitely negative utility to dying without a bequest. No amount of lifetime consumption would be sufficient to overcome this. This does not seem a plausible model for typical pension preferences. In our model, we call $a$ an adequacy level because the outcome $(\gamma,\tau)$ has the same utility as $(\tilde{\gamma},T)$ where $\tilde{\gamma}$ is equal to $\gamma$ up to time $\tau$ and equal to $a$ thereafter (this is because $f(a,U)=0$). Thus an investor is indifferent to dying or living at the adequacy level. The importance of the adequacy level in pension modelling is studied in [@ab-main]. The techniques of [@xing] allow one to prove that Epstein–Zin preferences with mortality are concave and monotone. Epstein–Zin preferences with mortality are invariant. Let $\phi:\Omega \to \Omega$ be a filtration automorphism. Let $(V,Z,\zeta) \in {\cal S}_2 \times L^2(W) \times L^2(\tilde{\lambda})$ be the solution to \[eq:bsde1\]. Then we have $$\ed (V\circ\phi)_t = f((\gamma\circ\phi)_t, (V\circ\phi)_t) {\1}_{t \leq \tau} \, \ed t - (Z\circ\phi)_t \ed (W\circ \phi)_t - \sum_{i=1}^n (\zeta\circ\phi)^i_t \, \ed (M^i\circ\phi)_t$$ for all $0 \leq t \leq T$ and $\quad V_T\circ\phi = U$. By the martingale representation theorem for processes with default jumps (see e.g. [@jeanblancEtAl]) we may find $\tilde{Z}_t \in L^2(W)$ and $\tilde{\zeta} \in L^2(\tilde{\lambda})$ such that $$\ed (V\circ\phi)_t = f((\gamma\circ\phi)_t, (V\circ\phi)_t) {\1}_{t \leq \tau} \, \ed t - \tilde{Z}_t \ed W_t - \sum_{i=1}^n \tilde{\zeta}_t^i \, \ed M^i_t,\quad 0 \leq t \leq T.$$ Hence the Epstein–Zin utility with mortality for $\gamma\circ\phi$ is equal to $(V\circ\phi)_0=V_0$ (the last equality follows since ${\cal F}_0=\{\emptyset,\Omega\}$ and so $V(\omega)_0$ is independent of $\omega$). Managing homogeneous funds {#sec:homogeneousModel} ========================== In this section we consider how to manage homogeneous funds. We call a fund homogeneous if all individuals in the fund have identical preferences, wealth and mortality. We will consider the management of heterogeneous funds in \[sec:heterogeneous\]. Since we assume that each individual has an identical mortality distribution we may define $p_t:=p^i_t$. Let us first consider the case of finite $n$. We wish to decide how to manage a collective pension fund where individual contributes an amount $X_0$ at time $0$. Individual $i$ will receive an income $\gamma^i_t$ at time $t$, with $\gamma^i_t=0$ if the individual is dead at that time. Any cash that is yet to be consumed is invested in the market. There is no bequest when an individual dies, all remaining cash is shared with the fund. We need to choose an objective for the fund itself. We informally outline two proposals which we will then explain in more detail. (i) Suppose that the individual gain function is law-invariant and so depends only on the distribution of outcomes. We may understand “distribution” to mean distribution in both probability and distribution across the population. In this way the individual gain function gives rise to an objective for the entire fund. We will call this the “distribution approach”. (ii) We follow the robust optimization approach to managing the fund: we maximize the infimum of the individual gain functions. Let us explain the mathematical detail required for the distribution approach. We suppose that the individual gain function is law-invariant. We define a discrete uniformly distributed random variable $\iota$ which takes values in $\{1, \ldots, n\}$. We write $(\Omega^\iota, \sigma^\iota, \P^\iota)$ for the probability space generated by $\iota$. We define a filtration ${\cal F}^\iota_{t \in {\R^+ \cup \{ \infty \}}}$ by $${\cal F}_t^\iota = \begin{cases} \{ \Omega^\iota, \emptyset \} & t < \infty \\ \sigma^\iota & t= \infty. \end{cases}$$ Thus $\Omega^\iota$ represents a random choice of individual made at time $\infty$. If we have a law-invariant individual gain function ${\cal J}_\iota$ defined relative to a probability space $\Omega$, we can then define a gain function relative to $\Omega \times \Omega^\iota$ by requiring $${\cal J}^D(\gamma)={\cal J}_\iota(\gamma^\iota, \tau_\iota). \label{eq:distributionObjective}$$ Note that since $\tau_\iota$ is not a stopping time, this gain function can only be given a meaning for law-invariant individual gain functions ${\cal J}_\iota$. The gain function for the robust approach is given by $${\cal J}^R(\gamma):=\inf_{i \in I_0} {\cal J}_i(\gamma^i, \tau_i). \label{eq:robustObjective}$$ These two approaches are not equivalent in general. Consider the Biblical problem faced by Solomon of distributing a child among two women who claim to be its mother. In the distribution approach, giving the child to a randomly selected woman would be optimal. In the robust approach, giving neither woman the child would be an equally optimal alternative. For concave individual gain functions, Solomon’s recommended approach of splitting the child in two would be ideal. Having decided on a gain function ${\cal J}$ for our fund, we can write down the associated optimization problem. We may also augment our probability space with a filtered probability space $\Omega^\perp$ so that any arbitrary decisions that need to be made (such as choosing a random woman to give the child) can be made using random variables defined on this space. We then take $\Omega = \Omega^M \times \Omega^L \times \Omega^\iota \times \Omega^\perp$ equipped with the product filtration ${\cal F}_t$ and product measure $\P$. As one might expect, under reasonable conditions, $\Omega^\perp$ proves to be irrelevant, and the optimal strategies can be taken to be $\Omega^M \times \Omega^L \times \Omega^\iota$ measurable. See Remark \[remark:omegaPerp\] below. Note that when working with Epstein–Zin preferences we will always assume $\Omega^\perp$ is trivial to ensure that the preferences are defined. We then wish to choose progressively measurable consumption streams $\gamma^i_t \geq 0$ with $t \in {\cal T}$ and investment proportions $\alpha^j_t$ in asset $j$ with $t \in \R^+$. We require $\sum_{j=1}^k \alpha^j_t\leq 1$. We have an inequality in this equation, because it is acceptable, if sub-optimal, to simply discard assets. We must choose these processes such that the total wealth of the fund is always non-negative. We write ${\cal A}$ for the resulting set of admissible controls $(\bm{\gamma},\bm{\alpha})$, and we will now give a fully precise mathematical description of this set. First let us write down the dynamics of the fund value. For continuous time consumption the fund value satisfies the SDE $$\ed F_t = \sum_{i=1}^k \alpha^i_s F_s \ed S^i_s - \sum_{i=1}^n \gamma^i_t \, \ed t \label{eq:fundValueCts}$$ with $F_0=\sum_{i=1}^n B_i$, where $B_i$ is the initial budget of individual $i$ (in this section we are assuming that all individuals have identical initial budget $B_i=X_0$, but the equations for the more general case will be useful later). For discrete time consumption, let $F_t$ denote the fund value before consumption and $\overline{F}_t$ denote the fund value after consumption. We then have the following budget equations for the dynamics of $F_t$ and $\overline{F}_t$. $$\begin{split} F_t &= \begin{cases} \sum_{i=1}^n B_i & \text{t = 0} \\ \lim_{h\to 0+}\overline{F}_{t-h} & t \in {\cal T}\setminus\{0\} \\ \overline{F}_t & \text{otherwise.} \end{cases} \\ \overline{F}_t &= \begin{cases} F_t - \sum_{i=1}^n \gamma^{i}_t \\ \overline{F}_{t^\prime} + \sum_{i=1}^k \int_{t^\prime}^t \alpha^i_s \overline{F}_s \, \ed S^i_s & t^\prime \in {\cal T} \text{ and } t^\prime\leq t < t^\prime+\delta t. \end{cases} \end{split} \label{eq:fundValue}$$ A strategy is $(\bm{\gamma},\bm{\alpha})$ is admissible if it ensures $F_t \geq 0$ and $\overline{F}_t \geq 0$ at all times. Hence $${\cal A}=\{(\bm{\gamma},\bm{\alpha})\in {\cal PM} \mid F_t \geq {0} \text{ and } \overline{F}_t \geq 0, \forall t \} \label{eq:admissibleControls}$$ where ${\cal PM}$ is the set of progressively measurable $\R^n \times \R^k$ valued processes for the probability space $\Omega$. Our objective is to compute $$v_n = \sup_{(\bm{\gamma},\bm{\alpha}) \in {\cal A}} {\cal J}(\gamma). \label{eq:fundObjective}$$ and to find $(\bm{\gamma}, \bm{\alpha})$ achieving this supremum. We will henceforth assume that our individual gain functions are concave. Since the market is positively homogeneous, given a strategy $(\bm{\gamma},\bm{\alpha})$ we may form a new strategy $(\bar{\bm{\gamma}},\bm{\alpha})$ which assigns the mean consumption to all survivors: $$\bar{\gamma}^i_t = \begin{cases} 0 & \tau_i < t \\ \frac{1}{n_t} \sum_{j=1}^n \gamma^j_t & \text{otherwise}. \end{cases}$$ The concavity of our individual gain functions ensures that for our fund’s gain function we have $${\cal J}(\bar{\gamma}^i,\alpha) \geq {\cal J}(\gamma^i,\alpha).$$ So we may assume without loss of generality that all survivors consume the same amount $\gamma_t$ at time $t$. Under this assumption we have that ${\cal J}^D={\cal J}^R$, so the distinction between the distributional approach and the robust approach will not in fact be important. Our motivation for introducing the two approaches is that we believe the distributional approach corresponds more closely to the intuitive notion of optimal investment for a collective fund, however the robust approach is required if we wish to use Epstein–Zin utility as the individual gain function. To justify our claim that the distributional approach is more intuitive we first note the example of Solomon above. A second justification is given by the concept of the “Veil of Ignorance” described in [@rawls]. This concept suggests that when making collective decisions one should make those decisions as though the identity of the individuals was unknown. Our filtered probability space $\Omega^\iota$ represents this veil of ignorance: the veil being lifted at time $\infty$, but the control $(\gamma, \alpha)$ is chosen in a state of ignorance. Having decided that the consumption, $\gamma_t$, should be the same for all individuals, we may now consider how to model infinite collectives ($n=\infty$). When performing accounting calculations in this case, we will perform all calculations on a per-individual basis. For example, rather than keep track of the total fund value which would be infinite, we keep track of the fund value per individual which will be finite. We will assume that a deterministic proportion of the original individuals dies over each time interval given by the integral of $p_t \, \DT$ over that interval. This assumption allows us to include mortality within our accounting. Let us express this precisely. Let $Y_t$ represent the fund value per individual at time $t$ before consumption or mortality, and let $\overline{Y}_t$ represent the fund value per individual after consumption. Then in the continuous time case we have $$\ed Y_t = \sum_{i=1}^k \alpha^i_s Y_s \ed S^i_s - \pi_t \gamma_t \, \ed t \label{eq:fundValuePerIndividualCts}$$ where $\pi_t$ is the proportion of individuals surviving to time $t$. In the discrete time case we have $$\begin{split} Y_t &= \begin{cases} X_0 & \text{t = 0} \\ \lim_{h\to 0+}\overline{Y}_{t-h} & t \in {\cal T}\setminus\{0\} \\ \overline{Y}_t & \text{otherwise.} \end{cases} \\ \overline{Y}_t &= \begin{cases} Y_t - \pi_t \gamma_t & t \in {\cal T} \\ \overline{Y}_{t^\prime} + \sum_{i=1}^k \int_{t^\prime}^t \alpha^i_s \overline{Y}_s \, \ed S^i_s & t^\prime \in {\cal T} \text{ and } t^\prime\leq t < t^\prime+\delta t. \end{cases} \end{split} \label{eq:fundValuePerIndividual}$$ For the case $n=\infty$, equations and define the process $Y_t$. In the case $n=\infty$, $\pi_t=1-F_\tau(t)$ is deterministic. For the case of finite $n$, equations and follow from and . In the case of finite $n$, $\pi_t=\frac{n_t}{n}$ is a random variable. In the case $n=\infty$ we take as the gain function for our fund $${\cal J}(\gamma):={\cal J}_1(\gamma, \tau_1). \label{eq:robustObjectiveInfinite}$$ This is reasonable since we have assumed that all living individuals receive the same consumption stream $\gamma_t$ and have isomorphic preferences. Alternatively if the individual gain function is law-invariant we may define a random variable $\tau_\iota$ which is measurable only at $\infty$ and which has distribution $p_t \ed t$. We may then write the gain function as $${\cal J}(\gamma):={\cal J}_\iota(\gamma, \tau_\iota). \label{eq:distributionObjectiveInfinite}$$ These two formulations will be equivalent, but seems a more intuitive formulation. We define the set ${\cal A}$ of admissible strategies in the case $n=\infty$ by saying that a strategy is admissible if $Y_t \geq 0$ for all time. We may now define $v_\infty$ via equation as before. We note that our approach to treating of the case $n=\infty$ is at present rather heuristic, but it will be justified rigorously via limiting arguments in the next section. Basic properties of collectivised investment {#sec:basicProperties} ============================================ Collectivisation is beneficial ------------------------------ It is intuitively clear that collectivisation should be beneficial. We give a formal proof within our model. The probability spaces we have defined depend upon the number of individuals, $n$, but if $n<m$ there is an obvious way to map random variables defined on the probability space for $n$ individuals to random variables defined on the probability space for $m$ individuals that preserves the random variables defining the market and the mortality of the first $n$ individuals. We shall assume henceforth that the individual gain functions are preserved under this mapping. This will happen automatically if we define our gain functions using a specification such as Epstein–Zin preferences with mortality with parameter values $(\alpha,\rho,b,a)$. \[thm:increasingValue\] If the individual gain functions ${\cal J}_i$ are concave then $$v_n \leq v_m \quad \text{if }n\leq m<\infty.$$ Moreover $v_n \leq v_\infty$ in complete markets. In place of assuming that the market is complete, one may instead require that admissible strategies are integrable. We recall that, by assumption, the individual gain functions ${\cal J}_i$ are invariant and isomorphic. We note that to each permutation of the individuals we can associate an automorphism of the probability space which permutes the individuals. Hence a strategy which is effective for one set of $n$ individuals will give rise to an isomorphic strategy for another set of $n$ individuals. Suppose that $m$ is finite. Suppose we set up $\binom{m}{n}$ funds corresponding to each possible choice of $n$ individuals from the full set of $m$ individuals and allocate the initial budget equally to each of these funds. Given an admissible strategy which achieves a value $v$ for the gain function for the first $n$ individuals, we can use it in each of these $\binom{m}{n}$ funds. By the concavity of the gain function, we see that the resulting strategy will have a value greater than or equal to $v$. The result follows. Now consider the case when $m=\infty$. Let $(\bm{\gamma},\bm{\alpha})$ be an admissible strategy for a collective of $n$ investors. Our budget constraint together with the requirement that the discounted asset prices are $\Q$-measure martingales ensures that $$\E_{\Q\times \P^L}\left( \frac{1}{n} \sum_{i=1}^n \int \frac{\gamma^i_t}{S^1_t} \, \DT \right) \leq X_0.$$ Note that $\gamma^i_t$ is non-negative. Hence by Fubini’s theorem we may define a stochastic process $\overline{\gamma}_t$ by $$\frac{\overline{\gamma}_t}{S^1_t}:=\E_{\P^L}\left( \frac{1}{n} \sum_{i=1}^n \int \frac{\gamma^i_t}{S^1_t} \, \DT \mid \bm{S}_t \right) \leq X_0.$$ where ${\bm{S}}_t$ is the vector of asset prices at time $t$. This will be progressively measurable with respect to the filtered probability space of the market $(\Omega^M)$ and moreover will satisfy $$\E_{\Q}\left( \int \frac{\overline{\gamma}_t}{S^1_t} \, \DT \right) \leq X_0.$$ It follows from our assumption that the market is complete that we can then find $\overline{\bm{\alpha}}$ which ensures that $(\overline{\gamma}, \overline{\bm{\alpha}})$ is an admissible strategy for an infinite collective. The concavity of the gain function now ensures that $${\cal J}_i(\gamma^i, \tau_i) \leq {\cal J}_{\iota}(\overline{\gamma}, \tau_\iota).$$ Hence $v_\infty \geq v_n$. If the market is not complete, we instead use the integrability of ${\bm{\alpha}}$ in order to define $\overline{\bm{\alpha}}$ by averaging. To prove that $v_n \to v_\infty$ as $n \to \infty$ we require a little more than concavity and invariance, one also needs some degree of continuity in the gain function when it is perturbed by a low probability event. As an example we prove the following in Appendix \[sec:ezConvergence\]. If preferences are given by Epstein–Zin preferences with mortality with $0<\rho<1$ and $\alpha<0$ then $v_n \to v_\infty$ as $n \to \infty$. \[thm:ezConvergence\] See [@ab-exponential] for the case of exponential utility. The proof of Theorem \[thm:increasingValue\] will continue to work if one introduces systematic longevity risk. The main challenge in extending \[thm:ezConvergence\] lies in defining the utility function in a richer probability space. Nevertheless, the continuity required to establish \[thm:ezConvergence\] can be expected for any gain function that yields a well-posed optimization problem. When is constant consumption optimal? ------------------------------------- Defined benefit pensions and annuities typically aim to provide constant consumption stream in real terms. It is therefore natural to ask when constant consumption is optimal. Our next result gives sufficient conditions. We assume for technical reasons that the asset price processes $S^i_t$ are all given by diffusion processes. We will call a market that satisfies this assumption a diffusion market. Constant consumption is optimal for von Neumann–Morgernstern preferences with $b=0$ in complete diffusion markets with $\P=\Q$ and $r_t=0$ when $n=\infty$. If $p_t=0$ for all times except $T-\delta t$ then constant consumption is optimal for all $n$. For example, in a Black–Scholes–Merton market with no drift and risk free rate of zero one has $\P=\Q$. \[thm:constantConsumption\] We will give our proof using symmetry arguments rather than direct calculation. Our aim in using this approach is to explain why this result feels “obvious”. First consider the case $n=\infty$. By the classification of standard probability spaces, the filtered probability space $(\Omega, {\cal F}_{t \in {\cal T}}, \P)$ is isomorphic to the Cartesian product $(S^1)^{\cal T}$ where $S^1$ is the circle of circumference $1$. Note that we are using the assumption that we are in a diffusion market to ensure that these probability spaces are standard and atomless. Thus rotations of the circles give rise to market isomorphisms (see [@armstrongClassification] for a definition of market isomorphism). This gives an action of the Lie group $(S^1)^{\cal T}$ on our market and hence on the space of admissible investment strategies. Given any strategy we may apply these rotations to obtain new, equivalent strategies. Choose an invariant metric on our Lie group, so that we may define the average strategy. Note that we are using market completeness here, just as we did in the proof of Theorem \[thm:increasingValue\], to ensure that the average strategy exists. By the concavity of our maximization problem and Jensen’s inequality, this averaged strategy will outperform the original strategy. But the averaged strategy will be invariant under the group action and hence deterministic. Since we may assume our strategy is deterministic, it will be a pure bond investment so $Y_{t+\delta t}=\tilde{Y}_t$. From our dynamics we have $$Y_{t+\delta t} = Y_t - (1-F_{\tau}(t))\gamma_t$$ Solving this difference equation yields the (intuitively obvious) budget constraint $$0 \leq X_0 = \int_0^T (1-F_{\tau}(t)) \gamma_t \, \DT. \label{eq:vnmBudgetConstraint}$$ Where $F_{\tau}$ is the distribution function of $\tau$. The expected utility is $$\int_0^T (1-F_{\tau}(t)) u(\gamma_t) \, \DT.$$ The result for $n=\infty$ is now easy to prove by brute force, but we wish to use symmetry. Given a measurable space with non-negative, standard measure $\mu$ we can consider the more general problem $$\begin{aligned} \underset{\gamma \in L^1(\mu)}{\text{maximize}} & \int u(\gamma) \, \mu \\ \text{subject to } & \gamma\geq 0 \\ \text{and} & \int \gamma \, \mu = X_0 \end{aligned} \label{eq:measureProblem}$$ The optimization we wish to solve is just the special case with $\mu=(1-F_\tau(t)) \, \DT$. If $\mu$ is isomorphic to an invariant measure $m$ on the circle $S^1$, we can use the rotation argument above to prove that the optimizer is given by constant $\gamma^*$, specifically $\gamma^*= \frac{X_0}{\int \mu}$. For more general $\mu$, suppose we are given a function $\gamma$ which satisfies the constraints. We let $\gamma \times \id$ be the function on the space with product measure $\mu \times \lambda$ where $\lambda$ is the Lebesgue measure on $[0,1]$. But $\mu \times \lambda$ is an atomless, positive probability measure so is isomorphic to an invariant measure $m$ on the circle. Hence $\gamma \times \id$ is outperformed by the constant function $\gamma^*$ on $\mu \times \lambda$. This can be projected to a constant function $\gamma^*$ on $\mu$. Since the objective functions on $\mu \times \lambda$ and $\mu$ are equal for functions which are constant on the $\lambda$ factor, this means that $\gamma$ itself is outperformed by $\gamma^*$. Hence the constant function $\gamma^*$ is optimal on $\mu$ itself. In the case when $p_t=0$ for all times except $T-\delta t$, the optimization problems for different values of $n$ are all equivalent, so the result follows. This result provides a mathematical explanation for the intuitive appeal of annuities and defined benefit pensions. However, there is the obvious caveat that the assumptions are extremely strong. One expects that with any slight weakening of these assumptions, constant cashflows will no longer be optimal. We prove such a converse in the case of Epstein–Zin preferences in the paper [@ab-ez]. One interpretation of this result that is worth noting is that it suggests taking the discount rate $b=0$ in our preferences will accord better with investor expectations than choosing other values of $b$. This accords with the theoretical discussion of discount rates in pension modelling in [@ab-main]. \[remark:omegaPerp\] We note that the same averaging argument can be used to show that the probability space $\Omega^\perp$ can be safely ignored in a complete market with concave preferences. Heterogeneous Funds {#sec:heterogeneous} =================== We now wish to consider how to manage consumption and investment when the individuals are not identical. For simplicity, let us assume that there are a finite number, $\ell$, of possible initial types of investor $\{\zeta_1, \ldots, \zeta_\ell\}$. We write ${\cal Z}$ for the set of types of investor. Each type $\zeta$ describes the initial capital, mortality distribution and preferences of the individual. Our argument will carry through with appropriate modifications to the case of a compact space of types rather than a finite set of types. However, giving such an account would increase the technical burden to little real purpose. It is hypothetically appealing to identify the optimal investment strategy for the heterogeneous fund. However, we would need to define optimality, and this will require us to specify preferences over outcomes between different types of investor. Describing such preferences seems challenging and the choice of preferences would be controversial. Instead we will take an axiomatic approach. Rather than find an optimal scheme for managing a fund we merely seek an [*acceptable*]{} management scheme. We will define the notion of acceptable axiomatically, and will show that for large funds all acceptable management schemes yield similar outcomes for the investors. We assume that the preferences of an individual of type $\zeta$ are given by a concave, invariant gain function ${\cal J}_\zeta(\gamma,\tau)$. Let us write $v(n,\zeta)$ for the value function for individuals of type $\zeta$ investing in this market in a collective of size $n$ as modelled in Section \[sec:homogeneousModel\]. We assume that $$\lim_{n\to \infty} v(n,\zeta) = v(\infty,\zeta).$$ Sufficient conditions to ensure this are described in the papers [@ab-exponential] and [@ab-ez]. The initial population is determined by a vector $\bm{\zeta} \in {\cal Z}^n$, where component $i$ of $\bm{\zeta}$ denotes the type of the $i$-th individual. A [*management scheme*]{} ${\cal M}$ is a function acting on vectors $\bm{\zeta}$ of arbitrary length $n$ and which yields a strategy $(\bm{\gamma}, \bm{\alpha})$ where $\bm{\gamma}_t$ is a vector of consumptions of length $n$ with $i$-th component being the consumption of individual $i$, and $\bm{\alpha}$ is an investment strategy. We require that the combined consumption and investment are self-financing. Thus $${\cal M}: \bigsqcup_{n=1}^\infty {\cal Z}^n \to \bigsqcup_{n=1}^\infty L^0(\R^n \times \R^k).$$ and ${\cal M}$ maps the $i$-th component of the first union into the $i$-th component of the second union. Our first axiom for ${\cal M}$ is one of fairness, which as we have seen is a property that arises automatically in any concave maximization problem. [I1]{} \[axiom:fairness\] All surviving individuals of type $\zeta$ consume the same amount. That is (i) if ${\cal M}(\bm{\zeta})=(\bm{\gamma}, \bm{\alpha})$ then $\gamma^i_t = \gamma^j_t$ when $\bm{\zeta}(i)=\bm{\zeta}(j)$, $t\leq \tau_i$ and $t\leq \tau_j$. (ii) If $\bm{\zeta}^\prime$ is obtained by permuting the elements of $\bm{\zeta}$, then ${\cal M}(\bm{\zeta}^\prime)$ is obtained by the corresponding permutation of the consumption streams of ${\cal M}(\bm{\zeta})$ Let the proportions of different individuals prevailing in the population be given by a vector of weights $\omega(\bm{\zeta})=(\omega_{\zeta_1}, \ldots, \omega_{\zeta_\ell})$ with $0<\omega_\zeta<1$, $\omega_\zeta$ rational and $\sum_{\zeta \in {\cal Z}} \omega_{\zeta}=1$. Let $\lcm(\omega)$ denote the lowest common multiple of the denominators of the fractions $\omega_i$, so we know that the population is some integer multiple of $\lcm(\omega)$. By Axiom \[axiom:fairness\] we may define $$a_{\cal M}(\omega,n,\zeta)$$ to be the value of the gain function achieved by an individual of type $\zeta$ for a population of size $n$. This is defined if $n$ is any integer multiple of $\lcm(\omega)$. If $a_{\cal M}(\omega, n, \zeta) < v(n \, \omega_\zeta, \zeta)$ then the investment strategy for the heterogeneous fund will not be able to attract investors of type $\zeta$ as they would be better off following the strategy of Section \[sec:homogeneousModel\]. Theorem \[thm:increasingValue\] then suggests that this would be to the detriment of all other investors in the heterogeneous fund, as increasing collectivisation should always be beneficial. These observations motivate the following axioms. [I2]{} A management scheme is monotone if: $$a_{\cal M}(\omega,m,\zeta) \leq a_{\cal M}(\omega,n,\zeta)$$ if $m\leq n$. \[axiom:montone\] [I3]{} A management scheme achieves the [*performance standard*]{} if $$a_{\cal M}(\omega,n,\zeta) \geq v(n \, \omega_\zeta, \zeta).$$ for all $n$ and $\zeta$. \[axiom:performance\] A management scheme is [*acceptable*]{} if it satisfies Axioms \[axiom:fairness\], \[axiom:montone\] and \[axiom:performance\]. By the monotonicity property, we may unambiguously define $$a_{\cal M}(\omega,\infty,\zeta) = \lim_{n\to \infty} a_{\cal M}(\omega,n,\zeta).$$ By the performance standard and our assumption on the convergence of $v(n,\zeta)$ as $n\to \infty$ we see that for any acceptable management scheme $$a_{\cal M}(\omega,\infty,\zeta) \geq v(\infty, \zeta). \label{eq:heterogeneousLowerBound}$$ The scheme of simply grouping all individuals of a given type together into a homogeneous fund and managing that according to the model of Section \[sec:homogeneousModel\] will yield an acceptable management scheme which achieves the lower bound . We call this the [*basic management scheme*]{}. We wish to show that the basic management scheme is asymptotically optimal in complete markets. The key observation is that collective investment provides no substantive benefit for the investment problem in a complete market without mortality. Let us consider how to write a collective investment problem for $n$ individuals without mortality investing in our market. We choose consumption $\gamma^i_t$ for individual $i$, and overall investment proportions $\alpha_t$. The total fund value will be given by the budget equations , and hence the set of admissible controls for this problem will be given by . The difference will be that the preferences of the individual will depend only upon cashflows received and not upon mortality, so we will suppose that for individual $i$ we have a gain function $\hat{\cal J}_i(\gamma^i_t)$ which depends only upon the cashflows. We define the value function $$\hat{v}_i := \sup_{{\cal A}_1} \hat{\cal J}_i(\gamma^i_t)$$ where ${\cal A}_n$ is the set of acceptable admissible controls for $n$ individuals. Although we have defined the set of admissible controls for $n$ individuals, we do not write down an optimization problem for $n$ individuals as we do not know what the objective should be across a heterogeneous collective. Our next result shows that in a complete market without mortality there is no real benefit in considering collectivised problems as, however we select an admissible control (by solving an optimization problem or otherwise), it will never bring a substantive advantage to any individual unless it also gives a substantive disadvantage to some other individual. In other words, one cannot pay Paul without robbing Peter. \[lemma:collectivisationNoHelp\] Suppose the gain functions $\hat{\cal J}_i(\gamma_t)$ are concave, monotone and do not saturate[^2] and satisfy $\hat{\cal J}_i(\gamma_t)>-\infty$ for positive cashflows $\gamma_t$ . Suppose that $\hat{\cal J}_i(\gamma_t) = -\infty$ whenever $\gamma_t$ is negative on a set of finite measure. Let $i_*$ be an individual, then for any $\epsilon_1>0$ there exists $\epsilon_2>0$ such that for any admissible investment consumption strategy for all the investors with $\gamma^i$ satisfying $$\hat{\cal J}_{i_*}(\gamma^{i_*}_t) \geq \hat{v}_{i_*} + \epsilon_1 \label{eq:optimalForEachI}$$ there is an investor $i$ such that $$\hat{\cal J}_i(\gamma^i_t) \leq \hat{v}_i - \epsilon_2.$$ Let us write $v_i(b)$ for the value function for individual $i$ as a function of their budget $b$. By Lemma below, $v_i$ is continuous as a function of $b$ for any $b > 0$. We write $B_i$ for the budget of individual $i$. In a complete market, we call a measurable non-negative cashflow $\gamma_t$ a derivatives contract and we call the discounted $\Q$-measure expectation of $\gamma_t$ the price of this contract. This price may be infinite, which means that the cashflows cannot be super-replicated by any admissible trading strategy. If the price is finite, the contract can be replicated by an admissible trading strategy with initial budget given by the price. Note that the requirement that $\gamma_t$ is non-negative ensures that the price of derivative contracts is additive: if negative infinities were allowed as prices this would not be the case. If we write ${\cal D}_b$ for the set of derivatives contracts of price less than or equal to $b$, we have: $$\hat{v}_i(b) = \sup_{\gamma \in {\cal D}_b} \hat{\cal J}(\gamma).$$ Here we have used our assumption that negative cashflows yield a value of $-\infty$ for the gain function. By the continuity of $v_{i_*}$, there is a price, $\delta$, such that if $\hat{\cal J}_{i_*}(\gamma^{i_*}_t)\geq v_{i_*} + \epsilon_1$, then the price of the derivative contract with payoff $\gamma^{i_*}_t$ is at least $B_{i_*}+\delta$. By the monotonicity and non-saturation of the gain functions, together with the concavity given by Lemma below $$\hat{v}_i\left(B_{i}-\frac{\delta}{\ell-1}\right) < \hat{v}_i(B_i).$$ Let $$\epsilon_2 = \inf_{i\neq i_*} \left\{\hat{v}_i(B_i)- \hat{v}_i\left(B_{i}-\frac{\delta}{2(\ell-1)}\right) \right\}.$$ Any derivative contract with cashflows $\gamma^i$ satisfying $$\hat{\cal J}_i(\gamma^i_t) \leq \hat{v}_i - \epsilon_2.$$ will cost at least $B_{i}-\frac{\delta}{2(\ell-1)}$. Hence the total cost of an investment strategy yielding all the cashflows $\gamma^i$ is at least $$B_{i}+\delta + \sum_{i\neq i_*} \left( B_{i}-\frac{\delta}{2(\ell-1)} \right) = \sum_i B_{i} + \frac{\delta}{2}$$ which is greater than the total budget and hence cannot be admissible. \[lemma:continuityOfV\] Consider investment without mortality in a homogeneous market. Suppose an individual’s gain function over consumption, $\hat{\cal J}(\gamma)$, is concave and monotone. Let ${\cal A}_b$ be the set of admissible consumption, investment strategies with initial budget $b$. Define $$\hat{v}(b) = \sup_{(\gamma,\alpha)\in{\cal A}_b} \hat{\cal J}_i(\gamma).$$ The function $v$ is concave and hence $v$ is continuous on any open set on which it is finite. Given two budgets $b_1$, $b_2>b$ then let $(\gamma_i,\alpha_i)$ be admissible strategies for each budget. So for any $\lambda \in [0,1]$, $(\lambda \gamma_1 + (1-\lambda) \gamma_2, \lambda \alpha_1 + (1-\lambda) \alpha_2)$ is an admissible strategy with budget $\lambda b_1 + (1-\lambda) b_2$. By the concavity of $\hat{\cal J}$, $$\hat{\cal J}(\lambda \gamma_1 + (1-\lambda) \gamma_2 ) \geq \lambda \hat{\cal J}(\gamma_1) + (1-\lambda) \hat{\cal J}(\gamma_2 ).$$ Hence $$\hat{v}(\lambda b_1 + (1-\lambda) b_2 ) \geq \lambda \hat{v}(b_1) + (1-\lambda) \hat{v}(b_2).$$ Lemma \[lemma:collectivisationNoHelp\] is only true in complete markets. If two individuals have different risk or consumption preferences in an incomplete market then it can often be beneficial to design a derivatives contract to the mutual advantage of both parties. This is the essential purpose of derivatives contracts and explains why there is a market for such contracts. We are now ready to prove the main result of this section. \[thm:acceptableFundTheorem\] For any acceptable management scheme in a complete market $$a_{\cal M}(\omega,\infty,\zeta) = v(\infty, \zeta)$$ so long as all individual gain functions are concave, monotone, invariant and do no saturate, and so long as $$\lim_{n \to \infty} v(n,\zeta) = v(\infty, \zeta)>-\infty. \label{eq:convergenceRequirement}$$ for any positive budget. We will prove the result in the case of discrete time consumption. The case of continuous time consumption is similar. By Axiom \[axiom:performance\], $\lim_{n \to \infty} a_{\cal M}(\omega, n, \zeta) \geq \lim_{n \to \infty} v(n \omega^\zeta, \zeta)$. By assumption $\lim_{n \to \infty} v(n, \zeta)= v(\infty,\zeta)$. Hence $a(\omega, \infty, \zeta) = \lim_{n \to \infty} a_{\cal M}(\omega, n, \zeta) \geq v(\infty, \zeta)$. Let us now define the meaning of an admissible strategy for an infinite heterogeneous collective where the types of each individual are given by the proportions $\omega$. We will assume that surviving individuals of a given type, $\zeta$, all consume $\gamma^\zeta_t$ at time $t$. Hence the total consumption per person (i.e. of any type, including both survivors and the deceased) is $$\sum_{\zeta \in {\cal Z}} \gamma^{\zeta}_t \pi^{\zeta}_t \omega^{\zeta}$$ where $\pi^{\zeta}_t$ denotes the proportion of individuals of type $\zeta$ who survive to time $t$. Note that $\pi^{\zeta}_t=1-F_{\tau_\zeta}(t)$ where ${\tau}_\zeta$ is a random variable distributed according to the time-of-death distribution for individuals of type $\zeta$. If we let $Y_t$ denote the fund value per person before consumption and $\overline{Y}_t$ denote the fund value per person after consumption we have the following budget equations for the dynamics of $Y_t$ and $\overline{Y}_t$. Let us write $B^\zeta$ for the initial budget of individuals of type $\zeta$. Then the fund value per person for the infinite heterogenous collective should be defined to follow the dynamics $$\begin{split} Y_t &= \begin{cases} \sum_{\zeta \in {\cal Z}} \omega^\zeta B^\zeta & \text{t = 0} \\ \lim_{h\to 0+}\overline{Y}_{t-h} & t \in {\cal T}\setminus\{0\} \\ \overline{Y}_t & \text{otherwise.} \end{cases} \\ \overline{Y}_t &= \begin{cases} Y_t - \sum_{\zeta \in {\cal Z}} \gamma^{\zeta}_t \pi^{\zeta}_t \omega^{\zeta} & t \in {\cal T} \\ \overline{Y}_{t^\prime} + \sum_{i=1}^k \int_{t^\prime}^t \alpha^i_s \overline{Y}_s \, \ed S^i_s & t^\prime \in {\cal T} \text{ and } t^\prime\leq t < t^\prime+\delta t. \end{cases} \end{split} \label{eq:budgetHeterogeneousCollective}$$ We may alternatively view the equations above as describing the dynamics of a fund of “virtual individuals” where each virtual individual represents the interests of an infinite fund of individuals all of type $\zeta$. To see this, we define $\hat{\gamma}^\zeta_t=\gamma^{\zeta}_t \pi^{\zeta}_t \omega^{\zeta}$. We say that virtual individual $\zeta$ consumes an amount $\hat{\gamma}^\zeta_t$ at each time $t \in {\cal T}$. We say that the initial budget of virtual individual $\zeta$ is $\omega^\zeta B^\zeta$. We may now view equation as giving the dynamics of the total fund value before consumption $Y_t$ of a collective of these heterogeneous virtual individuals. This observation will allow us to apply Lemma \[lemma:collectivisationNoHelp\] to the collective investment problem with mortality. Let us define a gain function (without mortality) $\hat{\cal J}_\zeta$ for the virtual individual $\zeta$ by $$\hat{\cal J}_\zeta(\hat{\gamma}^\zeta_t):={\cal J}_\zeta( \gamma^{\zeta}_t, {\tau}_\zeta)$$ Let us write $\hat{v}_\zeta(b)$ for the value function of a virtual individual with this gain function with an initial budget $b$. We see that $\hat{v}_\zeta( \omega^\zeta B_\zeta)=v(\infty, \zeta)$. Suppose $(\bm{\gamma},\bm{\alpha})$ is a strategy for the infinite heterogeneous collective and let $\zeta_*$ be a chosen type of individual. It follows by Lemma \[lemma:collectivisationNoHelp\] that for any given $\epsilon_1>0$ such that $$\hat{\cal J}_{\zeta_*}(\hat{\gamma}^{\zeta_*}_t) \geq v(\infty, \zeta_*) + \epsilon_1$$ there exists $\epsilon_2>0$ and a type $\zeta$ such that $$\hat{\cal J}_{\zeta}(\hat{\gamma}^\zeta_t) \leq v(\infty, \zeta) - \epsilon_2.$$ Hence given any $\epsilon_1>0$ such that $${\cal J}_{\zeta_*}(\gamma^{\zeta_*}_t, \tau_{\zeta_*}) \geq v(\infty, \zeta_*) + \epsilon_1 \label{eq:sigmaStarCondition}$$ there exists $\epsilon_2>0$ and a type $\zeta$ such that $${\cal J}_{\zeta}(\gamma^\zeta_t,\tau_{\zeta}) \leq v(\infty, \zeta) - \epsilon_2. \label{eq:epsilon2Condition}$$ Now let us suppose for a contradiction that for some $\epsilon_1>0$, $\zeta_*$ and $N$ we have $a(\omega, N, \zeta_*) \geq v(\infty, \zeta_*) + 2 \epsilon_1$. Note that we then have $a(\omega, n, \zeta_*) \geq v(\infty, \zeta_*) + 2 \epsilon_1$ for all $n \geq N$. We may then choose $\epsilon_2>0$ as described in the preceding paragraph. There are only finitely many types $\zeta$ and we know $\lim_{n\to \infty} a_{\cal M}(\omega,n,\zeta)\geq v(\infty, \zeta)$. So for sufficiently large $n$ we may assume that $$a_{\cal M}(\omega, n, \zeta) \geq v(\infty, \zeta)-\tfrac{1}{3} \epsilon_2.$$ Hence for such $n$ we may find an investment-consumption strategy $(\bm{\gamma}, \bm{\alpha})$ for a collective of $n$ individuals which yields the same consumption for all surviving individuals of a given type and which satisfies $${\cal J}_{\zeta_*}(\gamma^{\zeta_*}_t, \tau_{\zeta^*}) \geq v(\infty, \zeta_*) + \epsilon_1$$ and which also satisfies $${\cal J}_{\zeta}(\gamma^\zeta_t, _{\zeta}) \geq v(\infty, \zeta) - \tfrac{2}{3} \epsilon_2$$ for all types $\zeta$. Since the discounted asset prices $S^i_t$ are $\Q$-martingales we have $$\E_{\Q \times \P^L} \left( \sum_{\zeta \in {\cal Z}} \int e^{-rt} \omega^\zeta \gamma_t^\zeta \, \DT \right) \leq \sum_{\zeta \in {\cal Z}} \omega^\zeta B^\zeta.$$ Since the consumption $\gamma_t$ is non-negative we have by Fubini’s theorem that $$\overline{\gamma}^\zeta_t := \E_{\P^L} \left( \int e^{-rt} \omega^\zeta \gamma_t^\zeta \, \DT \mid \bm{S}_t \right)$$ is a progressively measurable process for the filtered probability space $(\Omega^M, {\cal F}_t, \P^M)$ and satisfies $$\sum_{\zeta \in {\cal Z}} \E_{\Q} \left( \int e^{-rt} \omega^\zeta \overline{\gamma}_t^\zeta \, \DT \right) \leq \sum_{\zeta \in {\cal Z}} \omega^\zeta B^\zeta.$$ By the complete market assumption we can then find an investment strategy $\overline{\bm{\alpha}}$ that funds $\overline{\bm{\gamma}}$. Hence we have found an admissible strategy $(\overline{\bm{\gamma}},\overline{\bm{\alpha}})$ for the infinite homogeneous collective which by the concavity of the preferences will satisfy $${\cal J}_{\zeta_*}(\overline{\gamma}^{\zeta_*}_t, \tau_{\zeta^*}) \geq v(\infty, \zeta_*) + \epsilon_1$$ and which also satisfies $${\cal J}_{\zeta}(\overline{\gamma}^\zeta_t, _{\zeta}) \geq v(\infty, \zeta) - \tfrac{2}{3} \epsilon_2$$ for all types $\zeta$. This contradicts equations and . We deduce that $a(\omega, N, \zeta) \leq v(\infty, \zeta)$ for all $\zeta$, which gives the result. The financial significance of Theorem \[thm:acceptableFundTheorem\] is that all acceptable strategies are asymptotically equivalent in a complete market, in particular this applies to the Black–Scholes–Merton market. The result is analogous to the classical result that in the Black–Scholes–Merton market any derivative has a unique price independent of the preferences of the investor. This too is an asymptotic result, in the sense that it assumes continuous time and zero transaction costs. In practice no two individuals are alike, and so our assumption of a finite number of distinct types can be criticised. However, simple modifications of our strategy for heterogeneous funds will yield good results if there are a large number of similar individuals. We give a concrete algorithm in [@ab-main] and show that it achieves approximately $98\%$ of the maximum benefit of collectivisation for a heterogeneous fund of only $100$ investors. It is natural to ask how our results extend if one incorporates systematic longevity risk. The simplest approach would be to assume that there is a complete market of contracts on the risk factors determining systematic longevity risk, in which case a similar result can be expected to hold. Longevity derivatives do exist, but at present the market is not very liquid. Our results suggest that each type of individual will have different views on the attractiveness of longevity risk, which would imply that the development of collective funds should result in a more liquid market in longevity derivatives. Proof of Theorem \[thm:ezConvergence\] {#sec:ezConvergence} ====================================== Recall that when working with Epstein–Zin preferences we assume $\Omega^\perp$ is trivial. To prove the existence of Epstein–Zin utility with mortality following the methods of [@xing] one first makes the change of variables $(\tilde{V}_t,\tilde{Z}_t, \tilde{\zeta}_t)=\alpha e^{-\frac{b \alpha t}{\rho}} (V_T,Z_t,\zeta_t)$ to transform \[eq:bsde1\] to the BSDE $$\ed \tilde{V}_t = F(t,\gamma_t, \tilde{V}_t) {\1}_{t \leq \tau} \, \ed t - \tilde{Z}_t \, \ed W_t - \sum_{i=1}^n \tilde{\zeta}_t^i \, \ed M^i_t, \quad 0 \leq t \leq T; \quad \tilde{V}_T = e^{ -\frac{bT\alpha}{\rho}} \alpha U \label{eq:bsde2}$$ where $$F(t,\gamma_t,v):= \frac{b \alpha \exp( -b t )}{\rho} \gamma_t^\rho v^{1-\frac{\rho}{\alpha}}.$$ To prove the existence and uniqueness of this BSDE, one next solves the family of BSDEs indexed by $m\in \R_{\geq 0}$, $$\ed \tilde{V}^m_t = F^m(t,\gamma_t, \tilde{V}^m_t) {\1}_{t \leq \tau} \, \ed t - \tilde{Z}_t \, \ed W_t - \sum_{i=1}^n \tilde{\zeta}_t^i \, \ed M^i_t, \quad 0 \leq t \leq T; \quad \tilde{V}_T = e^{ -\frac{bT\alpha}{\rho}} \alpha U \label{eq:bsde3}$$ where $$F^m(t,\gamma_t,v):=b \frac{\alpha}{\rho} \exp( -b t )(\gamma_t^\rho \wedge m) (v \wedge m)^{1-\frac{\rho}{\alpha}}.$$ The driver $F^m$ has the property that $y \to F^m(t,\gamma_t,y)$ is Lipschitz (since our parameter assumptions ensure $1-\frac{\rho}{\alpha}>1$) and so one can use the theory of [@dumitrescu] to obtain the existence of a unique solution to this BSDE. The comparison theorem ensures that $\tilde{V}^m$ is non-negative and decreasing in $m$. The proof of Proposition 2.2 in [@xing] then shows that the $\tilde{V}=\lim_{m \to \infty}\tilde{V}^m$ can be extended to give a solution to . Let us define $\EZ^{m}(\gamma,\tau):=\frac{1}{\alpha}\tilde{V}^m_0$. We have that $EZ^m(\gamma, \tau)$ is finite, non-positive and increasing in $m$ with limit given by the Epstein–Zin utility with mortality which we denote $\EZ(\gamma,\tau)$. Let $M < v_\infty$ and $\epsilon>0$ be given. We may find an admissible consumption stream $\gamma^\infty_t$ for the problem of an infinite collective with initial budget $B$ such that $\EZ(\gamma^\infty_t, \tau^\iota) \geq M$. Since $\EZ$ is concave and finite, for any admissible $\gamma$, the map $\lambda\to \EZ( \lambda \gamma^\infty_t, \tau^\iota)$ is continuous on $\R_{\geq 0}$. Hence we may choose $\lambda\in(0,1)$ such that $\EZ(\lambda \gamma^\infty_t, \tau^\iota) \geq M-\epsilon$. By the convergence of $\EZ^m$ in $m$, we may then find $m$ such that $$\EZ^m(\lambda \gamma^\infty_t, \tau^\iota) \geq M-2 \epsilon. \label{eq:choiceOfM}$$ Suppose that we have an initial budget of $B$ for the problem of a collective of $n$ investors. Recall that $n_t$ denotes the number of survivors at time $t$. Let $G_{n,t}$ be the event defined by $$G_{n,t} = \{\omega \mid n_s(\omega)\leq \frac{1}{\lambda} \E(n_s) \, \forall 0\leq s \leq t\}.$$ We write $\lnot G_{n,t}$ for its compliment. The consumption stream $\lambda \1_{G_{n,t}} \gamma_t^\infty$ will be admissible, as the consumption per survivor will always be less than $\gamma_t$. Define $F^{m,\infty}(t,v):=F^m(t,\lambda \gamma_t^\infty,v)$ and $F^{m,n}(t,v):=F^m(t,\lambda \1_{G_{n,t}} \gamma_t^\infty,v)$ for finite $n$. Write $(\tilde{V}^{m,n},\tilde{Z}^{m,n},\tilde{\zeta}^{m,n})$ for the solution to for the driver $F^{m,n}$ for both finite and infinite values of $n$, and define $\EZ^{m,n}:=\frac{1}{\alpha}\tilde{V}^{m,n}$ Differentiating the expression for $F^{m,n}(t,v)$ with respect to $v$, we obtain the following Lipschitz estimate for the drivers $$|F^{m,n}(t,y_1)-F^{m,n}(t,y_2)| \leq \left( 1 - \frac{\rho}{\alpha} \right) b \frac{|\alpha|}{\rho} m^{1-\frac{\rho}{\alpha}} |y_1-y_2|=:C_m |y_1-y_2|$$ for an appropriately defined constant $C_m$. Let us define $e=|\tilde{V}^{m,\infty}-\tilde{V}^{m,n}|$. By Proposition 2.4 of [@dumitrescu], we obtain the following bound on $e$: $$e^2 \leq \eta_m \E [ \int_0^T e^{\beta_m s} |F^{m,\infty}(s,\tilde{V}^{m,\infty}) - F^{m,n}(s,\tilde{V}^{m,n}) |^2 \, \ed s ],$$ where $\eta_m:=\frac{1}{C_m^2}$ and $\beta_m:=3 C_m^2 + 2 C_m$. Inserting the definitions of $F^{m,n}$ we obtain $$e^2 \leq \eta_m \E[ \int_0^T \frac{b \alpha}{\rho} e^{\beta_m s} ((\lambda \1_{\lnot G_{n,s}} \gamma^\infty_s)^\rho \wedge m)(\tilde{V}^{m,\infty} \wedge m)^{1-\frac{\rho}{\alpha}} \, \ed s.$$ Splitting the integral into two regions, we find that for $\delta\in(0,T)$, $$\begin{aligned} e^2 &\leq \eta_m \frac{ b \alpha m^{2-\frac{\rho}{\alpha}}}{\rho} e^{\beta_m T} \E[ \int_0^{T-\delta} \1_{\lnot G_{n,s}} \, \ed s + \int_{T-\delta}^T \ed s ] \nonumber \\ &\leq \eta_m \frac{ b \alpha m^{2-\frac{\rho}{\alpha}}}{\rho} e^{\beta_m T} \E[ \int_0^{T-\delta} \1_{\lnot G_{n,T-\delta}} \, \ed s + \delta ] \quad \text{as }\lnot G_{n,s} \subseteq \lnot G_{n,t} \text{ if } s\leq t \nonumber \\ &\leq \tilde{C}_m (T \P( \lnot G_{n,T-\delta} ) + \delta ) \label{eq:eBound}\end{aligned}$$ for an appropriately defined $C_m$. Choose $\delta$ such that $\tilde{C}_m \delta \leq \frac{1}{2}\alpha^2 \epsilon^2$. Now use the fact that $P(\lnot G_{n,T-\delta})\to 0$ as $n \to \infty$ (as follows readily from Lemma \[lemma:finiteTimePoints\], below) to choose $n$ such that $\tilde{C}_m (T \P( \lnot G_{n,T-\delta} ) \leq \frac{1}{2}\alpha^2 \epsilon^2$. By we will then have that $e \leq |\alpha| \epsilon$. We then have $$v_n \geq \EZ( \lambda \1_{G_{n,t}} \gamma^\infty_t ) \geq |\EZ^{m,n}| \geq |\EZ^{m,\infty}| - \frac{e}{|\alpha|} \geq |\EZ^{m,\infty}| - \epsilon \geq M - 3\epsilon.$$ where we have used in sequence: the definition of $v_n$; the convergence of the increasing sequence in $m$ given by $\EZ^{m,n}$; the definitions of $e$ and $\EZ^{m,n}$; our bound for $e$; equation . So by Theorem \[thm:increasingValue\], $\lim_{n \to \infty}v_n=v_\infty$. We now prove the Lemma used in the proof. Let $T^*$ be minimum time by which an individual is almost sure to have died. For a continuous mortality distribution, given a time point $0\leq t_0<T^*$, for any $\epsilon \in (0,1)$ there exists a finite set of points $t_i \in [0,t_0)$, indexed by $i \in I$ such that $$\P\left( \forall t \in [0,t_0] \,:\, n_t \leq \left( \frac{1}{1-\epsilon} \right)^2 \E(n_t) \right) \geq \P\left( \forall i \in I \,:\, n_{t_i} \leq \left( \frac{1}{1-\epsilon} \right) \E(n_{t_i}) \right) .$$ \[lemma:finiteTimePoints\] We define $t_i$ inductively. If $t_{i-1}=0$, we are done and take the index set $I=\{0,1,\ldots,i-1\}$. Otherwise define $$t_i = \inf \left\{ t \mid t=0 \text{ or } \E(n_{t}) \leq \frac{1}{1-\epsilon} \E(n_{t_{i-1}}) \right\}.$$ If $t_i\neq 0$ we see $\E(n_{t_{i-1}}) \geq \frac{1}{1-\epsilon} \E(n_{t_i})$, so for sufficiently large $i$ we must have $t_i=0$. Hence the index set, $I$, is finite. Given $t \in [0,t_0]$ we can find $i \in I$ with $t_i \leq t \leq t_{i-1}$. Suppose $$n_t > \left( \frac{1}{1-\epsilon}\right)^2 \E( n_t)$$ then we have $$n_{t} > \left( \frac{1}{1-\epsilon}\right)^2 \E( n_{t_i}) \geq \left( \frac{1}{1-\epsilon}\right)^2 \E( n_{t_{i-1}}) \geq \left( \frac{1}{1-\epsilon}\right) \E( n_{t}).$$ [^1]: [@xing] uses a slightly different parameterization. Their parameters $\delta$, $\gamma$ and $\psi$ are related to ours by $\delta=b$, $\gamma=1-\alpha$ and $\psi=\frac{1}{1-\rho}$. [^2]: The definitions of these terms were given for gain functions over cashflows with mortality, but the corresponding definitions for $\hat{\cal J}$ should be obvious
{ "pile_set_name": "ArXiv" }
--- abstract: 'In this paper, we generalize the idea from the method called “PCANet” [@chan2015pcanet] to achieve a new baseline deep learning model for image classification. Instead of using principal component vectors as the filter vector in “PCANet”, we use basis vectors in discrete Fourier analysis and wavelets analysis as our filter vectors. Both of them achieve comparable performance to “PCANet" in benchmark datasets. It is noticeable that our algorithms do not require any optimization techniques to get those basis.' bibliography: - 'bib.bib' ---
{ "pile_set_name": "ArXiv" }
--- author: - - - '[^1]' - '[^2]' - bibliography: - 'ref.bib' title: | Improvement of heavy-heavy current for calculation of\ $\bar{B}\to D^{(*)}\ell\bar{\nu}$ form factors using Oktay-Kronfeld heavy quarks --- Acknowledgements {#acknowledgements .unnumbered} ================ The research of W. Lee is supported by the Creative Research Initiatives Program (No. 2017013332) of the NRF grant funded by the Korean government (MEST). W. Lee would like to acknowledge the support from the KISTI supercomputing center through the strategic support program for the supercomputing application research (No. KSC-2015-G2-002). Computations were carried out on the DAVID clusters at Seoul National University. J.A.B. is supported by the Basic Science Research Program of the National Research Foundation of Korea (NRF) funded by the Ministry of Education (No. 2015024974). [^1]: [^2]: Speaker
{ "pile_set_name": "ArXiv" }
--- abstract: | The linear stability of a rotating, stratified, inviscid horizontal plane Couette flow in a channel is studied in the limit of strong rotation and stratification. Two dimensionless parameters characterize the flow: the Rossby number $\eps$, defined as the ratio of the shear to the Coriolis frequency and assumed small, and the ratio $s$ of the Coriolis frequency to the buoyancy frequency, assumed to satisfy $s \le 1$. An energy argument is used to show that unstable perturbations must have large, $O(\eps^{-1})$ wavenumbers. This motivates the use of a WKB-approach which, in the first instance, provides an approximation for the dispersion relation of the various waves that can propagate in the flow. These are Kelvin waves, trapped near the channel walls, and inertia-gravity waves with or without turning points. Although, the wave phase speeds are found to be real to all algebraic orders in $\eps$, we establish that the flow is unconditionally unstable. This is the result of linear resonances between waves with oppositely signed wave momenta. Three modes of instabilities are identified, corresponding to the resonance between (i) a pair of Kelvin waves, (ii) a Kelvin wave and an inertia-gravity wave, and (iii) a pair of inertia-gravity waves. Whilst all three modes of instability are active when the Couette flow is anticyclonic, mode (iii) is the only possible instability mechanism when the flow is cyclonic. We derive asymptotic estimates for the instability growth rates. These are exponentially small in $\eps$, of the form $\Im \omega = a \exp(-\Psi/\eps)$ for some positive constants $a$ and $\Psi$. For the Kelvin-wave instabilities (i), we obtain analytic expressions for $a$ and $\Psi$; the maximum growth rate, in particular, corresponds to $\Psi=2$. For the other types of instabilities, we make the simplifying assumption $s \ll 1$ and find that $\Psi=2.80$ for (ii) and $\Psi=\pi$ for (iii). The asymptotic results are confirmed by numerical computations. These reveal, in particular, that the instabilities (iii) have much smaller growth rates in cyclonic flows than in anticyclonic flows, in spite of having both $\Psi=\pi$. Our results, which extend those of [@kush-et-al] and [@yavn-et-al], highlight the limitations of the so-called balanced models, widely used in geophysical fluid dynamics, which filter out Kelvin and inertia-gravity waves and hence predict the stability of the Couette flow. They are also relevant to the stability of Taylor–Couette flows and of astrophysical accretion discs. author: - 'J. Vanneste$^1$ and I. Yavneh$^2$' bibliography: - 'mybib.bib' date: '\' title: Unbalanced instabilities of rapidly rotating stratified shear flows --- Introduction ============ Rapid rotation and strong density stratification characterise the dynamics of geophysical fluids, the atmosphere and the oceans in particular. Two dimensionless numbers are used to measure the importance of these two effects relative to nonlinear advection: the Rossby number $$\eps = \frac{U}{fL},$$ and the Froude number $$F = \frac{U}{N D}.$$ Here $U$ is a typical horizontal velocity, $f > 0$ is the Coriolis parameter, $N$ the Brunt–Väisälä frequency, and $L$ and $D$ are typical horizontal and vertical length scales. With $N>f$, as is realistically the case, the Rossby number estimates the maximum ratio between the typical frequency of the (slow) advective motion (given by $U/L$), and the frequency of inertia-gravity waves (bounded from below by $f$). Its smallness, explicity $\eps \ll 1$, has an important dynamical consequence, namely the weakness of the interaction between advective motion and inertia-gravity waves. This, together with the observation that inertia-gravity waves have generally weak amplitudes in the atmosphere and oceans, has led to development — and success — of the so-called balanced models, which filter out inertia-gravity waves completely. These models describe only the slow, large-scale dynamics, termed balanced because of its closeness to hydrostatic and geostrophic balance. They can be derived asymptotically, using power-series expansions in $\eps$, and in principle can achieve an arbitrary algebraic accuracy $O(\eps^n)$ [e.g., @warn97; @warn-et-al]. To understand balanced dynamics and its limitations more fully, it is important to identify and quantify the phenomena that balanced models fail to capture. Of particular interest are those unbalanced phenomena which occur in spite of the smallness of $\eps$ and cannot be suppressed by balancing the initial data. In the present paper we consider one such mechanism, namely the instability of balanced flows to unbalanced, gravity-wave-like perturbations. Since this type of instability is absent from balanced models of arbitrary high accuracy (which all have qualitively similar stability conditions; see @ren-shep), the growth rates can be expected to be $o(\eps^n)$ for all $n \ge 1$ or, in other words, to be beyond all orders in $\eps$, and typically exponentially small in $\eps$. Our results confirm this scaling and show that the instability bands, i.e., the range of unstable wavenumbers, are exponentially narrow. We note that unbalanced instabilities like the one examined in this paper are distinct from the mechanism of spontaneous generation of inertia-gravity waves sudied in @v-yavn04. Both mechanisms are exponentially weak, but whilst the exponentially small quantity is the growth rate for instabilities, it is the amplitude of the waves in the case of spontaneous generation. This difference may not be essential, however, if the unbalanced instabilities saturate at a level that decreases to zero with growth rate, as is typical. Another difference is the fact that the instabilities require an initial unbalanced perturbation, whilst spontaneous generation occurs from entirely balanced initial conditions. We emphasize that both mechanisms provide potential sources of inertia-gravity waves in the atmosphere and oceans. What the exponential smallness indicates in both cases is that the effectiveness of these sources is highly sensitive to the Rossby number. The specific flow whose stability we study is a horizontal Couette flow with velocity $(\Lambda y,0,0)$, modelled using the Boussinesq approximation with constant $N$, and an $f$-channel of width $2L$. See Figure \[fig:channel\] for an illustration. A natural definition of a (signed) Rossby number for this flow is the ratio $$\eps = \frac{\Lambda}{f}$$ of (minus) the basic-flow vorticity to the planetary vorticity. For $\eps > 0$ ($<0$), the shear is anticyclonic (cyclonic). The other dimensionless parameter characterising the flow can be taken to be the Prandlt ratio $$s = \frac{f}{N}.$$ We restrict our attention to $ s \le 1$ and note that in the atmosphere and oceans $s \ll 1$ generally holds. Because it is steady, the flow under consideration remains exactly balanced for all times, unlike generic time-dependent flows. Furthermore, it is stable in any balanced approximation, however accurate: this is because the shear is linear, and hence the potential vorticity constant, whilst balanced instabilities are inflectional instabilities,which require changes in the sign of the potential-vorticity gradient. Thus, with this flow, there are none of the difficulties in separating inertia-gravity waves from balanced motion that would appear for more complicated flows, and the analysis reduces to a straighforward linear stability analysis. The smallness of $\eps$ is of course exploited to derive asymptotic results. A number of authors have investigated gravity-wave-like instabilities of shear flows, although mostly in the context of two-dimensional (shallow-water or compressible-gas) models, in either parallel or cylindrical geometry [@sato81a; @sato81b; @sato82; @nara-et-al; @knes-kell; @ford94a; @balm99; @drit-v], and of isentropic models [@papa-prin and references therein]. The emphasis was not, however, on the small $\eps$ limit; indeed, in shallow water, flows with $\eps \ll 1$ and $F=O(\eps)$ are linearly stable as Ripa’s theorem indicates [@ripa83]. In contrast, the three-dimensional model examined here turns out to be always unstable, with growing modes whose horizontal and vertical wavenumbers scale like $\eps^{-1}$. Our analysis has nevertheless many common features with some of the works cited above, in particular the use of the WKB approximation. A common theme [in particular with @nara-et-al] is also the interpretation of the instabilities in terms of (linear) resonances between modes with differently signs of the conserved wave energy (or pseudoenergy) and wave momentum (or pseudomomentum) [see, e.g., @craik; @ripa90 and references therein]. In the presence of lateral boundaries, as is the case here, there are two types of unbalanced modes: inertia-gravity waves, which are oscillatory in the cross-stream direction, and Kelvin waves, which are trapped at each boundary. Instabilities involving the resonance of Kelvin waves have been studied recently by @kush-et-al for the model considered here, and by @yavn-et-al and @mole-et-al01 in the annular geometry of the (stratified) Taylor–Couette flow (see also @rudi-et-al and @dubr-et-al for astrophysical applications). For simple geometric reasons, these instabilities occur only for anticyclonic shears ($\Lambda>0$). @yavn-et-al and @mole-et-al01 also identified other modes of instability in anticyclonic shears. These can be associated with the resonance between Kelvin and inertia-gravity waves, and between inertia-gravity waves. The first mechanism is analogous to the mixed-mode instabilities examined by @saka89, @mcwi-et-al04, @mole-et-al05 and @plou-etal in a variety of contexts. As we show, the second mechanism is also active in cyclonic shears ($\Lambda<0$). Thus, we establish that the stratified horizontal Couette flow is unconditionally unstable. For all the instabilities that we study, the growth rates are exponentially small in $\eps$ because the resonant waves with differently signed wave momentum are localised exponentially in different sides of the channel. We provide both a qualitative description of the instabilities, based on the mode resonance and conservation laws, and quantitative results, based on the WKB approximation and numerical computations. The remainder of this paper is organized as follows. The linearized equations of motion governing the evolution of perturbations in the Couette flow are introduced in §\[sec:model\]. The conservation laws for the wave momentum and wave energy are also introduced there. The latter conservation law is used to show that the horizontal and vertical wavenumbers of growing perturbations must be $O(\eps^{-1})$ or larger. This motivates the WKB approach developed in §§\[sec:normalmodes\]–\[sec:inst\]. In §\[sec:normalmodes\] we formulate the eigenvalue problem for the normal modes of the system, then provide an approximate solution using a WKB expansion (§\[sec:wkb\]). To all orders in $\eps$, this leads to purely real eigenfrequencies or, in other words, to waves rather than growing modes. Instabilities with growth rates beyond all orders in $\eps$ are however possible, and we go on to show that they do occur. Focusing on the modes susceptible to be involved in instabilities, we give some details of the dispersion relation and structure of Kelvin waves (§\[sec:kelvin\]) and inertia-gravity-waves with turning points (§\[sec:igw\]). We then use arguments based on wave-momentum signature to show that the linear resonance between waves does lead to instabilities for both cyclonic and anticyclonic shears (§\[sec:momsig\]). Section \[sec:inst\] is devoted to the estimation of the instability growth rates. A detailed asymptotic estimate for Kelvin-wave instabilities, extending those of @kush-et-al and @yavn-et-al, is derived in §\[subsec:KW-KW\]. Rough estimates (focusing on the exponential dependence and ignoring order-one prefactors) are then obtained for the weaker types of instabilities (§§\[subsec:KW-IGW\]–\[subsec:IGW-IGW\]). These estimates are confirmed by the numerical solutions of the eigenvalue problem presented in §\[sec:numerics\]. The paper concludes with a Discussion in §\[sec:discussion\]. Model {#sec:model} ===== We consider small-amplitude perturbations to the Couette flow described in the Introduction and in Figure \[fig:channel\]. The corresponding linearized equations of motion can be written as $$\begin{aligned} D_t u - (f-\Lambda) v &=& - \partial_x p, {\label{eqn:lin1}} \\ D_t v + f u &=& - \partial_y p, \\ D_t w + \rho &=& - \partial_z p, \\ D_t \rho - N^2 w &=& 0, \\ \partial_x u +\partial_y v + \partial_z w &=&0 {\label{eqn:lin5}},\end{aligned}$$ where $(u,v,w)$ are the components of the velocity perturbation, $p$ is pressure perturbation, $\rho$ the buoyancy perturbation, and $D_t = \partial_t + \Lambda y \partial_x$. The material conservation $$D_t q = 0$$ of the perturbation potential vorticity $$q = (f - \Lambda) \partial_z \rho - N^2 (\partial_x v - \partial_y u)$$ follows readily. We restrict our attention to perturbations with vanishing potential vorticity, $q=0$, since this is a characteristic of unbalanced motion. (See @v-yavn04 for a study of the generation of inertia-gravity waves from perturbations with $q\not=0$.) With this restriction, the conservations of the wave energy (pseudoenergy) [\[eqn:energy\]]{} = ( + + y ) x y z and of the wave momentum (pseudomomentum) [\[eqn:moment\]]{} = x y z are readily derived, as detailed in Appendix \[app:a\]. The conservation of $\mathcal{E}$ constrains the structure of unstable perturbations. This is because exponentially growing modes must have vanishing $\mathcal{E}$ [see, e.g., @ripa90]. Completing the squares in [(\[eqn:energy\])]{}, we rewrite $\mathcal{E}$ as $$\begin{aligned} \mathcal{E} &=& \frac{1}{2} \int \! \! \int \! \! \int \left[ \left(u+\frac{\Lambda y \partial_z \rho}{N^2}\right)^2 + v^2 + \left(w - \frac{\Lambda y \partial_x \rho}{N^2} \right)^2 \right. \\ && \qquad \qquad \qquad \qquad \qquad \left. + \frac{\rho^2}{N^2} - \frac{\Lambda^2 y^2}{N^4}\left( (\partial_x \rho)^2 + (\partial_z \rho)^2 \right) \right] \, \d x \d y \d z.\end{aligned}$$ Clearly, instability can only occur if the perturbation satisfies $ \Lambda^2 y^2 \left[\partial_x \rho)^2 + (\partial_z \rho)^2 \right] > \rho^2 N^2 $ somewhere in the channel. In terms of horizontal and vertical wavenumbers $k$ and $m$, this gives the condition [\[eqn:subs1\]]{} &lt; L, which can be recognized as a subsonic condition: instability occurs only for modes whose phase speed is less than the maximum basic-flow velocity. With $s \le 1$ as assumed, the subsonic condition implies that $L (k^2 + m^2)^{1/2} \ge \eps^{-1}$, and therefore that modes involved in instabilies have asymptotically large wavenumbers. One interpretation of this result states that the Rossby number based on the wave scale, that is, $\Lambda L (k^2+m^2)^{1/2}/f$, is greater than unity for unstable modes. We note that for the shallow-water model with depth $H$, the subsonic condition analogous to [(\[eqn:subs1\])]{} is $(g H)^{1/2} < \Lambda L$ and does not involve the wavenumbers [@ripa83]. It is never satisfied for sufficiently small $\Lambda$ and thus, for order-one Burger number, for sufficiently small $\epsilon$. Thus the shallow-water analogue of our model is linearly stable in the limit $\eps \to 0$. Normal modes {#sec:normalmodes} ============ Let us now consider normal-mode solutions of the linearized equations of motion [(\[eqn:lin1\])]{}–[(\[eqn:lin5\])]{}. The subsonic condition [(\[eqn:subs1\])]{} suggests that the wavenumbers $k$ and $m$ should be rescaled by $\eps$. We therefore write the dependent variables in the form [\[eqn:normal\]]{} u(x,y,z,t) = (y/L) , with similar expressions for $v,w,p$ and $\rho$. Here $k$, $m$ and $\omega$ are dimensionless wavenumbers and frequency, with their dimensional counterparts given by $k/(\eps L)$, $m/(\eps s L)$ and $f \omega$, respectively. Without loss of generality we assume that $k>0$. Note that the non-dimensionalisation then implies that modes with $\omega > 0$ ($\omega<0$) propagate to the right (left) in anticyclonic shear and to the left (right) in cyclonic shear. In terms of the dimensionless $k$ and $m$, the subsonic condition [(\[eqn:subs1\])]{} reads [\[eqn:subs2\]]{} r = (s\^2 k\^2+m\^2)\^[1/2]{} &gt; 1. Introducing the normal modes [(\[eqn:normal\])]{} into [(\[eqn:lin1\])]{}–[(\[eqn:lin5\])]{} leads to a system of ordinary differential equations for $\hat{u}$, $\hat{v}$, $\hat{w}$, $\hat{p}$ and $\hat{\rho}$. These independent variables can be eliminated in favour of $\hat{p}$, leading in particular to [\[eqn:uwrp\]]{} = , = - w = - , where prime denotes differentiation with respect to the dimensionless variable $y/L$ which we henceforth denote simply by $y$. A second-order differential equation, already obtained by @kush-et-al, then follows. It reads [\[eqn:k1\]]{} \^2 ” - ’ - ( k\^2 + m\^2 ) = 0, where $$\hom=\omega-k y.$$ It is supplemented by the boundary conditions $\hat v = 0$, that is, [\[eqn:bc\]]{} ’ + = 0   y=1, where $\hat{c} = c - y = \omega / k - y$. Note that the singularities of [(\[eqn:k1\])]{} for $\hom^2 = 1- \eps$ are removable: in particular, they are absent from the equation for $\hat{u}$ equivalent to [(\[eqn:k1\])]{} and given in Appendix \[app:u\]. WKB approximation {#sec:wkb} ----------------- Together, [(\[eqn:k1\])]{} and [(\[eqn:bc\])]{} constitute an eigenvalue problem from which the dispersion relation giving $\omega$ as a function of $k$ and $m$ can be derived. Taking advantage of the small parameter $\eps$, this eigenvalue problem can be solved approximately using the WKB method. To this end, we first expand [(\[eqn:k1\])]{} in powers of $\eps$, with the frequency $$\omega = \omega_0 + \eps \omega_1 + \cdots$$ turning out to be real to all orders. Taking into account that $\hat{p}'=O(\eps^{-1})$ and $\hat{p}''=O(\eps^{-2})$, we rewrite [(\[eqn:k1\])]{} as [\[eqn:k2\]]{} \^2 p” - \^2 p - p’ + h p = O(\^2), where [\[eqn:lambda\]]{} \^2 = k\^2 + m\^2 . and $$h =\frac{2k^2}{\hom_0^2-1} + \frac{m^2}{1-s^2 \hom_0^2}+ 2 \omega_1 \frac{m^2 (1-s^2) \hom_0}{(1-s^2 \hom_0^2)^2}.$$ We introduce solutions of the form [\[eqn:wkb\]]{} p = (g\_ + g\_[1]{} + ) into [(\[eqn:k2\])]{} and find that $g_\pm$ satisfies [\[eqn:g’/g\]]{} = - + , where $\sigma=\mathrm{sgn} \, \eps$ equals $+1$ for an anticyclonic shear and $-1$ for a cyclonic shear. The solution can be written as [\[eqn:gpm\]]{} g\_= A ()\^[1/2]{} , where $A$ is an arbitrary complex constant. Note that this solution is single-valued near $\hom_0 = \pm 1$, consistent with the observation that the singularites of [(\[eqn:k1\])]{} for $\hom^2 = 1 - \eps$ are removable: the multi-valuedness caused by the square root factor in [(\[eqn:gpm\])]{} is cancelled by that of the integral in the argument of the exponential. We can classify the solutions [(\[eqn:wkb\])]{} according to the sign of $\lambda^2$ in the channel and distinguish: Kelvin waves (KWs) : , for which $\lambda^2 > 0$ for $-1 \le y \le 1$. These modes are trapped exponentially near one of the boundary, with $O(\eps)$ trapping scale. Inertia-gravity waves (IGWs) : , which satisfy $\lambda^2 < 0$ in at least part of the channel. There they have an oscillatory structure with $O(\eps)$ wavelength. We now derive approximate dispersion relations for both types of waves. Together with information on the signature of their wave momentum discussed in §\[sec:momsig\], these allow the prediction of instabilities associated with KW-KW, KW-IGW and IGW-IGW resonances. Asymptotic estimates for the growth rates of these instabilities are derived in §\[sec:inst\], where they are compared with numerical results. Kelvin waves {#sec:kelvin} ------------ We first consider WKB solutions to [(\[eqn:k1\])]{} for which $\lambda^2>0$. Two independent solutions can be written as $$\begin{aligned} p_- &=& \left(\frac{1-\hom_0^2}{\lambda}\right)^{1/2} \exp \left\{- |\eps|^{-1} \int_{-1}^y \left[\lambda(y')-\eps \frac{h(y')}{2 \lambda(y')}\right] \d y' \right\} \left[1+O(\eps)\right] {\label{eqn:p-}} \\ p_+ &=& \left(\frac{1-\hom_0^2}{\lambda}\right)^{1/2} \exp \left\{ - |\eps|^{-1} \int_y^1 \left[\lambda(y')-\eps \frac{h(y')}{2 \lambda(y')}\right] \d y' \right\} \left[1+O(\eps)\right]. {\label{eqn:p}}\end{aligned}$$ The dispersion relation is found from the boundary conditions in the form [\[eqn:dispfull\]]{} | [cc]{} (-1) p’\_-(-1)+ p\_-(-1) & (-1) p’\_+(-1)+p\_+(-1)\ (1) p’\_-(1)+ p\_-(1) & (1) p’\_+(1)+ p\_+(1) | = 0. Since the off-diagonal terms are exponentially small, the dispersion relation factorises to all orders into two branches corresponding to KWs trapped at each boundary. We denote by KW$_{\pm}$ the branch trapped at $y=\pm 1$, respectively; the corresponding frequency satisfies [\[eqn:dispex\]]{} (1) p’\_+(1)+ p\_+(1) = 0 (-1) p’\_-(-1)+ p\_-(-1) = 0 . At leading order in $\eps$, these two relations reduce to $$\sigma \hc_0(1) \lambda(1) + 1 = 0 \quad \mathrm{and} \quad -\sigma \hc_0(-1) \lambda(-1) + 1 =0.$$ with solutions [\[eqn:om0\]]{} \_0(1) = - \_0(-1) = . (In addition, there are spurious solutions $k \hc_0=\pm \sigma$.) Thus, the KWs localised near $y = \pm 1$ have the leading-order dispersion relation [\[eqn:c0\]]{} c\_0 = 1 - c\_0 = - 1 + . Higher-order approximations for the KW dispersion relation can be obtained by pursuing the expansion in powers of $\eps$, each leading to a purely real correction to [(\[eqn:om0\])]{}. Inertia-gravity waves {#sec:igw} --------------------- In the region where the IGW is oscillatory, two independent WKB solutions [(\[eqn:wkb\])]{} can be written as [\[eqn:pi\]]{} p = ()\^[1/2]{} { i||\^[-1]{} \^y }, where $\ell > 0$ is defined by $$\ell^2 = - \lambda^2.$$ Depending on the value of $\omega$, IGWs can have at most two turning points, i.e. points where $\lambda=\ell=0$, in the channel. These are located at [\[eqn:tpoints\]]{} y\_= c\_0 (1+\^2)\^[1/2]{}, =m/k, on either side of the ‘critical level’ $y=c_0$ where $\hom_0=0$. The mode structure is then oscillatory for $y< y_-$ and $y>y_+$, and exponential for $y_- < y < y_+$. Here, we concentrate on modes with at least one turning point since, as argued in §\[sec:momsig\] below, the presence of a turning point is necessary for instability. These IGWs are localised on one side of the channel and exponentially small on the opposite boundary. Let us consider one such IGW that is decaying exponentially with $y$ in $[y_-,y_+]$ and denote the corresponding solution by $p_-$. (Its counterpart, growing exponentially in $[y_-,y_+]$ and denoted by $p_+$, is readily deduced using the symmetry $(y,c) \mapsto (-y,-c)$.) In $[y_-,y_+]$, the solution $p_-$ can be written as [\[eqn:p–\]]{} p\_- \~A ()\^[1/2]{} {- ||\^[-1]{} \_[y\_-]{}\^y y’ }. The boundary condition [(\[eqn:bc\])]{} at $y=1$ is satisfied automatically to all orders in $\eps$. The form [(\[eqn:p–\])]{} breaks down in an $\eps^{2/3}$ neighbourhood of $y_-$, where it is replaced by an Airy function $\mathrm{Ai}$. In $[-1,y_-]$, the solution is given by a linear combination of the two solutions in [(\[eqn:pi\])]{}. The connection formula, which relates the two arbitrary constants to $A$ and is found by matching with the Airy function, gives (cf. Bender & Orszag, Eq. (10.4.16)) [\[eqn:p–2\]]{} p\_- \~2 A ()\^[1/2]{} { ||\^[-1]{} \_y\^[y\_-]{} y’ + }. The dispersion relation is then found by applying the boundary condition [(\[eqn:dispex\])]{} at $y=-1$, leading to $$-\sigma \hc(-1) \ell(-1) \cos S(y_-) + \sin S(y_-) = O(\eps),$$ where $$S(y_-)=|\eps|^{-1} \int_{-1}^{y_-} \left[ \ell(y') + \eps \frac{h(y')}{2 \ell(y')} \right] \, \d y' + \frac{\pi}{4}.$$ Solving for $S(y_-)$, we find $$S(y_-) = n \pi + \tan^{-1} [\sigma \hc(-1) \ell(-1)] + O(\eps)$$ where $n$ is an integer. At leading order this gives [\[eqn:igw0p\]]{} \_[-1]{}\^[y\_-]{} (y) y = n ||, which determines $c_0$ implicitly. The next order relation determines $c_1$. Let us write the dispersion relation [(\[eqn:igw0p\])]{} for $c_0$ in a convenient form. Define $\mu$ by $$c_0= -1 + \frac{(1+\delta^2)^{1/2}}{r} (\mu+1)$$ where $\delta=m/k$, so that $$y_-=-1 + \frac{(1+\delta^2)^{1/2}}{r} \mu.$$ The assumption that this turning point is inside the channel imposes the restriction $0<\mu<2 r (1+\delta^2)^{-1/2}$. Introducing the integration variable $Y$, with $y=-1+(1+\delta^2)^{1/2} Y/r$, reduces [(\[eqn:igw0p\])]{} to the expression [\[eqn:mu\]]{} \_0\^\^[1/2]{} Y = n ||, where $$\nu^2 = \frac{s^2(1+\delta^2)}{s^2+\delta^2}.$$ This defines implicitly a function $\mu(\delta,s,n|\eps|)$ with values in $[0,\nu^{-1}-1]$, from which $c_0$ is deduced. Taking both the solution $p_-$ and its symmetric $p_+$ into account, we find the two branches [\[eqn:igw0\]]{} c\_0 = 1 \[(,s,n ||)+1\], corresponding to modes exponentially small near $y=\mp 1$ and denoted by IGW$_\pm$, respectively. Again, higher-order approximations to the phase velocity can in principle be computed, leading to real corrections to $c_0$ in powers of $\eps$. Note that, at leading order in $\eps$, the dispersion relation is the same for both signs of $\eps$, that is, for both cyclonic and anticyclonic flows. An asymmetry only appears at higher order. For $n=O(1)$, $\mu \to 0$ as $\eps \to 0$, and the leading-order dispersion relation reduces to [\[eqn:igws\]]{} c\_0 \~1 , corresponding to $y_\pm \to \pm 1$. The small-$\mu$ behaviour of the left-hand side of [(\[eqn:mu\])]{} then suggests that the successive branches $n=1,\, 2, \cdots$ are $O(\eps^{2/3})$ apart. The asymptotic results [(\[eqn:c0\])]{} and [(\[eqn:igw0\])]{} provide a first approximation to the dispersion relation of KWs and IGWs. We have extended this by solving the eigenvalue problem [(\[eqn:k1\])]{}–[(\[eqn:bc\])]{} (or rather the equivalent formulation [(\[eqn:u\])]{}–[(\[eqn:bcu\])]{} in terms of $\hat{u}$) numerically. Our numerical solver is the same as the one used in @yavn-et-al, employing a second-order finite-volume discretization of [(\[eqn:u\])]{}–[(\[eqn:bcu\])]{}. For given physical parameters and wavenumbers, $m$ and $k$, we search for eigenfrequencies for which the matrix representing the discretized system is singular. The codes are implemented in MATLAB, with the search performed using the [*fminsearch*]{} function that employs the so-called Simplex algorithm. Dispersion relation {#subsec:disprel} ------------------- Figures \[fig:dispa\] and \[fig:dispc\] show the dispersion relation for anticyclonic and cyclonic flows, respectively. The parameters have been chosen as $\eps=0.1$, $\delta=2$ and $s=0.1$, but the qualitative features remain the same for a wide range of values. The numerical results (dotted curves) are compared with the asymptotic estimates (solid curves) to confirm the validity of the latter. For KWs, we have used an $O(\eps)$-accurate estimate which improves on [(\[eqn:c0\])]{} by adding the term $\eps c_1 = \mp \eps \sigma (2 r)$ derived in Appendix \[app:kwinst\]. For IGWs, we have used the estimate [(\[eqn:igw0\])]{}, corrected in the anticyclonic case by subtracting $\eps/2$ from the square bracket. This correction, which can be viewed as an experimentally determined $O(\eps)$ term in the expansion of $c$, is made for the clarity of the plot: without it, the $O(\eps)$ error in the dispersion relation is not significantly smaller than the $O(\eps^{2/3})$ distance between branches, and it is difficult to relate each asymptotic curve to its numerical counterpart. No corrections were necessary for the cyclonic shear, suggesting the dispersion relation [(\[eqn:igw0\])]{} is already $O(\eps)$-accurate in this case. To confirm this would require to continue the asymptotic developments to the next order in $\eps$; this is a daunting task which we have not attempted for IGWs. Figures \[fig:dispa\] and \[fig:dispc\] demonstrate the multiple intersections between the branches IGW$_{\pm}$ of the dispersion relation. In the anticyclonic case, there are additional intersections between the branches KW$_{\pm}$, and between KW$_\pm$ and IGW$_{\mp}$. (The KW do not appear in Figure \[fig:dispc\] for the cyclonic case because they have $|c|>1$.) The intersections, associated with the linear resonance between modes, are generically spurious: they result from the finite resolution of the plot for the numerical results, and from the limited accuracy for the asymptotic ones. There are in fact two possible behaviours: (i) mode conversion, when the phase velocities remaining real and the two curves, rather than intersecting, locally form the two branches of a hyperbola, or (ii) instability, when the phase velocities on the two branches become complex conjugate with non-zero imaginary parts. The two situations are distinguished by the signs of quadratic invariants, such as the wave momentum, along the colliding branches: (i) mode conversion occurs when both signs are the same, and (ii) instability occurs when the signs differ [e.g. @cair79]. We now show that the latter situation is the relevant one in our problem by examining the sign of the wave momentum for KWs and IGWs in the WKB approximation. Wave-momentum signature {#sec:momsig} ----------------------- IGWs and KWs have different leading-order approximations to their wave momentum. To see this, we introduce [(\[eqn:normal\])]{} and [(\[eqn:uwrp\])]{} into [(\[eqn:moment\])]{} and assume that $\omega$ is real. This gives $$\begin{aligned} \mathcal{M} &=& \frac{2N^2 m^2}{f^2 L^2 \eps^3} \int \left[ \frac{\eps(1-\eps)}{2(\hom^2-1+\eps)(1-s^2 \hom^2)} {\frac{\mathrm{d} |\hat p|^2}{\mathrm{d} y}} + \frac{k \hom (1-s^2+s^2 \eps)}{(\hom^2-1+\eps)(1-s^2 \hom^2)^2} |\hat p|^2 \right] \, \d y \nonumber \\ &=& \frac{2N^2 m^2}{f^2 L^2 \eps^3} \hat\mathcal{M},{\label{eqn:hm}}\end{aligned}$$ where the last line defines the dimensionless wave momentum $\hat{\mathcal{M}}$ which we will use henceforth. For IGWs, the first term is negligible: indeed, in the regions where $p$ oscillates rapidly, $\d |\hat p|^2 / \d y = O(1)$, while in the possible regions where $p$ decays exponentially, $\d |\hat p|^2 / \d y = O(\eps^{-1})$ only for a range of $y$ of size $\eps$; both types of regions thus contribute at $O(\eps)$ to $\hat{\mathcal{M}}$. This leads to the leading-order approximation [\[eqn:migw\]]{} \~ |p|\^2 y . Given that the denominator $(\hom^2-1)$ cancels with the same factor in $|\hat p|^2$ (see [(\[eqn:wkb\])]{}–[(\[eqn:gpm\])]{}), it is clear that instability involving IGWs implies that $\hom$ changes sign. It follows that there is at least one turning point in the channel, as announced, since the absence of turning points ($\ell^2>0$) implies that $|c|>1$. Assuming there are turning points, the sign of $\hat \mathcal{M}$ for the two types of IGWs considered in §\[sec:igw\] is then $$\hat \mathcal{M} < 0 \ \ \textrm{for IGW}_+ {\quad \textrm{and} \quad} \hat \mathcal{M} > 0 \ \ \textrm{for IGW}_-.$$ For KWs, the two terms in [(\[eqn:hm\])]{} have a similar, $O(\eps)$, order of magnitude. Using [(\[eqn:wkb\])]{}, we find that $$\hat{\mathcal{M}}_\pm \sim \frac{|\eps|}{2[\hom^2(\pm1)-1][1-s^2\hom^2(\pm1)]} \left[\pm \sigma + \frac{k(1-s^2)}{\lambda(\pm1) [1-s^2\hom^2(\pm1)]} \right] |\hat{p}(\pm1)|^2.$$ Using the dispersion relation for Kelvin waves, $\hom(\pm 1) = \mp \sigma k /r + O(\eps)$ in our non-dimensionalisation, and its consequence $\lambda(\pm 1) = r$ (see [(\[eqn:om0\])]{} and [(\[eqn:lampm1\])]{}), this reduces to $$\hat{\mathcal{M}}_\pm \sim \mp \frac{\eps r^4}{2 m^4} |\hat{p}(\pm1)|^2 \quad \textrm{for KWs},$$ leading to the following signs: $$\hat \mathcal{M} \lessgtr 0 \ \ \textrm{for KW}_+ {\quad \textrm{and} \quad} \hat \mathcal{M} \gtrless 0 \ \ \textrm{for KW}_- \quad \textrm{when} \ \ \eps \gtrless 0,$$ differing in the anticyclonic and cyclonic cases. With the wave-momentum signatures just obtained, it is clear from Figures \[fig:dispa\] and \[fig:dispc\] that the numerous intersections of branches correspond to waves with oppositely signed $\hat{\mathcal{M}}$. This establishes the existence of many modes of instability, both for anticyclonic and cyclonic shears. The main difference between the differently signed shears is that instabilities involving KWs only are possible only for anticyclonic shear. All the instabilities are associated with the interactions of modes exponentially localised on different sides of the channel. Therefore their interaction is exponentially weak and, as a consequence, the growth rates of the instabilities and range of unstable wavenumbers are exponentially small in $\eps$, as anticipated in the Introduction. As the asymptotic calculations of the next section show, such small growth rates are somewhat delicate to capture analytically. However, the interpretation in terms of interactions of waves with oppositely signed $\hat{\mathcal{M}}$ makes it possible to predict instability robustly, without detailed calculations. Instabilities {#sec:inst} ============= KW-KW instabilities {#subsec:KW-KW} ------------------- We start our study of the weak instabilities associated with mode interactions by deriving an estimate for the growth rate of the instability that arises through the resonance of KWs in anticyclonic shear. This instability has been examined in some detail by @kush-et-al and by @yavn-et-al. Because it is the strongest instability, with physical relevance in Taylor–Couette and accretion discs [see @dubr-et-al], we present here a complete asymptotic derivation of the growth rate. For the KW-IGW and IGW-IGW instabilities considered in §§\[subsec:KW-IGW\]–\[subsec:IGW-IGW\], we limit the derivation to the exponential behaviour of the growth rate as $\eps \to 0$. The method we now describe could however be applied to these instabilities as well, should a more accurate estimate be needed. To obtain the growth of the instability, we need to reconsider the dispersion relation [(\[eqn:dispfull\])]{} in the vicinity of the resonance point, taking into account exponentially small terms. Let $c_\star$ and $r_\star$ be the values of $r$ and $c$ at resonance. By symmetry, $c_\star = 0$. According to [(\[eqn:c0\])]{} (with $\sigma=1$ corresponding to the anticyclonic shear), $$r = r_\star = 1 + O(\eps).$$ Thus, resonance occurs on an ellipse with semi-axes $1/s$ and $1$ in the $(k,m)$-plane, and the instability region is an exponentially small annulus around this ellipse. It is best parameterized using the polar coordinates $(r,\theta)$, with $$s k = r \cos \theta \quad \textrm{and} \quad m = r \sin \theta.$$ Now, take $$c = C \quad \textrm{and} \quad r = r_\star + R,$$ where $C$ and $R$ are exponentially small. This can be introduced into the dispersion relation [(\[eqn:dispfull\])]{}; using the fact that $(c=0,r=r_\star)$ satisfy [(\[eqn:dispex\])]{}, a Taylor expansion leaves only terms that are exponentially small. In the coefficients of $C$ and $R$ in these terms, we can approximate $r_\star$ by its leading-order estimate $1$. Noting that, in this approximation, $$\lambda(\pm 1) \approx \lambda_\star(\pm 1) + \frac{(s^2 - \cos^2 \theta) R \pm \cos^2 \theta (1-s^2) C}{s^2 \sin^2 \theta},$$ we find the dispersion relation in the form [\[eqn:diskwkw1\]]{} ()\^2 (R\^2 - C\^2) = 4 \^[-2 /]{}, where $$\Psi = \int_{-1}^1 \left[\lambda_\star(y)-\eps \frac{h_\star(y)}{2\lambda_\star(y)}\right] \, \d y,$$ and the subscript $\star$ indicates evaluation at the resonance point. A consistent approximation of $\Psi$ requires to include the $O(\eps)$ contribution to $\lambda_\star$ in the first term of the integrand. To this end, we compute the KW dispersion relation to $O(\eps)$ in Appendix \[app:kwinst\] and find that $r_\star=1 + \eps /2 + O(\eps^2)$. This leads to $$\Psi = \Psi_0 + \eps \Psi_1,$$ where [\[eqn:psi0\]]{} \_0= \_[-1]{}\^1 \_0(y) y with $$\lambda_0 = \left[\frac{\cos^2 \theta (1-y^2)+s^2 \sin^2 \theta}{s^2(1-\cos^2 \theta y^2)} \right]^{1/2},$$ and [\[eqn:psi1\]]{} \_1= \_[-1]{}\^1 y, with $$h_0(y)= \frac{2 \cos^2 \theta}{\cos^2 \theta y^2 - s^2} + \frac{\sin^2 \theta}{1-\cos^2 \theta y^2}.$$ The second integral has to be interpreted as a Cauchy principal value at the singularities $y = \pm s/\cos \theta$ of $h_0(y)$ when these are in $[-1,1]$. With this result, the dispersion relation [(\[eqn:diskwkw1\])]{} can be rewritten as [\[eqn:kwgr\]]{} C = \^[1/2]{}, where $$\alpha = \frac{2 s^2 \sin^2 \theta \, \e^{- \Psi_1}}{s^2-\cos^2 \theta}.$$ Formula [(\[eqn:kwgr\])]{} is the first main result of this paper. It provides the leading-order asymptotics for the growth rate of the KW-KW instability (after multiplication by $k$) as $\eps \to 0$ and for arbitrary $s \le 1$). It also makes evident the exponential smallness of the growth rate and of the instability-band width. Its validity is confirmed in §\[sec:numerics\] where it is compared with numerical results. The minimum of $\Psi_0$, and hence the maximum growth rate, is attained for $\theta = \pi/2$, for which $\Psi_0 \sim 2$. Thus, at the crude level of exponential dependence on $\eps$, we obtain the estimate [\[eqn:kwkwro\]]{} \~-,   0, for the largest growth rate $\Im \omega=k \Im C$. Note that because $\theta = \pi/2$ implies that $k=0$ and hence $\omega=0$, the maximum growth rate is in fact achieved for $\theta$ slightly less than $\pi/2$; this does not affect the exponential dependence in [(\[eqn:kwkwro\])]{}, however (see below). Estimates more precise than [(\[eqn:kwkwro\])]{} can of course be inferred from [(\[eqn:kwgr\])]{}. Focusing on the limit $\theta \to \pi/2$, we note that $C$ depends on the relationship between $s$ and $\theta$. A distinguished limit is found for $s = O(\cos \theta) \ll 1$. This corresponds to the regime with $s \ll 1$ and $\delta=k/m=O(1)$, which we term the quasi-geostrophic regime, since it corresponds to the quasi-geostrophic scaling implying, in particular, the hydrostatic approximation ($k/m$ can be recognized as the square root of the Burger number based on the wave scale). Taking the limit $\theta \to \pi/2$ of [(\[eqn:psi0\])]{}–[(\[eqn:kwgr\])]{} with $k = s/\cos \theta$ fixed then yields $$\Psi_0 \sim 1 + \frac{1+k^2}{k} \tan^{-1} k {\quad \textrm{and} \quad} \alpha \sim 2 \frac{|1-k|^{k-1}}{|1+k|^{k+1}}. $$ The maximum of the imaginary part of the phase speed is then obtained for $k \to 0$ and given by $\Im c \sim 2 \exp(-2/\eps)$, consistent with @yavn-et-al’s equation (35). The maximum of the growth rate $\Im \omega$ is easily seen to be attained for $k=O(\eps^{1/2})$ and to be a factor $\eps^{1/2}$ smaller than the maximum of $\Im c$. In dimensional terms, this means that the horizontal and vertical scales are both large, but have different orders of magnitudes, scaling like $\eps^{-1/2}$ and $\eps^{-1}$, respectively. KW-IGW instabilities {#subsec:KW-IGW} -------------------- The KW-IGW instabilities occur for anticyclonic flows through the resonance of an IGW, which has one turning point and is localised on one side of the channel, with a KW localised on the other side. To estimate their growth rates, we can consider a solution consisting of a linear combination of the IGW$_-$ given by [(\[eqn:p–\])]{}–[(\[eqn:p–2\])]{} which is oscillatory near $y=-1$, and the KW$_+$ given by [(\[eqn:p\])]{}. (The other combination, of IWG$_+$ with KW$_-$, has the same growth rate, by symmetry.) A calculation similar to that carried out for KW-KW instabilities could in principle be performed to obtain the leading-order behaviour of the growth rate. However, this requires the derivation of the IGW dispersion relation accurate to $O(\eps)$ involving an inordinate amount of calculation. We shall therefore limit ourselves to the determination of the exponential behaviour of $\Im \omega$ (that is, to the determination of the constant $\Psi_0$ such that $\log \Im \omega \sim -\Psi_0/\eps$ as $\eps \to 0$) in the instability regions, and ignore the order-one prefactor in the expression of $\Im \omega$. As in the case of KW-KW instabilities, $\Psi_0$ is determined simply from the amplitude of the colliding modes at the boundary where they are exponentially small, given explicitly by $\exp(-\Psi_0/\eps)$. Note that $\Psi_0$ controls not only the exponential smallness of the growth rate but also that of the width of the instability bands. For simplicity we restrict our analysis to the quasi-geostrophic scaling $s\ll 1$, $\delta=O(1)$. For $s \ll 1$ and $\sigma=1$, the phase speeds of colliding KW$_+$ and IGW$_-$ branches given in [(\[eqn:c0\])]{} and [(\[eqn:igw0\])]{} reduce at leading order to $$c = 1 - \frac{1}{m} {\quad \textrm{and} \quad} c = -1 + \left(\frac{1}{k^2} + \frac{1}{m^2} \right)^{1/2},$$ respectively. The corresponding resonance condition $$\frac{1}{m} + \left(\frac{1}{k^2} + \frac{1}{m^2} \right)^{1/2} = 2,$$ that is, [\[eqn:kwigwres\]]{} k = ( )\^[1/2]{},   m&gt;1 defines a curve in the $(k,m)$ plane in the vicinity of which instabilities are concentrated. For KW-IGW instabilities, since there is a single turning point $y_-$ in the channel, $\Psi_0$ is given as [\[eqn:kwigwpsi\]]{} \_0 = \^1\_[y\_-]{} (y) y. The integrand $\lambda(y)$, given in [(\[eqn:lambda\])]{}, can be approximated by [\[eqn:kwigwlam\]]{} (y) = km \^[1/2]{}, with $y_\pm$ reducing to [\[eqn:kwigwtp\]]{} y\_= c ( + )\^[1/2]{} = { [l]{} 3 -2/m\ -1 . . Introducing [(\[eqn:kwigwres\])]{} and [(\[eqn:kwigwlam\])]{}–[(\[eqn:kwigwtp\])]{} into [(\[eqn:kwigwpsi\])]{} gives the expression $$\Psi_0 = \frac{1}{8[m(m-1)]^{1/2}} \left[(2m-1)^2\left( \pi + 2 \sin^{-1} \frac{1}{2m-1}\right) +4(m(m-1))^{1/2} \right].$$ The maximum growth rate of the KW-IGW instability is given by the minimum value of $\Psi_0$, found to be [\[eqn:kwigwkm\]]{} \_0 = 2.80   k = 1.04    m=1.30. Thus we obtain the asymptotics [\[eqn:kwigwro\]]{} \~-,   0, for the growth rate of KW-IGW instabilities. Comparison with [(\[eqn:kwkwro\])]{} then indicates that these are considerably weaker than the KW-KW instabilities. IGW-IGW instabilities {#subsec:IGW-IGW} --------------------- We now consider the instabilities that result from the resonance between IGWs. These are particularly important for cyclonic flows since they provide the only mode of instability in this case. In fact, as can be expected from the leading-order dispersion relation [(\[eqn:igw0\])]{}, the dominant behaviour of these instabilities is unaffected by rotation, so that the exponential dependence on $1/\eps$ is identical for anticyclonic and cyclonic shears. What differs between the two cases, however, is the order-one prefactor which we do not estimate analytically. IGW-IGW instabilities occur when a solution $p_-$ of the form [(\[eqn:p–\])]{}–[(\[eqn:p–2\])]{} is resonant with its counterpart $p_+$. The modes have then two turning points $y_\pm$ in the channel, leading to the necessary condition $r \ge (1+\delta^2)^{1/2}$ for the instability. We now estimate the factor $\Psi_0$ controlling the exponential smallness of the instability growth rates. As in the previous section, we restrict our attention to the quasi-gesotrophic scaling $s \ll 1$ and $\delta=O(1)$. We furthermore consider only the strongest IGW, associated with the (symmetric) resonance of the gravest ($n=1$) IGW modes, and for which $c=0$ to all orders in $\eps$. The resonance condition is therefore $$\frac{1}{k^2} + \frac{1}{m^2} =1.$$ Since for $n=O(1)$, the two turning points are $y_\pm = \pm 1$ at leading order in $\eps$, $\Psi_0$ is computed as $$\Psi_0 = k m \int_{y_-}^{y_+} [(y_+ - y)(y-y_-)]^{1/2} \, \d y = \frac{\pi km}{2}.$$ The minimum value is therefore [\[eqn:igwigwkm\]]{} \_0 =   k=m=, and the exponential scaling of the growth rate given by [\[eqn:igwigwro\]]{} \~- ,   0, for both anticyclonic and cyclonic flows. This is exponentially smaller than the growth rate for either the KW-KW or the KW-IGW instabilities [(\[eqn:kwkwro\])]{} or [(\[eqn:kwigwro\])]{}. Numerical computation of growth rates {#sec:numerics} ------------------------------------- We now present comparisons of the growth rate, or rather $\Im c$, computed numerically with the asymptotic results of §§\[subsec:KW-KW\]–\[subsec:IGW-IGW\]. The numerical method employed is that described in §\[subsec:disprel\] where $\Re c$ was considered. For the small values of $\eps$ examined here, $\Im c$ is very small and the bands of unstable wavenumbers are very narrow, so that very fine resolution in $y$ is needed to capture $\Im c$ accurately. In order to ensure high accuracy, we successively double the grid resolution until results are unchanged to at least four significant digits. This required grids of sizes ranging from about 250 mesh points for strong or moderate instabilities, to as many as 16000 mesh points for very weak instabilities. This may be improved upon by using nonuniform grids with high resolution only in regions where the solution changes fast. The search for the bands of instabilities in $(k,m)$ is quite delicate, but made possible by the excelllent approximations afforded by the asymptotic results. We start by considering the KW-KW instability of anticyclonic flows. Figure \[fig:profiles\] shows ${\Im} c$ as a function of $r$ for $\theta = \pi/2$ and $\epsilon = 0.3, 0.4, 0.5$ in the instability bands. The dots represent numerically computed values; the solid line are computed analytically using (\[eqn:kwgr\]). Note that we only know $r_\star$ to algebraic accuracy, while the bands are exponentially narrow. Hence, we use the numerical results for determining $r_\star$—the value of $r$ for which ${\Im} c$ is maximized. The narrowing of the instability band is clearly exhibited in the figure, and the small-$\epsilon$ analytical approximation quickly converges to the numerical results as $\epsilon$ becomes small. The dependence of $\Im c$ on $\theta$ is illustrated by Figure \[fig:kwkwtheta\] which compares numerical and asymptotic estimates for the maximum value of $\Im c$ as a function of $\theta$ for $s=0.1$ and $\eps=0.3, \, 0.4$ and $0.5$. The value of $\Im c$ in the quasi-geostrophic scaling $s \ll 1$, $\delta=O(1)$, that is, the limit $\theta \to \pi/2$, is also indicated. The Figure confirms the accuracy of the asymptotic estimate and shows the rapid decrease of $\Im c$ as $\theta$ decrases from $\pi/2$. Our results for all the types of instabilities are summarized by Figure \[fig:comparison2analyt\]. This compares asymptotic and numerically computed values of ${\Im} c$ as a function of $1/|\epsilon|$ for KW-KW, KW-IGW and IGW-IGW in anticyclonic flows, and IGW-IGW instabilities in cyclonic flows. The values of $\Im c$ displayed correspond to the maximum over $m$ and $k$ for fixed $s=0.1$. For KW-KW instabilities, the asymptotic estimates are obtained from [(\[eqn:kwgr\])]{}. For KW-GW and IGW-IGW instabilities, we use [(\[eqn:kwigwro\])]{} and [(\[eqn:igwigwro\])]{}, respectively. These give $\Im c$ only up to a multiplicative constant which we fix by matching the asymptotic and numerical results for the smallest values of $|\epsilon|$ shown in the figure. In the linear-logarithmic coordinates used, the numerical points line up with the predicted straight lines for larger $|\eps|$, thus confirming the validity of the asymptotic analysis. Further support is provided by the fact that the values of $k$ and $m$ for which $\Im c$ is maximised are close to the estimates [(\[eqn:kwigwkm\])]{} and [(\[eqn:igwigwkm\])]{}. Evidently, the match between the numerical and analytical results is quite good even for $\eps$ moderately small. We see that the instabilities become substantial for $\epsilon \approx 1$, especially KW-KW instabilities. Observe that, as predicted by the analysis, the decay of the growth rate in IGW-IGW instability as $\epsilon$ becomes small is the same for cyclonic and anticyclonic flows, and yet the growth rates of cyclonic flow are smaller by a factor of about 20. Thus the $O(1)$ prefactor in the asymptotics of $\Im c$ for IGW-IGW instabilities, ignored in [(\[eqn:igwigwro\])]{}, turns out to be numerically very different for anticyclonic and cyclonic flows. The smallness of this prefactor in the cyclonic case means that the instability remains exceedingly weak even for $\eps \approx 1$, and likely irrelevant in many physical situations. Discussion {#sec:discussion} ========== This paper examines the linear stability of a horizontal Couette flow of a rapidly rotating, strongly stratified, inviscid fluid. The main conclusion is that the flow is unconditionally unstable: unbalanced instabilities, associated with linear resonances between Kelvin and inertia-gravity waves, occur for arbitrarily small Rossby numbers $\eps=\Lambda/f$. The growing perturbations have small horizontal and vertical scales, with typical wavenumbers or spatial-decay rates of the order of $\eps^{-1}$. Physically, it is easy to understand why asymptotically small scales are a key ingredient of the instabilities. The phase locking between different waves which underlies the instability mechanisms requires the wave phase speed to be comparable to the basic flow velocity, and this only occurs for small-scale waves. The need for small vertical scales also explains why the instabilities examined in this paper have no direct counterparts in shallow-water flows; these are stable for small enough $|\eps|$ because of the inherent limitation in vertical structure imposed by the shallow-water approximation. Our conclusion that the rotating stratified Couette flow is always unstable is of course in sharp contrast with the one that may be drawn from balanced models. Regardless of their accuracy, which can be any power $\eps^n$, they predict the stability of flows without inflection points such as the Couette flow. There is no contradiction, however, since the growth rates found for the unbalanced instabilities are exponentially small in $|\eps|$. In practice, this exponential dependence means that the instabilities are exceedingly weak when $|\eps|$ is small, but can become important rather suddenly as $|\eps|$ increases towards $1$ and beyond. If the instabilities are to play a significant role in the breakdown of balance in geophysical flows, this will therefore be in a manner that is extremely sensitive to the Rossby number. In the literature, most attention has been paid to anticyclonic flows, and in particular to the coupled Kelvin-wave instability occuring in these flows. Our results clarify that cyclonic flows are also unstable, through an instability mechanism involving coupled inertia-gravity waves. This mechanism is also active in anticyclonic flows where, along with the instability mode mixing Kelvin and inertia-gravity waves, it provides an alternative to the well studied instability due to Kelvin-wave resonance [see @yavn-et-al; @mole-et-al01]. The focus on anticyclonic flows and Kelvin-wave instabilities is justified in practice by the fact that the associated growth rate is much larger than those of the other instability mechanisms, exponentially larger in fact in the limit $\eps \to 0$. The instability of the cyclonic flows is especially weak. This weakness is not completely accounted for by the exponential dependence on $1/\eps$, since this is the same for both anticyclonic and cyclonic flows whilst the growth rates obtained numerically are very different. We conclude, then, that the exponential dependence and the $O(1)$ prefactor conspire to make the instability of cyclonic flows extremely weak, even for moderate $|\eps|$. The WKB approach used in this paper could be extended to examine the instability in more general rotating stratified shear flows. Obvious applications are the stratified Taylor–Couette flow [@yavn-et-al; @mole-et-al01], which differs from the problem studied here by the presence of curvature terms, and the stability of accretion discs [@rudi-et-al; @dubr-et-al]. Additional physical effects that it would be of interest to study include different boundary conditions (in particular the case of infinite domains for which no Kelvin waves exist), viscous and thermal damping, and non-zero potential-vorticity gradients, leading to the existence of critical levels for neutral modes [cf. @balm99]. JV was funded by a NERC Advanced Research Fellowship. Conservation laws {#app:a} ================= Let $$M = (u \partial_z \rho - w \partial_x \rho)/N^2.$$ Denoting integration over the periodic domain in $x$ and $z$ by $${\langle \cdot \rangle} = \int \! \! \int \cdot \, \, \d x \d z,$$ we compute $$\begin{aligned} {\label{eqn:flux}} N^2 \partial_t {\langle M \rangle} &=& {\langle \partial_t u \partial_z \rho - \partial_z u \partial_t \rho - \partial_t w \partial_x \rho + \partial_x w \partial_t \rho \rangle} \nonumber \\ &=& {\langle (f - \Lambda) v \partial_z \rho - \partial_x p \partial_z \rho - N^2 \partial_z u w + \partial_z p \partial_x \rho \rangle} \\ &=& - N^2 \partial_y {\langle uv \rangle}, \nonumber\end{aligned}$$ where we have used integration by parts and periodicity extensively, and, for the last line, $q=0$ and the incompressibility equation. The conservation for the quadratic wave momentum (or pseudomomentum) $$\mathcal{M} = \int \! \! \int \! \! \int ( u \partial_z \rho - w \partial_x \rho ) \, \d x \d y \d z /N^2$$ follows by integration in $y$, using the boundary conditions $v=0$. The perturbation energy $\mathcal{E}'$, with density $|\mathbf{u}|^2/2 + \rho^2/(2N^2)$, is not conserved but satisfies $${\frac{\mathrm{d} \mathcal{E}'}{\mathrm{d} t}} = - \int \! \! \int \! \! \int \Lambda u v \, \d x \d y \d z.$$ Integrating by parts the right-hand side and using [(\[eqn:flux\])]{} gives a conservation law for the wave energy (or pseudoenergy) $$\mathcal{E} = \int \! \! \int \! \! \int \left(\frac{|\mathbf{u}|^2}{2} + \frac{\rho^2}{2 N^2} + \Lambda y \frac{u \rho_z -w \rho_x }{N^2} \right) \, \d x \d y \d z.$$ Note that the conservation of both $\mathcal{M}$ and $\mathcal{E}$ can also be derived from the exact conservation laws for momentum, energy and potential vorticity for the full system, that is, basic flow plus perturbation. Equation for $\hat{u}$ {#app:u} ====================== In §2, the eigenvalue problem satisfied by normal-mode solutions is formulated as the second-order differential equation [(\[eqn:k1\])]{} for $\hat p$ and its associated boundary condition [(\[eqn:bc\])]{} [cf. @kush-et-al]. An alternative formulation, employed by @yavn-et-al, uses $\hat u$ instead of $\hat p$ as the dependent variables. It has the advantage that the removable singularities that appear in [(\[eqn:k1\])]{} are absent. For completeness, we record this alternative formulation as [\[eqn:u\]]{} \^2 ( ’ )’ - ( + ) u = 0, where $$K = (1-s^2 \hom^2) k^2 + (1-\eps)^2 m^2.$$ The associated boundary conditions are [\[eqn:bcu\]]{} ’ + u = 0   y=1. This is the formulation used for the numerical computation of the normal modes. Kelvin-wave dispersion relation {#app:kwinst} =============================== In this Appendix, we derive the dispersion relation for KWs accurate to $O(\eps)$, as is necessary to obtain the leading-order asymptotics of the KW-instability growth rate. The dispersion relation for KW$_\pm$ valid to all orders in $\eps$ are given in [(\[eqn:dispex\])]{}. It is solved at leading order in §\[sec:kelvin\] to give [(\[eqn:c0\])]{}. At the next order, we find the two equations [\[eqn:disp1\]]{} c\_1 (1) g\_( 1) + \_0(1) g\_’(1) = 0, which allow the determination of the $O(\eps)$ contribtion to the frequency $\omega_1$. Note that the contributions of the $O(\eps)$ terms neglected in [(\[eqn:p-\])]{}–[(\[eqn:p\])]{} cancel in these two equations when [(\[eqn:om0\])]{} is taken into account. Equation [(\[eqn:g’/g\])]{} can be used to express the derivatives of $g_{\pm}$; the following results are therefore useful: $$\begin{aligned} \lambda(\pm 1)&=& r, {\label{eqn:lampm1}} \\ -\frac{\lambda'(\pm 1)}{2\lambda} &=& \frac{\pm \sigma (1-s^2)k^2 r}{m^2}, \nonumber \\ \frac{k \hom_0{(\pm1)}}{1-\hom_0^2(\pm1)} &=& \frac{\mp \sigma k^2 r}{r^2-k^2}, \nonumber \\ \frac{\mp \sigma h(\pm1)}{2 \lambda(\pm1)}&=& \mp \sigma r \left(\frac{1}{2}-\frac{k^2}{r^2-k^2}\right) + c_1 \frac{(1-s^2)k^2 r^2}{m^2}. \nonumber\end{aligned}$$ Using these and [(\[eqn:g’/g\])]{}, [(\[eqn:disp1\])]{} gives the first-order correction to the frequencies [(\[eqn:om0\])]{}, [\[eqn:c1\]]{} c\_1 = .
{ "pile_set_name": "ArXiv" }
--- abstract: 'We present a numerical scheme for calculating the first quantum corrections to the properties of static solitons. The technique is applicable to solitons of arbitrary shape, and may be used in 3+1 dimensions for multiskyrmions or other complicated solitons. We report on a test computation in 1+1 dimensions, where we accurately reproduce the analytical result with minimal numerical effort.' address: | ${}^*$Joseph Henry Laboratory, Princeton University, Princeton, NJ, USA 08540\ ${}^\dagger$DAMTP, Silver St,Cambridge, CB3 9EW, U.K. author: - 'Chris Barnes[^1], Neil Turok[^2]' title: A Technique for Calculating Quantum Corrections to Solitons --- ${\left(} \def$[)]{} $${\left[} \def$$[\]]{} §[[S]{}]{} Introduction ============ The quantum fate of static solitons has long been studied, and beautiful results have emerged in exactly solvable two dimensional field theories. In higher dimensions, very little is known except in certain very special supersymmetric theories, or in situations of spherical symmetry. In many theories, the quantum properties of the soliton-bearing sector are interesting, but beyond the reach of current calculations. One example of such a theory is the chiral model of nuclear physics, the solitonic sector of which describes nucleons and nuclei[@holst]. It is clear that quantum effects must be included if the theory is to match real nuclei, but hitherto only a very limited ‘collective coordinate’ quantization has been possible. There are many other 3+1 dimensional field theories with topological solitons, for example those bearing magnetic monopoles and vortex strings. It would be very useful to develop techniques to study quantum corrections to these solitons in both supersymmetric and non-supersymmetric contexts. The quantum corrections to a soliton’s classical properties may be expressed as an expansion in a dimensionless parameter, some power of the coupling constant, multiplied by Planck’s constant $\hbar$. At weak coupling, the corrections are small and may be calculated perturbatively in this small parameter. In this paper we outline a straightforward method for numerically computing the first quantum corrections, including all fluctuations about a static soliton, to first order in $\hbar$. In particular, we present a method for calculating the quantum correction to the soliton mass, although the technique is easily generalized to other quantum corrections. We apply the method to a simple two dimensional theory where the exact expression is known, and show that it is reproduced numerically with little computational cost. We also discuss the extension of the method to analogous calculations in four dimensions. There exist a set of standard techniques for calculating the quantum corrections to the mass, moments, and other properties of solitons [@rajar]. The techniques usually used require knowledge of the phase shifts associated with every possible meson scattering from the soliton. However in many cases of interest, for example multi-Skyrmions [@multi], the soliton field is not spherically symmetric and the necessary information would be extremely difficult to extract. In this paper we present a different approach, based upon quantizing the soliton in a finite box, which provides an infrared regulator and reduces the number of modes to be quantised to a discrete set. Our method centers upon a formula for the first quantum mass corrections due to Cahill [*et al.*]{}[@Cahill], which automatically removes the worst divergences. The quantum correction to the soliton mass is given as a trace over a complete set of modes. We compute this trace by summing first over the lowest normal modes of the soliton, and then over a plane wave basis up to a finite cutoff. The contribution of modes beyond this cutoff is included analytically using a derivative expansion. As we shall show, one per cent accuracy in the mass correction is achievable with very modest computational resources. We point toward the application of this technique to more interesting problems, where it appears feasible to calculate the quantum corrections for multisolitons in the Skyrme model [@skyrme], and other 3+1 dimensional theories. Fluctuations about classical solitons {#flucts} ===================================== We would like to quantize small fluctuations about some stable, static classical soliton. The field’s dynamics are determined by a Lagrangian, ${\cal L}(\phi,\d \phi)$, which gives classical equations of motion \_t\^2 \^a = F\^a(,,\^2) . [\[eq:motion\]]{} with the soliton  $\phi^a_{\rm st}({\mbox{\boldmath $x$}})$ as a time-independent solution. The soliton is a smooth, localized ‘lump’ in space, surrounded by a vacuum of constant field. The field $\phi^a(\xx,t)$ may be represented as the sum of the (c-number) classical soliton, $\phist(\xx)$, plus a quantum correction $\epsilon(\xx,t)$ which is an operator obeying canonical commutation relations. We work to first order in $\hbar$, which amounts to keeping terms up to second order in $\epsilon(\xx,t)$ in the the Hamiltonian. In this approximation, the quantum mechanical Hamiltonian is H\_T = E\_[cl]{} + :d\^d x\_2(\^a,\^a,\_i\^a,x): , where $E_{\rm cl} = \int d^d x V(\phist,\d \phist)$ is the classical soliton mass, and ${\cal H}_2$ quadratic in $\epsilon$ and its conjugate momentum $\pi$. Normal ordering is with respect to the trivial vacuum normal modes for $\epsilon$, the free mesons, so the Hamiltonian is zero in the soliton-free vacuum. We may express $\epsilon^a(\xx,t)$ in terms of normal modes about the classical soliton. It obeys a linear equation \_t\^2 \^a = -H\^2\_[ab]{} \^b [\[eq:fluc\]]{} with $H^2_{ab}$ a positive semi-definite differential operator depending on the soliton solution. The eigenvalues of $H^2$ are the frequencies squared $\omega_n^2$ of the normal modes about the soliton. We then write \^a(,t) = \_n $${e^{-i\omega_n t}\over \sqrt{2 \omega_n}} a_n \epsilon^a_n(\xx) + {\rm c.c.}$$  , where $\epsilon^a_n(\xx)$ is the $n$th classical normal mode, and $a_n$ is the associated annihilation operator. In terms of these normal mode operators, the Hamiltonian is H\_T = E\_[cl]{} + :\_n \_n a\_n\^a\_n: . where normal ordering is with respect to the trivial soliton-free vacuum. The ground state of this Hamiltonian in the presence of a soliton can be computed by normal ordering the second term with respect to the soliton normal modes. To do so, write the operators $a_n^\dagger,a_n$, in terms of $a^{a\dagger}({\mbox{\boldmath $k$}}),a^a({\mbox{\boldmath $k$}})$, the creation and annihilation operators for fluctuations about the soliton-free vacuum. a\_n\^= [12]{} \_[k]{}$$\tilde \epsilon^a_n (-{\mbox{\boldmath $k$}})a^{a\dagger}({\mbox{\boldmath $k$}}) \(\sqrt{\omega_n\over\omega_k} +\sqrt{\omega_k\over\omega_n} \) + \tilde \epsilon^a_n ({\mbox{\boldmath $k$}})a^a({\mbox{\boldmath $k$}}) \(\sqrt{\omega_n\over\omega_k} -\sqrt{\omega_k\over\omega_n} \)$$ , where $\omega_k = \sqrt{k^2+m^2}$ are the frequencies of elementary meson excitations. Applying this identity, one finds \_n a\^\_n a\_n &=& :\_n a\^\_n a\_n: + [14]{} \_[k]{}\_n\^a() \_n\^a (- ) [(\_n-\_k )\^2\_k]{}\ &=& :\_n a\^\_n a\_n: + [14]{}n |[(H-H\_0)\^2H\_0]{} | n , [\[eq:ccderiv\]]{} where $H$ is the single particle Hamiltonian for fluctuations about the soliton, i.e., the positive square root of $H^2$, and $H_0$ is the single particle Hamiltonian for fluctuations about the soliton-free vacuum. So, the total Hamiltonian may be written H\_T = E\_[cl]{} + \_n \_n a\^\_n a\_n +m, [\[eq:cc\]]{} where $\delta_m$, the quantum correction to the soliton mass, is m= -[14]{} [Tr]{} $${(H-H_0)^2 \over H_0}$$, [\[eq:regc\]]{} a result obtained by Cahill, Comtet and Glauber [@Cahill]. The trace is taken over any complete set of states. In one dimension, [(\[eq:cc\])]{} is finite and no further renormalization is needed. In three space dimensions it diverges logarithmically, and so requires regularization, which may be provided by inserting a factor $e^{-\epsilon H_0^2}$. The logarithmic divergence which results may then be subtracted with a local counterterm [@BTprog]. Nevertheless, the formula [(\[eq:regc\])]{} is still very useful in three dimensions, as it has removed both the quartic and quadratic divergences. To compute the trace (\[eq:regc\]) we need to be able to construct the three  terms $\<\alpha| e^{-\epsilon H_0^2}H |\alpha\>$, $\<\alpha|e^{-\epsilon H_0^2} H_0^{-1}H^2 |\alpha\>$, and $\<\alpha| e^{-\epsilon H_0^2}H_0 |\alpha\>$, where the states  $|\alpha\>$ are the elements of some basis for field perturbations. A moments thought reveals that the last two terms are simple to compute, since we have an explicit form for the operator $H^2$, and the action $H_0$ on any perturbation is trivial [(\[eq:H0times\])]{}. The difficulty lies with the first term, the trace of $H$ over its positive frequency subspace. This trouble arises because one does not have the $H$ operator in a calculationally useful form. The $H^2$ operator ------------------ In order to calculate the mass correction, [(\[eq:regc\])]{}, we need the $H^2$ operator, which can be extracted from the classical perturbation dynamics. To second order in $\epsilon^a$, the perturbation Lagrangian is (,) = \_t\^a I\_[ab]{}\_t\^b - \_i\^a $V_{ab;ij}\d_j+W_{ab;i}$\^b - \^a M\_[ab]{}\^b , where the tensors $I(\xx),V(\xx),W(\xx),M(\xx)$ are functions of the static soliton fields. The classical time evolution of $\epsilon^a(\xx,t)$ is \_t\^2 \^a &=& $$I^{-1}$$\^[ab]{}$V_{bc;ij}\d_{ij} +\d_iV_{bc;ij}\d_j + \d_iW_{bc;i} - M_{bc}$ \^c [\[eq:lineabove\]]{}\ &=& - H\^2 . [\[eq:timeevolution\]]{} And so the classical equations of motion for the perturbation yield an operator, $H^2$, whose eigenvalues are the squares of the normal mode frequencies of perturbations about the soliton. Low normal modes of the soliton {#normalmodes} ------------------------------- We have $H^2$; unfortunately its positive square root $H$ cannot be easily represented. So, one cannot calculate all terms in the trace [(\[eq:regc\])]{} directly. However, let us note several points. First, the main contribution in [(\[eq:regc\])]{} comes from the lowest frequency normal modes of the soliton; these are the modes where the difference between the $\< n| H |n\>$ and $\< n| H_0 | n\>$ are greatest. Indeed several authors, e.g. [@Scholtz], [@holz], go so far as to include [*only*]{} the contribution of the soliton zero modes to the trace. This is not an accurate approximation, however it [*is*]{} important to accurately compute the contributions of the lowest modes. Second,  $H^2$ shares eigenvectors with $H$. So, although we cannot directly construct $H$, we can find its eigenvectors and eigenvalues if we can extract them from $H^2$. Finally, it is clear that very short wavelength modes should be accurately described by the WKB approximation - their contribution to the renormalized mass correction should be small, local and analytically calculable via a derivative expansion. Our strategy is summarized in the following formula: $${\cal O}$$ &=& [Tr]{}\_n $${\cal O}$$ + [Tr]{}\_[\_[kk\_[max]{}]{}]{} $$(1-P){\cal O}$$ +[Tr]{}\_[\_[k&gt; k\_[max]{}]{}]{}$$(1-P){\cal O}$$ , [\[eq:strat\]]{}\ [where]{}& =& H\_0\^[-1]{} (H-H\_0)\^2  . Here $\{|n\rangle \}$ is a small set of precise, low frequency normal modes for the soliton, computed numerically as explained below, and used to take the first trace. The remaining part of the trace is computed using a simple plane wave basis $e^{i {\bf k.x}}$, corrected for overcounting using a projection operator $1-P$ with $P= \sum_n |n\rangle \langle n |$. The contribution from low momentum modes $k<k_{{\rm max}}$ is computed by straightforward matrix diagonalization. The final term from modes with $k>k_{{\rm max}}$ is computed analytically via a local derivative expansion as explained below. The lowest frequency normal modes of $H^2$ may be extracted using the real-time method of [@us], or by using $H^2$ to  drive a diffusion equation. The latter method has the advantage that for a limited set of modes the result converges more rapidly with integration time. For a fluctuation $|\epsilon\>$ about the static soliton, write \_| = - H\^2| . [\[eq:diffusion\]]{}Now, create an initial perturbation $|\epsilon(0)\>$. This may be written in terms of eigenvectors $|n\>$ of $H^2$, with $ H^2|n\> = \omega_n^2 |n\>$. |(0) = \_[n]{} \_n|n , so the solution of the diffusion equation [(\[eq:diffusion\])]{} is |() = \_n e\^[-\_n\^2 ]{}\_n |n . Starting in any state $|\epsilon\>$, and evolving it forward according to the diffusion equation [(\[eq:diffusion\])]{}, the resulting state rapidly becomes dominated by the lowest frequency mode. To find the lowest frequency normal modes of $H$, one starts with an arbitrary perturbation field $|\epsilon\>$ and evolves it forward according to the diffusion equation, until the resulting field is an acceptably pure eigenvector of $H^2$. Call this eigenvector $|1\>$ and store it. Next, take another perturbation, $|\epsilon'\>$, and project $|1\>$ out of it, and evolve it forward under the diffusion equation until one has another eigenvector of $H^2$. Orthonormalize this with respect to $|1\>$, and call it $|2\>$, and store it. And so on: in this way one builds up an archive spanning the lowest frequency normal modes of perturbations about the soliton. Individual perturbations in this archive will be a mixture of nearby normal modes of $H$, since evolving the diffusion equation forward a finite length of time only separates out modes with substantially different frequencies. However, once created, this archive of wavefunctions will span the space of the the lowest frequency normal modes of $H^2$, up to a cutoff frequency $\omega_c < \omega_{\max}$, the highest frequency probed. From this mixed archive of low normal modes, one can create a near-perfect archive by diagonalizing the matrix $\<i|H^2|j\>$, creating a new archive of perturbations, out of linear combinations of the old ones. Any new perturbation field $|i'\>$, will be a near machine-perfect eigenvector of $H^2$ if its frequency satisfies e\^[${\omega_i'}^2 - \omega_{\max}^2$T]{}1 , where $T$ is the finite time the diffusion equation evolved forward. All but the highest handful of modes extracted in this fashion will be near-perfect; those remaining will be discarded before further calculation. In the example shown later, we kept only the 9 lowest frequency modes (including the zero mode) as our set $\{|n\>\} $ of near-perfect low modes. The contribution of the near-perfect modes is easy to calculate, since H\^2|n = \_n\^2|nH|n = \_n|n [\[eq:eigdef\]]{}and $H_0$ acts on any perturbation via H\_0 | = \_[k]{} ||  . [\[eq:H0times\]]{}(Here the basis $|{\mbox{\boldmath $k$}}\>$ is an abbreviation for all possible plane wave perturbations of the field; for several component fields it carries an internal space index as well as a wavevector.) And so, in the basis of normal modes of $H$, it is straightforward to get each term in the trace. Calculating the trace over all fluctuations {#fourier_technique} =========================================== We wish to calculate the trace (\[eq:regc\]) over all modes, however it is not practical to compute all of them as above. Instead we proceed by supplementing our near-perfect eigenmodes with a trace over plane waves. For the set of plane waves we simply construct the matrix H\^2\_[kk’]{} = ’| H\^2 | for all plane waves ${\mbox{\boldmath $k$}},{\mbox{\boldmath $k$}}'$ satisfying $k \le k_{\rm max}$. We then diagonalize this matrix: H\^2\_[kk’]{} = [O]{}\^\_[kd]{} $$H^2_D$$\_[dd’]{} [O]{}\_[d’k’]{} , [\[eq:diag\]]{}where ${\cal O}$ is an orthogonal matrix, and define the action of $H$ via H\_[kk’]{} = [O]{}\^\_[kd]{} $$H^2_D$$\_[dd’]{}\^[12]{} [O]{}\_[d’k’]{} , [\[eq:diaga\]]{}The eigenvectors of this matrix give an approximate basis of normal modes for the soliton: |d = \_[|k|k\_]{} [O]{}\^\_[kd]{}| [\[eq:fourierbase\]]{} and the eigenvalues $\[H^2_D\]_{dd}$ yield the squared frequencies of these approximate normal modes. These normal modes will be as perfect eigenvectors of $H$ as can be constructed out of the limited region of Fourier space used. We may now straightforwardly compute each of the terms in the second trace in (\[eq:strat\]). Most simply, if  $k_{\max}$ is high enough, the trace will have converged and we are done. In order for this technique to provide an accurate mass correction, all terms in the trace, [(\[eq:regc\])]{}, which contribute significantly to the mass correction must be included. That is, the subspace of normal modes spanned by this limited perturbation basis must cover all modes for which $\<n|(H-H_0)^2/H_0|n\>$ is large. It is possible to arrange this, because the finite contibutions from both far infrared and ultraviolet scattering modes are vanishing as their wavelengths become greatly larger or smaller than the soliton size. Let $V_{\rm sol}$ be the soliton volume scale, and let $k_{\rm sol}$ be a characteristic soliton wavevector scale, $k_{\rm sol} \sim V_{\rm sol}^{-1/d}.$ All infrared modes, with wavelength $k<k_{\mbox{\tiny IR}}\ll k_{\rm sol}$ far away from the soliton give a total contribution $\sim k_{\mbox{\tiny IR}}^d\(V_{\rm sol}/V_{\rm box}\)$. After removing analytically calculable corrections (see the next section), all ultraviolet modes with wavelength $k>k_{\mbox{\tiny UV}}\gg k_{\rm sol}$ give a total contribution $\sim (1/k_{\mbox{\tiny UV}})^2$, if the Fourier transform of the soliton field dies away faster than $1/k^2$ for large $k$. (The solitons in which we are interested die exponentially with $k$.) So, the IR and UV cutoffs created by working in this truncated Fourier space create only controllable errors in the result. Example: Quantum correction to the mass of the $\phi^4$ kink {#example} ============================================================= As an example of this technique, we use it to calculate the first quantum mass correction to a well known example, the $\phi^4$ kink in 1+1 dimensions. This model has been thoroughly explored; see [@rajar] for example. In the infinite volume limit the kink solution, and all its perturbation modes are known analytically [@exact]. The calculation serves as a useful test since our numerical result can be compared with the exact analytic expression. Of course our technique can be straightforwardly be applied to [*any*]{} soliton bearing field theory in 1+1 dimensions. The Lagrangian for the $\phi^4$ kink is = [12]{}$(\d_t\phi)^2 - (\d_x\phi)^2$ + [12]{} m\^2 \^2 - [4]{}\^4 - [m\^44 ]{}  , with the usual double well potential in $\phi$, with minima at $\phi = \pm m/\sqrt{\lambda}$. The vacuum for this Lagrangian is $\phi$ resting at one or the other minimum. The kink is a field $\phi(x)$ which starts at one minimum at $x\to-\infty$, and smoothly crosses over the potential hump at $\phi=0$, to the other minimum as $x\to\infty$. Minimizing the energy of such field configurations, one finds $\phist(x)$ ( $ = m/\sqrt{\lambda}\tanh ( m x/\sqrt{2})$, for the continuum theory, but the precise form of the function is unimportant to us.) Expanding in small perturbations about the vacuum, $ \phi(x,t) = \phi_{\rm vac} + \epsilon(x,t) $, we obtain the $H_0^2$ operator: H\_0\^2 = -\_x\^2 + (m)\^2 . [\[eq:kinkH02\]]{}Similarly, expanding in small perturbations about the kink, we get H\^2 = -\_x\^2 + $3\lambda\phist^2(x) - m^2$ . [\[eq:kinkH2\]]{}which is the Schrodinger operator for a sech$^2$ potential. In order to calculate the mass correction, we discretise the system and and place it in a finite box. Equations [(\[eq:kinkH02\])]{},[(\[eq:kinkH2\])]{} are discretised by the simple replacement $\d^2_x\epsilon \to \(\epsilon_{i-1} - 2 \epsilon_i + \epsilon_{i+1}\)/\Delta_x^2$. Boundary conditions are handled by inserting a discontinuous jump of $2m/\sqrt{\lambda}$ (the distance between the two vacua) at the box boundary. This choice of boundary deforms the soliton but the deformation vanishes exponentially with box size. The perturbations obey periodic boundary conditions. [\[fig:kink\]]{} We begin by creating a stable soliton on the grid by relaxing kink-bearing initial conditions to a minimum energy configuration. The relaxed kink is shown in figure [\[fig:kink\]]{}. The field differs from its continuum value by no more than one part in $10^3$ at any point in the box. Once we have the static soliton, we compute the lowest normal modes as described above, solving the diffusion equation by the Crank-Nicholson technique [@NR]. After diagonalization of the $\<i|H^2|j\>$ matrix, we obtained an archive of the lowest normal modes of perturbations about the soliton, each a perfect eigenvector of the discrete $H^2$ operator to machine accuracy. A selection of these normal modes is shown in figure [\[fig:normal\]]{}. -- -- -- -- [\[fig:normal\]]{} Using these near-exact low normal modes, together with the set of approximate normal modes constructed from the Fourier basis as in section \[fourier\_technique\], we calculate terms in the mass correction trace,[(\[eq:regc\])]{}, using identities [(\[eq:eigdef\])]{},[(\[eq:H0times\])]{}, [(\[eq:diaga\])]{}. Including all Fourier modes in the box, we obtained a total mass correction $\delta m = -0.9468$. This is to be compared with the exact result (Rajaraman [@rajar]) of $\delta m = m/6\(\sqrt{3/2} - 18/(\pi\sqrt{2})\) = -0.9422$ for these values of parameters. Thus this technique generates the right mass correction to about 0.5%, as good as could be hoped from a $\sim 200$ point grid. As shown next, we do not need to include so many Fourier modes, since the high $k$ contribution may be computed analytically. Analytic Computation of High Momentum Contribution ================================================== The contribution of high $k$ Fourier modes to the mass correction can be calculated analytically using an expansion in $k^{-1}$. The result can then be used to improve the numerical result obtained using a limited range in Fourier space. The one dimensional theory studied here provides a test case for this approach. As we shall see, applying the lowest order correction makes the convergence with Fourier mode cutoff very rapid. As discussed above, the only nontrivial term is ${1\over 2} {\rm Tr} H$. We compute this using a heat kernel expansion as follows. First we set H\^2= -\_x\^2 + U(x) +M\^2 [\[eq:udef\]]{} with $M= \sqrt{2} m$, the meson mass, and $U=\partial^2 V /\partial \phi^2 -M^2$. The potential $U$ defined this way vanishes at infinity. Next we write H = [1]{} \_0\^ ( -[d dt]{} ) e\^[-t H\^2]{} [\[eq:heat\]]{} Performing the trace by summing over a complete set of plane wave states $e^{i kx}$ gives H = dx (-[d dt]{} ) e\^[-t]{} [\[eq:tr\]]{} where we used $\partial_x e^{i kx} = \partial_x +ik$. We now separate the exponent into a ‘free’ part $k^2+M^2$ and an ‘interaction’ part $I=(-2ik\partial_x -\partial_x^2 +U)$. We then expand $e^{-tI}$ out in powers of $I$, moving derivatives to the right noting that when a derivative reaches the extreme right, it gives zero. Thus we replace $I \rightarrow U(x)$, $I^2 \rightarrow U^2-U''-2ik\partial_x U$, $I^3 \rightarrow (-4 k^2) U''$ plus lower order terms in $k$. We then integrate over $t$, each power of $t$ giving a factor of $(k^2+M^2)^{-1}$. Keeping terms up to $k^{-3}$ we find H = dx $1 +{1\over 2} {U\over {k^2+M^2}} -{1\over 8} {U^2-U''\over (k^2+M^2)^2} -{1\over 4} {k^2 U''\over (k^2+M^2)^3} + ...$ [\[eq:exp\]]{} where we have ignored terms odd in $k$ since they cancel if we adopt a symmetric cutoff $k=|k|$. Substituting (\[eq:exp\]) into the Cahill et al. formula (9), the first two terms of (\[eq:exp\]) are cancelled by the last two terms in (9). The remainder gives the high $k$ contribution m\_k= -[18]{} \_k\^ \_[-]{}\^[+]{} dx (U\^2+U”) = -[3M\^3 8 ]{} k\^[-2]{} + o(k\^[-4]{}). [\[eq:hip\]]{} We may now use this formula to correct numerical results obtained with limited coverage in $k$ space. That is, we compute the last term in (\[eq:strat\]) analytically as above. In this calculation we may ignore the term involving the projection operator $P$ – as we shall argue, it vanishes exponentially as $k_{{\rm max}}$ is increased. To see this, note that the manipulations used in deriving (\[eq:hip\]) are actually still valid in the presence of the projector $P$, but with the introduction of a factor $\sum_n|\<k|n\>|^2$ under the $k$ integral. This involves sum of the squares of the Fourier transforms of the mode functions $|n\>$. But since the latter are smooth, the factor falls exponentially with $k_{{\rm max}}$. [\[fig:conv\]]{} Figure 3 shows the convergence of the ‘improved’ result as a function of the cutoff $k$. The horizontal line shows the full result obtained by including all modes in the box. The squares show the ‘improved’ result obtained for limited numbers of $k$ vectors, and $N$ labels the cutoff mode number, with $k=2 \pi N/L$. The ‘improved’ result is within one per cent of the full result when we include modes up to N=7, or $k\sim 5$. Our technique clearly works very well for the $\phi^4$ kink in one spatial dimension, and could be applied with similar ease to any other one dimensional scalar field theory with kinks. The convergence of the ‘improved’ formula for $\delta m$ with the cutoff in $k$ space augurs well for the prospect of performing analogous calculations in three dimensions. Just as here, the high $k$ contributions can be computed analytically. And if the convergence is similar, we can expect to obtain accurate results for the soliton mass corrections using of order $N^3 \sim 10^3$ modes in $k$ space, which should certainly be feasible. Conclusion ========== We have developed a technique for calculating the first quantum correction to the mass of a static soliton. This technique is implementable for arbitrary stable solitons within a broad class of Lagrangians. As an example, we applied this technique to the $\phi^4$ kink, where the result is known analytically: the numerical answer agreed well with the correct one. We hope to use this technique to calculate quantum corrections to the masses of more complicated solitons in three spatial dimensions. [99]{} J.F Donoghue, E. Golowich and B.R. Holstein, [*Dynamics of the Standard Model*]{}, Cambridge University Press (1992). T. H. R. Skyrme, [*Proc. Roy. Soc.*]{} [**260**]{} (1961) 127 E. Braaten, S. Townsend and L. Carson, Phys. Lett. [**B235**]{} (1990) 147; R. Battye and P. Sutcliffe, Imperial preprint 96-97/18, hep-th/9702089, Phys. Rev. Lett. (1997). C[.]{} Barnes, W[.]{}K[.]{} Baskerville, N[.]{} Turok, [*Phys[.]{} Rev[.]{} Lett[.]{}*]{} [**79(3)**]{}(1997) 367-370; C. Barnes, K. Baskerville and N. Turok, preprint hep-th/9704028, Phys. Lett. [**B**]{}, in press (1997). R. Rajaraman, , Chaps[.]{} 5-8, Elsevier, 1987. R. Dashen, B. Hasslacher and A. Neveu, Phys. Rev. [**D10**]{} (1974) 4130. A recent reference is A. Rebhan and P. van Nieuwenhuizen, hep-th/9707163. K. Cahill, A. Comtet, R. Glauber [*Phys[.]{} Lett[.]{} B*]{} [**64**]{}(1976) 283 F. Scholtz, B. Schwesinger, B. Geyer, [*Nuc[.]{} Phys[.]{} A*]{} [**561**]{}(1993) 542 M. Holzwarth, [*Phys[.]{} Lett[.]{} B*]{} [**291**]{}(1992) 218 P. M. Morse and H. F. Feshbach, [*Methods of Theoretical Physics*]{}, McGraw-Hill, 1953. W. Press, S. Teukolsky, W. Vetterling, B. Flannery [*Numerical Recipies, 2nd Ed[.]{}*]{}, Chapter 19, Cambridge University Press, 1988. C.Barnes and N.Turok, in progress. [^1]: email:[email protected] [^2]: email:[email protected]
{ "pile_set_name": "ArXiv" }
--- abstract: 'We extended McMillan’s Green’s function method to study the equilibrium spin current (ESC) in a ferromagnet/ferromagnet (FM/FM) tunnelling junction, in which the magnetic moments in both FM electrodes are not collinear. The single-electron Green’s function of the junction system is directly constructed from the elements of the scattering matrix which can be obtained by matching wavefunctions at boundaries. The ESC is found to be determined only by the Andreev-type reflection amplitudes as in the Josephson effect. The obtained expression of ESC is an exact result and at the strong barrier limit gives the same explanation for the origin of ESC as the linear response theory, that is, ESC comes from the exchange coupling between the magnetic moments of the two FM electrodes, ${\mathbf{J}}\sim{\mathbf{h}}_{l}\times{\mathbf{h}}_{r}$. In the weak barrier region, ESC cannot form spontaneously in a noncollinear FM/FM junction when there is no tunneling barrier between the two FM electrodes.' address: - '$^1$Department of Physics and Materials Science, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong, P.R. China' - '$^2$Department of Physics, Southeast University, Nanjing 210096, P.R. China' author: - 'J. Wang$^{1,2}$ and K.S. Chan$^{1}$' title: A Scattering method to the equilibrium spin current in a ferromagnet junction --- = 10000 introduction ============ Spin related transport in magnetic hybrid systems has been studied intensively for last two decades and considerable progress has been achieved since the discovery of the tunneling magnetoresistance (TMR) effect[@1] in ferromagnet (FM) tunnelling junctions. This effect originates from the electrical resistances of these junction being dependent strongly on whether the moments of adjacent magnetic layers are parallel or antiparallel. The reason is that the tunnelling electrons are scattered more strongly in an antiparallel magnetic configuration than in a parallel configuration. As a result, tunnelling junctions with moments in adjacent magnetic layers aligned antiparallelly have larger overall resistance than when the adjacent moments are aligned parallel, giving rise to TMR. The reverse effect of TMR is the spin transfer effect predicted independently by Slonczewski[@2] and Berger[@3] about a decade ago, in which a sufficiently large spin-polarized current injected from a normal metal (NM) into a FM layer can lead to magnetic moment reversal in the FM layer. Spin transfer torque occurs in magnetic multilayers with noncollinear moments and is due to the non-conservation of spin current through the interface between NM and FM. Due to the presence of noncollinearity, the component of spin current transverse to the magnetization of the layers is not transmitted across the interfaces between NM and FM; in other words, the discontinuity of spin current at the interface is the origin of the spin transfer effect that can result in magnetization precession or reversal.[@4; @5; @6; @7; @8] A large number of experiments have observed this spin transfer effect in magnetic multilayer structures,[@9; @10; @11] in which one of the FM layers is very thick and works as a polarizer of electric current with its fixed moment, and the moment of the thin FM layer may switch when a strong polarized electric current perpendicular to the layer plane flows through the layers. As the two magnetization directions in the FM/FM tunnelling junction are misaligned, an equilibrium spin current (ESC) can flow in the junction without any bias,[@7] this phenomena is analogous to the Josephson effect that the superconductor macroscopic phase difference between the two sides of a junction drives a supercurrent through the junction. The dissipationless ESC will disappear if two moments in the adjacent FM layers are collinear, parallel or antiparallel. The existence of ESC $\mathbf{J}$ has been verified by many authors using the linear response theory,[@12; @13; @14; @15; @16] and explained as the result of the exchange coupling between two magnetic moments ${\mathbf{h}}_r$ and ${\mathbf{h}}_{l}$, ${\mathbf{J}}\sim{\mathbf{h}}_{l}\times{\mathbf{h}}_{r}$, Thus the magnetization phase difference between the FM electrodes induces ESC in the FM junction as in the Josephson effect. In the case of a thin barrier between FMs or the strong coupling limit, the behavior of ESC is still unknown in the literature and worth studying. In this paper, we study the ESC flowing in a FM/FM junction with two fixed non-collinear magnetic moment using a simple quantum mechanical approach. The present work is motivated by two reasons. Firstly, dissipationless spin current in last several years has drawn considerable interest and also incurred much debate, such as the controversies in the spin Hall effect.[@17; @18; @19; @20] The study of ESC in FM/FM junctions may shed light on the spin current in spin-orbit coupled systems. Secondly, the well-known result of ESC ${\mathbf{J}}\sim{\mathbf{h}}_{l}\times{\mathbf{h}}_{r}$ in FM/FM junctions in the literature was obtained by using the linear response theory[@12; @13; @14; @15; @16] or heuristic derivation,[@21] thus an exact result of ESC is in order. By extending McMillan’s Green’s function method which was originally employed to study the Josephson effect in superconductor junction,[@22] we obtained an expression of ESC in the FM/FM junction at equilibrium. ESC is determined only by the Andreev-type reflection amplitudes,[@8; @23] which is similar to the Josephson effect.[@24; @25] It was also found that the exact ESC expression is slightly different from the linear response result; in other words, the high-order tunnelling processes are negligible and the linear response approximation has a wider range of application. When there is no barrier between the two FM electrodes, the ESC was found to vanish indicating that ESC is a pure quantum mechanical effect. This paper is organized in the following way. In Section II, a FM/FM junction model is given and the single electron Green’s function of the junction system is constructed from the scattering coefficients. An analytic expression of ESC in the FM/FM junction is obtained in the third section and some discussions are presented. A conclusion is drawn in the last section. Green’s function ================ The FM/FM tunnelling junction is depicted schematically in Fig. 1 and the layer between the FM electrodes can be either an insulator barrier or a normal metal. A simple free electron model is adopted to describe the FM/FM tunnelling junction and the Hamiltonian reads $${\cal H}=\frac{-{\hbar}^{2}{\nabla}^{2}}{2m_{e}}+V(x)- \theta(-x){\mathbf{h}}_{l}{\cdot}{\boldsymbol{\sigma}}-\theta(x-L){\mathbf{h}}_{r} {\cdot}{\boldsymbol{\sigma}},$$ where $m_e$ is the effective electron mass and assumed to be identical in all three regions, the two FM electrodes and the middle region. ${\mathbf{h}}_{l}$ and ${\mathbf{h}}_{r}$ are the internal molecular fields of the left and right FM, respectively. ${\boldsymbol{\sigma}}$ denotes the Pauli spin operator, and $\theta(x)$ is the step function. The potential energy $V(x)$ may take different values in different regions but remains constant in both FMs. Here the molecular fields ${\mathbf{h}}_{l}$ and ${\mathbf{h}}_{r}$ (magnetization with energy unit) are not collinear in our consideration and assumed to be fixed by an external magnetic field or other methods. Without loss of generality, we take the spin quantization axis of the system to be parallel to the magnetization of the left FM ${\mathbf{h}}_{l}$ and the direction of ${\mathbf{h}}_{r}$ of the right FM is described by the polar coordinate ($\theta$, $\phi$). In a free electron model, the energy dispersions of the two FMs can readily be solved and they are spin-dependent. In the left FM, the eigenvalue is $E_{\pm}=\frac{\hbar^{2}k^2}{2m_{e}}\pm h_{l}+U_{l}$ and the spinor is the eigenfunction of $\sigma_{z}$, and in the right FM, $E_{\pm}=\frac{\hbar^{2}k^2}{2m_{e}}\pm h_{r}+U_{r}$ and the spin eigenfunctions are $\psi_{+}^{l}=\left(% \begin{array}{c} \cos(\theta/2)e^{-i\phi/2} \\ \sin(\theta/2)e^{i\phi/2} \\ \end{array}% \right)$ and $\psi_{-}^{l}=\left(% \begin{array}{c} -\sin(\theta/2)e^{-i\phi/2} \\ \cos(\theta/2)e^{i\phi/2} \\ \end{array}% \right)$ where $U_{l} (U_{r})$ are the different potential energy in the left(right) FM, $\mathbf{k}$ is the wavevector of electron and $\pm$ is the spin index. The spatial eigenfunctions are plane waves. Due to spin splitting, there are four incoming wave functions with their corresponding outgoing wave functions in the left and right FM, as schematically shown in Fig. 2. $\Psi_{1}(x)$ and $\Psi_{2}(x)$ are the wavefunctions of the minority spin and majority spin electron injecting from the left FM, respetively, while $\Psi_{3}(x)$ and $\Psi_{4}(x)$ are those injecting from the right FM. For example, the wavefunction of the first type of scattering event $\Psi_{1}(x)$ is given by $$\Psi_{1}(x)= \left\{% \begin{array}{l} \exp(ik_{+}^{l}x)\left(% \begin{array}{c} 1 \\ 0 \\ \end{array}% \right)+a_{1}\exp(-ik_{+}^{l}x)\left(% \begin{array}{c} 1 \\ 0 \\ \end{array}% \right)+b_{1}\exp(-ik_{-}^{l}x)\left(% \begin{array}{c} 0 \\ 1 \\ \end{array}% \right), x<0 \\ c_{1}\exp(ik_{+}^{r})\left(% \begin{array}{c} \cos(\theta/2)e^{-i\phi/2} \\ \sin(\theta/2)e^{i\phi/2} \\ \end{array}% \right)+ d_{1}\exp(ik_{-}^{r})\left(% \begin{array}{c} -\sin(\theta/2)e^{-i\phi/2} \\ \cos(\theta/2)e^{i\phi/2} \\ \end{array}% \right),x>L \\ \end{array}% \right.$$ In this equation, $k_{\pm}^{l(r)}=\sqrt{2m(E-U_{l(r)}\mp{h_{l(r)}})/\hbar^{2}-k_{//}^{2}}$ is the spin-dependent wavevector along the $x$-direction in the left (right) FM electrode, $E$ is the single-electron energy which is conserved when electrons tunnel through the junction, ${\mathbf{k}}_{//}$ is the wavevector parallel to the interface between the different regions and assumed to be conserved in the quantum tunnelling process. Nevertheless, this is not a required condition for the following derivation. $\Psi_{1}$ describes a minority-spin electron coming from the left FM lead and being scattered in the middle region; the scattering coefficients $a_{1}$, $b_{1}$, $c_{1}$ and $d_{1}$ correspond to the normal reflection, Andreev-type reflection,[@8] transmission without branch crossing, and transmission with branch crossing,[@26] respectively. For the other $3$ scattering processes $\Psi_{i}$ as well as their coefficients $a_{i}$, $b_{i}$, $c_{i}$, and $d_{i}$ as shown in Fig. 2 have the same meaning . It is noted that we have omitted the parallel plane wave component $e^{i{\mathbf{k}}_{//}\vec{r}(y,z)}$ in the above equation. When we use the continuity of the derivatives of wavefucntions at the interfaces to determine these coefficients ( $a_i$, $b_i$, $c_i$, and $d_i$), the parallel momentum ${\mathbf{k}}_{//}$ should be explicitly taken into account. With four elementary scattering wavefunctions like those in Eq. 2 above, the single-electron Green’s function of the junction system can be worked out by the McMillan formula,[@22] which has been further developed by Kashiwaya and Tanaka.[@25] This formula has been used to treat the Josephoson current in a superconductor junction and relate directly the super current to the Andreev reflection amplitudes. The Green’s function $G^{r}(x,x')$ is proportional to the direct product of the left-going wavefunctions (processes $i=3$, $4$ in Fig. 2) and the right-going wavefunctions (processes $i=1$, $2$), $G^{r}\sim \Psi_{L}(x)\hat{\Psi}_{R}(x')$ for $x\leq x'$ and $G^{r}\sim \Psi_{R}(x)*\hat{\Psi}_{L}(x')$ for $x\geq x'$, where the hat ‘$\wedge$’ denotes the conjugate process to the elementary scattering one shown in Fig. 2, and these conjugate scattering wavefunctions can be obtained by determining the Hermitian conjugate of only the spinor part of the wavefunction not including the spatial part of the wavefunction. The reflection and transmission coefficients $\tilde{a}_i$, $\tilde{b}_i$, $\tilde{c}_i$, and $\tilde{d}_i$ in four conjugate processes have the relations $\tilde{a}_{i}(\phi)=a_{i}(-\phi)$, $\tilde{b}_{i}(\phi)=b_{i}(-\phi)$, $\tilde{c}_{i}(\phi)=c_{i}(-\phi)$, and $\tilde{d}_{i}(\phi)=d_{i}(-\phi)$ ($i=1 \cdots 4$), where $\phi$ is the azimuthal angle of the magnetization of the right FM. With these scattering wavefunctions of the elementary processes as well as their conjugate processes, the Green’s function is then constructed in a linear combination as $$G^{r}(x,x',E)=\left\{% \begin{array}{l} \alpha_{1}\Psi_{3}(x)\hat{\Psi}_{1}(x')+\alpha_{2}\Psi_{3}(x)\hat{\Psi}_{2}(x') +\alpha_{3}\Psi_{4}(x)\hat{\Psi}_{1}(x')+\alpha_{4}\Psi_{4}(x)\hat{\Psi}_{2}(x'), x\leq x' \\ \beta_{1}\Psi_{1}(x)\hat{\Psi}_{3}(x')+\beta_{2}\Psi_{2}(x)\hat{\Psi}_{3}(x') +\beta_{3}\Psi_{1}(x)\hat{\Psi}_{4}(x')+\beta_{4}\Psi_{2}(x)\hat{\Psi}_{4}(x'), x\geq x' \\ \end{array}% \right.$$ Here $G^{r}(x,x',E)$ is implicitly a function of the parallel momentum $k_{//}$. The prefactors $\alpha_{i}$ and $\beta_{i}$ ($i=1\cdots 4$) can be determined by the boundary conditions that the Green’s function fulfills, $$\frac{\partial}{\partial x}G^{r}(x,x',E)|_{x=x'+0}-\frac{\partial}{\partial x}G^{r}(x,x',E)|_{x=x'-0}=\frac{2m_{e}}{\hbar^2}\left(% \begin{array}{cc} 1 & 0 \\ 0 & 1 \\ \end{array}% \right),$$ $$G^{r}(x,x',E)|_{x=x'+0}=G^{r}(x,x',E)|_{x=x'-0}.$$ With these two equations, we can directly solve the prefactors $\alpha_{i}$ and $\beta_{i}$ in Eq. 3, which are independent of the spatial position $x$. After some direct algebra, they read $$\begin{aligned} \alpha_{1}=\frac{z_{+}c_{4}}{c_{3}c_{4}-d_{3}d_{4}},\ \ \alpha_{2}=\frac{-z_{-}d_{4}}{c_{3}c_{4}-d_{3}d_{4}}, \nonumber \\ \alpha_{3}=\frac{-z_{+}d_{3}}{c_{3}c_{4}-d_{4}d_{4}},\ \ \alpha_{4}=\frac{z_{-}c_{3}}{c_{3}c_{4}-d_{4}d_{4}}, \nonumber \\ \beta_{1}=\frac{z_{+}\tilde{c}_{4}}{\tilde{c}_{3}\tilde{c}_{4}-\tilde{d}_{3}\tilde{d}_{4}},\ \ \beta_{2}=\frac{-z_{-}\tilde{d}_{4}}{\tilde{c}_{3}\tilde{c}_{4}-\tilde{d}_{3}\tilde{d}_{4}}, \nonumber \\ \beta_{3}=\frac{-z_{+}\tilde{d}_{3}}{\tilde{c}_{3}\tilde{c}_{4}-\tilde{d}_{3}\tilde{d}_{4}},\ \ \beta_{4}=\frac{z_{-}\tilde{c}_{3}}{\tilde{c}_{3}\tilde{c}_{4}-\tilde{d}_{3}\tilde{d}_{4}},\end{aligned}$$ where $z_{\pm}=\frac{m_{e}}{i\hbar^{2}k_{\pm}^{l}}$. In these solutions, we have employed some detailed balance conditions to facilitate our derivation, such as $a_{i}(\phi)=a_{i}(-\phi)$ ($i=1\cdots 4$) and $k_{-}^{l}\tilde{b}_{1}=k_{+}^{l}b_{2}$. Substituting these prefactors into Eq. 3, we obtained the Green’s function in the left FM electrode ($x$,$x'<0$) as $$\begin{aligned} G^{r}(x,x',E)=\left(% \begin{array}{cc} G_{11} & G_{12} \\ G_{21} & G_{22} \\ \end{array}% \right), \nonumber \\ G_{11}=z_{+}\exp(ik_{+}^{l}|x-x'|)+a_{1}z_{+}\exp(-ik_{+}^{l}x-ik_{+}^{l}x'),\nonumber \\ G_{12}=z_{-}b_{2}\exp(-ik_{+}^{l}x-ik_{-}^{l}x'),\nonumber \\ G_{21}=z_{+}b_{1}\exp(-ik_{+}^{l}x'-ik_{-}^{l}x),\nonumber \\ G_{22}=z_{-}\exp(ik_{-}^{l}|x-x'|)+a_{2}z_{-}\exp(-ik_{-}^{l}x-ik_{-}^{l}x').\nonumber \\\end{aligned}$$ The Green’s functions in other regions such as the middle region between the two FMs can also be constructed in a similar manner according to Eq. 3 with solutions given in Eq. 6. Although the Green’s function is obtained here for the FM/FM junction, this method can also be applied to nonmagnetic junctions, and we believe the derivation procedure is universal in mesoscopic junction systems. When the retarded Green’s function of the studied system is worked out, we can in principle calculate the physical observables that we want. Since we focus on an equilibrium tunnelling junction, the lesser Green’s function is easily obtained by the formula $$G^{<}(x,x',E)=\left[G^{a}(x,x',E)-G^{r}(x,x',E)\right]f(E)$$ where $G^{a}(x,x',E)$ is the advanced Green’s function and $f(E)$ is the Fermi-Dirac distribution function. The lesser Green’s function is related to the spectral function of the electron and is very useful for calculating some physical quantities as shown later. ESC in the FM/FM junction ========================= In this section, we first present a general formula of the spin current in an FM and then utilize the Green’s function obtained in Sec. II to work out the ESC in a FM/FM tunnelling junction. Then as an example, we calculated ESC of a simple FM/FM junction with a delta function insulator barrier between the two FM electrodes. The local spin density $\vec{s}({\mathbf{r}},t)$ at position $\mathbf{r}$ and time $t$ is defined as $$\vec{s}({\mathbf{r}},t)=\Psi^{\dagger}({\mathbf{r}},t)\hat{\vec{s}}\Psi({\mathbf{r}},t),$$ where $\Psi({\mathbf{r}},t)$ is a two-component wavefunction and $\hat{\vec{s}}=\frac{\hbar}{2}\hat{\vec{\sigma}}$ with $\hat{\vec{\sigma}}$ denoting the Pauli spin matrices. Taking the time-derivative of $\vec{s}({\mathbf{r}},t)$ and using the Schrodinger equation for a single FM, the continuity equation of spin current density is given by $$\frac{\partial \vec{s}}{\partial t}+\nabla\cdot {\mathbf{J}}_{s}+{\mathbf{S}}=0,$$ where the respective spin current density ${\mathbf{J}}_s$ and the source term $\mathbf{S}$ are $$\begin{aligned} {\mathbf{J}}_s=\frac{\hbar^{2}}{4m_{e}i}\left[\Psi^{\dagger}{\boldsymbol {\sigma}}\nabla\Psi- (\nabla\Psi)^{\dagger}{\boldsymbol {\sigma}}\Psi\right],\nonumber \\ {\mathbf{S}}=\Psi^{\dagger}({\boldsymbol\sigma}\times{\mathbf{h}})\Psi,\end{aligned}$$ where $\mathbf{h}$ is the magnetization of the FM leads (for example, ${\mathbf{h}}_{l}$, ${\mathbf{h}}_{r}$). ${\mathbf{J}}_s$ is the usual definition of spin current density and $\mathbf{S}$ is referred to as the spin source term or spin torque that leads to spin rotation when the spin is not collinear with the magnetization $\mathbf{h}$. Therefore this spin source term gives rise to a spin current flow which is not included in ${\mathbf{J}}_s$. Using the lesser Green’s function in Eq. 8, $G^{<}(xt,x't')=i\langle\Psi^{\dagger}(x't')\Psi(xt)\rangle$ where $\langle\cdots\rangle$ is the quantum statistical average, the spin current density ${\mathbf{J}}_s$ and the source term $\mathbf{S}$ above in a steady state can be rewritten as $${\mathbf{J}}_s=\frac{-\hbar^{2}}{4m_{e}}\lim_{x'\rightarrow x}\left\{\frac{\partial}{\partial x'}-\frac{\partial}{\partial x}\right\}\int\frac{dE}{2\pi}Tr\left[{\boldsymbol\sigma} G^{<}(x',x)\right],$$ $${\mathbf{S}}=\lim_{x'\rightarrow x}\int\frac{dE}{2\pi}Tr\left[({\mathbf{h}}\times{\boldsymbol\sigma}) G^{<}(x',x)\right],$$ where the trace is taken over the spin space. Substituting the Green’s function given by Eq. 7 and Eq. 8 into the above equation, we obtain the ESC expression in the FM/FM junction as $$\begin{aligned} J_{s}^{x}=\int\frac{dE}{4\pi}\sum_{k_{//}} Re\left\{(k_{+}^{l}-k_{-}^{l})(\frac{b_{2}}{k_{-}^{l}}-\frac{b_{1}}{k_{+}^{l}})\exp[-i(k_{+}^{l}+k_{-}^{l})x] \right\}f(E),\end{aligned}$$ $$\begin{aligned} J_{s}^{y}=-\int\frac{dE}{4\pi}\sum_{k_{//}} Im\left\{(k_{+}^{l}-k_{-}^{l})(\frac{b_{2}}{k_{-}^{l}}+\frac{b_{1}}{k_{+}^{l}})\exp[-i(k_{+}^{l}+k_{-}^{l})x] \right\}f(E)\end{aligned}$$ $$\begin{aligned} J_{s}^{z}=0.\end{aligned}$$ Here the summation over $k_{//}$ includes the contribution of all transverse modes to the spin current, where $k_{//}$ is assumed to be conserved in the quantum tunnelling process. According to the above equation, the spin current density is determined by the Andreev-type reflection coefficients $b_{1}$ and $b_{2}$; this result is very similar to the formula of the Josephson current which is directly related to the Andreev reflection coefficients.[@24; @25] The spin current density in Eq. 14 is exponentially $x$-dependent so that the integral over energy and transverse density of states will make it disappear in bulk FM not far away from interface. This phenomena is same as the spin transfer effect that the transmitted spin current is also dependent on position and will vanish after summing all possible transverse modes with different wavevectors. Stiles and Zagwill[@6] estimated the characteristic length of this spatial spin precession was about $1/k_f$ ($k_{f}$ Fermi wavevector) because only the electrons near the Fermi energy contribute to the electric current. For our case in Eq. 14, all the band electrons contribute to ESC so that ${\mathbf{J}}_s$ is expected to decay much more quickly with distance from interface. Apart from the spin current density ${\mathbf{J}}_s$, which can be calculated directly from its definition, additional spin current flow arises from the source term $\mathbf{S}$ according to Eq. 10. According to the Gaussian theorem, another spin current ${\mathbf{J}}_{soc}$ can be defined from this source term as $$\begin{aligned} J_{soc}^{(x)}=\int_{x}^{0}S^{x}dx =\int\frac{dE}{4\pi}\sum_{k_{//}} Re\left\{(k_{+}^{l}-k_{-}^{l})(\frac{b_{2}}{k_{-}^{l}}-\frac{b_{1}}{k_{+}^{l}})(1-\exp[-i(k_{+}^{l}+k_{-}^{l})x] )\right\}f(E)\end{aligned}$$ $$\begin{aligned} J_{soc}^{y}=\int_{x}^{0}S^{y}dx=-\int\frac{dE}{4\pi}\sum_{k_{//}} Im\left\{(k_{+}^{l}-k_{-}^{l})(\frac{b_{2}}{k_{-}^{l}}+\frac{b_{1}}{k_{+}^{l}})(1-\exp[-i(k_{+}^{l}+k_{-}^{l})x] )\right\}f(E).\end{aligned}$$ The $z$-component of ${\mathbf{J}}_{soc}$ is zero. The $x$-dependent part of ${\mathbf{J}}_{soc}$ is just the $-{\mathbf{J}}_{s}$ in Eq. 14, so that the sum of ${\mathbf{J}}_{soc}$ and ${\mathbf{J}}_{s}$ is independent of $x$ and keep constant from the interface to the bulk FM, which does not disappear away from the interface, $$\begin{aligned} J^{x}=\int\frac{dE}{4\pi}\sum_{k_{//}} Re\left\{(k_{+}^{l}-k_{-}^{l})(\frac{b_{2}}{k_{-}^{l}}-\frac{b_{1}}{k_{+}^{l}})\right\}f(E),\end{aligned}$$ $$\begin{aligned} J^{y}=-\int\frac{dE}{4\pi}\sum_{k_{//}} Im\left\{(k_{+}^{l}-k_{-}^{l})(\frac{b_{2}}{k_{-}^{l}}+\frac{b_{1}}{k_{+}^{l}})\right\}f(E).\end{aligned}$$ When a spin current ${\mathbf{J}}_{s}$ enters the bulk FM, it will be absorbed by the lattice in FM because the magnetization $\mathbf{h}$ rotates the spin and the sum of ${\mathbf{J}}_{soc}$ and ${\mathbf{J}}_{s}$ therefore remains constant. As an example, we calculated a simple case of the FM/FM tunnelling junction, the potential energy in both FM is $U_{l}=U_{r}=0$, and the insulator barrier is described by a delta barrier $U_{0}\delta(x)$ where $U_{0}$ denotes the strength of the barrier, The magnitudes of the magnetizations in two FM are equal $h_{l}=h_{r}$. Using simple quantum mechanics method, we obtain the Andreev-type reflection amplitudes as $$\begin{aligned} b_{1}=2(k_{-}-k_{+})k_{+}\sin\theta\exp(i\phi)/A,\nonumber \\ b_{2}=2(k_{-}-k_{+})k_{-}\sin\theta\exp(-i\phi)/A, \nonumber \\ A=k_{-}^{2}+k_{+}^{2}+6k_{+}k_{-}-2k_{0}^{2}+4ik_{0}(k_{+}+k_{-})-(k_{+}-k_{-})^2\cos\theta,\end{aligned}$$ where $k_{\pm}=k_{\pm}^{l(r)}$ and $k_0=mU_{0}/\hbar^2$. Substituting these quantities into Eq. 16, the total ESC is then given by $${\mathbf{J}}=\int\frac{dE}{4\pi}\sum_{k_{//}}Im[1/A(E)](k_{+}-k_{-})f(E) \hat{\mathbf{h}}_{l}\times\hat{\mathbf{h}}_{r}$$ where $\hat{\mathbf{h}}_{l}(0,0,1)$ and $\hat{\mathbf{h}}_{r}(\sin\theta\cos\phi,\sin\theta\sin\phi,\cos\theta)$ are just the unit vectors of the magnetizations in the left and right FM. The ESC depends on the directions of the two magnetizations not only through the cross product term $\hat{\mathbf{h}}_{l}\times\hat{\mathbf{h}}_{r}$ but also on the quantity $A$ in Eq. 17. In the finite barrier case, we can neglect the $\cos\theta$ term in $A$ so that Eq. 18 can reproduce the result obtained with the linear response theory, i.e., the exchange coupling between two magnetization leading to ESC. At the opposite limit, zero barrier between two FMs $U_{0}=0$, Eq. 18 indicates there is no ESC flowing in the junction. This is the pure quantum mechanics effect, the reflected spin direction by a barrier rotates with respect to the incident spin direction and an imaginary part can appear in the quantity $A$ of Eq. 17, which was regarded as[@6] one of reasons that spin current is not conserved at the interface when a polarized charge current flows through a NM/FM junction. For ESC discussed here, the reflections by a barrier leading to an additional phase to the incident wavefunction which is necessary for the formation of ESC in the noncollinear FM/FM junction. It is quite different from the Josephson effect between two superconductors, in which a cooper pair carries supercurrent through the junction even with zero barrier. summary ======= We have studied ESC flowing in the noncollinear FM/FM tunnelling junction by extending McMillan Green’s function method. The single-electron Green’s function of the junction can be constructed by the scattering coefficients, reflection and transmission amplitudes which can be directly calculated by simple quantum mechanics method. In the derived formula of spin current density, we found that the exact result of ESC does not have a big difference from that obtained by the linear response theory, in other words, the high-ordered tunnelling process contribute little to ESC. It was also found that when there is no barrier between two FMs, the ESC will disappear for the reflected spin state from a finite barrier has an additional phase which is crucial for the formation of ESC. ACKNOWLEDGEMENT This work is supported by City University of Hong Kong Strategic Research Grant (Project No. 7001619). M. N. Baibich, J. M. Broto, A. Fert, F. Nguyen Van Dau, F. Petroff, P. Eitenne, G. Creuzet, A. Friederich, and J. Chazelas, Phys. Rev. Lett. [**61**]{}, 2472 (1988) J. C. Slonczewski, J. Magn. Magn. Mater. [**159**]{}, L1 (1996). L. Berger, Phys. Rev. B [**54**]{}, 9353 (1996). C. Heide, P.E. Zilberman, and R.J. Elliott, Phys. Rev. B [**63**]{}, 064424 (2001); C. Heide, Phys. Rev. Lett. [**87**]{}, 197201 (2001); Phys. Rev. B [**65**]{}, 054401 (2002); X. Waintal, E.B. Myers, P.W. Brouwer, and D.C. Ralph, Phys. Rev. B [**62**]{}, 12 317 (2000). A. Brataas, Yu.V. Nazarov, and G.E.W. Bauer, Phys. Rev. Lett. [**84**]{}, 2481 (2000); A. Brataas, Yu.V. Nazarov, and G.E.W. Bauer, Eur. Phys. J. B [**22**]{}, 99 (2001). M.D. Stiles and A. Zangwill, J. Appl. Phys. [**91**]{}, 6812 (2002); Phys. Rev. B [**66**]{}, 014407 (2002). Y. Tserkovnyak, A. Brataas, G. E. W. Bauer, and B. I. Halperin, Rev. Mod. Phys. [**77**]{}, 1375 (2005). A. Brataas, G. E. W. Bauer, and P. J. Kelly, cond-mat/0602151. E. B. Myers, D. C. Ralph, J. A. Katine, R. N. Louie, and R. A. Buhrman, Science [**285**]{}, 867 (1999); E. B. Myers, F. J. Albert, J. C. Saneky, E. Bonet, R. A. Buhrman, and D. C. Ralph, Phys. Rev. Lett. [**89**]{}, 196801 (2002). J. A. Katine, F. J. Albert, R. A. Buhrman, E. B. Myers, and D. C. Ralph, Phys. Rev. Lett. [**84**]{}, 3149 (2000); S. I. Kiselev, J. C. Sankey, I. N. Krivorotov, N. C. Emley, R. J. Schoelkopf, R. A. Buhrman, and D. C. Ralph, Nature (London) [**425**]{}, 380 (2003). M. Tsoi, A. G. M. Jansen, J. Bass, W. C. Chiang, V. Tsoi, and P. Wyder, Nature (London) [**406**]{}, 46 (2000). J. Wang and K. S. Chan, Phys. Rev. B [**74**]{}, 035342 (2006). Nogueira and K.H. Bennemann, Europhys. Lett. [**67**]{}, 620 (2004). Y.L. Lee and Y.W. Lee, Phys. Rev. B [**68**]{}, 184413 (2003). H.Katsura, N. Nagaosa, A.V. Balatsky, Phys. Rev. Lett. [**95**]{}, 057205 (2005). M.Braun, J. Konig, and J. Martinek, Superlattices and Microstrutures [**37**]{}, 333 (2005). S. Murakami, N. Nagaosa, and S.-C. Zhang, Science [**301**]{}, 1368 (2003); J. Sinova, D. Culcer, Q. Niu, N.A. Sinitsyn, T. Jungwirth, and A.H. MacDonald, Phys. Rev. Lett. [**92**]{}, 126603 (2004). S. Zhang and Z. Yang, Phys. Rev. Lett. [**94**]{}, 066602 (2005); A. Khaetskii, Phys. Rev. Lett. [**96**]{}, 056602 (2006) J. Shi, P. Zhang, D. Xiao, and Q. Niu, Phys. Rev. Lett. [**96**]{}, 076604 (2006); Y. Wang, K. Xia, Z. B. Su, and Z. Ma, Phys. Rev. Lett. [**96**]{}, 066601 (2006). J. Wang, K. S. Chan, and D. Y. Xing, Phys. Rev. B [**73**]{}, 033316 (2006); J. Wang, K. S. Chan, Phys. Rev. B [**72**]{}, 045331 (2005). J. Slonczewski, Phys. Rev. B, [**39**]{}, 6995 (1989). W. L. McMillan, Phys. Rev. [**175**]{}, 559 (1968). A. F. Andreev, Sov. Phys. JETP [**19**]{}, 1228, (1964). A. Furusaki and M. Tsukada, Solid State Commun. [**78**]{}, 299 (1991). S. Kashiwaya and Y. Tanaka, Rep. Prog. Phys. [**63**]{}, 1641 (2000). G.E. Blonder, M. Tinkham, and T.M. Klapwijk, Phys. Rev. B [**25**]{}, 4515 (1982).
{ "pile_set_name": "ArXiv" }
--- abstract: 'Using the plasma model for the metal dielectric function we have calculated the electromagnetic fluctuation induced forces on a free standing metallic film in vacuum as a function of the film size and the plasma frequency. The force for unit area is attractive and for a given film thickness it shows an intensity maximum at a specific plasma value, which cannot be predicted on the basis of a non retarded description of the electromagnetic interaction. If the film is deposited on a substrate or interacts with a plate, both the sign and the value of the force are modified. It is shown that the force can change sign from attraction to repulsion upon changing the substrate plasma frequency. A detailed comparison between the force on the film boundaries and the force between film and substrate indicates that, for $50-100nm$ thick films, they are comparable when film-substrate distance is of the order of the film thickness.' address: - '$^1$ CNR/INFM-National Research Center on nanoStructures and bioSystems at Surfaces (S3), Via Campi 213/A, I-41100 Modena, Italy' - '$^2$ Dipartimento di Fisica, Università di Modena e Reggio Emilia, Via Campi 213/A, I-41100 Modena, Italy' author: - 'A. Benassi$^{1,2}$, C. Calandra$^2$' title: A study of the electromagnetic fluctuation induced forces on thin metallic films --- Introduction ============ Electromagnetic fluctuation induced forces have been the subject of several investigations both in the non retarded small distance limit (dispersion forces) and in the retarded large distance case (Casimir forces)[@lifshitz; @dzyaloshinskii; @mahanty; @milloni; @bordag; @lamoreaux]. Since the basic work from Lifshitz[@lifshitz], studies have been focused mainly on the determination of the forces between two semi-infinite planar media[@dzyaloshinskii] or between a sphere and a planar medium[@derjaguin]. Even if model calculations have been performed for special geometries[@bordag; @lamoreaux; @barton], forces on realistic systems have been obtained starting from the above mentioned configurations.\ For several technological applications, multilayer systems, obtained by depositing thin films of different species onto a given substrate, are important and theoretical approaches have been devised to determine van der Waals forces between laminated media[@podgornik; @white]. Casimir forces between moving parts have been considered as possible source of instabilities in micro and nano-devices, where the components are in close proximity[@serry; @buks; @zhao]. In many of these systems the situations of interest correspond to interaction between parallel interfaces of films and plates with micro or submicro-size and submicro-distances. While some of these studies have been performed using simplified models for the interaction, like the assumption that the interaction is correctly represented by the force between ideal metallic plates, it has been pointed out that a realistic description of adhesion or stiction phenomena has to account for the dependence of the force upon the shape and optical properties of the components[@white; @barcenas].\ In spite of this large amount of work, a detailed study of the behaviour of the electromagnetic fluctuation induced forces in unsupported or deposited conductive films, as a function of their size and optical parameters has not been published. The interest has been focused mainly on the interaction between two-dimensional films, for which forces are supposed to show a peculiar dependence upon the film distance[@sernelius; @boestrom1; @boestrom2]. Less interest has been given to the study of the forces on the film boundaries due to vacuum fluctuations, which are present even in an isolated film and depend upon its size and properties.\ We can formulate the problem as follows: suppose we have a simple metal, whose dielectric function can be expressed as $$\epsilon(\omega)=1-\frac{\Omega_{p}^{2}}{\omega(\omega+i/\tau)} \label{prima}$$ where $\Omega_{p}$ is the plasma frequency and $\tau$ is the relaxation time. If we consider an isolated metal film of thickness $d$, at $T=0^{\circ}K$ we know that, in the limit $\Omega_{p}\rightarrow 0$ the force on the film (the electromagnetic pressure on the film boundaries or the force between them) vanishes and the same happens in the limit of infinite plasma frequency (perfect metal case). The vanishing of the force is due to the peculiar values of film reflectivity in the two limits, it cannot stay identically zero for physical values of the reflectivity.[^1] The problem of what sort of behaviour has the force between these two limits has not been investigated: clearly, for a given film thickness, it must reach at least one maximum of intensity. The questions to be answered are: (i) what is the behaviour of the force as a function of $\Omega_{p}$, in particular at which plasma frequencies are force maxima obtained, (ii) how do such maxima depend upon the film thickness, (iii) how is the behaviour of the force modified when the film is deposited onto a substrate, (iv) how does the electromagnetic force on the free standing film compare with the film-substrate interaction, that may be responsible of adhesion and stiction phenomena.\ To provide answers to some of these questions, we have used the Lifshitz approach to the electromagnetic fluctuation induced forces to study the force on metallic films. Here we report on some general results that can be obtained using the plasma model model dielectric functions. Our purpose is not to achieve a precise description of the force for specific systems, since an accurate evaluation with an intrinsic force uncertainty of few per cent requires a precise determination of the Drude parameters, which are very sensitive to the sample condition [@lambrecht; @pirozhenko]. Rather we want to illustrate some general trends that can be understood using a model description of the dielectric properties. Force on isolated metallic films {#ontometal} ================================ We consider metallic films of thickness $d$ ranging from $10 nm$ to a few hundred $nm$. For typical metallic densities the electronic distribution deviates significantly from the bulk behaviour when the size of the film is less than about ten times the Fermi wavelength $\lambda_{F}$. Taking $\lambda_{F}\simeq5$ Å, the bulk description is expected to become inaccurate when $d$ is of the order of $50$ Å. For such ultrathin film quantum size effects are known to be important[@rogers1; @rogers2; @loly; @lindgren1; @lindgren2].\ ![\[refframe3\] Notation for three layers system](fig1.ps){width="3cm"} We adopt a local description of the electromagnetic properties of the metal based on a dielectric function $\epsilon(\omega)$ i.e. neglecting the wavenumber dependence of the dielectric response. This local theory is expected to be less accurate in thin films than for half space or bulk systems. Recent calculations have shown that non local corrections to electromagnetic induced forces for typical metallic densities are of the order of a few tenth of a per cent, suggesting that the local theory can be appropriate in the interpretation of the experimental data[@esquivel]. The expression of the force per unit area $F$ at absolute temperature $T$ in a configuration with two semi-infinite planar media of dielectric functions $\epsilon_{1}(\omega)$ and $\epsilon_{2}(\omega)$ respectively separated by a film of thickness $d$ and dielectric function $\epsilon_{3}(\omega)$ (figure \[refframe3\]) is given by $$F=-\frac{1}{\pi\beta}\int_{0}^{\infty}k dk \sum_{n}'\bigg\{ \frac{1-Q_{TM}(i\Omega_{n})}{Q_{TM} (i\Omega_{n})}+ \frac{1-Q_{TE}(i\Omega_{n})}{Q_{TE}(i\Omega_{n})} \bigg\}\gamma_{3} \label{exact}$$ here $k$ is the modulus of a two dimensional wave vector parallel to the plates, $\beta=1/k_{B}T$, $k_{B}$ is the Boltzmann constant, $\Omega_{n}=2 \pi n / \hbar \beta$ is the Matsubara frequency corresponding to the $n$-th thermal fluctuation mode, the prime on the summation indicates that the $n=0$ term is given half weight. $Q_{TM}$ and $Q_{TE}$ refer to transverse magnetic (TM) and transverse electric (TE) modes respectively and are given by $$\begin{aligned} Q_{TM}(i\omega)=1&-\frac{(\epsilon_{1}\gamma_{3}-\epsilon_{3}\gamma_{1}) (\epsilon_{2}\gamma_{3}-\epsilon_{3}\gamma_{2})}{(\epsilon_{1}\gamma_{3}+\epsilon_{3}\gamma_{1}) (\epsilon_{2}\gamma_{3}+\epsilon_{3}\gamma_{2})}e^{-2d\gamma_{3}}\\ Q_{TE}(i\omega)&=1-\frac{(\gamma_{3}-\gamma_{1}) (\gamma_{3}-\gamma_{2})}{(\gamma_{3}+\gamma_{1}) (\gamma_{3}+\gamma_{2})}e^{-2d\gamma_{3}}\end{aligned}$$ with $$\gamma_{i}^2=k^{2}+\frac{\Omega_{n}^{2}}{c^{2}}\epsilon_{i}(i\Omega_{n}) \label{gamma}$$ and the dielectric functions are evaluated at the frequency $i\Omega_{n}$.\ We performed the calculation of the force per unit area on a free standing metal film using equation (\[exact\]) at $T=300^{\circ}K$, taking $\epsilon_{1}=\epsilon_{2}=1$ and $\epsilon_{3}=1-\Omega_{3}^{2}/\omega^{2}$. We neglect for simplicity relaxation time effects. Although they are important in determining the infrared response of metals and the calculation we report can be extended to a complex dielectric function, we focus our interest mainly on general trends in the behaviour of electromagnetic fluctuation induced forces. For realistic calculations on specific materials relaxation time effects have to be included and in metal films they can affect the force intensity.\ Notice that there is a basic difference between the electromagnetic fluctuation induced forces between two semi-infinite metals and those on the boundaries of a film. This is clearly seen if one considers an ideal (perfectly reflecting) metal, corresponding to an infinite plasma frequency: the interaction between two semi-infinite systems is expressed by the well known Casimir force $F(d)=\hbar\pi^{2}c/240 d^{4}$, while for a film of finite thickness, the force on the boundaries vanishes.\ ![\[fig1\] Force as a function of film plasma frequency: the force increases on decreasing the film thickness $d$. The calculation were performed summating the first thousand Matsubara frequencies.](fig2.ps){width="8cm"} Figure \[fig1\] displays the calculated behaviour of $F$ as a function of the plasma frequency for films of different thickness ranging from $10$ to $100 nm$. Notice that at finite temperatures the force vanishes in the large plasma frequency limit, while for $\Omega_{p}\rightarrow 0$ there is a finite contribution from transverse magnetic modes. This contribution is due to the $m=0$ term of the sum over Matsubara frequencies and it depends linearly upon $T$. It is seen that the force is attractive (it tends to contract the film) and it shows a maximum and a tail at high plasma frequency. As expected from the general behavior of the electromagnetic induced forces as a function of the distance, the maximum intensity reduces as a function of $d$, while its frequency moves to higher values. The dots in the figure correspond to the free electron plasma frequency for sp-bonded simple metals. This should not be seen as an accurate prediction of the force value for metals. It indicates only that the force on real metal films may fall on both sides of the maximum, depending upon the film thickness.\ Notice that retardation effects are essential to obtain the maximum in the theoretical curve. This can be understood by a simple calculation of the force on a free standing metal film in the van der Waals (small $d$) regime at $T=0^{\circ}K$. In this case we have[@dzyaloshinskii; @bergstroem] $$F=-\frac{\hbar}{8\pi^{2}d^{3}}\int_{0}^{\infty}\frac{(\epsilon_{3}(i\xi)-1)^{2}} {(\epsilon_{3}(i\xi)+1)^{2}}d\xi \label{smalld}$$ which leads to $$F=-\frac{\hbar\Omega_{s}}{32\pi d^{3}} \label{nomax}$$ with $\Omega_{s}=\Omega_{3}/\sqrt{2}$ frequency of the surface plasmon. Equation (\[nomax\]) does not show any maximum as a function of $\Omega_{3}$. This is not surprising since the above expression is valid under the condition that $d$ is much smaller than the plasma wavelength, therefore is appropriate in the small plasma frequency regime only.\ ![\[fig2\] Maximum value of the force and plasma frequency at which it occurs, as a function of distance: the fitting functions are $-64.05-3.98ln(d)$ (triangles) and $19.83-0.98ln(d)$ (stars).](fig3.ps){width="8cm"} The behaviour of the maximum frequency as a function of $d$ is given in figure \[fig2\]: it is shown that in the range of thickness we have considered, the maximum frequency falls like $d^{-1}$, while the intensity maximum falls as $d^{-4}$, as expected for the interaction in the retarded regime. The behaviour of the force maximum, that is displaced to larger values for smaller thicknesses, can be understood by noticing that the attraction arises from the interaction between the surface plasmons at the two film boundaries[@kampen; @gerlach; @intravaia]. At a given film thickness the interaction is screened by the electron gas with increasing efficiency as the plasma frequency increases. For large electron density $\Omega_{3} \rightarrow\infty$, the force goes to zero and one surface does not feel the presence of the other. The maximum in the force results from the balance between the surface plasmons interaction and the screening effects. In particular for small $d$ a higher electron density is required to screen the attractive force.\ Some interesting comments can be made on these data. The first concerns the unsupported film stability: the force tends to shrink the film and it has to be equilibrated by some repulsive interaction, most likely provided by the force built up by the valence electron rearrangement at the surfaces. Second we notice that the force can be tuned significantly by changing the electron density of the metal: this effect could be useful in engineering the film properties for specific applications. The film-ideal metal substrate interaction ========================================== To show how these conclusions are modified when the metal film is interacting with a substrate, we display in figure \[fig4\] the behaviour of the force per unit area on a film of $d=100nm$ thickness deposited onto a perfectly reflecting substrate, (corresponding to the configuration with $\epsilon_{2}=1$ and $\epsilon_{1}$ equal to infinity), as a function of film plasma frequency. This is a very simplified description of a bi-metallic interface, based on the assumption of the validity of the continuum model, that neglects all the details of the interactions between the atoms at the interface. It is expected to hold when the size of the film is large compared to the interface region (typically a few angstroms) so that the interface plays a minor role in determining the electromagnetic force. Notice that the force becomes repulsive and nearly double in intensity, although it shows the same qualitative behaviour with a maximum and a long asymmetric tail at large frequency values. It comes from the difference between the electromagnetic force per unit area on the substrate side and that on the vacuum side ![\[fig4\] Force as a function of film plasma frequency: a change in sign occurs when the isolated metallic film is placed on a perfectly reflecting substrate (ideal metal).](fig4.ps){width="8cm"} ![\[refframe5\] Notation for five layers system](fig5.ps){width="3cm"} The behaviour of the force can be understood by noticing that at $T=0^{\circ}K$ the exact calculation in the non-retarded limit gives the simple result: $$F(d)=\frac{\hbar\Omega_p}{32\pi d^{3}}\sqrt{2}$$ showing the change of sign and the increased force value. This result is consistent with the behaviour of the London dispersion forces between dissimilar materials separated by a gap, that has been reported since many years [@mahanty; @israelachvili; @french] . In this case the force is known to be repulsive when $\epsilon_{1} \gtrless \epsilon_{2} \gtrless \epsilon_{3}$ and attractive when $\epsilon_{1} \gtrless \epsilon_{2} \lessgtr \epsilon_{3}$ within a wide frequency range.\ It is interesting to understand how the force between film boundaries in a multilayer system is modified as a function of the film-substrate distance. For the case of a perfectly reflecting substrate, one can determine the range of distances over which the sign of the force changes. To this aim one has to extend equation (\[exact\]) to a configuration with more than tree planar media. In practice this amounts to replace the functions $Q_{TM}$ and $Q_{TE}$ by those appropriate to a multi-layer configuration. For a five layer system the appropriate expressions were derived by Zhou and Spruch[@zhou]: $$\begin{aligned} Q_{TM}=Q_{TM1}Q_{TM2}\qquad Q_{TE}=Q_{TE1}Q_{TE2}\\ Q_{TM 1,TE 1}=\frac{\rho_{13}^{TM,TE}-\rho_{14}^{TM,TE}e^{-2\gamma_{1}d_{1}}}{1-\rho_{13}^{TM,TE}\rho_{14}^{TM,TE}e^{-2\gamma_{1}d_{1}}} e^{-\gamma_{3}d}\\ Q_{TM 2,TE 2}=\frac{\rho_{23}^{TM,TE}-\rho_{25}^{TM,TE}e^{-2\gamma_{2}d_{2}}}{1-\rho_{23}^{TM,TE}\rho_{25}^{TM,TE}e^{-2\gamma_{2}d_{2}}} e^{-\gamma_{3}d}\\ \rho_{mn}^{TE}=\frac{\gamma_{m}-\gamma_{n}}{\gamma_{m}+\gamma_{n}}\qquad \rho_{mn}^{TM}=\frac{\gamma_{m}\epsilon_{n}-\gamma_{n}\epsilon_{m}}{\gamma_{m}\epsilon_{n}+\gamma_{n}\epsilon_{m}} \label{5layers}\end{aligned}$$ where $\gamma_{i}$ is again given by (\[gamma\]) and the new indexes refers to figure \[refframe5\].\ For the study of substrate-metal film interaction, we take $\epsilon_{4}$ equal to infinity, $\epsilon_{1}=\epsilon_{2}=\epsilon_{5}=1$ while $\epsilon_{3}$ is the metallic film dielectric function (\[prima\]). Since the configuration depends upon two parameters, the size $d$ of the film and the film-substrate distance $d_{1}$, one can define the force $F$ between the film boundaries, given by the derivative of the free energy with respect to $d$, and the force $F'$, obtained by deriving the free energy with respect to $d_{1}$, giving the interaction between the film and the substrate. Figure \[new1\] shows the behaviour of the force $F$ on the film boundaries as a function of the film-substrate distance for a $100 nm$ film. It is seen that the force remains constant and attractive if the distance $d_{1}$ is larger than the film thickness $d$; at lower distances the force decreases until it becomes repulsive. In other words, the film starts feeling a difference between the pressure from the metal substrate side and the external vacuum pressure, when the film-substrate distance is comparable with its thickness.\ Discussions on device stability refer usually to the interaction between film and substrate (here we use the word substrate to indicate a structure of much larger size than the film, it could be a plate in a device), which gives rise to an attractive force $F'$. To show how this interaction behaves as a function of the ideal film-substrate distance, we have calculated $F'$ using equations (\[5layers\]). It turns out to be attractive for any value of the film plasma frequency and, at distances smaller than the film size, it is considerably more intense than the force $F$ on the film. This force is responsible of the change in sign observed in figure \[fig4\]: if the film is close to the substrate, the difference between the attractive force on the film boundaries tends to stretch the film, causing a repulsive force between them.\ The behaviour of the film-substrate force $F'$ in the range of distances $d_{1}$ below the film thickness, where the substrate effect is more significant, is illustrated by the results shown in figure \[new2\] for a $100nm$ film with plasma frequency $\Omega_{3}=5\cdot 10^{15} rad/s$ and a perfectly reflecting substrate.\ Notice that in this range of distances the force $F'$ increases like $d_{1}^{-x}$ with $3<x<4$, (the simple $d_{1}^{-4}$ behaviour at all distances is characteristic of the interaction between ideal metal plates only and it is appropriate for real metals only at large distances). At $100nm$ distance this force is approximately $-4.8N/m^{2}$, (the Casimir force between ideal metals at the same distance is of the order of $-10N/m^{2}$), while the force on the film boundaries is approximately $-1N/m^{2}$. The gray curve in the figure displays the calculated force per unit area for a semi-infinite metal interacting with a perfectly reflecting semi-infinite substrate. It can be seen that it does not deviate significantly from the curve for the $100nm$ film. At higher distances the attractive force decreases while the force on the film remains approximately constant.\ We report in the same figure the calculated $F'$ for a $10nm$ film: in this case the force versus distance behaviour is rather different, showing a significantly higher exponent than in the $100nm$ case ($3.52$ rather than $3.29$). Clearly this behaviour cannot be understood using arguments based on results for semi-infinite systems: for a semi infinite metal interacting with an ideal substrate one would expect the exponent $x$ to become closer to $3$ upon decreasing the distance. The fact that it results to be significantly higher is a direct consequence of the finite thickness of the film. Indeed, as first pointed out by Zhou and Sprunch[@zhou] higher negative exponents characterize the interaction in the presence of film of very small thickness. An important consequence of this behaviour is that the calculated $F'$ at $10nm$ distance (approximately $-7511.7 N/m^{2}$) is considerably higher than the force $F$ on the film boundaries (approximately $0.001 N/m^{2}$).\ We can conclude that the interaction of a metal film with a perfectly reflecting substrate leads to an attractive film-substrate force and, at short distances, to a repulsive force on the film boundaries. For $50-100nm$ thick films these forces are approximately of the same order when the film-substrate distance is comparable with the film size. In the low distance range ($1-10nm$) the force on the film can be neglected and the attractive film-substrate interaction prevails in intensity. These considerations are expected to be important for systems, like microswitches, that consist of two conducting electrodes, where one is fixed and the other one is able to move, being suspended by a mechanical spring. The stability of the system may depend upon the electromagnetic induced force acting on the mobile film [@Palasantzas3; @Palasantzas]. ![\[new1\] Force on the film boundaries as a function of the film-substrate distance, the film plasma frequency is $\Omega_{3}=5\cdot 10^{15}rad/s$.](fig6.ps){width="8cm"} ![\[new2\] Film-substrate force as a function of the film-substrate distance between ideal metal substrate and real metal film, film thickness $100nm$ (triangles) and $10nm$ (crosses). The fitting functions are $-51.5-3.29ln(d_{1})$ (dotted line) and $-55.7-3.52ln(d_{1})$ (dashed line). The gray curve is the force between two semi-infinite systems, an ideal metal and a real metal with the same plasma frequency of the film.](fig7.ps){width="8cm"} ![\[fig5a\] Force on the film boundaries as a function of substrate plasma frequency calculated for different film plasma frequencies, in the calculation about two thousands Matsubara frequencies have been used.](fig8.ps){width="8cm"} ![\[fig5b\] Force on the film boundaries as a function of film plasma frequency calculated for different substrate frequencies, in the calculation about two thousands Matsubara frequencies have been used.](fig9.ps){width="8cm"} The bimetallic interfaces ========================= The situation changes if we consider a more realistic description of the substrate. Referring to figure \[refframe3\] this corresponds to take $\epsilon_{1}=1-\Omega_{1}^{2}/\omega^{2}$. Figure \[fig5a\] shows the behaviour of the force per unit area on a $100 nm$ metal film deposited onto various metal substrates as a function of the substrate plasma frequency. Notice that the force is attractive when $\Omega_{1}<\Omega_{3}$ and is repulsive in the opposite case. For $\Omega_{1}\gg\Omega_{3}$ we get the repulsive force corresponding to a perfectly reflecting substrate. The change in the sign it can be easily understood by considering the force in the small $d$ limit, i.e in the non retarded regime. At $T=0^{\circ}K$ the force calculated from equation (\[smalld\]) is simply given by $$F=\frac{\hbar}{32 \pi d^{3}}\frac{\Omega_{s}(\Omega_{1}^{2}-\Omega_{3}^{2})} {\bar{\Omega}(\bar{\Omega}+\Omega_{s})} \label{maxexist}$$ where $$\bar{\Omega}=\sqrt{(\Omega_{1}^{2}+\Omega_{3}^{2})/2}$$ is the interface plasmon frequency obtained from the relation $\epsilon_{1}(\omega)=-\epsilon_{3}(\omega)$. Note that $F$ shows the expected change from the repulsive to the attractive behaviour.\ Figure \[fig5b\] shows curves of the force on films deposited onto different substrates as a function of the film plasma frequencies. The curves show two extrema: on the repulsive side a maximum, that increases in intensity and moves to higher frequency upon increasing the substrate plasma frequency; on the attractive side a minimum which decreases upon increasing $\Omega_{1}$ and shifts to higher frequency values. This behaviour is consistent with the previous conclusions concerning the ideal substrate: as the plasma frequency $\Omega_{1}$ increases the repulsive force on the film becomes dominant.\ It is interesting to see how the extrema behave upon varying the film thickness. As shown in figure \[fig6\], the intensity of the repulsive maximum falls like $d^{-3}$, in the range of distances we are considering, while for the attractive minimum it falls approximatively as $d^{-4}$. Indeed the occurrence of the maximum can be understood on the basis of the short distance formula (\[maxexist\]), which gives a $d^{-3}$ dependence of the force, while the behaviour of the attractive part is mainly due to retarded interactions. ![\[fig6\] $F(\Omega_{3})$ minimum and $F(\Omega_{3})$ maximum as a function of distance $d$, the fitting functions are respectively $-68.36-4.17ln(d)$ (stars) and $-49.10-3.05ln(d)$ (triangles), calculation were performed at fixed $\Omega_{1}=5\times 10^{15} rad/s$](fig10.ps){width="8cm"} . These results lead to the conclusion that the electromagnetic fluctuation induced forces can give contribute of opposite sign, and with different dependence upon the film size, to the deposited film stability.\ As in the case of the ideal substrate, we can study the electromagnetic fluctuation induced force $F_{1}$ between the film and the substrate as a function of the film-substrate distance. Based on the previous analysis we expect the film-substrate force to be attractive and to lead to a repulsive or attractive force between the film boundaries depending upon the difference between the plasma frequencies: if $\Omega_{1}\gg\Omega_{3}$ the situation is similar to the ideal substrate case, while for $\Omega_{1}\ll\Omega_{3}$ the force on the film is only weakly modified by the interaction. The various case are illustrated in figure \[new4\].\ ![\[new4\] Film boundaries force as a function of film-substrate distance, comparison between an ideal substrate (continuous line), a real metal substrate with plasma frequency $10^{16}rad/s$ (dot-dash line) and $10^{15}rad/s$ (dashed line). The film plasma frequency is $5\cdot 10^{15}rad/s$.](fig11.ps){width="8cm"} ![\[new5\] Film-substrate force as a function of the film-substrate distance for real metals with $5\cdot 10^{15}rad/s$, film thickness $100nm$ (triangles) and $10nm$ (crosses). The fitting functions are $-51.4-3.25ln(d_{1})$ (dotted line) and $-53.3-3.35ln(d_{1})$ (dashed line). The gray curve is the force between two semi-infinite bulks, two real metals with the same plasma frequency.](fig12.ps){width="8cm"} To see how our results depend upon the film thickness we have compared the calculated curves for film-substrate force as a function of the distance $d_{1}$ with the electromagnetic induced bulk-bulk interaction. We have compared the bulk-bulk interaction with film-substrate interaction for typical values of the plasma frequency. As shown in figure \[new5\] the results seem not to depend significantly upon the film thickness for $100nm$ films, while size effects become important for $d$ of the order of $10nm$.\ It is clear, from these calculations that, in the nanometric distances range, the adoption of the simple force expression appropriate to ideal plates is not correct. Both the sign and the intensity of the force may result wrong, if material properties and thickness effects are not properly accounted in the theory. Discussion and Conclusions {#conclusions} ========================== We have presented a rather complete set of results based on a continuum dielectric model to illustrate trends in the behaviour of the electromagnetic fluctuation induced forces on free-standing and supported metal films, which allow to identify the conditions under which the force is attractive or repulsive and how it depends upon the film thickness and the interacting substrate (plate) properties.\ We have shown that both the sign and the intensity of the force between a film and a plate depend upon the difference in the plasma frequencies and can be modified upon changing the carrier density. This is in line with the recent proposal of modulating the Casimir force between a metal and a semiconductor plate by illuminating the semiconducting material, i.e. by enhancing the electron plasma and creating a hole plasma in the semiconductor plate[@klimchitskaya2]. We expect that any experimental system that allows to change the difference in plasma frequencies can be used to modulate the electromagnetic force.\ An adequate description of the electronic properties in thin film does, in general, require consideration of the changes in the electron energy levels resulting from the confinement of the electrons. Since these effects are observed for film thicknesses of several nanometers, one can argue that the results of the present paper may be significantly modified if quantum size effects are taken into account. To clarify this point we have calculated the dielectric permittivity of metallic films adopting the particle in a box model[@wood], in which independent electrons are confined by a surface potential of a given length scale $d$ along the $z$ direction, with the eigenvalue spectrum: $$E_{\bf{k},n}=\frac{\hbar^{2}\bf{k}^{2}}{2 m}+E_{0}n^{2}\qquad n=1,2,3...$$ here ${\bf k}$ is a two dimensional wavevector and $E_{0}=\hbar^{2}\pi^{2}/2 m d^{2}$. The electron confinement leads to the quantization of the transverse component of the momentum and formation of lateral sub-bands. The surface effects is built in the eigenstates: $$\psi_{{\bf k},n}({\bf r})=\sqrt{\frac{2}{A\times d}}sin(\frac{n \pi}{d}z)e^{i {\bf k}\cdot\bf{\rho}}$$ where $A\times d$ is the film volume and $\bf{\rho}$ is the positive vector in the $xy$ plane. Under such conditions the film dielectric tensor is given by: $$\begin{aligned} \epsilon_{\alpha,\alpha}(\omega)=1-\frac{\Omega_{p}^{2}}{\omega^{2}}-\frac{8\pi e^{2}}{A\times d\quad m^{2} \omega^{2}}\sum_{{\bf k},n}\sum_{{\bf k}',n'}f_{0}\times \\ \nonumber \times (\epsilon_{\bf{k},n})\frac{ (\epsilon_{\bf{k},n}-\epsilon_{\bf{k}',n'})\vert\langle\psi_{{\bf k},n}\vert\hat{p_{\alpha}}\vert\psi_{{\bf k}',n'} \rangle\vert^{2}} {(\epsilon_{\bf{k},n}-\epsilon_{\bf{k}',n'})^{2}-(\hbar\omega)^{2}}\end{aligned}$$ This expression differs from the model dielectric function in several respects: (i) it has a tensor character with $\epsilon_{x,x}=\epsilon_{y,y}\neq\epsilon_{z,z}$, (ii) the plasma frequency $\Omega_{p}$ depends upon the film density, which, at a fixed chemical potential at $T=0^{\circ}K$, changes as a function of the film thickness, (iii) it accounts for transitions between lateral sub-bands (the Fermi momentum being constant the number of occupied sub-band increases upon increasing the film size). It can be easily shown that these transitions do not affect the lateral components of the dielectric tensor. They modify the low frequency behaviour of $\epsilon_{z,z}$. The model has been used to interpret optical and transport properties of thin films[@jalochowski1; @jalochowski2; @jalochowski3; @rogacheva].\ ![\[fig13\] Force as a function of film plasma frequency: comparison between Drude model (dashed line) and particle in a box model (continuous line).](fig13.ps){width="8cm"} We have calculated the electromagnetic induced forces on free-standing metals films of different Fermi energy. Figure \[fig13\] shows the results of a calculation for a $50 nm$ film. One can compare them with those plotted in figure \[fig2\]. It can be noticed that, although the value of the force is modified by the inclusion of size effects, the behaviour as a function of the free electron plasma frequency remains the same. Similar results have have been obtained in other cases and will be reported elsewhere, in a more detailed study of quantum size effects on electromagnetic induced forces.\ We conclude that the main trend of the results given in the present paper is not modified by quantum size effects for thickness above $10nm$.\ The present theory can be improved along two main lines. Inclusion of bulk relaxation effects, both in the continuum dielectric theory and in the particle in a box model, is expected to modify the calculated value of the force. A more accurate description of surface electromagnetic field, that treats the modifications to the Fresnel optics caused by the surface, may also lead to appreciable changes specially for film size of the order of few nanometers. AB thanks *CINECA Consorzio Interuniversitario* ([www.cineca.it]{}) for funding his Ph.D. fellowship. Reference {#reference .unnumbered} ========= [10]{} E.[M. Lifshitz]{}. , 2:73, 1956. I.[E. Dzyaloshinskii]{}, E.[M. Lifshitz]{}, and L.[P. Pitaevskii]{}. , 10:165, 1958. J. Mahanty and B.[N. Ninham]{}. . Academic Press, New York, 1976. P.[W. Milloni]{}. . Academic Press, New York, 1994. M. Bordag, U. Mohideen, and V.[M. Mostephanenko]{}. , 353:1, 2003. S.[K. Lamoreaux]{}. , 68:201, 2005. B.[V. Derjaguin]{} and I.[I. Abrikosova]{}. , 3:819, 1957. G.[ Barton]{}. , 34:4083, 2001. R. Podgornik and V.[A. Parsegian]{}. , 120:3410, 2004. L.[R. White]{}, R.[R. Dagastine]{}, P.[M. Jones]{}, and Yiao-Tee Hsia. , 97:104503, 2005. F.[M. Serry]{}, D.[ Walliser]{}, and G.[J. Maclay]{}. , 84:2501, 1998. E.[ Buks]{} and M.[L. Roukes]{}. , 63:033402, 2001. Y.[-P. Zhao]{}, L.[S. Wang]{}, and T.[X. Yu]{}. , 17:519, 2003. J.[ Barcenas]{}, L.[ Reyes]{}, and R.[ Esquivel-Sirvent]{}. , 87:263106, 2005. B.E.[ Sernelius]{} and P. Bjork. , 57:6592, 1998. M. Boström and B.E.[ Sernelius]{}. , 61:2204, 2000. M. Boström and B.E.[ Sernelius]{}. , 62:7523, 2000. A.[ Lambrecht]{} and S.[ Reynaud]{}. , 8:309, 2001. I.[ Pirozhenko]{}, A.[ Lambrecht]{}, and [V.B.Svetovoy]{}. , 8:238, 2006. J.[P. [Rogers III]{}]{}, P.[H. Cutler]{}, T.[R. Feuchtwang]{}, N. Miskovski, and A.[A. Lucas]{}. , 141:61, 1984. J.[P. [Rogers III]{}]{}, P.[H. Cutler]{}, T.[R. Feuchtwang]{}, and A.[A. Lucas]{}. , 181:436, 1987. P.[D. Loly]{} and J.[B. Pendry]{}. , 16:423, 1986. S.[A. Lindgren]{} and L. Wallden. , 59:3003, 1987. S.[A. Lindgren]{} and L. Wallden. , 61:2894, 1988. R. Esquivel-Sirvent and V.[B. Svetovoy]{}. , 72:045443, 2005. L. Bergström. , 70:125, 1987. N.G.[ van Kampen]{}, B.R.A.[ Nijboer]{}, and K.[ Schram]{}. , 26A:307, 1968. E.[ Gerlach]{}. , 4:393, 1971. F.[ Intravaia]{} and A.[ Lambrecht]{}. , 94:110404, 2005. J.N. Israelachvili. . Academic Press, London U.K., 1991. R.[H. French]{}. , 83:2117, 2000. F.[ Zhou]{} and L.[ Spruch]{}. , 52:297, 1995. G.[ Palasantzas]{}. , 20:1321, 2006. G.[ Palasantzas]{}. , 97:126104, 2005. G.L.[ Klimchitskaya]{}, U.[ Mohideen]{}, and V.M.[ Mostepanenko]{}. , 40:F841, 2007. D.M.[ Wood]{} and N.W.[ Ashcroft]{}. , 25:6255, 1982. M.[ Jalochowski]{}, E.[ Bauer]{}, H.[ Knoppe]{}, and G.[ Lilienkamp]{}. , 45:13607, 1992. M.[ Jalochowski]{}, H.[ Knoppe]{}, G.[ Lilienkamp]{}, and E.[ Bauer]{}. , 46:4693, 1992. M.[ Jalochowski]{}, M.[ Hoffmann]{}, and E.[ Bauer]{}. , 51:7231, 1995. E.I.[ Rogacheva]{}, O.N.[ Nashchekina]{}, S.N.[ Grigorov]{}, M.A.[ Us]{}, M.S.[ Dresselhaus]{}, and S.B.[ Cronin]{}. , 13:1, 2002. [^1]: This can be easily understood by noticing that in the non-retarded regime the force on the free standing metal film is the same as the force between two semi-infinite plates of the same metal separated by a distance equal to the film thickness (see equations (\[smalld\]) and (\[nomax\]) in the text), which is obviously attractive and different from zero.
{ "pile_set_name": "ArXiv" }
--- abstract: '[If $A$ is an AF algebra and $\alpha \in {\operatorname{Aut}}(A)$, it is shown that AF embeddability of the crossed product, $A\times_{\alpha}\Bbb Z$, is equivalent to $A\times_{\alpha}\Bbb Z$ being stably finite. This equivalence follows from a simple K-theoretic characterization of AF embeddability. ]{}' --- xy [**On the AF Embeddability of Crossed Products of AF Algebras by the Integers**]{} [**Nathanial P. Brown**]{}\ [**Purdue University**]{} This paper is concerned with the question of when the crossed product of an AF algebra by an action of $\Bbb Z$ is itself AF embeddable. It is well known that quasidiagonality and stable finiteness are hereditary properties. That is, if $A$ and $B$ are C$^{*}$-algebras with $A \subset B$ and $B$ has either of these properties, then so does $A$. Since AF algebras enjoy both of these properties we have that quasidiagonality and stable finiteness are geometric obstructions to AF embeddability. For crossed products of AF algebras by $\Bbb Z$, these turn out to be the only obstructions.\ If $A$ is an AF algebra, then we may easily describe an algebraic obstruction to the AF embeddability of $A\times_{\alpha}\Bbb Z$.\ [**Definition 0.1**]{} If $A$ is an AF algebra and $\alpha \in {\operatorname{Aut}}(A)$ then we denote by $H_{\alpha}$ the subgroup of $K_{0} (A)$ given by all elements of the form $\alpha_{*}(x) - x$ for $ x \in K_{0} (A)$.\ It follows from the Pimsner-Voiculescu six term exact sequence (\[PV\]) that if $A$ is unital and AF then $K_{0} (A\times_{\alpha}\Bbb Z$) = $K_{0} (A)/H_{\alpha}$. Now, if $B$ is unital and stably finite (in particular, if $B$ is AF) and $p \in M_{n}(B)$ is a projection then $[p]$ must be a nonzero element of $K_{0} (B)$. Thus, if $A\times_{\alpha}\Bbb Z$ embeds into B (or if $A\times_{\alpha}\Bbb Z$ is already stably finite) then every projection in the matrices over $A\times_{\alpha}\Bbb Z$ must give a nonzero element in $K_0 (A\times_{\alpha}\Bbb Z)$. In particular, since $A \hookrightarrow A\times_{\alpha}\Bbb Z$, and we have observed that $K_0 (A\times_{\alpha}\Bbb Z) = K_0(A)/H_{\alpha}$, we then conclude that $H_{\alpha} \cap K_{0}^{+} (A) = \{0\}$, where $K_{0}^{+} (A)$ is the positive cone of $K_{0} (A)$. Thus, an algebraic obstruction to AF embeddability of a crossed product would be if $H_{\alpha}$ contained a positive element of $K_0 (A)$. It turns out that (so long as $A$ is AF) this is the only algebraic obstruction. Indeed, the main result of this paper is the following.\ [**Theorem 0.2**]{} If $A$ is an AF algebra (not necessarily unital) and $\alpha \in {\operatorname{Aut}}(A)$ then the following are equivalent. 1. $A\times_{\alpha}\Bbb Z$ is AF embeddable 2. $A\times_{\alpha}\Bbb Z$ is quasidiagonal 3. $A\times_{\alpha}\Bbb Z$ is stably finite 4. $H_{\alpha} \cap K_{0}^{+} (A) = \{0\}$ The implications $(1) \Rightarrow (2) \Rightarrow (3) \Rightarrow (4)$ are relatively straightforward (whether or not $A$ is unital). In fact, if $A$ is any C$^{*}$-algebra with a countable approximate unit consisting of projections, then the implications $(1) \Rightarrow (2) \Rightarrow (3) \Rightarrow (4)$ hold. However the remaining implication, $(4) \Rightarrow (1)$, will require some technical machinery (see Section 3 for the details of all the implications). This paper consists of five sections. In Section 1 we collect a few useful facts and set some of our notation. In Section 2 we develop the key technical tools. Using the Rohlin property for automorphisms of AF algebras (see Definition 2.1) we will show that certain commutative diagrams at the level of K-theory lift to commutative diagrams on the algebras. In recent years, the Rohlin property has been studied intensively. There are notions of the Rohlin property for more general group actions (\[Oc\], \[Na\]) and this notion for actions of $\Bbb Z$ has been used in many interesting applications (\[R$\o$\], \[Ki2\], \[BKRS\], \[Co2\], \[EK\], just to name a few). All of the embedding results presented here could be viewed as more applications of the Rohlin property. In Section 3 we prove a useful K-theoretic characterization of AF embeddability (different than that given in Theorem 0.2) and use this to prove Theorem 0.2. In Section 4 we present several applications of our results. Among other things, we will prove a natural generalization of a result of Voiculescu that states that if some power of an automorphism of an AF algebra is approximately inner then the corresponding crossed product is AF embeddable (see Theorem 3.6 in \[Vo\]). We will also recover Pimsner’s criterion for AF embeddability of crossed products of $C(X)$ by $\Bbb Z$, in the case $X$ is compact, metrizable and totally disconnected. In Section 5 we show that our constructions can be done in such a way as to yield rationally injective maps on $K_0(A\times_{\alpha}\Bbb Z)$ and thus injective maps when $K_0(A\times_{\alpha}\Bbb Z)$ is assumed to be torsion free. The key technical results presented here were established while the author participated in the Monbusho Summer Program for Young Foreign Researchers in Sapporo, Japan. The author would like to thank Monbusho and the NSF for sponsoring this program and allowing him the opportunity to participate in it. The author is particularly indebted to his host, Professor Akitaka Kishimoto (nihongo de Kishimoto Sensei), for the many invaluable discussions on crossed products, dimension groups and most importantly, the Rohlin property. Indeed, these notes owe a great deal to the experience and insights shared with the author during his stay in Sapporo. The author would also like to thank Professor Hiroshi Takai for a very profitable discussion at Tokyo Metropolitan University. A few of the results presented here are (partial) answers to questions raised at that meeting. Finally, the author expresses his deepest thanks to his advisor and mentor, Professor Marius Dadarlat, for his support, encouragement and the countless hours of patiently answering questions. [**Section 1:  Preliminaries**]{} In this section we will recall a few facts and present and easy lemma that will be useful later on. [*Throughout this paper, A and B will always denote (not necessarily unital) AF algebras*]{}. Following the notation and terminology in \[Da\], we will let $K_{0}^{+}(A)$ and $\Gamma(A)$ be the positive cone and scale, respectively, of $K_0(A)$. We will say that a group homomorphism, $\theta$: $K_0(A)$ $\rightarrow$ $K_0(B)$, is a [*contraction*]{} if $\theta(\Gamma(A)$) $\subset \Gamma(B)$. We will say $\theta$ is [*faithful*]{} if it is contractive and $\Gamma(A) \ \cap Ker(\theta$) = $\{0\}$. Both of the following two facts are essentially contained in Elliot’s original classification of AF algebras (\[El\]). Section IV.4 in \[Da\] has a very nice treatment of this result. Moreover, the following two facts follow easily from Lemma IV.4.2 and the proof of Elliot’s Theorem (Theorem IV.4.3) in \[Da\]. Fact 1.1   [*If $\theta$: $K_0(A)$ $\rightarrow$ $K_0(B)$ is a faithful group homomorphism then there exists a \*-monomorphism, $\varphi$: A $\rightarrow$ B, with $\varphi_{*} = \theta$*]{}. Fact 1.2   [*If $\varphi , \sigma$: A $\rightarrow$ B are \*-homomorphisms, where A is finite dimensional, then $\varphi_{*} = \sigma_{*} \Longleftrightarrow \varphi = Adu \circ \sigma$ for some unitary, u $\in$ B.*]{} Recall that when $A$ is unital, $A\times_{\alpha}\Bbb Z$ is defined to be the universal C$^{*}$-algebra generated by $A$ and a unitary $u$, subject to the relation $Adu(a) = \alpha(a)$, for all $a \in A$. We will call $u$ the [*distinguished*]{} unitary in $A\times_{\alpha}\Bbb Z$. Since one always has the freedom to tensor with other AF algebras when proving AF embeddability, the following lemma will be useful. [**Lemma 1.3**]{}   If $A$ and $B$ are unital, $\alpha \in {\operatorname{Aut}}(A)$ and $\beta \in {\operatorname{Aut}}(B)$ are given, then there is a natural embedding $A\otimes B\times_{\alpha\otimes \beta}\Bbb Z$ ${\Large \hookrightarrow}$ ($A\times_{\alpha}\Bbb Z)\otimes (B\times_{\beta}\Bbb Z$) [**Proof**]{}   Let $u \in A\times_{\alpha}\Bbb Z$, $v \in B\times_{\beta}\Bbb Z$, and $w \in A\otimes B\times_{\alpha\otimes \beta}\Bbb Z$ be the respective distinguished unitaries. Then define a covariant representation by $w \longmapsto u\otimes v$\ $a\otimes b \longmapsto a\otimes b$ Let $\varphi: A\otimes B\times_{\alpha\otimes\beta}\Bbb Z \rightarrow (A\times_{\alpha}\Bbb Z)\otimes (B\times_{\beta} \Bbb Z$) be the induced \*-homomorphism and $D$ = range($\varphi$). To prove that $\varphi$ is injective (see Thm. 4 in \[La\]), we must provide automorphisms, $\rho(\xi) \in {\operatorname{Aut}}(D)$, for all $\xi \in \Bbb C$ with $|\xi|$ = 1, such that $\rho(\xi)(a\otimes b) = a\otimes b$ and $\rho(\xi)(u\otimes v) = \xi(u\otimes v)$. So, take $\rho_{A}(\xi) \in {\operatorname{Aut}}(A\times_{\alpha}\Bbb Z)$ such that $\rho_{A}(\xi)(a) = a, \ \forall \ a \in A,$ and $\rho_{A}(\xi)(u) = \xi u$. Then let $\rho(\xi) = \rho_{A}(\xi) \otimes id_{B\times_{\beta}\Bbb Z}$ and it is easy to check that $\rho(\xi)(D) = D$ and $\rho(\xi)|_{D}$ satisfies the desired properties. $\Box$ [**Section 2: The Rohlin Property**]{} We now state the definition of the Rohlin property in the unital case. The appropriate definition in the non-unital case can be found in \[EK\]. [**Definition 2.1**]{}   If $B$ is unital and $\beta \in {\operatorname{Aut}}(B)$ then $\beta$ satisfies the Rohlin property if for every $k \in \Bbb N$ there are positive integers $k_{1}, \ldots , k_{m} \geq k$ satisfying the following condition:  For every finite subset, $\cal F$ $\subset B$, and every $\epsilon >$ 0, there exist projections $e_{i,j}, \ i = 1, \ldots , m, \ j = 0, \ldots , k_{i} - 1$ in $B$ such that $\sum_{i=1}^{m} \sum_{j=0}^{k_{i} - 1} e_{i,j} = 1_{B}$ $\| \beta(e_{i,j}) - e_{i,j+1} \| < \epsilon$ $\| [x,e_{i,j}] \| < \epsilon $ for $i = 1, \ldots , m, \ j = 0, \ldots , k_{i} - 1$ and all $x \in \cal F$ (where $e_{i,k_{i}} = e_{i,0}$). For each integer $i$ above, the projections $\{e_{i,j}\}_{j = 0}^{k_i - 1}$ are called a [*Rohlin tower*]{}. [**Example 2.2**]{}   It follows from Theorem 1.3 in \[Ki2\] that every UHF algebra admits an automorphism with the Rohlin property. We now give a concrete example of such an automorphism which will be used repeatedly. Let $M_n$ be the $n\times n$ matrices over $\Bbb C$ with canonical matrix units $E_{i,j}^{(n)}$. Now let $u_n \in M_n$ be the unitary matrix such that $Adu_n (E_{i,i}^{(n)}) = E_{i + 1, i + 1}^{(n)}$ (with addition modulo $n$). Now consider the Universal UHF algebra $\cal U = \otimes_{n \geq 1} M_n$ and the automorphism $\sigma = \otimes_{n \geq 1} Adu_n \in {\operatorname{Aut}}(\cal U)$. We claim that $\sigma$ has the Rohlin property. It is important to note that the integers $k_1, \ldots, k_m$ in the definition of the Rohlin property must be fixed and must work for all finite subsets and for all choices of $\epsilon$. We will show that for this particular example, one may always take $m = 1$ and $k_1 = k$. So, let $\cal F \subset \cal U$ be a finite subset and $\epsilon > 0$ be given. Now, take some large $m \in \Bbb N$ such that $\cal F$ is nearly contained (within $\epsilon /2$ will suffice) in $ \otimes_{n = 1}^{m}M_n \subset \cal U$. Now take $m^{\prime} > m$ such that $k$ divides $m^{\prime}$, say $m^{\prime} = sk$. Then let $e_j = \sum_{t = 0}^{s - 1} E_{j + tk, j + tk}^{(m^{\prime})} \in \cal U$ (where we have identified $M_{m^{\prime}}$ with it’s unital image in $\cal U$), for $1 \leq j \leq k$. Evidently we have $\sigma(e_j) = e_{j + 1}$, $\sum_{j = 1}^{k} e_j = 1_{\cal U}$ and the $e_j$’s nearly commute with $\cal F$ since they commute with $\otimes_{n = 1}^{m} M_n$ by construction. We would also like to point out that $\cal U\times_{\sigma}\Bbb Z$ is AF embeddable (and actually embeds back into $\cal U$) by Lemma 2.8 in \[Vo\]. In fact every crossed product of a UHF algebra (by $\Bbb Z$) is AF embeddable since every automorphism of a UHF algebra is approximately inner (cf. Theorem 3.6 in \[Vo\]). However, this particular example is also limit periodic (see Definition 5.1) and thus we do not need the full power of Theorem 3.6 in \[Vo\] to deduce AF embeddability. [**Remark 2.3**]{}   If $B_0 \subset B$ is a unital, finite dimensional subalgebra and $\Psi : B \rightarrow B_{0}^{\prime} \cap B$ is a conditional expectation then it is readily verified that $\| \Psi(b) - b \|$ is small whenever $b$ almost commutes with the set of matrix units from which $\Psi$ is constructed. Thus, if the finite subset, $\cal F$, in the definition of the Rohlin property is taken to be the matrix units of $B_0$ then we have that the projections, $\{e_{i,j}\}$, are nearly contained in $B_{0}^{\prime} \cap B$ and so we can assume without loss of generality that $\{e_{i,j}\} \subset B_{0}^{\prime} \cap B$. (see Lemma III.3.1 in \[Da\]). More generally, if $\cal F$ is contained in a finite dimensional subalgebra then the Rohlin towers may be chosen to commute with $\cal F$. Throughout this section we will mainly be concerned with unital algebras. However, this is only out of convenience and it should be noted that all of the results in this section have analogues in the non-unital case (cf. \[EK\]). We now state a theorem due to Evans and Kishimoto that illustrates the usefulness of the Rohlin property when dealing with crossed products. Actually, Theorem 4.1 in \[EK\] is much stronger than the following. [**Theorem 2.4**]{}(Evans and Kishimoto)   If $A$ is unital, $\alpha, \beta \in {\operatorname{Aut}}(A)$, $\alpha_{*} = \beta_{*}$ on $K_0(A)$, and both $\alpha$ and $\beta$ satisfy the Rohlin property, then $A\times_{\alpha}\Bbb Z \cong A\times_{\beta}\Bbb Z$. [**Remark 2.5**]{}   The following corollary follows immediately from condition (4) of Theorem 0.2 (even without the assumption that $A$ is unital). However, it is easily verified that if $\alpha$ has the Rohlin property then $\alpha\otimes\beta$ does also. Thus using Theorem 2.4 and Lemma 1.3 we may easily prove homotopy invariance of AF embeddability. [**Corollary 2.6**]{}   If $A$ is unital and $\alpha, \beta \in {\operatorname{Aut}}(A)$ are homotopic then $A\times_{\alpha}\Bbb Z$ is AF embeddable $\Longleftrightarrow A\times_{\beta}\Bbb Z$ is AF embeddable. [**Proof**]{}    Let $\cal U$ be the Universal UHF algebra and $\sigma \in {\operatorname{Aut}}(\cal U)$ be the automorphism with the Rohlin property described in Example 2.2. Then $\alpha\otimes \sigma$ is homotopic to $\beta\otimes \sigma$. Hence, from Theorem 2.4 we have $A\otimes \cal U\times_{\alpha\otimes \sigma}\Bbb Z \cong A\otimes \cal U\times_{\beta\otimes \sigma}\Bbb Z$ The conclusion now follows from Lemma 1.3 since the crossed product, $\cal U\times_{\sigma}\Bbb Z$ is AF embeddable. $\Box$ The following stabilization lemma was essentially proved in \[EK\] (without the assumption of a unit). Our formulation is somewhat different, but our proof is just a more detailed version of that given in \[EK\]. [**Lemma 2.7**]{}   Let $B$ be unital , $\beta \in {\operatorname{Aut}}(B)$ have the Rohlin property and $k_{1} \geq k_{2} \geq \ldots \geq k_{m} \geq k$ be integers satisfying the definition of the Rohlin property. If $B_{0} \subset B_{1} \subset B_{2}$ are finite dimensional subalgebras of $B$ with $1_{B} \in B_{i}, \ i=0,1,2$ and $\beta^{-j}(B_{i}) \subset B_{i+1}, \ \ $for $ \ i=0,1 \ $ and $ 0 \leq j \leq k_{1}$ then for every unitary, $u \in B_{2}^{\prime} \cap B$, there exists a unitary, $v \in B_{0}^{\prime} \cap B$ such that $\| u - v\beta(v^{*}) \| \leq 4/(k - 1)$. [**Proof**]{}  (cf. \[EK\], Lemma 3.2)  Let $u \in B_{2}^{\prime} \cap B$ be given. Then define\ $\tilde{u}_{0} = 1_B, \ \tilde{u}_{1} = u, \ \tilde{u}_{j} = u\beta(u)\cdots\beta^{j-1}(u),$   for $j \geq 2$ Notice that our hypotheses imply that for $0 \leq j \leq k_{1}$ we have $\tilde{u}_{j} \in B_{1}^{\prime} \cap B$ and for every unitary $w \in B_{1}^{\prime} \cap B$ we have that $\beta^{j}(w) \in B_{0}^{\prime} \cap B, \ j = 0, \ldots , k_{1}$. Now, since $B_{1}^{\prime} \cap B$ is a unital AF subalgebra of $B$ (cf. Exercise 7.7.5 in \[Bl\]), we can find a unital, finite dimensional subalgebra, $B_{3} \subset B_{1}^{\prime} \cap B$, such that there exist unitaries, $u_{i} \in B_{3}$ with $\| u_{i} - \tilde{u}_{i} \| < \epsilon$, for some small $\epsilon > 0$. Then let $w^{(t)}_{k_{i}}, i = 1, \ldots , m$, be paths of unitaries in $ B_{3}$ such that $w^{(0)}_{k_{i}} = 1_{B} \ \ \ \ \ , \ \ \ \ \ w^{(1)}_{k_{i}} = u_{k_{i}}$ $\| w^{(s)}_{k_{i}} - w^{(t)}_{k_{i}} \| \leq \pi|s - t| \ , \ s,t \in [0,1]$ We have already observed that $\beta^{j}(w^{(t)}_{k_{i}}) \in B_{0}^{\prime} \cap B$ for $0 \leq j \leq k_1$. So, (by Remark 2.3) take a set of Rohlin towers, $\{e_{i,j}\} \subset B_{0}^{\prime} \cap B$. We may also assume without loss of generality that the Rohlin towers approximately commute (with error also bounded by $\epsilon$) with the $\tilde{u}_j$’s, $\beta(\tilde{u}_j)$’s and $\beta^{j}(w^{(1 - \frac{j}{k_i - 1})}_{k_i})$’s below. So we define $$v^{\prime} = \sum_{i=1}^{m} \sum_{j=0}^{k_{i} - 1} \tilde{u}_{j}\beta^{j} (w^{(1 - \frac{j}{k_{i} - 1})}_{k_{i}})e_{i,j}$$ Now, $v^{\prime}$ is not a unitary. However, it is easy to verify that $\| v^{\prime}v^{\prime*} - 1_{B} \| < \epsilon C$ and $\| v^{\prime*}v^{\prime} - 1_B \| < \epsilon C$, where $C$ is a constant depending only on $k_1, \ldots, k_m$. For the remainder of the proof, $C$ will always denote a constant depending only on $k_1, \ldots, k_m$. We make no effort to keep track of the best constants as this detail is not needed for the proof. So we have that $v^{\prime}$ is invertible and hence the unitary, $v$, in it’s polar decomposition actually lives in $B_{0}^{\prime} \cap B$ (since $v^{\prime} \in B_{0}^{\prime} \cap B$ also). However, the same estimates imply that $v^{\prime}$ is also close to this unitary since $|v^{\prime}|$ will be close to $1_B$. More precisely, if we write $v^{\prime} = v|v^{\prime}|$ then from elementary spectral theory we have $\| v - v^{\prime} \| < 1 - \sqrt{1 + \epsilon C}$. Notice that the definition of the Rohlin property implies the following estimates. $\| \beta(e_{i,j})e_{i,j + 1} - e_{i, j + 1} \| \leq \epsilon$ $\| \beta(e_{i,j})e_{k,l} \| \leq \epsilon$,  unless $k = i$ and $l = j + 1$ Thus a straightforward calculation yields $$\| v^{\prime}\beta(v^{\prime*}) - \sum_{i = 1}^{m}\sum_{j = 0}^{k_i - 1}\tilde{u}_{j}\beta^{j}(w_{k_i}^{(1 - \frac{j}{k_i - 1})})e_{i,j}\beta^{j}(w_{k_i}^{(1 - \frac{j - 1}{k_i - 1})*}) \beta(\tilde{u}_{j - 1}^{*}) \| \leq \epsilon C$$ where for each $i$, the $j = 0$ term in the sum above is actually given by $\tilde{u}_0 u_{k_i}e_{i,0}\beta^{k_i}(w_{k_i}^{(1 - \frac{k_i - 1}{k_i - 1})*})\beta(\tilde{u}_{k_i - 1}^{*}) = u_{k_i}e_{i,0}\beta(\tilde{u}_{k_i - 1}^{*}) \approx \tilde{u}_{k_i} e_{i,0}\beta(\tilde{u}^{*}_{k_i - 1})$ The following observations, will be needed to get the desired estimate. 1. $\| e_{i,j}\beta^{j}(w_{k_i}^{(1 - \frac{j - 1}{k_i - 1})*}) \beta(\tilde{u}_{j - 1}^{*}) - \beta^{j}(w_{k_i}^{(1 - \frac{j - 1}{k_i - 1})*}) \beta(\tilde{u}_{j - 1}^{*})e_{i,j} \| \leq 2\epsilon$ 2. $\| \beta^{j}(w_{k_i}^{(1 - \frac{j}{k_i - 1})}) \beta^{j}(w_{k_i}^{(1 - \frac{j -1}{k_i - 1})*}) - 1_{B} \| \leq \frac{\pi}{k_i - 1}$ <!-- --> 1. $\tilde{u}_j \beta(\tilde{u}_{j - 1}^{*}) = u$  for $j \geq 1 $ Using the estimates above, we have the following approximations. $$\begin{aligned} v^{\prime}\beta(v^{\prime*}) & \stackrel{\epsilon C}{\approx} & \sum_{i = 1}^{m}\sum_{j = 0}^{k_i - 1}\tilde{u}_{j}\beta^{j}(w_{k_i}^{(1 - \frac{j}{k_i - 1})})e_{i,j}\beta^{j}(w_{k_i}^{(1 - \frac{j - 1}{k_i - 1})*}) \beta(\tilde{u}_{j - 1}^{*}) \\ & \stackrel{\epsilon C}{\approx} & \sum_{i = 1}^{m}\sum_{j = 0}^{k_i - 1}\tilde{u}_{j}\beta^{j}(w_{k_i}^{(1 - \frac{j}{k_i - 1})})\beta^{j}(w_{k_i}^{(1 - \frac{j - 1}{k_i - 1})*}) \beta(\tilde{u}_{j - 1}^{*})e_{i,j} \ \ \ \ by \ \ (a) \\ & \stackrel{\frac{\pi}{k - 1} + \epsilon C}{\approx} & \sum_{i = 1}^{m}\sum_{j = 0}^{k_i - 1}\tilde{u}_{j} \beta(\tilde{u}_{j - 1}^{*})e_{i,j} \hspace{50mm} by \ \ (b) \\ & \stackrel{\epsilon C}{\approx} & u \hspace{80mm} by \ \ (c). \end{aligned}$$ Recall that the only thing keeping the last approximation from being an equality is the fact that for each $i$, the $j = 0$ term above is actually given by $u_{k_i} \beta (\tilde{u}_{k_i - 1}^{*})e_{i,0} \stackrel{\epsilon}{\approx} \tilde{u}_{k_i} \beta(\tilde{u}_{k_i - 1}^{*})e_{i,0} = ue_{i,0}$. Finally, combining all of these estimates we have that $\| u - v^{\prime}\beta(v^{\prime*}) \| < \frac{\pi}{k - 1} + \epsilon C$ and hence $\| u - v\beta(v^{*}) \| < \frac{\pi}{k - 1} + \epsilon C + 2(1 - \sqrt{1 + \epsilon C})$, where we have the liberty to take $\epsilon$ as small as we like. $\Box$ We now turn to the main tool of this paper. The following proposition is the key to Proposition 3.1, which in turn is the key to the rest of the embedding results presented here. [**Proposition 2.8**]{}   Let $A, B$ be unital, $\alpha \in {\operatorname{Aut}}(A)$, $\beta \in {\operatorname{Aut}}(B)$ and $\varphi: A \rightarrow B$ a unital, \*-monomorphism. Assume also that $\beta$ satisfies the Rohlin property and $\beta_{*} \circ \varphi_{*} = \varphi_{*} \circ \alpha_{*}$, i.e. that we have commutativity in the diagram $K_{0}(A) \stackrel{\varphi_{*}}{\longrightarrow} K_{0}(B)$ $\alpha_{*}$ [$\downarrow$]{}             $\beta_{*}$ $K_{0}(A) \stackrel{\varphi_{*}}{\longrightarrow} K_{0}(B)$ Then there exist unitaries, $v \in A, u \in B$, and a unital \*-monomorphism,\ $\varphi^{\prime}: A \rightarrow B$, with $\varphi_{*}^{\prime} = \varphi_{*}$ and commutativity in $A \stackrel{\varphi^{\prime}}{\longrightarrow} B$ [$Adv \circ \alpha$]{} [$\downarrow$]{}        [$\downarrow$]{} [$Adu \circ \beta$]{} $A \stackrel{\varphi^{\prime}}{\longrightarrow} B$ [**Proof**]{}   Let $\{ A_{i} \}_{i=0}^{\infty}$ be an increasing nest of finite dimensional subalgebras whose union is dense in $A$ and $A_{0} = \Bbb C 1_{A}$. Then $\{ \alpha(A_{i}) \}$ is a nest with the same properties and thus we can find a unitary $v \in A$ (with $v$ as close to $1_{A}$ as we like) such that (see Theorem III.3.5 in \[Da\]) $Adv\circ\alpha(\bigcup A_{i}) = \bigcup A_{i}$ To ease our notation somewhat, we will henceforth denote $Adv\circ\alpha$ also by $\alpha$ (and just remember that the $v$ in the statement of the proposition has already been chosen). Now, let $m_i = 2^{i + 3} + 1$ and applying the definition of the Rohlin property to $m_i$ we get a finite set of integers, $\{k_{1}^{(i)}, \ldots, k_{s_i}^{(i)}\}$, with $k_{t}^{(i)} \geq m_i$ for $1 \leq t \leq s_i$. Then let $m_{i}^{\prime} = max\{k_{1}^{(i)}, \ldots, k_{s_i}^{(i)}\}$. Since we have arranged that $\bigcup\alpha(A_{i}) = \bigcup A_{i}$, by passing to a subsequence we may further assume that for $i \geq 2$ we have that $\alpha^{-j}(A_{i - 1}) \subset A_{i} , \ 0 \leq j \leq m_{i}^{\prime}$ $\alpha^{-j}(A_{i - 2}) \subset A_{i - 1} , \ 0 \leq j \leq m_{i}^{\prime}$ We will now inductively construct sequences of unitaries, $u_{i} \in B$, automorphisms, $\beta_{i} \in {\operatorname{Aut}}(B)$, and unital \*-monomorphisms, $\varphi_{i}: A \rightarrow B$, with the following properties 1. $\| u_{i} - 1_{B} \| \leq 1/2^{i}$ 2. $\beta_{i + 1} = {\rm Adu}_{i + 1}\circ\beta_{i}$ 3. $\varphi_{i*} = \varphi_{*}$ 4. $\varphi_{i + 1}|_{\alpha(A_{i - 2})} = \varphi_{i}|_{\alpha(A_{i - 2})} \ $ where $ \ \Bbb C 1_{A} = A_{0} = A_{-1} = A_{-2}$ 5. $\beta_{i}\circ\varphi_{i}|_{A_{i}} = \varphi_{i}\circ\alpha|_{A_{i}}$ where $u_{0} = 1_{B}, \ \varphi_{0} = \varphi$ and $\beta_{0} = \beta$. Note that each $\beta_i$ will also have the Rohlin property. So, assume that we have found the desired $u_{i}, \ \beta_{i}, \ \varphi_{i}$ for $0 \leq i \leq n$ and we will show how to construct $u_{n + 1}, \ \beta_{n + 1}$ and $ \varphi_{n + 1}$. By construction, we have $\beta_{n*}\circ\varphi_{n*} = \varphi_{n*}\circ\alpha_{*}$ and thus restricting to $A_{n + 1}$ we have that $\beta_{n*}\circ\varphi_{n*}$ and $\varphi_{n*}\circ\alpha_{*}$ agree as maps from $K_{0}(A_{n + 1}) \rightarrow K_{0}(B)$. Thus, (by Fact 1.2) we can find a unitary $w_{n + 1} \in B$ such that $Adw_{n + 1}\circ\beta_{n}\circ\varphi_{n}|_{A_{n + 1}} = \varphi_{n}\circ\alpha|_{A_{n + 1}}$ But, by assumption, we already have $\beta_{n}\circ\varphi_{n}|_{A_{n}} = \varphi_{n}\circ\alpha|_{A_{n}}$ and hence $w_{n + 1} \in \varphi_{n}(\alpha(A_{n}))^{\prime} \cap B$. Now, since we have control of the iterates of $A_{n - 2}$ and $A_{n - 1}$ (under $\alpha$) we claim that $\beta_{n}^{-j}(\varphi_{n}(\alpha(A_{n - 1}))) \subset \varphi_{n}(\alpha(A_{n})) \ \ , \ \ 0 \leq j \leq m_{n}^{\prime}$ $\beta_{n}^{-j}(\varphi_{n}(\alpha(A_{n - 2}))) \subset \varphi_{n}(\alpha(A_{n - 1})) \ \ , \ \ 0 \leq j \leq m_{n}^{\prime}$ To see this we first note that since $\alpha^{-1}(A_{n - 1}) \subset A_n$, $(5)$ implies $\beta^{-1}_{n}(\varphi_{n}(\alpha(A_{n -1}))) = \varphi_{n}(A_{n - 1}) \subset \varphi(\alpha(A_n))$ Similarly we have $$\begin{aligned} \beta^{-2}_n(\varphi_n(\alpha(A_{n - 1}))) & = & \beta^{-1}(\beta^{-1}(\varphi_n(\alpha(A_{n - 1}))))\\ & = & \beta^{-1}(\varphi_n(A_{n - 1}))\\ & = & \beta^{-1}(\varphi_n(\alpha(\alpha^{-1}(A_{n - 1}))))\\ & = & \varphi_n(\alpha^{-1}(A_{n - 1})) \ \ \ \ ( since \ \ \alpha^{-1}(A_{n - 1}) \subset A_n)\end{aligned}$$ but, $\alpha^{-2}(A_{n - 1}) \subset A_n$ implies that $\varphi_n(\alpha^{-1}(A_{n - 1})) \subset \varphi_n(\alpha(A_n))$. By repeating the above argument, one can show that $\beta_{n}^{-j}(\varphi_n(\alpha(A_{n - 1}))) = \varphi_n(\alpha^{-j + 1}(A_{n - 1}))$ for $0 \leq j \leq m_{n}^{\prime}$. But then $\alpha^{-j}(A_{n -1}) \subset A_n$ implies that $\varphi_n(\alpha^{-j + 1}(A_{n -1})) \subset \varphi_n(\alpha(A_n))$, for $0 \leq j \leq m_{n}^{\prime}$. The same argument works for the iterates of $\varphi_n(\alpha(A_{n - 2}))$ as well. Thus, by Lemma 2.7, we may take a unitary, $v_{n + 1} \in \varphi_{n}(\alpha(A_{n - 2}))^{\prime} \cap B$ such that $\| w_{n + 1} - v_{n + 1}\beta_{n}(v_{n + 1}^{*}) \| \leq \frac{4}{m_{n} - 1} = \frac{1}{2^{n + 1}}$. So, we define $\varphi_{n + 1} = Adv_{n + 1}^{*}\circ\varphi_{n}$, $ u_{n + 1} = v_{n + 1}^{*}w_{n + 1}\beta_{n}(v_{n + 1})$ $ \beta_{n + 1} = Adu_{n + 1}\circ\beta_{n}$ It is now easy to check that we have satisfied all of the required properties. The proof of the proposition is now complete as it is easy to see that $u = \lim_{n} u_{n}\cdots u_{1}$ is a well defined unitary in $B$. Also, it is easy to see (by condition 4) that the $\varphi_{i}$’s converge (in the point-norm topology) to a unital \*-monomorphism, $\varphi^{\prime}$, and clearly $Adu\circ\beta\circ\varphi^{\prime} = \varphi^{\prime}\circ\alpha$. $\Box$ [**Remark 2.9**]{}   It would be of independent interest to know if Proposition 2.6 can be proved without the assumption that $\beta$ satisfy the Rohlin property. However, the author was unable to either prove or provide a counterexample to such a claim. [**Corollary 2.10**]{}   Under the assumptions of Proposition 2.8 we also have that $A\times_{\alpha}\Bbb Z$ embeds into $B\times_{\beta}\Bbb Z$. [**Proof**]{}    Essentially what we have done in Proposition 2.8 is embedded $A$ into $B$ in such a way that the automorphism $Adv \circ \alpha$ extends to an automorphism (namely, $Adu \circ \beta$) of $B$. In general it is true that if $C, \ D$ are C$^{*}$-algebras with $C \subset D$, $\gamma \in {\operatorname{Aut}}(D)$ and $\gamma (C) = C$ then there is a natural inclusion $C\times_{\gamma|_{C}}\Bbb Z \hookrightarrow D\times_{\gamma}\Bbb Z.$ Thus the conclusion follows from the isomorphisms $A\times_{\alpha}\Bbb Z \cong A\times_{Adv\circ\alpha}\Bbb Z, \ \ B\times_{\beta}\Bbb Z \cong B\times_{Adu\circ\beta}\Bbb Z$ and the fact that $Adu \circ \beta$ is an extension of $Adv \circ \alpha$ under the embedding $\varphi^{\prime}$. $\Box$ [**Section 3: Characterizing AF Embeddability**]{} If $\varphi: A \rightarrow B$ is a \*-homomorphism, we will denote by $\tilde{A}$ and $\tilde{B}$ the C$^{*}$-algebras obtained by adjoining a (possibly new) unit and we will let $\tilde{\varphi}$ be the (unique) unital extension of $\varphi$. [**Proposition 3.1**]{}   Let $A$ and $B$ be given (not necessarily unital), $\alpha \in {\operatorname{Aut}}(A), \beta \in {\operatorname{Aut}}(B)$ and $\varphi: A \rightarrow B$ be a \*-monomorphism. Further assume that $\beta_{*}\circ\varphi_{*} = \varphi_{*}\circ\alpha_{*}$ and that $\tilde{B}\times_{\tilde{\beta}}\Bbb Z$ is AF embeddable. Then $A\times_{\alpha}\Bbb Z$ is also AF embeddable. [**Proof**]{}   Note that after unitizing everything we still have $\tilde{\beta_{*}}\circ\tilde{\varphi_{*}} = \tilde{\varphi_{*}}\circ\tilde{\alpha_{*}}$. Let $\cal U$ be the universal UHF algebra and $\sigma \in {\operatorname{Aut}}(\cal U$) be the automorphism (with the Rohlin property) defined in Example 2.2. Now we let $\tilde{\beta}^{\prime} = \tilde{\beta}\otimes \sigma$ and let $\tilde{\varphi}^{\prime}: A \rightarrow \tilde{B}\otimes \cal U$ be given by $\tilde{\varphi}^{\prime}(a) = \tilde{\varphi}(a)\otimes 1_{\cal U}$. Then $\tilde{\beta}^{\prime}$ satisfies the Rohlin property and it is easy to check that we still have commutativity at the level of K-theory, i.e. $\tilde{\beta}_{*}^{\prime}\circ\tilde{\varphi}_{*}^{\prime} = \tilde{\varphi}_{*}^{\prime}\circ\tilde{\alpha}_{*}$. Now, it is always true that $A\times_{\alpha}\Bbb Z$ $\hookrightarrow$ $\tilde{\rm A}\times_{\tilde{\alpha}}\Bbb Z$ and by Corollary 2.10 we have that $\tilde{A}\times_{\tilde{\alpha}}\Bbb Z$ $\hookrightarrow$ $\tilde{B}\otimes \cal U\times_{\tilde{\beta}^{\prime}}\Bbb Z$. But then by Lemma 1.3 we also have that $\tilde{B}\otimes \cal U\times_{\tilde{\beta}^{\prime}}\Bbb Z$ $\hookrightarrow$ ($\tilde{B}\times_{\tilde{\beta}}\Bbb Z)\otimes (\cal U\times_{\sigma}\Bbb Z$) where each of the algebras on the right hand side are AF embeddable. $\Box$ [**Corollary 3.2**]{}   If $\alpha \in {\operatorname{Aut}}(A)$ and $\theta : K_0 (A) \rightarrow G$ is a positive group homomorphism where $G$ is a dimension group (cf. \[Ef\]), $Ker (\theta) \cap \Gamma(A) = \{0\}$ and $H_{\alpha} \subset Ker(\theta)$ then $A\times_{\alpha}\Bbb Z$ is AF embeddable. [**Proof**]{}   Since there exists an AF algebra, $B$, with $K_0 (B) = G$ and $\Gamma(B) = G^{+}$, we have that $\theta$ is a faithful homomorphism and thus by Fact 1.1 we may find a \*-monomorphism, $\varphi : \ A \rightarrow B$, with $\varphi_{*} = \theta$. But we may now take $\beta \in {\operatorname{Aut}}(B)$ to be the identity and the conclusion follows from Proposition 3.1. $\Box$ [**Remark 3.3**]{}   The hypotheses of Corollary 3.2 are easily seen to be necessary as well. The proof of this necessity is essentially contained in the introduction. We nearly have the necessary tools to prove Theorem 0.2 now. However, for the implication (4) $\Rightarrow$ (1) we need to provide a candidate AF algebra for the desired embedding of $A\times_{\alpha}\Bbb Z$. The following key lemma due to Spielberg will give us our candidate. We will not prove this lemma (see Lemma 1.14 in \[Sp\]) but would like to point out that it depends on the Effros-Handelman-Shen Theorem (\[EHS\]). [**Lemma 3.4**]{}(Spielberg)  If $G$ is a dimension group and $H$ is a subgroup of $G$ with $H \cap G^{+} = \{0\}$, then there exists a positive group homomorphism $\theta: G \rightarrow G^{\prime}$ (where $G^{\prime}$ is also a dimension group) with 1. $H \subset Ker(\theta)$ 2. $Ker(\theta) \cap G^{+} = \{0\}$ [**Proof of Theorem 0.2**]{} $(1) \Rightarrow (2)$ is obvious (whether or not $A$ is unital). $(2) \Rightarrow (3)$ If $A$ is unital then so is $A\times_{\alpha}\Bbb Z$ and thus quasidiagonality implies that every isometry in $A\times_{\alpha}\Bbb Z$ is actually a unitary. But this implies that the identity of $A\times_{\alpha}\Bbb Z$ is not equivalent to any proper subprojection and hence $A\times_{\alpha}\Bbb Z$ is finite. Since matrix algebras over quasidiagonal algebras are again quasidiagonal, the same argument shows that $A\times_{\alpha}\Bbb Z$ is stably finite. If $A$ is non-unital and $A\times_{\alpha}\Bbb Z$ is quasidiagonal then so is $\widetilde{A\times_{\alpha}\Bbb Z}$. Thus by the above argument we have that $\widetilde{A\times_{\alpha}\Bbb Z}$ is stably finite and thus $A\times_{\alpha}\Bbb Z$ inherits this property also. $(3) \Rightarrow (4)$ We have already shown this in the introduction when $A$ is unital. However, the following is a very elementary argument which does not depend on the Pimsner-Voiculescu six term exact sequence and which actually holds for much general algebras than just AF algebras. Since $A$ has a countable approximate unit consisting of projections, we do not need to pass to the unitization of $A$ when computing $K_0 (A)$ (see Proposition 5.5.5 in \[Bl\]). Now, assume that $H_{\alpha} \cap K_{0}^{+}(A) \supsetneqq \{0\}$, i.e. there is some $x \in K_0 (A)$ and some projection, $r \in M_n (A)$ such that $\alpha_{*} (x) - x = [r] \neq 0$. Now, write $x = [p] - [q]$, where (without loss of generality) $p, \ q \in M_n (A)$ are projections. We will not keep track of the sizes of matrices that we are dealing with since this is not important for our argument. So, rewriting the equation $\alpha_{*} (x) - x = [r]$ we get $[\alpha (p)] + [q] = [p] + [\alpha (q)] + [r]$ Now, since AF algebras enjoy cancellation, we can find a partial isometry (in the matrices over A), $v$, with support projection $diag(p, \ \alpha(q), \ r)$ and range projection $diag(\alpha (p), \ q, \ 0)$. If we now move to $A\times_{\alpha}\Bbb Z$ and let $u$ be the distinguished unitary (i.e. $uau^{*} = \alpha (a), \forall \ a \in A$) then it is easy to verify that $diag(pu^{*}, \ uq, \ 0)$ is a partial isometry with support projection $diag(\alpha (p), \ q, \ 0)$ and range projection $diag(p, \ \alpha (q), \ 0)$. Thus multiplying these two partial isometries we get a partial isometry (in the matrices over $A\times_{\alpha}\Bbb Z$) from $diag(p, \ \alpha (q), \ r)$ to the proper subprojection $diag(p, \ \alpha(q), \ 0)$ and thus $A\times_{\alpha}\Bbb Z$ is not stably finite. $(4) \Rightarrow (1)$ Letting $H = H_{\alpha}$ in Lemma 3.4, this implication now follows from Corollary 3.2. $\Box$ [**Remark 3.5**]{}   Note that the implications $(1) \Rightarrow (2) \Rightarrow (3)$ hold for any C$^{*}$-algebra, $A$, while the only property of AF algebras that we really used in the implication $(3) \Rightarrow (4)$ was the fact that one need not pass to the unitization of $A$ when computing $K_0(A)$ (i.e. cancellation is not necessary in the argument above). [**Section 4: Applications**]{} We now introduce a simple K-theoretical condition which is sufficient to ensure AF embeddability of the corresponding crossed product. This condition gives a generalization of all the embedding theorems in \[Vo\] as all of the automorphisms considered there satisfy the following. [**Definition 4.1**]{}   If $\alpha \in {\operatorname{Aut}}(A)$ then $\alpha$ satisfies the [*finite orbit property*]{}, (FOP), if for every $x \in K_{0}(A)$ there exists an integer, $0 \neq n \in \Bbb N$, such that $\alpha_{*}^{n}(x) = x.$ [**Theorem 4.2**]{}   If A (not necessarily unital) is given, $\alpha \in {\operatorname{Aut}}(A)$ and $\alpha$ satisfies (FOP) then $A\times_{\alpha}\Bbb Z$ is AF embeddable. [**Proof**]{}   We will show the contrapositive. If $A\times_{\alpha}\Bbb Z$ is not AF embeddable then $H_{\alpha} \cap K_{0}^{+} (A) \neq \{0\}$. So, let $x \in K_0(A)$ be chosen so that $0 \neq \alpha_{*}(x) - x \in K_{0}^{+}(A)$. Note that since $\alpha_{*}$ is an isomorphism and $K_{0}^{+}(A) \cap (-K_{0}^{+}(A)) = 0$ we have the following two facts. $i) If \ 0 \neq [p] \in K_{0}^{+}(A) \ then \ 0 \neq \alpha_{*}([p]) = [\alpha(p)] \in K_{0}^{+}(A)$ $ii) If \ 0 \neq [p], [q] \in K_{0}^{+}(A) \ then \ 0 \neq [p] + [q] \in K_{0}^{+}(A)$ Thus we have that $0 \neq \alpha_{*}^{2}(x) - \alpha_{*}(x)$ and hence $0 \neq (\alpha_{*}^{2}(x) - \alpha_{*}(x)) + (\alpha_{*}(x) - x) = \alpha_{*}^{2}(x) - x \in K_{0}^{+}(A)$. Arguing similarly we have that $0 \neq \alpha_{*}^{j}(x) - x$, for all nonzero $j \in \Bbb N$. Thus $\alpha$ does not satisfy (FOP). $\Box$ [**Remark 4.3**]{}   Recall that Theorem 3.6 in \[Vo\] states that if there exists a nonzero integer, $n$, such that $\alpha_{*}^{n} = id_{A*}$ then the corresponding crossed product is AF embeddable. The following example shows that Theorem 4.2 is a generalization of this result. Let $X = \{1, 1/2, 1/3, 1/4, \ldots\} \cup \{0\}$ and let $\varphi: X \rightarrow X$ be the homeomorphism which leaves 0 fixed, interchanges 1 and 1/2, cyclicly permutes 1/3, 1/4 and 1/5, cyclicly permutes 1/6, 1/7, 1/8 and 1/9 and so on (with increasing lengths of cycles). Then it is easy to see that $\varphi$ satisfies (FOP), but there is no power of $\varphi$ which is approximately inner since the only inner automorphism is the identity map. It is not hard to show that (FOP) is not a necessary condition for AF embeddability (see Remark 4.8). However, automorphisms satisfying (FOP) do admit a nice characterization. It should be pointed out that the following proposition will not be needed in the remainder of this paper. Note that $\alpha$ satisfies (FOP) if and only if $\tilde{\alpha}$ satisfies (FOP) and thus there is no harm in passing to unitizations when dealing with this property. Abusing notation a little, we will also denote by $\alpha_{*}$ the image of an automorphism in the quotient group ${\operatorname{Aut}}(A)_{*} = {\operatorname{Aut}}(A)/\overline{Inn(A)}$. We now show that (FOP) characterizes all the well behaved automorphisms in ${\operatorname{Aut}}(A)_{*}$. [**Proposition 4.4**]{}   Let $A$ be unital and $\alpha \in {\operatorname{Aut}}(A)$. Then the following are equivalent: 1. $\alpha$ satisfies (FOP) 2. There exists a sequence, $n_{k} > 0$, such that $d(\alpha^{n_{k}}, Inn(A)) \rightarrow$ 0, as $k \rightarrow \infty$ (where $d$ is any metric on ${\operatorname{Aut}}(A)$ which induces the point-norm topology). 3. The sequence $\{\alpha_{*}^{n}\}_{n \geq 0}$ has a convergent subsequence in ${\operatorname{Aut}}(A)_{*}$. [**Proof**]{}   (1) $\Rightarrow$ (2)   Let $\{A_{i}\}$ be any increasing nest of finite dimensional subalgebras whose union is dense in $A$. It suffices to show that for each $i \in \Bbb N$ we can find integers $n_{i}$ (with $n_{i} < n_{i + 1}$) and unitaries $u_{i} \in A$ such that $\alpha^{n_{i}}|_{A_{i}} = Adu_{i}|_{A_{i}}$. So let $e_{1}, \ldots , e_{m_{i}}$ be minimal projections in $A_{i}$ such that $\{ [e_{1}], \ldots, [e_{m_{i}}] \}$ generate $K_{0}(A_{i})$. Now, take an integer $n_{i}$ such that $\alpha^{n_{i}}_{*}$(\[$e_{j}$\]) = \[$e_{j}$\], for $1 \leq j \leq m_{i}$ (note that $n_{i}$ may be chosen larger than any specified number). Thus, the restriction of $\alpha^{n_{i}}$ to $A_{i}$ agrees on K-theory with the natural inclusion map $A_{i} \hookrightarrow A$. Thus we may find the desired unitary , $u_{i} \in A$. \(2) $\Rightarrow$ (3)   Clearly $\alpha_{*}^{n_{i}} \rightarrow id_{A*}$ \(3) $\Rightarrow$ (4)   Assume there exists a subsequence $\{m_{k}\}$ and an element $\beta_{*} \in {\operatorname{Aut}}(A)_{*}$ such that $\alpha_{*}^{m_{k}} \rightarrow \beta_{*}$. Then defining $n_{k} = m_{k} - m_{k - 1}$ we have that $\alpha_{*}^{n_{k}} = \alpha_{*}^{m_{k}} \circ \alpha_{*}^{-m_{k - 1}} \rightarrow \beta_{*} \circ \beta_{*}^{-1} = id_{A*}$ Since $\Gamma(A)$ generates $K_0(A)$, it suffices to check (FOP) on this set. So, let $p \in A$ be any projection. Then let $\{a_{i}\}$ be any sequence which is dense in the unit ball of A, with $a_{1}$ = p. Now recall that the metric on Aut($A$) is defined as follows $$d(\sigma,\gamma) = \sum_{n = 1}^{\infty} \frac{1}{2^{n}} \| \sigma(a_{i}) - \gamma(a_{i}) \| , \ \ {\rm for} \ \sigma, \gamma \in {\operatorname{Aut}}(A)$$ where choosing a different sequence just gives an equivalent metric (cf. \[Ar\]). Now, let $U_{1/2}$ = $\{ \sigma \in {\operatorname{Aut}}(A): \text{there exists} \ \gamma \in Inn(A) \text{ such that} \ d(\sigma,\gamma) < 1/2\}$. Note that if $\sigma \in U_{1/2}$ then $\sigma_{*}([p]) = [p]$, since $a_{1} = p$ and $\| \sigma(p) - \gamma(p) \| < 1$ for some $\gamma \in Inn(A)$. Now, recall that by definition of the topology on ${\operatorname{Aut}}(A)_{*}$ (cf. \[Po\]) we have that $U_{1/2 *} = \{ \sigma_{*} : \sigma \in U_{1/2} \}$ is an open set in ${\operatorname{Aut}}(A)_{*}$ (obviously containing $id_{A*}$). Thus, there exists a $k \in \Bbb N$ large enough that $\alpha_{*}^{n_{k}} \in U_{1/2 *}$. Thus, there exists $\sigma \in U_{1/2}$ with $\alpha_{*}^{n_{k}} = \sigma_{*}$ and hence $\alpha_{*}^{n_{k}}([p]) = \sigma_{*}([p]) = [p]$. $\Box$ We will now show how to recover (the AF case of) a result of Pimsner on the AF embeddability of crossed products of commutative C$^{*}$-algebras. (See \[Pi\]) [**Definition 4.5**]{} If $X$ is a compact, metrizable space and $\varphi$ is a homeomorphism of $X$. Then $x \in X$ is called [*pseudo-nonwandering*]{} if for every $\epsilon > 0$, there exists a set of points $x_0, \ldots, x_{n + 1}$ with $x = x_0 = x_{n + 1}$ and $d(\varphi(x_i),x_{i + 1}) < \epsilon$ for $0 \leq i \leq n$. The set of all pseudo-nonwandering points will be denoted by $X(\varphi)$. [**Lemma 4.6**]{}(Pimsner) If $\varphi: X \rightarrow X$ is a homeomorphism of the compact metrizable space, $X$, then $x \notin X(\varphi) \Leftrightarrow$ there exists an open set, $U \subset X$ such that $\varphi(\overline{U}) \subset U$ and $x \in U\backslash\varphi(\overline{U})$. As we are interested in the case when $C(X)$ is AF, we will assume from now on that $X$ is also totally disconnected. Then we can find a basis of clopen sets, say $\{V_n\}$. We now claim that in this case, the open set $U$ in Lemma 4.2 can be taken to be clopen. To see this, we assume that there exists an open set $U \subset X$ satisfying the two properties of the lemma. Now, since $\varphi(\overline{U})$ is a compact subset of $U$, we can find a finite subset of the clopen basis, say $V_{n_1}, \ldots, V_{n_k}$, such that $x \notin \cup_{i = 1}^{k}V_{n_i}$\ $\varphi(\overline{U}) \subset \cup_{i = 1}^{k}V_{n_i} \subset U$ Then letting $V = \varphi^{-1}(\cup_{i = 1}^{k}V_{n_i})$, we get that $V$ is a clopen set with the same properties as $U$. Recall that when $X$ is totally disconnected, $K_0 (C(X)) = C(X, \Bbb Z)$, the group of continuous functions from $X$ to $\Bbb Z$, with positive cone given by the nonnegative functions. [**Theorem 4.7**]{}(Pimsner)  If $\varphi: X \rightarrow X$ is a homeomorphism of the compact, totally disconnected metric space $X$, then $C(X)\times_{\varphi}\Bbb Z$ is AF embeddable if and only if $X(\varphi) = X$, where the $\varphi$ appearing in the crossed product denotes the corresponding automorphism of $C(X)$, i.e. $\varphi(f) = f \circ \varphi^{-1}$. [**Proof**]{} $(\Rightarrow)$ We prove the contrapositive. So assume that $X(\varphi) \neq X$. Then by Lemma 4.6 and the discussion which follows , we can find a clopen set, $V$, such that $\varphi(V)$ is properly contained in $V$. So, if $P \in C(X)$ is the projection with support $V$, then we have that $P \circ \varphi^{-1}$ is a projection in $C(X)$ which is dominated by $P$. Thus as elements in $C(X, \Bbb Z)$ we have that $\varphi_{*}([P]) - [P]$ is a nonzero function in $-K_{0}^{+}(C(X))$ and hence $H_{\varphi} \cap K_{0}^{+}(C(X)) \neq 0$. Thus by Theorem 0.2 we have that $C(X)\times_{\varphi}\Bbb Z$ is not AF embeddable. $(\Leftarrow)$ Assume that $X(\varphi) = X$. Now take $f \in C(X, \Bbb Z)$ and assume that $f \circ \varphi^{-1} - f \geq 0$. Let $\{s_1, \ldots, s_k, -t_1, \ldots, -t_j\}$ be the range of $f$, where the $s_i, \ t_i$ are all nonnegative integers and $s_i > s_{i + 1}, \ t_l > t_{l + 1}$ (and perhaps $s_k = 0$). Then define the clopen sets $E_i = f^{-1}(s_i), \ F_l = f^{-1}(t_l)$ and notice that these sets form a pairwise disjoint, finite clopen cover of $X$. Letting $P_E$ denote the characteristic function of a clopen set, $E$, we can write $f = \sum_{i = 1}^{k}s_iP_{E_i} - \sum_{l = 1}^{j}t_lP_{F_l}$ and thus $$f \circ \varphi^{-1} - f = \sum_{i = 1}^{k} s_iP_{\varphi(E_{i})} + \sum_{l = 1}^{j} t_{l}P_{F_l} - \sum_{i = 1}^{k} s_iP_{E_{i}} - \sum_{l = 1}^{j} t_{l}P_{\varphi(F_j)}$$ Now, given $x \in E_1$ we have that the second summation above vanishes at $x$. Also, since $f \circ \varphi^{-1} - f \geq 0$, we have a unique index $i$ such that $f \circ \varphi^{-1}(x) - f(x) = s_i - s_1 \geq 0$ However, as $s_1 > s_i$, for $2 \leq i \leq k$, we conclude that $s_i = s_1$ and hence $\varphi(E_1) \supset E_1$. But this implies that $\varphi(E_{1}^{c}) \subset E_{1}^{c}$, where $E^{c}$ denotes the complement. But then by hypothesis (and Lemma 4.2) we see that $\varphi(E_1) = E_1$. Repeating this argument we get that $\varphi(E_i) = E_i$ for $1 \leq i \leq k$. An obvious adaptation of this argument shows equality for the $F_l$’s. Hence $f \circ \varphi^{-1} - f = 0$, i.e. $H_{\alpha} \cap K_{0}^{+} (C(X)) = 0$. $\Box$ [**Remark 4.8**]{}   We may now give a simple example showing that in general, (FOP) is not a necessary condition for AF embeddability. Let $X$ be the one point compactification of $\Bbb Z$ and define $\varphi : X \rightarrow X$ to be the homeomorphism taking $n \mapsto n + 1$ and $\infty \mapsto \infty$. As Pimsner points out in \[Pi\], every point of $X$ will be pseudo-nonwandering (and hence $C(X)\times_{\varphi}\Bbb Z$ will be AF embeddable) while it is not hard to see that every element of $K_0(C(X)) = C(X, \Bbb Z)$ will have infinite orbit under the iterates of $\varphi_{*}$. We now show that the existence of enough $\alpha_{*}$-invariant states is a sufficient condition for AF embeddability. For convenience we will assume that A is unital and, again following the terminology in \[Da\], we define a [*state*]{} on $K_0(A)$ to be a contractive group homomorphism, $\tau$: $K_0(A)$ $\rightarrow (\Bbb R,\Bbb R_{+}, [0, 1])$, with $\tau([1_{A}]) = 1$. There is a 1-1 correspondence between the states on $K_0(A)$ and the tracial states on the algebra, $A$ (Theorem IV.5.3 in \[Da\]). Recall that a (tracial) state is said to be [*faithful*]{} if the only positive element in it’s kernel is the zero element. We will denote by $S_{\alpha}$ the set of all [*$\alpha_{*}$-invariant*]{} states, i.e. those states for which $\tau\circ\alpha_{*} = \tau$. Note that when $A$ is unital, $S_{\alpha}$ is never empty (since $\Bbb Z$ is amenable). The author would like to thank Professors L.G. Brown and N.C. Phillips for pointing out a substantial simplification in the proof of the following theorem. [**Theorem 4.9**]{}   Let $A$ be unital and $\alpha \in {\operatorname{Aut}}(A)$ be given. Assume that for every projection, $p \in A$, there exists a state, $\tau \in S_{\alpha}$, such that $\tau([p]) > 0$. Then $A\times_{\alpha}\Bbb Z$ is AF embeddable. In particular, if $S_{\alpha}$ contains a faithful state or if $A\times_{\alpha}\Bbb Z$ admits a faithful tracial state then it is AF embeddable. [**Proof**]{}   Again, we prove the contrapositive. So assume that there exists an element $x \in K_0(A)$ and a projection $q \in M_n(A)$ such that $\alpha_{*}(x) - x = [q] \neq 0$. Then clearly $\tau([q]) = 0$, for every $\tau \in S_{\alpha}$. But since $\Gamma(A)$ generates $K_{0}^{+}(A)$, we can find projections $p_1, \ldots ,p_n \in A$ such that $[q] = [p_1] + \ldots + [p_n]$ and thus we see that $\tau([p_i]) = 0$ for every $\tau \in S_{\alpha}$ and $1 \leq i \leq n$. $\Box$ We are indebted to Kishimoto Sensei for pointing out the following corollary. [**Corollary 4.10**]{}   If $A$ is simple and unital then $A\times_{\alpha}\Bbb Z$ is AF embeddable for every $\alpha \in {\operatorname{Aut}}(A)$. [**Proof**]{}   This follows from Theorem 4.9 since every state on a simple dimension group is faithful.(cf. \[Ef\], Corollary 4.2) $\Box$ [**Remark 4.11**]{}   Corollary 4.10 shows the contrast between crossed products of unital and non-unital algebras. Indeed, it was first shown in \[Cu\] that the stabilizations of the Cuntz algebras, $\cal O_{n} \otimes \Bbb K$, are isomorphic to crossed products of (non-unital) simple AF algebras. Thus crossed products of non-unital simple AF algebras can be purely infinite and hence not AF embeddable. As a final application, we present a few more conditions which characterize AF embeddability. [**Theorem 4.12**]{}   If $A$ is an AF algebra (not necessarily unital) and $\alpha \in {\operatorname{Aut}}(A)$ then the following are equivalent 1. $A\times_{\alpha}\Bbb Z$ is AF embeddable 2. $\tilde{\rm A}\times_{\tilde{\alpha}}\Bbb Z$ is AF embeddable 3. For every $m \in \Bbb Z, A\times_{\alpha^{m}}\Bbb Z$ is AF embeddable 4. There exists an integer, $m \neq$ 0, such that $A\times_{\alpha^{m}}\Bbb Z$ is AF embeddable 5. There exists an AF algebra, $B$, and a group homomorphism\ $\theta: K_{0}( A\times_{\alpha}\Bbb Z) \rightarrow$ $K_0(B)$ with $i) \ \theta([p]) \in \Gamma(B)$ and $ii) \ [p] \in Ker(\theta) \Rightarrow p = 0$, for every projection $p \in A$. [**Proof**]{}   $(1)\Leftrightarrow(2)$ It is easy to see that $K_0(\tilde{A}) = K_0(A)\oplus \Bbb Z$\ $H_{\tilde{\alpha}} = H_{\alpha}\oplus 0$ Thus it is clear that $H_{\tilde{\alpha}} \cap K_{0}^{+}(\tilde{A}) = 0 \Leftrightarrow H_{\alpha} \cap K_{0}^{+}(A) = 0$. $(1) \Leftrightarrow (5)$ The hypotheses of (5) imply that $\theta \circ i_{*}$ is a faithful homomorphism (where $i: A \hookrightarrow A\times_{\alpha}\Bbb Z$ is the natural inclusion) with $\alpha_{*}(x) - x \in Ker(\theta \circ i_{*}$), for all $x \in \Gamma(A)$. However, since $\Gamma(A)$ generates $K_0(A)$ we see that $H_{\alpha} \subset Ker(\theta\circ i_{*})$. Thus the faithfulness of $\theta \circ i_{*}$ implies that $H_{\alpha} \cap K_{0}^{+}(A) = \{0\}$. The converse is trivial. By the equivalence of (1) and (2), we may assume for the remainder of the proof that A is unital. (1)$\Rightarrow$(3)   If $u \in A\times_{\alpha}\Bbb Z$ and $v \in A\times_{\alpha^{m}}\Bbb Z$ are the distinguished unitaries, then it is routine to check that the covariant representation $v \longmapsto u^{m}$ $a \longmapsto a$ defines an embedding: $A\times_{\alpha^{m}}\Bbb Z$ $\hookrightarrow$ $A\times_{\alpha}\Bbb Z$. \(3) $\Rightarrow$ (4) is immediate \(4) $\Rightarrow (1)$   Assume $\varphi: A\times_{\alpha^{m}}\Bbb Z$ $\longrightarrow B$ is an embedding into an AF algebra, $B$. Notice that in $K_0(B)$ we have that for every projection, $p \in A, [\varphi(p)] = [\varphi(upu^{*})] = [\varphi(\alpha^{m}(p))]$. We will assume that m is positive, for if m is negative, it will be clear how to adapt the following argument. Now, define $B_{m} = \oplus_{0}^{m - 1}B$ and let $\beta \in {\operatorname{Aut}}(B_{m}$) be the backwards cyclic permutation of the summands of $B_{m}$. That is, $\beta(b_{0}\oplus \cdots \oplus b_{m - 1}) = b_{1}\oplus \cdots \oplus b_{m -1} \oplus b_{0}$. Now, identifying $A$ with it’s image in $A\times_{\alpha^{m}} \Bbb Z$, we define $\psi_{i} = \varphi|_{A}\circ\alpha^{i}, 0 \leq i \leq m-1$. Then we define $\psi = \oplus_{0}^{m - 1} \psi_{i}: A \rightarrow B_{m}$, and it is clear that $\psi$ is a \*-monomorphism. It is also clear that $\beta$ satisfies (FOP). Finally, using the fact that in $K_0(B)$ we know \[$\varphi(p)] = [ \varphi(\alpha^{m}(p))]$, one easily checks that $\beta_{*} \circ \psi_{*} = \psi_{*} \circ \alpha_{*}$. Thus the conclusion follows from Theorem 4.2 and Proposition 3.1. $\Box$ [**Remark 4.13**]{}   It follows from Lemma 2.8 in \[Vo\] that if $\cal U = \otimes_{n \geq 1}M_n$ is the Universal UHF algebra, then condition (5) implies that the crossed product can be embedded into $B\otimes \cal U$. [**Remark 4.14**]{}   The author believes the equivalence of (1) and (2) to be a nontrivial application of Theorem 0.2 as he has been unable to find a direct proof of $(1) \Rightarrow (2)$. [**Remark 4.15**]{}   The equivalence of (1), (3) and (4) was proved in \[Pi\] when $A$ is assumed to be any (not necessarily AF) unital, abelian C$^{*}$-algebra. [**Section 5: K-Theory**]{} In this section we will show that our previous methods can be used to get rationally injective maps on $K_0(A\times_{\alpha}\Bbb Z)$ (whenever $A\times_{\alpha}\Bbb Z$ is AF embeddable). That is, we will show that if $A\times_{\alpha}\Bbb Z$ is AF embeddable, then one can find a \*-monomorphism, $\varphi : A\times_{\alpha}\Bbb Z \rightarrow B$, where $B$ is AF and $\varphi_{*} : K_0(A\times_{\alpha}\Bbb Z) \rightarrow K_0(B)$ is rationally injective. The author presented the main results of Sections 2, 3 and 4 in the Functional Analysis Seminar at Purdue University. He would like to thank Professor Larry Brown for asking the questions that led to this section of the paper. The following definition is taken from \[Vo\]. [**Definition 5.1**]{}   If $\beta \in {\operatorname{Aut}}(B)$ then $\beta$ is called $limit \ periodic$ if there exists an increasing nest, $\{B_n\}$, of finite dimensional subalgebras (with $ B = \overline{\bigcup B_n}$) and a sequence of positive integers, $d_n \in \Bbb N$, such that $\beta(B_n) = B_n$ and $\beta^{d_n}|_{B_n} = id_{B_n}$. In Lemma 2.8 of \[Vo\] it is shown that crossed products of AF algebras by limit periodic automorphisms are AF embeddable. The following lemma may be well known to the experts, but the author is not aware of a specific reference and thus includes a proof for completeness. [**Lemma 5.2**]{}   If $B$ is unital and $\beta \in \overline{Inn(B)}$ is limit periodic (with respect to $\{B_n\}$ and with integers $\{d_n\}$) then the embedding of $B\times_{\beta} \Bbb Z$ into $B\otimes \cal U^{\prime}$ (where $\cal U^{\prime}$ is a UHF algebra) constructed in Lemma 2.8 of \[Vo\] induces an injective map on $K_0(B\times_{\beta} \Bbb Z)$. [**Sketch of Proof**]{}   (cf. Lemma 2.8 in \[Vo\]). By the Pimsner-Voiculescu six term exact sequence, we have that $i_{*} : K_0(B) \rightarrow K_0(B\times_{\beta} \Bbb Z)$ is an isomorphism, where $i : B \rightarrow B\times_{\beta} \Bbb Z$ is the natural inclusion map. Letting $\varphi : B\times_{\beta} \Bbb Z \rightarrow B\otimes \cal U^{\prime}$ be the embedding constructed in \[Vo\] we have the following commutative diagrams. $\xymatrix{ B \ar[r]^i \ar[dr]^{\varphi} & B\times_{\beta} \Bbb Z \ar[d]^{\varphi} \\ & B\otimes \cal U^{\prime} }$ $\xymatrix{ K_0(B) \ar[r]^{i_{*}} \ar[dr]^{\varphi_{*}} & K_0(B\times_{\beta} \Bbb Z) \ar[d]^{\varphi_{*}} \\ & K_0(B\otimes \cal U^{\prime}) }$ Thus, it suffices to show that $\varphi_{*} : K_0(B) \rightarrow K_0(B\otimes \cal U^{\prime})$ is an injective map. Now, as in the proof of Lemma 2.8 in \[Vo\], we define $U_n = M_{d_n}(\Bbb C)$ and take maps $\varphi_{n} : B_n \rightarrow B_n \otimes U_n$ with commutativity in the diagram $$\begin{CD} B_n @>j_n>> B_{n + 1} \\ @V\varphi_n VV @V\varphi_{n + 1}VV \\ B_n \otimes U_n @>>> B_{n + 1} \otimes U_{n + 1} \end{CD}$$ where we don’t need to know what map the lower arrow is given by, $j_n$ is the natural inclusion map and $\varphi_n$ takes $b \mapsto \sum_{j = 1}^{d_n} \beta^{j}(b)\otimes e_{j,j}$, where $e_{i,j}$ are the standard matrix units of $M_{d_n}$. Since $\varphi$ is defined as the limit of the maps $\varphi_n$, we have the following commutative diagram $$\begin{CD} K_0(B_n) @>j_{n*}>> K_0(B) \\ @V\varphi_{n*} VV @V\varphi_{*}VV \\ K_0(B_n \otimes U_n) @>\Psi_{n*}>> K_0(B \otimes \cal U^{\prime}) \end{CD}$$ where $\Psi_n : B_n \otimes U_n \rightarrow B \otimes \cal U^{\prime}$ is the induced map and actually, $\cal U^{\prime} = \otimes_{n \geq 1} M_{d_n}$. From this diagram we see that it suffices to show the following assertion. $Claim$: If $x \in K_0(B_n)$ and $\varphi_n(x) = 0$ then $j_{n*}(x) = 0$. So, assume we have $x = [p] - [q] \in K_0(B_n)$ (where $p, \ q$ are projections in the matrices over $B_n$) such that $\varphi_{n*}(x) = 0$. Then with the identification $K_0(B_n \otimes U_n) = K_0(M_{d_n}(B_n)) = K_0(B_n)$, we have that (in $K_0(B_n)$) $\varphi_n(x) = \sum_{j = 1}^{d_n} [\beta^{j}(p)] - \sum_{j = 1}^{d_n} [\beta^{j}(q)] = 0$. However, by assumption we have that in $K_0(B)$, $[r] = [\beta^{j}(r)]$, for every projection, $r$, in the matrices over $B$ and for every $j \in \Bbb Z$. Thus, passing to $K_0(B)$ the above formula becomes $d_n([p] - [q]) = 0$ or $d_n(j_{n*}(x)) = 0$ and thus (since $K_0(B)$ is torsion free) $j_{n*}(x) = 0$. $\Box$ We now present an analogue of the lemma of Spielberg (Lemma 3.4) which gives more control on K-theory. The main idea in the proof is similar to the proof of Lemma 1.14 is \[Sp\]. [**Lemma 5.3**]{}   If $(G, \ G^{+})$ is a torsion free, ordered group then there exists a cone $\tilde{G}^{+} \supset G^{+}$ such that $(G, \ \tilde{G}^{+})$ is totally ordered and hence is a dimension group. [**Proof**]{}   Let $\cal L = \{ E \subset G : i) \ E \supset G^{+} \ \ ii) \ E \text{ is a semigroup} \ \ iii) \ E\cap (-E) = 0 \ \}$ be partially ordered by inclusion. It is easy to check that the hypotheses of Zorn’s Lemma are satisfied and thus we may choose a maximal element, $\tilde{G}^{+} \in \cal L$. Now, choose $x \in G\backslash \tilde{G}^{+}$ and consider the semigroup $\{ nx + e : n \in \Bbb N, \ e \in \tilde{G}^{+} \}$. This semigroup clearly (properly) contains $\tilde{G}^{+} $, and thus by maximality we get that there exist nonnegative integers, $n, \ m \in \Bbb N$ (at least one of which is nonzero) , and elements, $e, \ f \in \tilde{G}^{+} $ , such that $nx + e = -(mx + f) \neq 0$. Thus $(n + m)x \in -\tilde{G}^{+} $. We now claim that $(G, \ \tilde{G}^{+})$ is unperforated. So assume that we have found some nonzero element, $x \in G$, such that $kx \in \tilde{G}^{+}$ while $x \notin \tilde{G}^{+} $. Then the argument above (since $x \notin \tilde{G}^{+} $) provides us with a positive integer (letting $j = n + m$, from above) such that $jx \in -\tilde{G}^{+} $. Thus, $(kj)x \in \tilde{G}^{+} \cap -\tilde{G}^{+} = 0$. But this is a contradiction to the hypothesis that $G$ is torsion free and hence $(G, \ \tilde{G}^{+})$ is unperforated. However, if we now take $x \in G\backslash \tilde{G}^{+} $ then we can find a positive integer $j$ such that $jx \in -\tilde{G}^{+} $ which implies (since $(G, \ \tilde{G}^{+} )$ is unperforated) that $x \in -\tilde{G}^{+} $. Thus $(G, \ \tilde{G}^{+} )$ is totally ordered. $\Box$ [**Remark 5.4**]{}   If $A\times_{\alpha}\Bbb Z$ is AF embeddable then it is easy to see that $K_0(A\times_{\alpha}\Bbb Z) = K_0(A)/H_{\alpha}$ is an ordered group with the cone $K_{0}^{+}(A) + H_{\alpha}$. Clearly, the only thing that needs to be checked is that $K_{0}^{+}(A) + H_{\alpha} \cap -(K_{0}^{+}(A) + H_{\alpha}) = \{0\}$. However, this follows easily from the fact that $K_{0}^{+}(A) \cap H_{\alpha} = \{0\}$ by assumption (and Theorem 0.2). [**Theorem 5.5**]{}    Assume that $A\times_{\alpha}\Bbb Z$ is AF embeddable and $H$ is a subgroup of $K_0(A\times_{\alpha}\Bbb Z) = K_0(A)/H_{\alpha}$ with the property that $H \cap (K_{0}^{+}(A) + H_{\alpha}) = \{0\}$ and $K_0(A\times_{\alpha}\Bbb Z)/H$ is torsion free. Then one may choose the AF embedding of $A\times_{\alpha}\Bbb Z$ such that the kernel of the induced map on $K_0(A\times_{\alpha}\Bbb Z)$ is precisely $H$. [**Proof**]{}   To ease our notation somewhat, we begin by defining $G = K_0(A\times_{\alpha}\Bbb Z)/H$\ $G^{+} = (K_{0}^{+}(A) + H_{\alpha}) + H$ where $K_{0}^{+}(A) + H_{\alpha}$ was shown to be a cone of $K_0(A\times_{\alpha}\Bbb Z) = K_0(A)/H_{\alpha}$ in Remark 5.4 and thus $G^{+}$ is just the natural image of this cone in $G$. Now we claim that $(G, \ G^{+})$ is an ordered group. It is clear that we have $G^{+} + G^{+} \subset G^{+}$ and $G = G^{+} - G^{+}$ and thus we only have to check that $G^{+} \cap -(G^{+}) = \{0\}$. So assume there are elements $x, \ y \in K_{0}^{+}(A)$ such that $(x + H_{\alpha}) + H = -(y + H_{\alpha}) + H$ (in G). This implies that $(x + y) + H_{\alpha} \in H \cap (K_{0}^{+}(A) + H_{\alpha}) = \{0\}$ (in $K_0(A\times_{\alpha}\Bbb Z)$) and hence $x + y \in H_{\alpha} \cap K_{0}^{+}(A) = \{0\}$ (in $K_0(A)$). But, this implies that $x = y = 0$. Thus we have that $(G, \ G^{+})$ is a torsion free, ordered group and hence by Lemma 5.3 we can find a cone $\tilde{G}^{+} \supset G^{+}$ and an AF algebra, $B$, such that $(K_0(B), \ K_{0}^{+}(B), \ \Gamma(B)) = (G, \ \tilde{G}^{+}, \ \tilde{G}^{+})$. Note that if we let $\pi : K_0(A) \rightarrow K_0(B) = G$ be the canonical projection then $\pi$ is a contractive, faithful group homomorphism. Hence, by Fact 1.1, there exists a \*-monomorphism, $\varphi : A \rightarrow B$ with $\varphi_{*} = \pi$. The crucial properties in the remainder of the proof will be, 1. $\varphi : A \rightarrow B$ is a unital \*-monomorphism 2. $K_0(B) = K_0(A\times_{\alpha}\Bbb Z)/H$ 3. $Ker(\varphi_{*}) = Ker(\pi)$ where $\pi$ is the composition of the canonical projection maps $K_0(A) \rightarrow K_0(A\times_{\alpha}\Bbb Z) \rightarrow K_0(A\times_{\alpha}\Bbb Z)/H$. We now show that we may assume that $A$ and $B$ are unital and that $\varphi : A \rightarrow B$ is a unital injection. By the split exactness of the sequence $0 \rightarrow A\times_{\alpha}\Bbb Z \rightarrow \tilde{A}\times_{\tilde{\alpha}}\Bbb Z \rightarrow C(\Bbb T) \rightarrow 0$ we have that $K_0(\tilde{A}\times_{\tilde{\alpha}}\Bbb Z) = K_0(A\times_{\alpha}\Bbb Z) \oplus \Bbb Z$ and the image of $H$ under this identification is $\tilde{H} = H \oplus 0 \subset K_0(\tilde{A}\times_{\tilde{\alpha}}\Bbb Z)$. We now claim that properties 1), 2) and 3) above still hold with the unitizations of $A, \ B$ and $\varphi$. Now we recall that $K_0(\tilde{A}) = K_0(A) \oplus \Bbb Z$ and $K_0(\tilde{B}) = K_0(B) \oplus \Bbb Z$ and hence $ker(\tilde{\varphi}_{*}) = ker(\varphi_{*}) \oplus 0 \subset K_0(A) \oplus \Bbb Z$. Also, we observe that $K_0(\tilde{A}\times_{\tilde{\alpha}}\Bbb Z)/\tilde{H} = (K_0(A\times_{\alpha}\Bbb Z) \oplus \Bbb Z)/(H \oplus 0) = K_0(B) \oplus \Bbb Z = K_0(\tilde{B})$ with all these identifications being natural. Thus properties 1) and 2) are satisfied. Finally, it is easy to see that the kernel of the composition of the projections maps $K_0(\tilde{A}) \rightarrow K_0(\tilde{A}\times_{\tilde{\alpha}}\Bbb Z) \rightarrow K_0(\tilde{A}\times_{\tilde{\alpha}}\Bbb Z)/\tilde{H}$ is precisely $Ker(\varphi_{*}) \oplus 0 = Ker(\tilde{\varphi}_{*})$ since the above sequence is really just $K_0(A) \oplus 0 \rightarrow K_0(A\times_{\alpha}\Bbb Z) \oplus \Bbb Z \rightarrow (K_0(A\times_{\alpha}\Bbb Z) \oplus \Bbb Z)/(H \oplus 0)$ Thus it suffices to prove the theorem with $\tilde{A}\times_{\tilde{\alpha}}\Bbb Z$ and $\tilde{H}$ and hence we may assume that $A$ and $B$ are unital and $\varphi$ is a unit preserving, injective \*-homomorphism with properties 1), 2) and 3) above. Now, by the proof of Proposition 3.1 (and Proposition 2.8) we may further assume that, $\varphi : A \rightarrow B\otimes \cal U$, where $\cal U$ is the Universal UHF algebra and we have commutativity in the diagram $$\begin{CD} A @>\varphi>>B\otimes \cal U \\ @VAdv\circ\alpha VV @VVAdu\circ(id_{B}\otimes\sigma) V \\ A @>\varphi>>B\otimes \cal U \end{CD}$$ where $v \in A$, $u \in B\otimes \cal U$ are unitaries and $\sigma \in {\operatorname{Aut}}(\cal U)$ is the automorphism (with the Rohlin property) from Example 2.2. Note that it was necessary to first arrange that $\varphi$ be a unital map in order to appeal to Proposition 2.8 and that we have not changed the kernel of $\varphi_{*}$ in doing so (although we now have that $K_0(B) = K_0(A\times_{\alpha}\Bbb Z)/H$ sits injectively inside $K_0(B\otimes \cal U)$). Note also that $id_{B}\otimes\sigma$ is limit periodic. Now, from the Pimsner-Voiculescu six term exact sequence we have that $K_0(B\otimes\cal U) = K_0(B\otimes\cal U \times_{id_{B}\otimes\sigma}\Bbb Z)$. Thus commutativity in the diagram $$\begin{CD} A @>i>> A\times_{Adv\circ\alpha}\Bbb Z \\ @V\varphi VV @VV\tilde{\varphi} V \\ B\otimes\cal U @>>> B\otimes\cal U \times_{Adu\circ(id_{B}\otimes\sigma)}\Bbb Z \end{CD}$$ implies commutativity in the diagram $$\begin{CD} K_0(A) @>i_{*}>> K_0(A\times_{Adv\circ\alpha}\Bbb Z) \\ @V\varphi_{*} VV @VV\tilde{\varphi}_{*} V \\ K_0(B\otimes\cal U) @>\cong>> K_0(B\otimes\cal U \times_{Adu\circ (id_{B}\otimes\sigma)}\Bbb Z) \end{CD}$$ Where the $\tilde{\varphi}$ on the right side of the first diagram is now the natural extension of $\varphi$ to the crossed products (i.e. it no longer denotes the unital extension) and $i$ is the natural inclusion map. Finally, since $Ker(\varphi_{*}) = Ker(\tilde{\varphi}_{*}\circ i_{*}) = Ker(\pi)$ where $\pi$ was the composition of the maps $K_0(A) \rightarrow K_0(A\times_{\alpha}\Bbb Z) \rightarrow K_0(A\times_{\alpha}\Bbb Z)/H = K_0(B) \hookrightarrow K_0(B\otimes\cal U)$ it is now a routine exercise to verify that $Ker(\tilde{\varphi}_{*}) = H$ (under the identifications $K_0(A\times_{Adv\circ\alpha}\Bbb Z) = K_0(A\times_{\alpha}\Bbb Z)$ and $K_0(B\otimes\cal U \times_{Adu(\circ id_{B}\otimes\sigma)}\Bbb Z) = K_0(B\otimes\cal U \times_{id_{B}\otimes\sigma}\Bbb Z)$). Thus the proof is complete since $id_{B}\otimes\sigma$ is limit periodic and hence from Lemma 5.2 we have that $K_0(B\otimes\cal U \times_{id_{B}\otimes\sigma}\Bbb Z)$ gets embedded into $K_0(B\otimes\cal U\otimes \cal U^{\prime})$ where $\cal U^{\prime}$ is the UHF algebra constructed in Lemma 5.2. $\Box$ [**Corollary 5.6**]{}   If $A\times_{\alpha}\Bbb Z$ is AF embeddable, then one may choose an embedding which induces a rationally injective map on $K_0(A\times_{\alpha}\Bbb Z)$. [**Proof**]{}   Let $T(A\times_{\alpha}\Bbb Z) \subset K_0(A\times_{\alpha}\Bbb Z) = K_0(A)/H_{\alpha}$ denote the torsion subgroup. To apply Theorem 5.5 we only need to see that $T(A\times_{\alpha}\Bbb Z) \cap (K_{0}^{+}(A) + H_{\alpha}) = \{0\}$. But this follows easily from the fact that $K_{0}^{+}(A) \cap H_{\alpha} = \{0\}$. $\Box$ [**Remark 5.7**]{}   It is easy to prove that if $K_0(A)$ is a divisible group then $K_0(A\times_{\alpha}\Bbb Z)$ is always torsion free. However, it not hard to construct an example where $A\times_{\alpha}\Bbb Z$ is AF embeddable and $K_0(A\times_{\alpha}\Bbb Z)$ has torsion. [**Remark 5.8**]{}   It is clear that, in general, our constructions will not yield isomorphisms on K-theory since we must tensor with some UHF algebra to get the desired embeddings. [99]{} W.B. Arveson, [*Notes on extensions of C$^{*}$-algebras*]{}, Duke Math. J. [**44**]{} (1977), 329 - 355. B. Blackadar, [*K-theory for operator algebras*]{}, Springer-Verlag, New York (1986). O. Bratteli, [*Inductive limits of finite dimensional C$^{*}$-algebras*]{}, Trans. Amer. Math. Soc. [**171**]{} (1972), 195 - 234. O. Bratteli, D.E. Evans and A. Kishimoto, [*The Rohlin property for quasi-free automorphisms of the Fermion algebra*]{}, Proc. London Math. Soc. [**71**]{} (1995), 675 - 694. O. Bratteli, A. Kishimoto, M. R$\o$rdam and E. St$\o$rmer, [*The crossed product of a UHF algebra by a shift*]{}, Ergod. Th. Dynam. Sys. [**13**]{} (1993), 615 - 626. A. Connes, [*Periodic automorphisms of hyperfinite factors of type II$_{1}$*]{}, Acta Sci. Math. [**39**]{} (1977), 33 - 66. A. Connes, [*Outer conjugacy class of automorphisms of factors*]{}, Ann. Scient. Ec. Nor. Sup. [**8**]{} (1975), 383 - 420. J. Cuntz, [*Simple C$^{*}$-algebras generated by isometries*]{}, Comm. Math. Phys. [**57**]{} (1977), 173 - 185. K.R. Davidson, [*C$^{*}$-algebras by Example*]{}, Fields Inst. Monographs vol. 6, Amer. Math. Soc., (1996). E.G. Effros, [*Dimensions and C$^{*}$-algebras*]{}, Conf. Board Math. Sci., Reg. Conf. Ser. Math. [**46**]{}, (1981). E.G. Effros, D. Handelman and C.-L. Shen, [ *Dimension groups and their affine representations*]{}, Amer. J. Math. [**102**]{} (1980), 385 - 407. G.A. Elliot, [*On the classification of inductive limits of sequences of semisimple finite dimensional algebras*]{}, J. Algebra [**38**]{} (1976), 29 - 44. D.E. Evans and A. Kishimoto, [*Trace scaling automorphisms of certain stable AF algebras*]{}, Hokkaido Math. J. [**26**]{} (1997), 211 - 224. R.H. Herman and A. Ocneanu, [*Stability for integer actions on UHF C$^{*}$-algebras*]{}, J. Funct. Anal. [**59**]{} (1984), 132 - 144. A. Kishimoto, [*The Rohlin property for shifts on UHF algebras and automorphisms of Cuntz algebras*]{}, J. Funct. Anal. [**140**]{} (1996), 100 - 123. A. Kishimoto, [*The Rohlin property for automorphisms of UHF algebras*]{}, J. reine angew. Math. [**465**]{} (1995), 183 - 196. A. Kishimoto, [*A Rohlin property for one-parameter automorphism groups*]{}, Comm. Math. Phys. [**179**]{} (1996), 599 - 622. M.B. Landstad, [*Duality theory for covariant systems*]{}, Trans. Amer. Math. Soc. [**248**]{} (1979), 223 - 267. H. Nakamura, [*The Rohlin property for actions of $\Bbb Z^{2}$ on UHF algebras*]{}, preprint. A. Ocneanu, [*A Rohlin type theorem for groups acting on von Neumann algebras*]{}, Topics in Modern Operator Theory, Birkhäuser Verlag (1981), 247 - 258. M. Pimsner, [*Embedding some transformation group C$^{*}$-algebras into AF algebras*]{}, Ergod. Th. Dynam. Sys. [**3**]{} (1983), 613 - 626. M. Pimsner and D. Voiculescu, [*Exact sequences for K-groups and Ext-groups of certain crossed products of $C^{*}-algebras$*]{}, J. Oper. Th. [**4**]{} (1980), 93 - 118. L.S. Pontryagin, [*Topological groups*]{}, Selected Works vol. 2, Gordon and Breach, (1966). M. R$\o$rdam, [*Classification of certain infinite simple C$^{*}$-algebras*]{}, J. Funct. Anal. [**131**]{} (1995), 415 - 458. J.S. Spielberg, [*Embedding C$^{*}$-algebra extensions into AF algebras*]{}, J. Funct. Anal. [**81**]{} (1988), 325 - 344. D. Voiculescu, [*Almost inductive limit automorphisms and embeddings into AF algebras*]{}, Ergod. Th. Dynam. Sys. [**6**]{} (1986), 475 - 484. N.E. Wegge-Olsen, [*K-theory and C$^{*}$-algebras*]{}, Oxford Univ. Press, (1993).
{ "pile_set_name": "ArXiv" }
--- abstract: 'We characterize inverse limits of nilsystems in topological dynamics, via a structure theorem for topological dynamical systems that is an analog of the structure theorem for measure preserving systems. We provide two applications of the structure. The first is to nilsequences, which have played an important role in recent developments in ergodic theory and additive combinatorics; we give a characterization that detects if a given sequence is a nilsequence by only testing properties locally, meaning on finite intervals. The second application is the construction of the maximal nilfactor of any order in a distal minimal topological dynamical system. We show that this factor can be defined via a certain generalization of the regionally proximal relation that is used to produce the maximal equicontinuous factor and corresponds to the case of order $1$.' address: - | Laboratoire d’analyse et de mathématiques appliquées, Université de Paris-Est, Marne la Vallée & CNRS UMR 8050\ 5 Bd. Descartes, Champs sur Marne\ 77454 Marne la Vallée Cedex 2, France - | Department of Mathematics, Northwestern University\ 2033 Sheridan Road Evanston\ IL 60208-2730, USA - | Departamento de Ingeniería Matemática, Universidad de Chile & Centro de Modelamiento Matemático UMI 2071 UCHILE-CNRS\ Casilla 170/3 correo 3\ Santiago, Chili. author: - Bernard Host - Bryna Kra - Alejandro Maass date: 'May 13, 2009' title: Nilsequences and a structure theorem for topological dynamical systems --- [^1] Introduction ============ Nilsequences ------------ The connection between ergodic theory and additive combinatorics started in the 1970’s, with Furstenberg’s beautiful proof of Szemerédi’s Theorem via ergodic theory. Furstenberg’s proof paved the way for new combinatorial results via ergodic methods, as well as leading to numerous developments within ergodic theory. More recently, the interaction between the fields has taken a new dimension, with ergodic objects being imported into the finite combinatorial setting. Some objects at the center of this interchange are nilsequences and the nilsystems on which they are defined. They enter, for example, in ergodic theory into convergence of multiple ergodic averages [@HK1] and into the theory of multicorrelations [@BHK]. In number theory, they arise in finding patterns in the primes (see  [@GT] and the companion articles [@GT2] and [@GT3]). In combinatorics, they are used to find intricate patterns in subsets of integers with positive upper density [@FrWi]. Nilsequences are defined by evaluating a function along the orbit of a point in the homogeneous space of a nilpotent Lie group. In a variety of situations, nilsequences have been used to test for a lack of uniformity of a function. Yet, the local properties of nilsequences are not well understood. It is difficult to detect if a given sequence is a nilsequence, particularly if one only knows [*local*]{} information about the sequence, meaning properties that can only be tested on finite intervals. We recall the definition of a nilsequence. A [*basic $d$-step nilsequence*]{} is a sequence of the form $(f(T^nx)\colon n\in{{\mathbb Z}})$, where $(X, T)$ is a $d$-step nilsystem, $f\colon X\to {{\mathbb C}}$ is a continuous function, and $x\in X$. A [*$d$-step nilsequence*]{} is a uniform limit of basic $d$-step nilsequences. (See Section \[sec:nilsystems\] for the definition of a nilsystem.) We give a characterization of nilsequences of all orders that can be tested locally, generalizing the work in [@HM] that gives such an analysis for $2$-step nilsequences. We look at finite portions, the “windows”, of a sequence and we are interested in finding a copy of the same finite window up to some given precision. To make this clear, we introduce some notation. For a sequence ${\mathbf{a}}= (a_n\colon n\in{{\mathbb Z}})$, integers $k,j,L$, and a real $\delta > 0$, if each entry in the window $[k-L, k+L]$ is equal to the corresponding entry in the window $[j-L, j+L]$ up to an error of $\delta$, then we write $$\label{eq:approx} {\mathbf{a}}_{[k-L, k+L]} =_\delta {\mathbf{a}}_{[j-L, j+L]}\ .$$ The characterization of almost periodic sequences (which are exactly $1$-step nilsequences) by compactness can be formulated as follows: The bounded sequence ${\mathbf{a}}= (a_n\colon n\in{{\mathbb Z}})$ of complex numbers is almost periodic if and only if for all $\varepsilon > 0$, there exist an integer $L\geq 1$ and a real $\delta > 0$ such that for any $k,n_1,n_2 \in{{\mathbb Z}}$ whenever ${\mathbf{a}}_{[k-L, k+L]} =_\delta {\mathbf{a}}_{[k+n_1-L, k+n_1+L]}$ and ${\mathbf{a}}_{[k-L, k+L]} =_\delta {\mathbf{a}}_{[k+n_2-L, k+n_2+L]}$ then $|a_k-a_{k+n_1+n_2}|< \varepsilon$. We give a similar characterization for a $(d-1)$-step nilsequence ${\mathbf{a}}$: if in every interval of a given length the translates of the sequence ${\mathbf{a}}$ along finite sums (i.e. cubes) of any sequence ${\mathbf{n}}= (n_1, \ldots, n_{d})$ are $\delta$-close to the original sequence [*except*]{} possibly at the sum $n_1+\ldots +n_{d}$, then we also have control over the distance between ${\mathbf{a}}$ and the translate by $n_1+\ldots +n_{d}$. The general case is: \[theorem:nilseq\] Let ${\mathbf{a}}=(a_n: n\in {{\mathbb Z}})$ be a bounded sequence of complex numbers and let $d\geq 2$ be an integer. The sequence ${\mathbf{a}}$ is a $(d-1)$-step nilsequence if and only if for every $\varepsilon>0$ there exist an integer $L\geq 1$ and real $\delta >0$ such that for any $(n_1,\ldots,n_{d})\in{{\mathbb Z}}^d$ and $k\in{{\mathbb Z}}$, whenever $${\mathbf{a}}_{[k+\epsilon_1n_1 +\ldots +\epsilon_dn_d -L, k+\epsilon_1n_1 +\ldots +\epsilon_dn_d +L]} =_\delta {\mathbf{a}}_{[k-L, k+L]}$$ for all choices of $\epsilon_1, \ldots, \epsilon_d\in\{0,1\}$ other than $\epsilon_1= \ldots = \epsilon_d = 1$, then we have $|a_{k+n_1+\ldots+n_d} - a_k| < \varepsilon$. In fact, we can replace the approximation in  in both the hypothesis and conclusion by any other approximation that defines pointwise convergence and have the analogous result. A structure theorem for topological dynamical systems ----------------------------------------------------- We prove a structure theorem for topological dynamical systems that gives a characterization of inverse limits of nilsystems. Theorem \[theorem:nilseq\] follows from this structure theorem, exactly as it does in the case for $d=2$ in [@HM], where the proof of this implication can be found. The structure theorem for topological dynamical systems can be viewed as an analog of the purely ergodic structure theorem of [@HK1]. We introduce the following structure: \[def:parallelepipeds\] Let $(X,T)$ be a topological dynamical system and let $d\geq 1$ be an integer. We define ${{\bf Q}}{^{[d]}}(X)$ to be the closure in $X^{2^d}$ of elements of the form $$(T^{n_1\epsilon_1+\ldots+n_d\epsilon_d}x\colon\epsilon = (\epsilon_{1}, \ldots, \epsilon_{d})\in\{0,1\}^d) \ ,$$ where ${\mathbf{n}}=(n_1, \ldots, n_d)\in{{\mathbb Z}}^d$, $x\in X$, and we denote a point of $X^{2^d}$ by $(x_\epsilon\colon\epsilon\in\{0,1\}^d)$. When there is no ambiguity, we write ${{\bf Q}}{^{[d]}}$ instead of ${{\bf Q}}{^{[d]}}(X)$. An element of ${{\bf Q}}{^{[d]}}(X)$ is called a [*(dynamical) parallelepiped of dimension $d$*]{}. As examples, ${{\bf Q}}{^{[2]}}$ is the closure in $X^{4}$ of the set $$\{(x,T^mx,T^nx,T^{n+m}x)\colon x \in X, m,n\in {{\mathbb Z}}\}$$ and ${{\bf Q}}{^{[3]}}$ is the closure in $X^{8}$ of the set $$\begin{gathered} \bigl\{(x,T^mx,T^nx,T^{m+n}x,T^px,T^{m+p}x,T^{n+p}x,T^{m+n+p}x)\colon \\ x \in X, m,n,p\in {{\mathbb Z}}\bigr\}\ .\end{gathered}$$ In each of these, the indices $m,n$ and $m,n,p$ can be taken in ${{\mathbb N}}$ rather than ${{\mathbb Z}}$, giving rise to the same object. This is obvious if $T$ is invertible, but can also be proved without the assumption of invertibility. Thus, throughout the article, we assume that all maps are invertible. We use these parallelepipeds structures to characterize nilsystems: \[theorem:main\] Assume that $(X,T)$ is a transitive topological dynamical system and let $d\geq 2$ be an integer. The following properties are equivalent: 1. \[it:change-one\] If ${\mathbf{x}},{\mathbf{y}}\in{{\bf Q}}{^{[d]}}(X)$ have $2^{d}-1$ coordinates in common, then ${\mathbf{x}}={\mathbf{y}}$. 2. If $x,y\in X$ are such that $(x, y, \ldots, y)\in{{\bf Q}}{^{[d]}}(X)$, then $x=y$. 3. \[it:nil\] $X$ is an inverse limit of $(d-1)$-step minimal nilsystems. (For definitions of all the objects, see Section \[section:parallele\].) We note that the use of both $d$ and $d-1$ is necessary throughout the article, and this leads us to use whichever is notationally more convenient at various times in the proofs. The first property clearly implies the second, since $(y, y, \ldots, y)\in{{\bf Q}}{^{[d]}}(X)$ for all $y\in X$. The second property implies that the system is distal (see Section \[section:parallele\]). The second property plus the assumption of distality implies the first property (see Section \[section:paradistal\]), which together give that the first two properties are equivalent. Systems satisfying these properties play a key role in the article and so we define: A transitive system satisfying either of the first two equivalent properties of Theorem \[theorem:main\] is called a [*system of order $d-1$*]{}. The implication  $\Rightarrow$ in Theorem \[theorem:main\] follows from results in [@HK3] and is reviewed here in Proposition \[prop:Qnil\]. The implication $\Rightarrow$ is proved in Section \[sec:final\], using completely different methods from that used in [@HM] for $d=3$, and proceeds by introducing an invariant measure on $X$. The regionally proximal relation and generalizations ---------------------------------------------------- We give a second application of Theorem \[theorem:main\] in topological dynamics. The study of maximal equicontinuous factors is classical (see, for example [@Aus]). The maximal equicontinuous factor is the topological analog of the Kronecker factor in ergodic theory and recovers the continuous eigenvalues of a system. There are several ways to construct this factor, but the standard method is as a quotient of the [*regionally proximal relation*]{}. The first step in generalizing this relation was carried out in [@HM], where the concept of a double regionally proximal relation is introduced and is used in the distal case to define the maximal $2$-step nilfactor. In this article we generalize this relation for higher levels and for $d\geq 1$ we define the [*regionally proximal relation of order $d$*]{}, referring to it as ${{\bf RP}}{^{[d]}}$. While these generalizations were motivated by the study of abstract parallelepipeds in additive combinatorics [@HK2], they require new techniques. Although we defer the definition of the regionally proximal relation of order $d$ until Section \[section:parallele\], we summarize its uses. \[prop:orderRP\] Assume that $(X,T)$ is a transitive topological dynamical system and that $d \geq 1$ is an integer. If the regionally proximal relation of order $d$ on $X$ is trivial, then the system is distal. In a distal system, we show that ${{\bf RP}}{^{[d]}}$ is an equivalence relation and that it defines the [*maximal $d$-step topological nilfactor*]{} of the system. \[theorem:maxd\] Assume that $(X,T)$ is a distal minimal system and that $d \geq 1$ is an integer. Then the regionally proximal relation of order $d$ on $X$ is a closed invariant equivalence relation and the quotient of $X$ under this relation is its maximal $d$-step nilfactor. The maximal $d$-step (topological) nilfactor is the topological analog of the ergodic theoretic factor ${\mathcal Z}_d$ constructed in [@HK1]. These ergodic factors are characterized by inverse limits of $d$-step nilsystems. In this direction, we prove in the distal case that ${{\bf RP}}{^{[d]}}$ is trivial if and only if the system itself is an inverse limit of $d$-step nilsystems. To prove Theorem \[theorem:maxd\] we show in Proposition \[prop:quaotient1\] that the quotient of $X$ under ${{\bf RP}}{^{[d]}}$ is its maximal factor of order $d$. From Theorem \[theorem:main\], we deduce that notions of a system of order $d$ and an inverse limit of $d$-step nilsystems are equivalent, giving us the conclusion. We conjecture that the hypothesis of distality in Theorem \[theorem:maxd\] is superfluous, but were unable to prove this. Guide to the paper ------------------ The article is divided into two somewhat distinct parts. In the first part (Sections \[section:parallele\] and \[section:paradistal\]), we develop the topological theory of parallelepipeds and the associated theory of generalized regionally proximal relations. With the topological methods developed in these sections, we are able to prove all but the implication “(1) $\Rightarrow$ (3)” of Theorem \[theorem:main\]. In Section \[section:parallele\], we state the properties of parallelepiped structures and the relation with generalized regionally proximal pairs and show how the conditions of Theorem \[theorem:main\] imply that the system is distal. In Section \[section:paradistal\], we prove that in the distal case, the main structural properties of parallelepipeds (the “property of closing parallelepipeds”) allows us to show that first two conditions in Theorem \[theorem:main\] are equivalent and to show that regionally proximal relation of order $d$ gives rise to the maximal factor of order $d$. The proof of the remaining implication is carried out in Section \[sec:final\] and relies heavily on ergodic theoretic notions of Section \[section:preergodic\]. However, the interaction of the topological and measure theoretic structures plays a key role in the analysis, and it is only via measure theoretic methods that we are finally able to obtain the general topological results. Background ========== Topological dynamical systems ----------------------------- A [*transformation*]{} of a compact metric space $X$ is a homeomorphism of $X$ to itself. A [*topological dynamical system*]{}, referred to more succinctly as just a [*system*]{}, is a pair $(X,T)$, where $X$ is a compact metric space and $T\colon X\to X$ is a transformation. We use $d_X(\cdot,\cdot)$ to denote the metric in $X$ and when there is no ambiguity, we write $d(\cdot, \cdot)$. We also make use of a more general definition of a topological system. That is, instead of just a single transformation $T$, we consider commuting homeomorphisms $T_1,\ldots,T_k$ of $X$ or a countable abelian group of transformations. We summarize some basic definitions and properties of systems in the classical setting of one transformation. Extensions to the general case are straightforward. A [*factor*]{} of a system $(X,T)$ is another system $(Y,S)$ such that there exists a continuous and onto map $p\colon X\to Y$ satisfying $S \circ p= p \circ T$. The map $p$ is called a [*factor map*]{}. If $p$ is bijective, the two systems are [*(topologically) conjugate*]{}. In a slight abuse of notation, when there is no ambiguity, we denote all transformations (including ones in possibly distinct systems) by $T$. A system $(X,T)$ is [*transitive*]{} if there exists some point $x\in X$ whose orbit $\{T^n x\colon n\in{{\mathbb Z}}\}$ is dense in $X$ and we call such a point a [*transitive point*]{} . The system is [*minimal*]{} if the orbit of any point is dense in $X$. This property is equivalent to saying that $X$ and the empty set are the only closed invariant sets in $X$. Distal Systems {#sec:distal} -------------- The system $(X,T)$ is [*distal*]{} if for any pair of distinct points $x,y \in X$, $$\label{eq:distal} \inf_{n\in {{\mathbb Z}}} d(T^n x,T^n y) >0\ .$$ In an arbitrary system, pairs satisfying property  are called [*distal pairs*]{}. The points $x$ and $y$ are [*proximal*]{} if $\liminf_{n\to \infty} d(T^nx,T^ny)=0$. The following proposition summarizes some basic properties of distal systems: (See Auslander [@Aus], chapters 5 and 7) \[prop:distal\] 1. The Cartesian product of a finite family of distal systems is a distal system. 2. If $(X,T)$ is a distal system and $Y$ is a closed and invariant subset of $X$, then $(Y,T)$ is a distal system. 3. A transitive distal system is minimal. 4. A factor of a distal system is distal. 5. Let $p\colon X\to Y$ be a factor map between the distal systems $(X,T)$ and $(Y,T)$. If $(Y,T)$ is minimal, then $p$ is an open map. Up to the obvious changes in notation, this proposition holds for systems with a countable abelian group of transformations acting on the space $X$. For later use, we note the following lemma on distal systems: \[lem:extendcontinu\] Let $(X,T)$ and $(Y,T)$ be two minimal systems and assume that $(Y,T)$ is distal. If $X_1$ is a nonempty invariant subset of $X$ and $\Phi\colon X_1\to Y$ is a continuous map on $X_1$ with the induced topology and commuting with the transformations $T$, then $\Phi$ has a continuous extension to $X$. Let $\Gamma\subset X\times Y$ be the graph of $\Phi$: $$\Gamma=\{(x,\Phi(x))\colon x\in X_1\}\ .$$ Let $\overline\Gamma$ be the closure of $\Gamma$ in $X\times Y$. We claim that $\overline\Gamma$ is the graph of some map $\Phi'\colon X\to Y$. The projection of $\overline\Gamma$ on $X$ is a closed invariant subset of $X$ containing $X_1$, and by minimality this projection is equal to $X$. Assume that $x\in X$ and $y,y'\in Y$ are such that $(x,y)$ and $(x,y')$ belong to $\overline\Gamma$. Let $x_1\in X_1$ and chose a sequence $(n_i)_{i\in{{\mathbb N}}}$ of integers such that $T^{n_i}x\to x_1$ and such that the sequences $(T^{n_i}y)_{i\in{{\mathbb N}}}$ and $(T^{n_i}y')_{i\in{{\mathbb N}}}$ converge in $Y$, to the points $z$ and $z'$, respectively, as $i\to\infty$. Then $(x_1,z)$ and $(x_1,z')$ belong to $\overline\Gamma\cap(X_1\times Y)$. On the other hand, since $\Phi$ is continuous on $X_1$, we have that $\overline\Gamma\cap(X_1\times Y)=\Gamma$ and thus $z=\Phi(x_1)=z'$. Since $(Y,T)$ is distal, we conclude that $y=y'$ and we have that $\overline{\Gamma}$ is the graph of a map $\Phi'\colon X\to Y$. The restriction of $\Phi'$ to $X_1$ is equal to $\Phi$ and because its graph is closed, $\Phi'$ is continuous. Finally, since $X_1$ is invariant and nonempty, it is dense in $X$. By minimality and density, we conclude that $\Phi'\circ T=T\circ\Phi'$. Nilsystems and nilsequences {#sec:nilsystems} --------------------------- Let $d\geq 1$ be an integer and assume that $G$ is a $d$-step nilpotent Lie group and that $\Gamma\subset G$ is a discrete, cocompact subgroup of $G$. The compact manifold $X = G/\Gamma$ is a [*$d$-step nilmanifold*]{} and $G$ acts naturally on $X$ by left translations: $x\mapsto \tau.x$ for $\tau \in G$. If $T$ is left multiplication on $X$ by some fixed element of $G$, then $(X,T)$ is called a [*$d$-step nilsystem*]{}. A $d$-step nilsystem is an example of a distal system. In particular if the nilsystem is transitive, then it is minimal. Also, the closed orbit of a point in a $d$-step nilsystem is topologically conjugate to a $d$-step nilsystem. See [@AGH], [@Pa], and [@Le] for proofs and general references on nilsystems. We also make use of inverse limits of nilsystems and so we recall the definition of an inverse limit of systems (restricting ourselves to the case of sequential inverse limits). If $(X_i, T_i)_{i\in{{\mathbb N}}}$ are systems and $\pi_{i}\colon X_{i+1}\to X_i$ are factor maps, the [*inverse limit*]{} of the systems is defined to be the compact subset of $\prod_{i\in{{\mathbb N}}} X_i$ given by $$\{(x_i)_{i\in{{\mathbb N}}}\colon\pi_i(x_{i+1}) = x_i\} \ .$$ It is a compact metric space endowed with the distance $$d(x,y) = \sum_{i\in{{\mathbb N}}} 1/2^id_i(x_i,y_i) \ .$$ We note that the maps $T_i$ induce a transformation $T$ on the inverse limit. Many properties of the systems $(X_i, T_i)$ also pass to the inverse limit, including minimality, distality, and unique ergodicity. We return to the definition of a nilsequence: If $(X = G/\Gamma, T)$ is a $d$-step nilsystem, where $T$ is given by multiplication by the element $\tau\in G$, $f\colon X\to {{\mathbb C}}$ is a continuous function, and $x\in X$, the sequence $(f(\tau^n.x)\colon n\in{{\mathbb Z}})$ is a [*basic $d$-step nilsequence*]{}. A uniform limit of basic $d$-step nilsequences is a [*nilsequence*]{}. Equivalently, a $d$-step nilsequence is given by $(f(T^nx)\colon n\in{{\mathbb Z}})$, where $(X, T)$ is an inverse limit of $d$-step nilsystems, $f\colon X\to {{\mathbb C}}$ is a continuous function and $x\in X$. The two statements in the definition are shown to be equivalent in Lemma 14 in [@HM]. Moreover, in the definition of a $d$-step nilsequence, we can assume that the system is minimal. Namely, considering the closed orbit of $x_0$, this is a transitive and so minimal system. The $1$-step nilsystems are translations on compact abelian Lie groups and $1$-step nilsequences are exactly almost periodic sequences (see [@Pa]). Examples of $2$-step nilsequences and a detailed study of them are given in [@HK4]. Dynamical Parallelepipeds: first properties {#section:parallele} =========================================== Notation -------- Let $X$ be a set, let $d\geq 1$ an integer, and write $[d]=\{1,2,\dots,d\}$. We view $\{0,1\}^d$ in one of two ways, either as a sequence $\epsilon =\epsilon_1\ldots\epsilon_d$ of $0$’s and $1$’s written without commas or parentheses; or as a subset of $[d]$. A subset $\epsilon$ corresponds to the sequence $(\epsilon_1,\ldots,\epsilon_d) \in {\{0,1\}^d}$ such that $i \in \epsilon$ if and only if $\epsilon_i=1$ for $i \in [d]$. If ${\mathbf{n}}=(n_1, \ldots, n_d)\in{{\mathbb Z}}^d$ and $\epsilon\subset[d]$, we define $${\mathbf{n}}\cdot\epsilon = \sum_{i=1}^dn_i\epsilon_i = \sum_{i \in \epsilon} n_i \ .$$ We denote $X^{2^d}$ by $X{^{[d]}}$. A point ${\mathbf{x}}\in X{^{[d]}}$ can be written in one of two equivalent ways, depending on the context: $${\mathbf{x}}=(x_\epsilon\colon\epsilon \in {\{0,1\}^d}) = (x_\epsilon\colon\epsilon\subset [d]) \ .$$ For $x\in X$, we write $x{^{[d]}} = (x, x, \ldots, x)\in X{^{[d]}}$. The diagonal of $X{^{[d]}}$ is $\Delta{^{[d]}}=\{x{^{[d]}} \colon x \in X\}$. A point ${\mathbf{x}}\in X{^{[d]}}$ can be decomposed as ${\mathbf{x}}=({\mathbf{x}}',{\mathbf{x}}'')$ with ${\mathbf{x}}',{\mathbf{x}}'' \in X{^{[d-1]}}$, where ${\mathbf{x}}' = (x_{\epsilon 0}\colon \epsilon \in \{0,1\}^{d-1})$ and ${\mathbf{x}}'' = (x_{\epsilon 1}\colon \epsilon \in \{0,1\}^{d-1})$. We can also isolate the first coordinate, writing $X_*{^{[d]}}=X^{2^d-1}$ and then writing a point ${\mathbf{x}}\in X{^{[d]}}$ as ${\mathbf{x}}= (x, {\mathbf{x}}_*)$, where ${\mathbf{x}}_* = (x_\epsilon\colon\epsilon\neq\emptyset)\in X_*{^{[d]}}$. The [*faces*]{} of dimension $r$ of a point in ${\mathbf{x}}\in X{^{[d]}}$ are defined as follows. Let $J\subset [d]$ with $|J|=d-r$ and $\xi \in \{0,1\}^{d-r}$. The elements $(x_\epsilon \colon \epsilon \in\{0,1\}^d , \ \epsilon_J=\xi)$ of $X{^{[r]}}$ are called [*faces* ]{} of dimension $r$ of ${\mathbf{x}}$, where $ \epsilon_J=(\epsilon_i\colon i \in J)$. Thus any face of dimension $r$ defines a natural projection from $X{^{[d]}}$ to $X{^{[r]}}$, and we call this the [*projection along this face*]{}. Identifying $\{0,1\}^d$ with the set of vertices of the Euclidean unit cube, a Euclidean isometry of the unit cube permutes the vertices of the cube and thus the coordinates of a point $x\in X{^{[d]}}$. These permutations are the [*Euclidean permutations*]{} of $X{^{[d]}}$. Examples of Euclidean permutations are permutations of digits, meaning a permutation of $\{0,1\}^d$ induced by a permutation of $[d]$, and symmetries, such as replacing $\epsilon_i$ by $1-\epsilon_i$ for some $i$. For $d=2$, an example of a digit permutation is the map $(00, 01, 10, 11)\mapsto (00, 10, 01, 11)$ and an example of a symmetry is the map $(00, 01, 10, 11)\mapsto (01, 00, 11, 10)$. Dynamical parallelepipeds {#subsec:dynparallel} ------------------------- We recall that ${{\bf Q}}{^{[d]}}$ is the closure in $X^{2^d}$ of elements of the form $$(T^{n_1\epsilon_1+\ldots+n_d\epsilon_d}x\colon\epsilon\in\{0,1\}^d) \ ,$$ where ${\mathbf{n}}=(n_1, \ldots, n_d)\in{{\mathbb Z}}^d$ and $x\in X$ (Definition \[def:parallelepipeds\]). It follows immediately from the definition that ${{\bf Q}}{^{[d]}}$ contains the diagonal. Some other basic structural properties of ${{\bf Q}}{^{[d]}}$ are: 1. \[item:faces\] Any face of dimension $r$ of any ${\mathbf{x}}\in{{\bf Q}}{^{[d]}}$ belongs to ${{\bf Q}}{^{[r]}}$. (This condition is trivial for $d=2$.) 2. \[item:Euclidean\] ${{\bf Q}}{^{[d]}}$ is invariant under the Euclidean permutations of $X{^{[d]}}$. 3. \[item:QdtoQd+1\] If ${\mathbf{x}}\in{{\bf Q}}{^{[d]}}$, then $({\mathbf{x}},{\mathbf{x}})\in{{\bf Q}}{^{[d+1]}}$. \[lemma:Qfactors\] Let $d\geq 1$ be an integer, $(X,T)$ and $(Y,T)$ be systems, and $\pi:X\to Y$ be a factor map. Then ${{\bf Q}}{^{[d]}}(Y)$ is the image of ${{\bf Q}}{^{[d]}}(X)$ under the map $\pi^{[d]}:=\pi \times \ldots \times \pi$ ($2^d$ times). We can rephrase the definition of ${{\bf Q}}{^{[d]}}$ using some groups of transformations on $X{^{[d]}}$. We define: Let $(X,T)$ be a system and $d\geq 1$ be an integer. The *diagonal transformation* of $X{^{[d]}}$ is the map given by $(T{^{[d]}}{\mathbf{x}})_\epsilon = Tx_\epsilon$ for every ${\mathbf{x}}\in X{^{[d]}}$ and every $\epsilon\subset[d]$. For $ j\in [d]$, the [*face transformation*]{} $T_j{^{[d]}}: X {^{[d]}} \to X{^{[d]}}$ is defined for every ${\mathbf{x}}\in X{^{[d]}}$ and $\epsilon \subset [d]$ by: $$T_j{^{[d]}}{\mathbf{x}}= \begin{cases} (T_j{^{[d]}}{\mathbf{x}})_\epsilon = Tx_\epsilon & \text{ if } j \in \epsilon \\ (T_j{^{[d]}}{\mathbf{x}})_\epsilon = x_\epsilon & \text{ if } j \notin \epsilon\ . \end{cases}$$ The *face group of dimension $d$* is the group ${{\mathcal F}}{^{[d]}}(X)$ of transformations of $X{^{[d]}}$ spanned by the face transformations. The *parallelepiped group of dimension $d$* is the group ${{\mathcal G}}{^{[d]}}(X)$ spanned by the diagonal transformation and the face transformations. We often write ${{\mathcal F}}{^{[d]}}$ and ${{\mathcal G}}{^{[d]}}$ instead of ${{\mathcal F}}{^{[d]}}(X)$ and ${{\mathcal G}}{^{[d]}}(X)$, respectively. For ${{\mathcal G}}{^{[d]}}$ and ${{\mathcal F}}{^{[d]}}$, we use similar notations to that used for $X{^{[d]}}$: namely, an element of either of these groups is written as $S = (S_\epsilon\colon \epsilon\in\{0,1\}^d)$. In particular, ${{\mathcal F}}{^{[d]}} = \{S\in{{\mathcal G}}{^{[d]}}\colon S_\emptyset = {\bf Id}\}$. We note that the group ${{\mathcal G}}{^{[d]}}$ satisfies the three properties –  above, with ${{\bf Q}}{^{[d]}}$ replaced by ${{\mathcal G}}{^{[d]}}$. Moreover, for $S\in{{\mathcal F}}{^{[d]}}$, we have that $(S,S)\in{{\mathcal F}}{^{[d+1]}}$. As well, ${{\mathcal F}}{^{[d]}}$ is invariant under digit permutations. The following lemma follows directly from the definitions: \[lemma:no-name\] Let $(X,T)$ be a system and let $d\geq 1$ be an integer. Then ${{\bf Q}}{^{[d]}}$ is the closure in $X{^{[d]}}$ of $$\{S x{^{[d]}}\colon S \in{{\mathcal F}}{^{[d]}}, x\in X\}\ .$$ If $x$ is a transitive point of $X$, then ${{\bf Q}}{^{[d]}}$ is the closed orbit of $x{^{[d]}}$ under the group ${{\mathcal G}}{^{[d]}}$. Definition of the regionally proximal relations ----------------------------------------------- In this section, we discuss the relation ${{\bf RP}}{^{[d]}}$ and its relation to ${{\bf Q}}{^{[d+1]}}$. \[def:RP\] Let $(X,T)$ be a system and let $d\geq 1$ be an integer. The points $x,y\in X$ are said to be [*regionally proximal of order $d$*]{} if for any $\delta > 0$, there exist $x', y'\in X$ and a vector ${\mathbf{n}}= (n_1, \ldots, n_d)\in {{\mathbb Z}}^d$ such that $d(x,x')< \delta$, $d(y, y')<\delta$, and $$d(T^{{\mathbf{n}}\cdot\epsilon}x', T^{{\mathbf{n}}\cdot\epsilon}y') <\delta \text{ for any nonempty } \epsilon\subset [d] .$$ (In other words, there exists $S\in{{\mathcal F}}{^{[d]}}$ such that $d(S_\epsilon\cdot x', S_\epsilon\cdot y')<\delta$ for every $\epsilon\neq\emptyset$.) We call this the [*regionally proximal relation of order $d$*]{} and denote the set of regionally proximal points by ${{\bf RP}}{^{[d]}}$ (or by ${{\bf RP}}{^{[d]}}(X)$ in case of ambiguity). Since ${{\bf RP}}{^{[d+1]}}$ is finer than ${{\bf RP}}{^{[d]}}$, we have defined a nested sequence of closed and invariant relations. \[lemma:RPandQQ\] Assume that $(X,T)$ is a transitive system and that $d\geq 1$ is an integer. Then $(x,y)\in{{\bf RP}}{^{[d]}}$ if and only if there exists ${\mathbf{a}}_* \in X_*{^{[d]}}$ such that $$(x, {\mathbf{a}}_*, y, {\mathbf{a}}_*) \in {{\bf Q}}{^{[d+1]}}$$ Assume that $(x,y)\in{{\bf RP}}{^{[d]}}$. Let $\delta>0$ and let $x',y'$ and $S$ be as in the definition of regionally proximal points. As transitive points are dense in $X$, there exists a transitive point $z$ with $d(z,x')<\delta$ and, for every $\epsilon\neq\emptyset$, $d(S_\epsilon\cdot z, S_\epsilon\cdot x')<\delta$. There exists an integer $k$ such that $d(T^k z,y')<\delta$ and that, for every $\epsilon\neq\emptyset$, $d(S_\epsilon \cdot T^k z,S_\epsilon \cdot y')<\delta$. We have that $d(z,x)<2\delta$, $d(T^k z,y)<2\delta$ and $d(S_\epsilon\cdot \ T^k z, S_\epsilon\cdot \ z)<3\delta$. Define ${\mathbf{z}}\in X{^{[d+1]}}$ by $z_{\epsilon 0} = S_\epsilon\cdot z$ and $z_{\epsilon 1} = S_\epsilon\cdot T^k z$ for $\epsilon\in\{0,1\}^d$. Then ${\mathbf{z}}=(S,S)(T{^{[d+1]}}_{d+1})^kz{^{[d+1]}}$ and thus this point belongs to ${{\bf Q}}{^{[d+1]}}$. We have that $d(z_\emptyset,x)<2\delta$, $d(z_{00\ldots 01},y)<2\delta$ and $d(z_{\epsilon 0},z_{\epsilon 1})<3\delta$ for every $\epsilon\in\{0,1\}^d$ different from $\emptyset$. Letting $\delta\to 0$ and passing to a subsequence, we have a point of ${{\bf Q}}{^{[d+1]}}$ of the announced form. Conversely, if $(x,{\mathbf{a}}_*,y,{\mathbf{a}}_*) \in{{\bf Q}}{^{[d+1]}}$ with ${\mathbf{a}}_* \in X_*{^{[d]}}$, then for every $\delta > 0$, there exist $z\in X$, ${\mathbf{n}}\in{{\mathbb Z}}^{d}$, and $p\in{{\mathbb Z}}$ such that $d(z,x)<\delta$, $d(T^pz,y)< \delta$, and $d(T^{{\mathbf{n}}\cdot \epsilon}z, a_\epsilon) < \delta$ and $d(T^{{\mathbf{n}}\cdot\epsilon + p}z, a_\epsilon)< \delta$ for every nonempty $\epsilon\subset [d]$. Thus $(x,y)\in{{\bf RP}}{^{[d]}}$. \[lemma:stupid\] Assume that $(X,T)$ is a transitive system and that $d\geq 1$ is an integer. The relation ${{\bf RP}}{^{[d]}}(X)$ is a closed, symmetric relation that is invariant under $T$. If $\phi\colon X\to Y$ is a factor map and if $(x, y)\in{{\bf RP}}{^{[d]}}(X)$, then $(\phi(x), \phi(y))\in{{\bf RP}}{^{[d]}}(Y)$. This follows immediately from the definition and Lemma \[lemma:RPandQQ\]. If the first property of Theorem \[theorem:main\] holds, then the relation ${{\bf RP}}{^{[d]}}$ is trivial: if $(x,{\mathbf{a}}_*, y, {\mathbf{a}}_*)\in{{\bf Q}}{^{[d+1]}}$, then $(x, {\mathbf{a}}_*)\in{{\bf Q}}{^{[d]}}$ and so $(x, {\mathbf{a}}_*, x, {\mathbf{a}}_*)\in{{\bf Q}}{^{[d+1]}}$. By the first property of Theorem \[theorem:main\], $x=y$. Reduction to the distal case {#sec:reduce} ---------------------------- We show that systems verifying the conditions of Theorem \[theorem:main\] are distal. \[prop:rpsequal\] Assume $(X,T)$ is a transitive system and that $d\geq 1$ is an integer. If $x$ and $y$ are proximal and the closed orbit of $y$ is a minimal set, then $(x,y,y, \ldots, y)\in{{\bf Q}}{^{[d]}}$. First we claim that for every $\eta > 0$, there exists $n\in{{\mathbb N}}$ such that $d(T^nx,y) < \eta$ and $d(T^ny,y) < \eta$. Since $x$ and $y$ are proximal, there exists a sequence $(m_i\colon i\geq 1)$ and a point $z\in X$ such that $T^{m_i}x\to z$ and $T^{m_i}y\to z$. We have that $z$ belongs to the closed orbit of $y$, which is minimal, and so $y$ belongs to the closed orbit of $z$. Thus there exists $p$ such that $d(T^pz, y)< \eta/2$. By continuity of $T^p$, for $i$ sufficiently large we have that $d(T^{m_i+p}x, y) < \eta$ and $d(T^{m_i+p}y, y) < \eta$. Setting $n = m_i+p$ for some sufficiently large $i$, we have $n$ that satisfies the claim. Fix $\delta > 0$. Applying the claim for $\eta = \delta/d$, we find some $n_1$ such that $d(T^{n_1}x, y) < \delta/d$ and $d(T^{n_1}y, y) < \delta/d$. Taking $\eta$ with $0< \eta < \delta/d$ such that $d(T^{n_1}u,T^{n_1}v) \leq \delta/d$ when $d(u,v)\leq \eta$, and then taking $n_2$ associated to this $\eta$, from the claim we have that: $d(T^{n_1\epsilon_1 + n_2\epsilon_2}x, y) < 2\delta/d$ and $d(T^{n_1\epsilon_1 + n_2\epsilon_2}y, y) < 2\delta/d$ for all $\epsilon_1, \epsilon_2\in\{0,1\}^2$ other than $\epsilon_1=\epsilon_2 =0$. Thus by induction, there is a sequence of integers $n_1, \ldots, n_d$ such that $d(T^{{\mathbf{n}}\cdot\epsilon}x, y) < \delta$ for all $\emptyset\neq\epsilon\subset[d]$ . Taking $\delta\to 0$, we have the statement of the proposition. \[cor:two\_distal\] Assume that $(X,T)$ is a transitive system. If the second property of Theorem \[theorem:main\] holds, then $X$ is distal. We first show that any point in $X$ is minimal, i.e. its closed orbit is minimal, and so the system is minimal. Every $x\in X$ is proximal to some minimal point $y$ (see [@Aus]). By the previous proposition and the hypothesis, $x=y$ and so $x$ is a minimal point. Applying the proposition to any pair of proximal points, the statement follows. Parallelepipeds in distal systems {#section:paradistal} ================================= Minimal distal systems and parallelepiped structures ---------------------------------------------------- \[lemma:Qminimal\] Let $(X,T)$ be a minimal distal system and let $d\geq 1$ be an integer. Then $({{\bf Q}}{^{[d]}}, {{\mathcal G}}{^{[d]}})$ is a minimal distal system. Since $(X,T)$ is distal, so is the system $(X{^{[d]}}, {{\mathcal G}}{^{[d]}})$. Since ${{\bf Q}}{^{[d]}}$ is a closed and invariant subset of $X{^{[d]}}$ under the face transformations, the system $({{\bf Q}}{^{[d]}}, {{\mathcal G}}{^{[d]}})$ is also distal. By the second part of Lemma \[lemma:no-name\], the system is transitive and thus is minimal. Using the Ellis semigroup, Eli Glasner [@glasner] showed us a proof that this lemma holds without the assumption of distality. \[prop:equiv\] Let $(X,T)$ be a minimal distal system and let $d\geq 1$ be an integer. The relation $\sim_{d-1}$ defined on ${{\bf Q}}{^{[d-1]}}$ by $${\mathbf{x}}\sim_{d-1} {\mathbf{x}}' \text{ if and only if the element } ({\mathbf{x}}, {\mathbf{x}}')\in X{^{[d]}} \text{ belongs to } {{\bf Q}}{^{[d]}}$$ is an equivalence relation. By Property  of Section \[subsec:dynparallel\], we have that the relation is symmetric and by Property , it is reflexive. We are left with showing that the relation is transitive. Let ${\mathbf{u}}, {\mathbf{v}}, {\mathbf{w}}\in{{\bf Q}}{^{[d-1]}}$ and assume that $({\mathbf{u}},{\mathbf{v}})\in{{\bf Q}}{^{[d]}}$ and $({\mathbf{v}}, {\mathbf{w}})\in{{\bf Q}}{^{[d]}}$. Choose $z\in X$. By Lemma \[lemma:Qminimal\], the system $({{\bf Q}}{^{[d]}}, {{\mathcal G}}{^{[d]}})$ is minimal and so it is the closed orbit of $z{^{[d]}}$ under the group ${{\mathcal G}}{^{[d]}}$. There exists a sequence $(S_{i}\colon i\geq 1)$ such that $S_{i}({\mathbf{u}}, {\mathbf{v}})\to z{^{[d]}}=(z{^{[d-1]}}, z{^{[d-1]}})$ as $i\to \infty$. Writing $S_i = (S'_i, S''_i)$ with $S'_i, S''_i\in {{\mathcal G}}{^{[d-1]}}$, we have that $S'_i{\mathbf{u}}\to z{^{[d-1]}}$ and $S''_i{\mathbf{v}}\to z{^{[d-1]}}$. Passing to a subsequence if needed, we can assume that $S''_i{\mathbf{w}}$ converges to some point $\hat{{\mathbf{z}}}\in X{^{[d-1]}}$ as $i\to\infty$. We have that $$(S''_i, S''_i)({\mathbf{v}},{\mathbf{w}})\to (z{^{[d-1]}}, \hat{{\mathbf{z}}})\in X{^{[d]}}\ .$$ But for each $i\in{{\mathbb N}}$, $(S''_i, S''_i)\in {{\mathcal G}}{^{[d]}}$ and thus $(z{^{[d-1]}}, \hat{{\mathbf{z}}})$ belongs to the closed orbit of $({\mathbf{v}},{\mathbf{w}})$ under ${{\mathcal G}}{^{[d]}}$ and so $(z{^{[d-1]}}, \hat{{\mathbf{z}}})\in{{\bf Q}}{^{[d]}}$. On the other hand, $S_i({\mathbf{u}}, {\mathbf{w}}) = (S'_i{\mathbf{u}}, S''_i{\mathbf{w}})$ converges to $(z{^{[d-1]}}, \hat{{\mathbf{z}}})$ and this point belongs to the closed orbit of $({\mathbf{u}}, {\mathbf{w}})$ under ${{\mathcal G}}{^{[d]}}$. By distality this orbit is minimal and so it follows that $({\mathbf{u}}, {\mathbf{w}})$ also belongs to the orbit closure of $(z{^{[d-1]}}, \hat{{\mathbf{z}}})$. In particular, $({\mathbf{u}}, {\mathbf{w}})\in{{\bf Q}}{^{[d]}}$ and the relation $\sim_{d-1}$ is transitive. \[cor:trans\] Let $(X,T)$ be a minimal distal system and let $d\geq 1$ be an integer. If ${\mathbf{x}},{\mathbf{y}}\in{{\bf Q}}{^{[d+1]}}$ and $x_\epsilon=y_\epsilon$ for all $\epsilon\neq\emptyset$, then $(x_\emptyset,y_\emptyset)\in{{\bf RP}}{^{[d]}}$. We write ${\mathbf{x}}=(x_\emptyset, {\mathbf{a}}_*,{\mathbf{z}})$ with ${\mathbf{a}}_*\in X{^{[d]}}_*$ and ${\mathbf{z}}\in{{\bf Q}}{^{[d]}}$. By hypothesis, ${\mathbf{y}}=(y_\emptyset,{\mathbf{a}}_*,{\mathbf{z}})$ and by transitivity of relation $\sim_{d+1}$, we have that $(x_\emptyset,{\mathbf{a}}_*,y_\emptyset,{\mathbf{a}}_*)\in{{\bf Q}}{^{[d+1]}}$. We conclude via Lemma \[lemma:RPandQQ\]. Completing parallelepipeds -------------------------- For $x\in X$ and $d\geq 1$, write $${{\bf Q}}{^{[d]}}(x)=\{{\mathbf{y}}\in{{\bf Q}}{^{[d]}}\colon y_\emptyset =x\}\ .$$ In this section, we show: \[prop:QK\] For $x\in X$ and $d\geq 1$, ${{\bf Q}}{^{[d]}}(x)$ is the closed orbit of $x{^{[d]}}$ under the action of the group ${{\mathcal F}}{^{[d]}}$. Proposition \[prop:QK\] follows from the more general Proposition \[prop:important\] below. In this section (and only in this section), we make use of yet another notation for the points of $X{^{[d]}}$: For $\epsilon \subset [d]$, define $$\sigma_d(\epsilon) = \sum_{k=1}^d\epsilon_k2^{k-1} \ .$$ For $0\leq j < 2^d$, set $$E(d,j) = \{\epsilon\subset[d]\colon \sigma_d(\epsilon)\leq j\} \ .$$ For $x\in X$ and $d\geq 1$, let ${{\bf K}}{^{[d]}}(x)$ denote the closed orbit of $x{^{[d]}}$ under ${{\mathcal F}}{^{[d]}}$. We remark that ${{\bf K}}{^{[d]}}(x)$ is minimal under the action of ${{\mathcal F}}{^{[d]}}$. Moreover, if $d\geq 2$ and ${\mathbf{y}}\in{{\bf K}}{^{[d-1]}}(x)$, then $({\mathbf{y}},{\mathbf{y}})\in{{\bf K}}{^{[d]}}(x)$. As well, ${{\bf K}}{^{[d]}}(x)$ is invariant under digit permutations. \[prop:important\] Assume that $d\geq 1$ is an integer and let $0\leq j< 2^d$. Assume that ${\mathbf{x}}\in X{^{[d]}}$ satisfies the hypothesis - for every $r$ and every face $F$ of dimension $r$ of $\{0,1\}^{d}$ included in $E(d,j)$, the projection of ${\mathbf{x}}$ along $F$ belongs to ${{\bf Q}}{^{[r]}}$. Then there exists ${\mathbf{w}}\in {{\bf K}}{^{[d]}}(x_\emptyset)$ such that $w_\epsilon=x_\epsilon$ for every $\epsilon\in E(d,j)$. First we remark that this proposition implies Proposition \[prop:QK\]. Indeed, if ${\mathbf{x}}\in{{\bf Q}}{^{[d]}}(x)$ then $x_\emptyset=x$. Moreover, ${\mathbf{x}}$ satisfies the hypothesis $H(d,2^d-1)$ and thus agrees with a point of ${{\bf K}}{^{[d]}}(x)$ on $E(d,2^d-1)=[d]$. For $d=1$, the result is obvious since ${{\bf K}}{^{[1]}}(x_\emptyset)=\{x_\emptyset\}\times X$. For $d>1$ and $j=0$, there is nothing to prove. We proceed by induction: take $d>1$ and $j>0$ and assume that the result holds for $d-1$ and all values of $j$ and for $d$ and $j'<j$. Assume that ${\mathbf{x}}\in X{^{[d]}}$ satisfies the hypothesis $H(d,j)$ and write $x=x_\emptyset$. ### We first make a reduction. We assume that the result holds under the additional hypothesis - ${\mathbf{x}}$ is of the form $x_\epsilon = x_\emptyset$ for $\epsilon\in E(d, j-1)$ and we show that it holds in the general case. Assume that ${\mathbf{x}}$ satisfies $H(d,j)$. By the induction hypothesis, there exists ${\mathbf{v}}\in {{\bf K}}{^{[d]}}(x)$ such that $v_\epsilon=x_\epsilon$ for all $\epsilon\in E(d,j-1)$. By minimality, the point $x{^{[d]}}$ lies in the closed ${{\mathcal F}}{^{[d]}}$-orbit of ${\mathbf{v}}$, meaning that there exists a sequence $(S_\ell\colon \ell\geq 1)$ in ${{\mathcal F}}{^{[d]}}$ such that $S_\ell{\mathbf{v}}\to x{^{[d]}}$. Passing to a subsequence, we can assume that $S_\ell{\mathbf{x}}\to{\mathbf{x}}'$. We have that $x_\epsilon' = x$ for all $\epsilon\in E(d, j-1)$ and ${\mathbf{x}}'$ satisfies property $(*)$. Property $H(d,j)$ is invariant under the action of ${{\mathcal F}}{^{[d]}}$ and under passage to limits. Thus since ${\mathbf{x}}'$ lies in the closed ${{\mathcal F}}{^{[d]}}$-orbit of ${\mathbf{x}}$, ${\mathbf{x}}'$ satisfies $H(d,j)$. Using the result of the proposition with the additional assumption of $(*)$, we have that there exists ${\mathbf{v}}'\in {{\bf K}}{^{[d]}}(x)$ such that $v_\epsilon'=x'_\epsilon$ for $\epsilon\in E(d,j)$. Since the system is distal and ${\mathbf{x}}'$ belongs to the closed ${{\mathcal F}}{^{[d]}}$-orbit of ${\mathbf{x}}$, we also have that ${\mathbf{x}}$ belongs to the closed ${{\mathcal F}}{^{[d]}}$-orbit of ${\mathbf{x}}'$. There exists a sequence $(S_\ell'\colon \ell\geq 1)$ such that $S_\ell'{\mathbf{x}}'\to{\mathbf{x}}$. Passing to a subsequence, we have that $S_\ell'{\mathbf{v}}'\to{\mathbf{u}}$. Thus ${\mathbf{u}}\in {{\bf K}}{^{[d]}}(x)$ and $u_\epsilon=x_\epsilon$ for $\epsilon\in E(d,j)$. ### We now assume ${\mathbf{x}}$ satisfies $H(d,j)$ and $(*)$ and assume that $j\neq 2^{d}-1$. Again, we write $x = x_\emptyset$. Let $\eta\in\{0,1\}^{d}$ be defined by $\sigma_{d}(\eta) = j$. By hypothesis, there exists some $k$ with $1\leq k \leq d$ such that $\eta_k = 0$. Choose $k$ to be the largest $k$ with this property. Define the map $\Phi\colon\{0,1\}^{d-1}\to\{0,1\}^{d}$ by $$\Phi(\epsilon) = \epsilon_1\ldots\epsilon_{k-1}0\epsilon_{k}\ldots\epsilon_{d-1} \ .$$ Setting $$\theta = \eta_1\ldots\eta_{k-1}1\ldots 1\in\{0,1\}^{d-1}\ ,$$ we have that $\Phi(\theta) = \eta$. Set $i = \sigma_{d-1}(\theta)$. It is easy to check that for $\alpha\in\{0,1\}^{d-1}$, $$\label{eq:iff} \sigma_{d-1}(\alpha) < i \text{ if and only if } \sigma_{d}(\Phi(\alpha)) < \sigma_{d}(\Phi(\theta) )= j \ .$$ In particular, $\Phi(E(d-1,i))\subset E(d, j)$. Define ${\mathbf{u}}\in X{^{[d-1]}}$ to be the projection of ${\mathbf{x}}$ on $X{^{[d-1]}}$ along the face defined by $\epsilon_k = 0$. In other words, $$u_\epsilon = x_{\Phi(\epsilon)}\ , \epsilon\in\{0,1\}^{d-1}.$$ Moreover, if $F$ is a face of $\{0,1\}^{d-1}$, then $\Phi(F)$ is a face of $\{0,1\}^{d}$. Since ${\mathbf{x}}$ satisfies $H(d,j)$, we have that ${\mathbf{u}}$ satisfies $H(d-1, i)$. We have that $u_\emptyset = x$ and by the induction hypothesis, there exists ${\mathbf{v}}\in {{\bf K}}{^{[d-1]}}(x)$ with $v_\epsilon = u_\epsilon$ for all $\epsilon\in E(d-1,i)$. Define the map $\Psi\colon\{0,1\}^{d} \to \{0,1\}^{d-1}$ by $$\Psi(\epsilon) = \epsilon_1\ldots\epsilon_{k-1}\epsilon_{k+1}\ldots\epsilon_{d} \ .$$ By definition, $\Psi\circ\Phi$ is the identity and $\Psi(\eta) = \theta$. On the other hand, $\Phi\circ\Psi(\epsilon) = \epsilon_1\ldots\epsilon_{k-1}0\epsilon_{k+1}\ldots\epsilon_d$. In particular, $$\sigma_d(\Phi\circ\Psi(\epsilon)) \leq \sigma_d(\epsilon) \text{ for every } \epsilon\in\{0,1\}^d \ .$$ Define ${\mathbf{w}}\in X{^{[d]}}$ by $w_\epsilon = v_{\Psi(\epsilon)}$ for $\epsilon\in\{0,1\}^{d}$. In other words, ${\mathbf{w}}$ is obtained by duplicating ${\mathbf{v}}$ on two opposite faces. We check that ${\mathbf{w}}\in {{\bf K}}{^{[d]}}(x)$. To see this, let ${\mathbf{v}}'$ be obtained from ${\mathbf{v}}$ by the digit permutation that exchanges the digits $k-1$ and $d-1$. Then ${\mathbf{v}}'\in{{\bf K}}{^{[d-1]}}(x)$ and so $({\mathbf{v}}',{\mathbf{v}}')\in {{\bf K}}{^{[d]}}(x)$. We obtain ${\mathbf{w}}$ from the point $({\mathbf{v}}',{\mathbf{v}}')$ by the digit permutation that exchanges the digits $k$ and $d$. We claim that $\Psi(E(d,j-1))\subset E(d-1,i-1)$. To show this, we take $\epsilon\in E(d, j-1)$ and distinguish two cases. First assume there exists some $m$ with $k+1\leq m \leq d$ with $\epsilon_m=0$. Then one of the $d-k$ last coordinates of $\Psi(\epsilon) = 0$ and by definition of $\theta$, $\sigma_{d-1}(\Psi(\epsilon)) < \sigma_{d-1}(\theta) = i$. Now assume that here is no such $m$. Because $\sigma_{d}(\epsilon)< \sigma_{d}(\eta)$ and $\eta_k=0$, we have that $\epsilon_k = 0$. Then $\Phi(\Psi(\epsilon)) = \epsilon $. Thus $$\sigma_{d}(\Phi(\Psi(\epsilon))) = \sigma_{d}(\epsilon) < j$$ and applying  with $\alpha = \Psi(\epsilon)$, we have that $\sigma_{d-1}(\Psi(\epsilon))< i$. This proves the claim. We check that ${\mathbf{w}}$ satisfies the conclusion of the proposition. First for $w_\eta$, we have that $w_\eta = v_{\Psi(\eta)} = v_{\theta} = u_{\theta}$ since $\theta\in E(d-1, i)$, and $u_\theta = x_{\Phi(\theta)} = x_\eta$. Thus $w_\eta = x_\eta$. Next, if $\epsilon\in E(d,j-1)$, then $x_\epsilon = x$. On the other hand, $w_\epsilon = v_{\Psi(\epsilon)} = u_{\Psi(\epsilon)}$, where the last equality holds because by the claim we have $\Psi(\epsilon) \in E(d-1,i-1)$. But $u_{\Psi(\epsilon)} = x_{\Phi\circ\Psi(\epsilon)} = x$, because $\sigma_d(\Phi\circ\Psi(\epsilon)) \leq \sigma_d(\epsilon)\leq j-1$. This ${\mathbf{w}}$ is as announced. ### We are left with considering the case that $j=2^{d}-1$ The hypothesis $H(d, 2^{d}-1)$ means that ${\mathbf{x}}=(x, x, \ldots, x, y)\in {{\bf Q}}{^{[d]}}$ and we have to show that this lies in ${{\bf K}}{^{[d]}}(x)$. We start with a general property. Writing a point ${\mathbf{x}}\in X{^{[d]}}$ as ${\mathbf{x}}= ({\mathbf{x}}',{\mathbf{x}}'')$, define the projection $\phi\colon {{\bf K}}{^{[d]}}(x)\to{{\bf Q}}{^{[d-1]}}$ by $\phi({\mathbf{x}}) = {\mathbf{x}}''$. The range of $\phi$ is invariant under the group ${{\mathcal G}}{^{[d-1]}}$ and thus by Lemma \[lemma:Qminimal\], it is equal to ${{\bf Q}}{^{[d-1]}}$. By distality, the map $\phi$ is open. Assume $(x, x, \ldots, x,y)\in{{\bf Q}}{^{[d]}}$. Write ${\mathbf{v}}= (x, \ldots, x, y)\in X{^{[d-1]}}$. Let $\delta > 0$. Since $(x{^{[d-1]}}, x{^{[d-1]}})\in {{\bf K}}{^{[d]}}(x)$, by the openness of $\phi$, there exists $\delta'$ with $0< \delta' <\delta$ such that if ${\mathbf{u}}\in{{\bf Q}}{^{[d-1]}}$ is $\delta'$-close to $x{^{[d-1]}}$, there exists ${\mathbf{z}}$ that is $\delta$-close to $x{^{[d-1]}}$ and $({\mathbf{z}}, {\mathbf{u}})\in {{\bf K}}{^{[d]}}(x)$. Since $(x{^{[d-1]}}, {\mathbf{v}})\in{{\bf Q}}{^{[d]}}$, there exists ${\mathbf{u}}\in {{\bf Q}}{^{[d-1]}}$ and $n\in{{\mathbb Z}}$ such that ${\mathbf{u}}$ is at most distance $\delta'$ from $x{^{[d-1]}}$ and $(T{^{[d-1]}})^n{\mathbf{u}}$ is at most distance $\delta$ from ${\mathbf{v}}$. Taking ${\mathbf{z}}$ as above, we have that $({\mathbf{z}}, (T{^{[d-1]}})^n{\mathbf{u}})\in {{\bf K}}{^{[d]}}(x)$ and is $\delta$-close to $(x{^{[d-1]}}, {\mathbf{v}})$. Letting $\delta$ go to $0$, we have that $(x{^{[d-1]}}, {\mathbf{v}})\in {{\bf K}}{^{[d]}}(x)$. The next result follows directly from Proposition \[prop:important\] and the definition of ${{\bf Q}}{^{[d]}}$. It shows that ${{\bf Q}}{^{[d]}}$ verifies properties that are generalizations of the $2-$and $3-$dimensional parallelepiped structures as defined in [@HM]. In particular, ${{\bf Q}}{^{[d]}}$ satisfies the “property of closing parallelepipeds”. This plays a key role in our study of the first condition in Theorem \[theorem:main\]. \[prop:closing\] Let $(X,T)$ be a minimal distal system and let $d \geq 1$ be an integer. Assume that $x_\epsilon$, $\epsilon\subset [d]$ with $\epsilon \neq [d]$, are points in $X$ such that the face $(x_\epsilon \colon j \notin \epsilon)$ belongs to ${{\bf Q}}{^{[d-1]}}$ for each $j \in [d]$. Then there exists $x_{[d]}\in X$ such that $(x_\epsilon \colon\epsilon \subset [d]) \in{{\bf Q}}{^{[d]}}$. Although we have given the last coordinate in the statement of this proposition a particular role, using Euclidean permutations the analogous statement holds for any other fixed coordinate, provided that the corresponding faces lie in ${{\bf Q}}{^{[d-1]}}$. Strong form of the regionally proximal relation ----------------------------------------------- \[cor:replace\] Let $(X,T)$ be a minimal distal system and let $d \geq 1$ be an integer. Let $x,y\in X$ and ${\mathbf{b}}_*\in X{^{[d+1]}}_*$ with $(x,{\mathbf{b}}_*)\in{{\bf Q}}{^{[d+1]}}$. Then $(y,{\mathbf{b}}_*)\in{{\bf Q}}{^{[d+1]}}$ if and only if $(y,x,x,\ldots,x)\in{{\bf Q}}{^{[d+1]}}$. We write ${\mathbf{u}}=(x,{\mathbf{b}}_*)$, ${\mathbf{v}}=(y,{\mathbf{b}}_*)$, and ${\mathbf{y}}=(y,x,x,\ldots,x)\in X{^{[d+1]}}$. By Proposition \[prop:important\], we have that ${\mathbf{u}}$ belongs to ${{\bf K}}{^{[d+1]}}(x)$ and, by minimality, there exists a sequence $(S_n\colon n\geq 1)$ in ${{\mathcal F}}{^{[d+1]}}$ such that $S_n{\mathbf{u}}\to x{^{[d+1]}}$. Then $S_n{\mathbf{v}}\to {\mathbf{y}}$ and ${\mathbf{y}}$ belongs to the closed orbit of ${\mathbf{v}}$ under ${{\mathcal F}}{^{[d+1]}}$. By distality, this last property implies that ${\mathbf{v}}$ belongs to the closed orbit of ${\mathbf{y}}$. Since ${{\bf Q}}{^{[d+1]}}$ is closed and invariant under ${{\mathcal F}}{^{[d+1]}}$, we have that ${\mathbf{y}}\in{{\bf Q}}{^{[d+1]}}$ if and only if ${\mathbf{v}}\in{{\bf Q}}{^{[d+1]}}$. \[cor:RPRPs\] Let $(X,T)$ be a minimal distal system and let $d \geq 1$ be an integer. Let $x,y\in X$. Then $(x,y)\in{{\bf RP}}{^{[d]}}$ if and only if $(y,x,x,\ldots,x)\in{{\bf Q}}{^{[d+1]}}={{\bf K}}{^{[d+1]}}(y)$. For ${\mathbf{a}}^*\in X{^{[d]}}_*$, apply the preceding corollary with ${\mathbf{b}}_*=({\mathbf{a}}_*,x,{\mathbf{a}}_*)$ and use Lemma \[lemma:RPandQQ\]. The combination of the previous corollaries allows to prove that each coordinate in a parallelepiped of ${{\bf Q}}{^{[d]}}$ can be replaced by another point that is regionally proximal of order $d$ with it and the resulting point is still a parallelepiped. We finish with a comment about the regionally proximal relation of order $d$. In [@Aus], Corollary 10, chapter 9, Auslander (see also Ellis [@Ellis]) proves that in the definition of the regionally proximal relation, the point $x'$ (see Definition \[def:RP\] with $d=1$) can be taken to be $x$. The same result can be stated for the regionally proximal relation of order $d$ in the distal case. In fact, a similar argument to the one used to prove Lemma \[lemma:RPandQQ\] allows us to show that: $(x,y,\ldots,y)\in{{\bf Q}}{^{[d+1]}}={{\bf K}}{^{[d+1]}}(x)$ if and only if for any $\delta >0$ there exist $y' \in X$ and a vector ${\mathbf{n}}=(n_1,\ldots,n_d) \in {{\mathbb Z}}^d$ such that for any nonempty $\epsilon \subset [d]$ $$d(y,y')< \delta \ , \ d(T^{ {\mathbf{n}}\cdot \epsilon }x,y)<\delta, \text{ and } d(T^{ {\mathbf{n}}\cdot \epsilon}y',y)< \delta$$ Summarizing ----------- ### We show that the second property in Theorem \[theorem:main\] implies the first one. Assume that the transitive system $(X,T)$ satisfies the second property. By Corollary \[cor:two\_distal\], the system is distal. If ${\mathbf{x}},{\mathbf{y}}\in{{\bf Q}}{^{[d+1]}}$ agree on all coordinates other than the coordinate indexed by $\emptyset$, then ${\mathbf{x}}={\mathbf{y}}$ by Corollary \[cor:replace\]. By permutation of coordinates we deduce that the first property of Theorem \[theorem:main\] is satisfied. The first two properties of this theorem are thus equivalent. From the above discussion, Proposition \[prop:orderRP\] follows: these properties mean that the relation ${{\bf RP}}{^{[d]}}$ is trivial. ### \[prop:quaotient1\] Let $(X,T)$ be a minimal distal system and let $d\geq 1$ be an integer. Then the relation ${{\bf RP}}{^{[d]}}$ is a closed invariant equivalence relation on $X$. The quotient of $X$ under this equivalence relation is the maximal factor of order $d$ of $X$. The second statement means that this quotient is a system of order $d$ and that every system of order $d$ which is a factor of $X$ is a factor of this quotient. In order to prove the first statement, we are left with showing that the relation is transitive. Assume that $(x,y)$ and $(y,z)\in{{\bf RP}}{^{[d]}}$. By Corollary \[cor:RPRPs\] applied to the pair $(x,y)$, $(y,x,x,\ldots,x)\in{{\bf Q}}{^{[d+1]}}$. By Corollary \[cor:replace\] applied to the pair $(y,z)$, $(z,x,x,\ldots,x)\in{{\bf Q}}{^{[d+1]}}$ and by Corollary \[cor:RPRPs\] again, $(x,z)\in{{\bf RP}}{^{[d]}}$. We show now the second part of the proposition. Let $Y$ be the quotient of $X$ under the equivalence relation ${{\bf RP}}{^{[d]}}$ and let $\phi$ denote the factor map. Let $(a, b)\in{{\bf RP}}{^{[d]}}(Y)$. Then $(a, b, b, \ldots, b)\in{{\bf Q}}{^{[d+1]}}(Y)$. By Lemma 3.1, there exists ${\mathbf{x}}\in {{\bf Q}}{^{[d+1]}}(X)$ satisfying $\phi{^{[d+1]}}({\mathbf{x}}) = (a, b, b, \ldots, b)$. Write $x_\emptyset = x$ and $x_{00\ldots 01} =y$. For every $\epsilon\neq\emptyset$, $\phi(x_\epsilon) = b = \phi(y)$. Thus $(x_\epsilon, y)\in{{\bf RP}}{^{[d]}}(X)$. Using Corollary \[cor:RPRPs\] and Corollary \[cor:replace\], we can replace $x_\epsilon$ by $y$ in ${\mathbf{x}}$ and obtain an element of ${{\bf Q}}{^{[d+1]}}(X)$. Doing this for all $\epsilon\neq\emptyset$, we have that $(x, y, y,\ldots, y)\in{{\bf Q}}{^{[d+1]}}(X)$. By Corollary \[cor:RPRPs\], this means that $(x,y)\in{{\bf RP}}{^{[d]}}(X)$. Thus that $\phi(x) = \phi(y)$ and so $a=b$. Let $W$ be a system of order $d$ and let $\psi\colon X\to W$ be a factor map. Take $Y$ and $\phi$ as above and let $x,y\in X$. If $\phi(x) = \phi(y)$, then $(x,y)\in{{\bf RP}}{^{[d]}}(X)$. Thus by Corollary \[lemma:stupid\], $(\psi(x), \psi(y))\in{{\bf RP}}{^{[d]}}(W)$ and thus $\psi(x) = \psi(y)$. ### In order to complete the proofs of Theorems \[theorem:main\] and \[theorem:maxd\], we are left with showing that the notions of a system of order $d$ and an inverse limit of $d$-step minimal nilsystems are equivalent. In one direction, a result from Appendix B of [@HK3], translated into our current vocabulary, states that a $d$-step minimal nilsystem is a system of order $d$. This property easily passes to inverse limits, and so we have: \[prop:Qnil\] Let $(X,T)$ be an inverse limit of minimal $(d-1)$-step nilsystems and let $d\geq 2$ be an integer. Then $(X,T)$ is a system of order $d-1$. We are left with showing the converse, which is: \[th:converse\] Assume that $(X,T)$ is a transitive system of order $d-1$. Then it is an inverse limit of $d-1$-step minimal nilsystems. We recall that the hypothesis of this theorem means that if ${\mathbf{x}},{\mathbf{y}}\in{{\bf Q}}{^{[d]}}$ have $2^{d}-1$ coordinates in common, then ${\mathbf{x}}={\mathbf{y}}$. In particular, this implies that the system is distal and minimal. The proof of this theorem is carried out in the next two sections. Ergodic preliminaries {#section:preergodic} ===================== The result of Theorem \[th:converse\] is established in the next section using invariant measures on $X$. In this section, we summarize the background material and give some preliminary results. Inverse limits of nilsystems {#subsec:invelimit} ---------------------------- A [*measure preserving system*]{} is defined to be a quadruple $(X, {{\mathcal B}}, \mu, T)$, where $(X, {{\mathcal B}}, \mu)$ is a probability space and $T\colon X\to X$ is a measure preserving transformation. In general, we omit the $\sigma$-algebra ${{\mathcal B}}$ from the notation and write $(X,\mu, T)$. Throughout, we make use both of the vocabulary of topological dynamics and of ergodic theory, leading to possible confusion. In general, it is clear from the context whether we are referring to a measure preserving system or a topological system, and so we just refer to either as a [*system*]{}. Topological factor maps were already defined. We recall that an *ergodic theoretic factor map* between the measure preserving systems $(X, \mu, T)$ and $(X', \mu', T)$ is a measurable map $\pi\colon X\to X'$ (defined almost everywhere), mapping the measure $\mu$ to $\mu'$ and commuting with the transformations (almost everywhere). If the map $\pi$ is invertible (almost everywhere), we say that the two systems are [*isomorphic*]{}. Inverse limits of nilsystems in the topological sense were discussed in Section \[sec:nilsystems\]. We make this notion precise in the measure theoretic sense, in this case also we consider only sequential inverse limits. A $d$-step nilsystem $(X,T)$, endowed with its Haar measure $\mu$, is ergodic if and only if $(X,T)$ is a minimal topological system; in this case, $\mu$ is its unique invariant measure. Therefore, every inverse limit (in the topological sense) of $d$-step minimal nilsystems is uniquely ergodic. Now, let $(X,\mu,T)=\varprojlim (X_j,\mu_j,T_j)$ be an inverse limit in the ergodic theoretic sense of a sequence of $d$-step ergodic nilsystems. Recall that each nilsystem $(X_j,T)$ is endowed with its Borel $\sigma-$algebra and $\mu_j$ its Haar measure. This means that for every $j\in{{\mathbb N}}$, there exist ergodic theoretic factor maps $\pi_j\colon (X_{j+1},\mu_{j+1},T)\to(X_j,\mu_j,T)$ and $p_j\colon (X,\mu,T)\to(X_j,\mu_j,T)$ satisfying $\pi_j\circ p_{j+1}=p_j$ for every $j$ such that the Borel $\sigma$-algebra ${{\mathcal B}}$ of $X$ is spanned by the $\sigma$-algebras $p_j{^{-1}}({{\mathcal B}}_j)$, where ${{\mathcal B}}_j$ denotes the Borel $\sigma$-algebra of $X_j$. Every ergodic theoretic factor map between ergodic nilsystems is equal almost everywhere to a topological factor map. A short proof of this fact is given in the Appendix (Theorem \[fact:factonil\]). Therefore, the factor maps $\pi_j$ in the definition of an inverse limit (in the ergodic sense) can be assumed to be topological factor maps. It follows that $(X,\mu,T)$ can be identified with the topological inverse limit. This allows us, in the sequel, to not distinguish between the notions of topological and ergodic theoretic inverse limits of $d$-step ergodic nilsystems. Ergodic uniformity seminorms and nilsystems {#subsec:mud} ------------------------------------------- Let $(X,\mu,T)$ be an ergodic system. For points in $X{^{[d]}}$ and transformations of these spaces we use the same notation as in the topological setting. In Section 3 of [@HK1], a measure $\mu{^{[d]}}$ on $X{^{[d]}}$ and a seminorm ${|\!|\!|\cdot|\!|\!|}_d$ on $L^\infty(\mu)$ are constructed. We recall the properties of these objects: Assume $(X, \mu, T)$ is an ergodic system and that $d\geq 1$ is an integer. 1. The measure $\mu{^{[d]}}$ is invariant and ergodic under the action of the group ${{\mathcal G}}{^{[d]}}$. 2. Each one dimensional marginal of $\mu{^{[d]}}$ is equal to $\mu$ and each of its two dimensional marginals (meaning the image under the map ${\mathbf{x}}\mapsto(x_\epsilon,x_\theta)$ for $\epsilon\neq\theta\subset[d]$) is equal to $\mu\times\mu$. 3. If $p\colon(X,\mu,T)\to(Y,\nu,T)$ is a factor map then, $\nu{^{[d]}}$ is the image of $\mu{^{[d]}}$ under the map $p{^{[d]}}\colon X{^{[d]}}\to Y{^{[d]}}$. For every $f\in L^\infty(\mu)$, the *$d$-th seminorm ${|\!|\!|f|\!|\!|}_d$ of $f$* is defined by $$\label{eq:seminorm} {|\!|\!|f|\!|\!|}_d^{2^d}=\int\prod_{\epsilon\subset[d]}f(x_\epsilon)\,d\mu{^{[d]}}({\mathbf{x}})\ .$$ We have that: \[fact:seminorm\] Assume that $(X, \mu, T)$ is an ergodic system and let $d\geq 1$ be an integer. 1. \[it:seminormintegral\] For every $f\in L^\infty(\mu)$, $\Bigl|\int f\,d\mu\Bigr|\leq{|\!|\!|f|\!|\!|}_d$. 2. If $p\colon(X,\mu,T)\to(Y,\nu,T)$ is a factor map, then ${|\!|\!|f|\!|\!|}_d={|\!|\!|f\circ p|\!|\!|}_d$ for every function $f\in L^\infty(\nu)$. We summarize some of the main results of [@HK1]: \[fact:structure\] Assume that $(X,\mu,T)$ is an ergodic system and that $d\geq 1$ is an integer. The following properties are equivalent: 1. $(X,\mu,T)$ is measure theoretically isomorphic to an inverse limit of $(d-1)$-step ergodic nilsystems. 2. ${|\!|\!|\cdot|\!|\!|}_d$ is a norm on $L^\infty(\mu)$ (equivalently ${|\!|\!|f|\!|\!|}_d=0$ implies that $f=0$). 3. There exists a measurable map $J\colon X{^{[d]}}_*\to X$ such that $x_\emptyset=J(x_\epsilon\colon \emptyset\neq\epsilon\subset[d])$ for $\mu{^{[d]}}$-almost every ${\mathbf{x}}\in X{^{[d]}}$. Using these properties, it follows that any measure theoretic factor of an inverse limit of $(d-1)$-step nilsystems is isomorphic in the ergodic theoretic sense to an inverse limit of $(d-1)$-step nilsystems. \[fact:cubiclimit\] Assume that $(X,\mu, T)$ is an ergodic system, $d\geq 1$ is an integer, and $f_\epsilon\in L^\infty(\mu)$ for $\emptyset\neq\epsilon\subset[d]$. The averages $$\label{eq:cubiclimit} \frac 1{N^d}\sum_{0\leq n_1,\dots,n_d<N} \prod_{\substack{\epsilon\subset[d]\\ \epsilon\neq\emptyset}} f_\epsilon(T^{{\mathbf{n}}\cdot\epsilon }x)$$ converge in $L^2(\mu)$ as $N\to+\infty$. Letting $F$ denote the limit of these averages, we have that for every $g\in L^\infty(\mu)$, $$\label{eq:limcubiclimit} \int g(x)F(x)\,d\mu(x)=\int g(x_\emptyset) \prod_{\substack{\epsilon\subset[d]\\ \epsilon\neq\emptyset}} f_\epsilon(x_\epsilon)\,d\mu{^{[d]}}({\mathbf{x}})\ .$$ \[fact:boudlim\] Let $(X, \mu, T)$, $d$, $f_\epsilon$, $\emptyset\neq\epsilon\subset[d]$, and $F$ be as in Theorem \[fact:cubiclimit\]. Then $${\Vert F\Vert}_{L^\infty(\mu)}\leq \prod_{\substack{\epsilon\subset[d]\\ \epsilon\neq\emptyset}} {\Vert f_\epsilon\Vert}_{L^{2^d-1}(\mu)}\ .$$ Let $g\in L^\infty(\mu)$ and choose a function $h$ with $h^{2^d-1}=g$. By  and the Hölder Inequality, $$\bigl|\int gF\,d\mu\bigr|\leq \Bigl(\prod_{\substack{\epsilon\subset[d]\\ \epsilon\neq\emptyset}} \int |h(x_\emptyset)f_\epsilon(x_\epsilon)|^{2^d-1}\,d\mu{^{[d]}}({\mathbf{x}}) \Bigr)^{1/2^d-1}\ .$$ Since each two dimensional marginal of $\mu{^{[d]}}$ is equal to $\mu\times\mu$, this can be rewritten as $$\Bigl(\prod_{\substack{\epsilon\subset[d]\\ \epsilon\neq\emptyset}} \int |h(x)f_\epsilon(y)|^{2^d-1}\,d\mu(x)\,d\mu(y)\Bigr)^{1/2^d-1} ={\Vert g\Vert}_{L^1(\mu)} \prod_{\substack{\epsilon\subset[d]\\ \epsilon\neq\emptyset}} {\Vert f_\epsilon\Vert}_{L^{2^d-1}(\mu)}$$ and the result follows. Dual functions -------------- Here again, $(X,\mu,T)$ is an ergodic system. Following the notation and terminology of [@HK3], for every $f\in L^\infty(\mu)$, the limit function $$\label{eq:defdual} \lim_{N\to+\infty}\frac 1{N^d}\sum_{n_1,\dots,n_d=0}^{N-1} \prod_{\emptyset\neq \epsilon \subset d }f(T^{{\mathbf{n}}\cdot \epsilon}x)$$ is called the *dual function of order $d$ of $f$* and is written ${{\mathcal D}}_df$. It is worth noting that ${{\mathcal D}}_df$ is only defined as an element of $L^2(\mu)$, and thus is defined almost everywhere. By  and , for every $f\in L^\infty(\mu)$ we have that $$\label{eq:duality} \int f\,{{\mathcal D}}_df\,d\mu={|\!|\!|f|\!|\!|}_d^{2^d}\ .$$ It follows from Lemma \[fact:boudlim\] that: \[fact:boundedDd\] If $(X, \mu, T)$ is an ergodic system and $d\geq 1$ is an integer, then for every $f\in L ^\infty(\mu)$: $${\Vert {{\mathcal D}}_d f\Vert}_{L^\infty(\mu)}\leq {\Vert f\Vert}_{L^{2^d-1}(\mu)}^{2^d-1}\ .$$ Moreover, the map ${{\mathcal D}}_d$ extends to a continuous map from $L^{2^d-1}(\mu)$ to $L^\infty(\mu)$. We remark that if $0\leq f\leq g$, then $0\leq{{\mathcal D}}_d f \leq{{\mathcal D}}_d g$. \[fact:positive\] If $(X, \mu, T)$ is an ergodic system and $d\geq 1$ is an integer, then for every $A\subset X$ we have ${{\mathcal D}}_d{{\bf 1}}_A(x)>0$ for $\mu$-almost every $x\in A$. Let $B=\{x\in A\colon {{\mathcal D}}_d{{\bf 1}}_A(x)=0\}$. By part  of Lemma \[fact:seminorm\] and , since ${{\mathcal D}}_d{{\bf 1}}_B\leq{{\mathcal D}}_d{{\bf 1}}_A$ we have that $$\mu(B)^{2^d}\leq {|\!|\!|{{\bf 1}}_B|\!|\!|}_d^{2^d} = \int_B{{\mathcal D}}_d{{\bf 1}}_B(x)\,d\mu(x)\leq \int_B{{\mathcal D}}_d{{\bf 1}}_A(x)\,d\mu(x)=0\ .$$ Thus $\mu(B)=0$. Using the definition  of the dual function, we immediately deduce: \[fact:dualfactor\] Let $p\colon (X,\mu,T)\to (X',\mu',T)$ be a measure theoretic factor map. For every $f\in L^\infty(\mu')$ we have $({{\mathcal D}}_df)\circ p={{\mathcal D}}_d(f\circ p)$. By Theorem \[fact:cubiclimit\], it follows that: \[fact:suppotymud\] Let $(X,T)$ be a minimal topological dynamical system and $\mu$ be an invariant ergodic measure on $X$. Then the measure $\mu{^{[d]}}$ is concentrated on the subset ${{\bf Q}}{^{[d]}}$ of $X{^{[d]}}$. \[fact:Ddzero\] Let $(X,T)$ be a minimal system of order $d-1$ and let $\mu$ be an invariant ergodic measure on $X$. Let $d_X$ denote a distance on $X$ defining the topology of this space and for every $x\in X$ and $r>0$, let $B(x,r)$ denote the ball centered at $x$ of radius $r$ with respect to the distance $d_X$. Then for every $\eta>0$, there exists $\delta > 0$ such that for every $x\in X$, ${{\mathcal D}}_d{{\bf 1}}_{B(x,\delta)}=0$ $\mu$-almost everywhere on the complement of $B(x,\eta)$ By definition of a system of order $d-1$, the last coordinate of an element of ${{\bf Q}}{^{[d]}}$ is a function of the other ones. Using the symmetries of ${{\bf Q}}{^{[d]}}$, we have that the same property holds with the first coordinate substituted for the last one. Therefore, writting ${{\bf Q}}{^{[d]}}_*$ for ${{\bf Q}}{^{[d]}}$ without the first coordinate, there exists a map $J\colon {{\bf Q}}{^{[d]}}_*\to X$ such that for every ${\mathbf{x}}\in{{\bf Q}}{^{[d]}}$, $$x_\emptyset=J(x_\epsilon\colon \epsilon\subset[d],\ \epsilon\neq\emptyset)\ .$$ The graph of this map is the closed subset ${{\bf Q}}{^{[d]}}$ of ${{\bf Q}}{^{[d]}}_*\times X$ and thus is continuous. Fix $\eta>0$. Since $J$ is uniformly continuous and satisfies $J(x,\dots,x)=x$ for every $x$, there exists $\delta>0$ such that for every $x\in X$, the set $$(X\setminus B(x,\eta))\times B(x,\delta)\times\dots\times B(x,\delta)$$ has empty intersection with ${{\bf Q}}{^{[d]}}$. Thus by Lemma \[fact:suppotymud\] it has zero $\mu{^{[d]}}$-measure. By Theorem \[fact:cubiclimit\] and the definition of ${{\mathcal D}}_d{{\bf 1}}_B$, we have that $$\int{{\bf 1}}_{X\setminus B(x,\eta)}{{\mathcal D}}_d{{\bf 1}}_{B(x,\delta)}\,d\mu=0\ . \qed$$ Systems with continuous dual functions -------------------------------------- It is convenient to give a name to the following, although we only make use of it within proofs: \[def:propdual\] Let $(X,T)$ be a minimal system and let $\mu$ an ergodic invariant measure on $X$. We say that $(X,T,\mu)$ has property ${{\mathcal P}(d)}$ if whenever $f_\epsilon$, $\emptyset\neq\epsilon\subset[d]$, are continuous functions on $X$, the averages  converge everywhere and uniformly. If this property holds, then in particular, for every continuous function $f$ on $X$, the averages   converge everywhere and uniformly for every continuous function $f$ on $X$. The limit of these averages coincides almost everywhere with the function ${{\mathcal D}}_df$ defined above and so we also denote it by ${{\mathcal D}}_df$. \[fact:uniform\] Let $(X,T)$ be an inverse limit of minimal $(d-1)$-step nilsystems and let $\mu$ be the invariant measure of this system. Then $(X,\mu,T)$ has property ${{\mathcal P}(d)}$. Assume first that $(X,\mu,T)$ is a $(d-1)$-step ergodic nilsystem. In [@HK3] (Corollary 5.2), the convergence of the averages  is shown to hold everywhere and this convergence is uniform when the functions $f_\epsilon$, $\emptyset\neq\epsilon\subset [d]$, are continuous. Assume now that $(X,\mu,T)$ is as in the statement. Every continuous function on $X$ can be approximated uniformly by a continuous function arising from one of the nilsystems which are factors of $X$. By density, the result also holds in this case. We now establish some properties of systems with property ${{\mathcal P}(d)}$. We write ${{\mathcal C}}(X)$ for the algebra of continuous functions on $X$. We always assume that ${{\mathcal C}}(X)$ is endowed with the norm of uniform convergence. By Lemma \[fact:boundedDd\] and density: \[fact:continuousDd\] Assume that the ergodic system $(X,\mu,T)$ has property ${{\mathcal P}(d)}$. - For every $f\in L^{2^d-1}(\mu)$, the function ${{\mathcal D}}_df$ is equal $\mu$-almost everywhere to a continuous function on $X$, which we also denote by ${{\mathcal D}}_df$, called the dual function of $f$. - The map $f\mapsto{{\mathcal D}}_df$ is continuous from $L^{2^d-1}(\mu)$ to ${{\mathcal C}}(X)$. \[fact:factoPd\] Let $(X,\mu,T)$ be a system with property ${{\mathcal P}(d)}$, $(Y,T)$ be a minimal system, $p\colon X\to Y$ a topological factor map, and $\nu$ be the image of $\mu$ under $p$. Then $(Y,T,\nu)$ has property ${{\mathcal P}(d)}$. Let $f_\epsilon$, $\emptyset\neq\epsilon\subset[d]$, be continuous functions on $Y$. Then the averages $$\frac 1{N^d}\sum_{0\leq n_1,\dots,n_d<N} \prod_{\substack{\epsilon\subset[d]\\ \epsilon\neq\emptyset}} f_\epsilon(T^{{\mathbf{n}}\cdot\epsilon }p(x))$$ converge uniformly on $X$ and thus the averages  converge uniformly on $Y$. Using a measure {#sec:final} =============== In this section, we prove Theorem \[th:converse\] which completes the proof of Theorem \[theorem:main\]: any transitive system $(X,T)$ of order $d-1$ is an inverse limit of $(d-1)$-step minimal nilsystems. By Corollary \[cor:two\_distal\], $(X,T)$ is distal and thus is minimal. The method we use is completely different from that used in [@HM] for $d=3$, and proceeds by introducing an invariant measure on $X$. We start by reducing the proof of Theorem \[th:converse\] to the following: \[prop:iso\] Let $(X,T)$ be a minimal system of order $d-1$, $\mu$ be an invariant ergodic measure on $X$, and let $(Y,T)$ be an inverse limit of minimal $(d-1)$-step nilsystems with Haar measure $\nu$. Let $\Psi\colon (Y,\nu,T)\to(X,\mu,T)$ be a measure theoretic isomorphism. Then $\Psi$ coincides $\nu$-almost everywhere with a topological isomorphism. By Lemma \[fact:suppotymud\], the measure $\mu{^{[d]}}$ is concentrated on the subset ${{\bf Q}}{^{[d]}}$ of $X{^{[d]}}$. Since $(X,T)$ is a system of order $d-1$, there exists a continuous map $J\colon {{\bf Q}}{^{[d]}}_*\to X$ such that $$x_\emptyset=J(x_\epsilon\colon \epsilon\subset[d],\ \epsilon\neq\emptyset)\text{ for every }{\mathbf{x}}\in{{\bf Q}}{^{[d]}}$$ and so this property holds $\mu{^{[d]}}$-almost everywhere. (Again, ${{\bf Q}}{^{[d]}}_{*}$ denotes ${{\bf Q}}{^{[d]}}$ without the first coordinate.) By Theorem \[fact:structure\], $(X,\mu,T)$ is isomorphic in the ergodic theoretic sense to an inverse limit $(Y,\nu,T)$ of $(d-1)$-step ergodic nilsystems. By Proposition \[prop:iso\], $(X,T)$ and $(Y,S)$ are isomorphic in the topological sense and we are finished. Proof of Proposition \[prop:iso\] {#subsec:endrpoof} --------------------------------- To prove Proposition \[prop:iso\], we start with a lemma: \[lem:Phicontinuous\] Let $(Y,\nu,T)$ be a system with Property ${{\mathcal P}(d)}$, $(X,T)$ be a minimal system of order $d-1$, $\mu$ be an invariant probability measure on $X$, and $\Psi\colon (Y,\nu,T)\to(X,\mu,T)$ be a measure theoretic factor map. Then $\Psi$ agrees $\nu$-almost everywhere with some topological factor map. We can assume that there exists a Borel invariant subset $Y_0$ of full measure and that $\Psi$ is a Borel map from $Y_0$ to $X$, mapping the measure $\nu$ to the measure $\mu$ and such that $\Psi(Tx)=T\Psi(x)$ for every $x\in Y_0$. We claim that: \[fact:openopen\] For every open subset $U$ of $X$, there exists an open subset $\widetilde U$ of $Y$ equal to $\Psi{^{-1}}(U)$ up to a $\nu$-negligible set. To see this, if $(X,\mu)$ is a probability space and $A,B\subset X$, write $A\subset_\mu B$ if $\mu(A\setminus B)=0$. The notations $A\supset_\mu B$ and $A=_\mu B$ are defined similarly. Assume that $U\neq\emptyset$, as otherwise the claim holds trivially. Let $x\in U$. By Lemma \[fact:Ddzero\] there exists an open subset $W_x$ containing $x$ and included in $U$ such that the set $$U_x:=\{x\in X\colon{{\mathcal D}}_d{{\bf 1}}_{W_x}>0\}$$ satisfies $$\label{eq:UxU} U_x\subset_\mu U\ .$$ Define $$\widetilde U_x= \bigl\{y\in Y\colon {{\mathcal D}}_d({{\bf 1}}_{W_x}\circ\Psi)(y)>0\bigr\}\ .$$ By Lemmas \[fact:continuousDd\] and \[fact:dualfactor\], $\widetilde U_x$ is an open subset of $Y$ and $$\widetilde U_x=_\nu \Psi{^{-1}}(U_x)\ .$$ We have that $U$ is the union of the open sets $W_x$ for $x\in U$. Since $U$ is $\sigma$-compact, there exists a countable subset $\Gamma$ of $U$ such that the union $\bigcup_{x\in\Gamma}U_x$ is equal to $U$. Define $$\widetilde U=\bigcup_{x\in\Gamma}\widetilde U_x\ .$$ Then $$\widetilde U=_\nu \Psi{^{-1}}\bigl(\bigcup_{x\in\Gamma}U_x\bigr)\ .$$ By , $\widetilde U\subset_\nu \Psi{^{-1}}(U)$. By Lemma \[fact:positive\], for every $x\in\Gamma$ we have that $ U_x\supset_\mu W_x$. Thus $\widetilde U_x\supset_\nu \Psi{^{-1}}(W_x)$ and $$\widetilde U\supset_\nu\bigcup_{x\in\Gamma}\Psi{^{-1}}(W_x)= \Psi{^{-1}}\bigl(\bigcup_{x\in\Gamma}W_x\bigr)=\Psi{^{-1}}(U)\ .$$ This completes the proof of the claim. \[fact:continuousalmost\] There exists an invariant subset $Y_1$ of full measure such that the restriction of $\Psi$ to $Y_1$ (endowed with the induced topology) is continuous. To prove this claim, we let $(U_j\colon j\geq 1)$ be a countable basis for the topology of $X$. For every $j\geq 1$, by Claim \[fact:openopen\] there exists an open subset $\widetilde U_j$ of $Y$ such that the symmetric difference $$Z_j:= \widetilde U_j \ \Delta \ \Psi{^{-1}}(U_j)$$ has zero $\nu$-measure. Define $$Y_1=Y_0\setminus \bigcup_{n\in{{\mathbb Z}}}\bigcup_{j\geq 1} T^{n}Z_j\ .$$ (Recall that $Y_0$ is the invariant subset of $Y$ where the map $\Phi$ is defined.) For every $j\geq 1$, $\Psi{^{-1}}(U_j)\cap Y_1=\widetilde U_j\cap Y_1$. Every nonempty open subset $U$ of $X$ is the union of some of the sets $U_j$, and if $\widetilde U$ is the union of the corresponding sets $\widetilde U_j$ we have that $\Psi{^{-1}}(U)\cap Y_1=\widetilde U\cap Y_1$. This proves the claim. We combine these results to complete the proof of Lemma \[lem:Phicontinuous\]. Since $(Y,T)$ is minimal, the measure $\nu$ has full support in $Y$ and the subset $Y_1$ given by Claim \[fact:continuousalmost\] is dense in $Y$. Since $(X,T)$ is distal, the result now follows from Lemma \[lem:extendcontinu\]. Using this, we return to the proposition: There exist a Borel invariant subset $Y_0$ of $Y$ of full measure, a Borel invariant subset $X_0$ of full measure, and a Borel bijection $\Psi\colon Y_0\to X_0$ with Borel inverse, mapping $\nu$ to $\mu$ and commuting with the transformations. Recall that $(X,T)$ is a system of order $d-1$ and that $(Y,\nu,S)$ satisfies property ${{\mathcal P}(d)}$. By Lemma \[lem:Phicontinuous\], there exist a subset $Y_1$ of $Y_0$ of full measure and a topological factor map $\Phi\colon Y\to X$ that coincides with $\Psi$ on $Y_1$. By Lemma \[fact:factoPd\], $(X,\mu,T)$ has property ${{\mathcal P}(d)}$. Recall that $(Y,T)$ is a system of order $d-1$. Using Lemma \[lem:Phicontinuous\] again, there exist a subset $X_1$ of $X_0$ of full measure and a topological factor map $\Theta\colon X\to Y$ that coincides with $\Psi{^{-1}}$ on $X_1$. The subset $Y_1\cap \Psi{^{-1}}(X_1)$ has full measure in $Y$ and for $y$ in this set, we have $\Theta\circ\Phi(y)=y$. Since the measure $\nu$ has full support, this equality holds everywhere and $\Theta\circ\Phi={\bf Id}_Y$. By the same argument, $\Phi\circ\Theta={\bf Id}_X$ and we are done. Rigidity properties of inverse limits of nilsystems {#sec:inutile} =================================================== In this Section, we assume that $d>1$ is an integer and establish some “rigidity” properties of inverse limits of $(d-1)$-step nilsystems, meaning some continuity properties. A property of nilsystems of this type (Theorem \[fact:factonil\]) was used in Section \[subsec:invelimit\] in the discussion on the definition of inverse limits, and so the reader may be concerned about a possible vicious circle in the argument. The way to avoid this is to first carry out the results in this section for nilsystems, and not inverse limits of nilsystems. This suffices to establish the property needed in Section \[subsec:invelimit\]. Then it is easy to check that the same proofs extend to the general case. Throughout the remainder of this section, we assume that $(X,T)$ is an inverse limit of minimal $(d-1)$-step nilsystems and that $\mu$ is the invariant measure of this system. We recall that $(X,T)$ is a system of order $d-1$ and has property ${{\mathcal P}(d)}$ of continuity of dual functions (Proposition \[fact:uniform\]). We first give a slight improvement of Lemma \[fact:Ddzero\], maintaining the same notation: \[fact:Ddzero2\] For every $x\in X$ and every neighborhood $U$ of $X$, there exists a neighborhood $V$ of $x$ such that if $f$ is a continuous function on $X$ whose support lies in $V$, then the support of the function ${{\mathcal D}}_df$ is contained in $U$. Pick $\eta>0$ such that the ball $B(x,2\eta)$ is contained in $U$. Let $\delta$ be as in Lemma \[fact:Ddzero\] and let $V=B(x,\delta)$. Assume that $f\in{{\mathcal C}}(X)$ has support contained in $V$ and assume that $|f|\leq 1$. We have that $|{{\mathcal D}}_df|\leq{{\mathcal D}}_d|f|\leq {{\mathcal D}}_d{{\bf 1}}_{B(x,\delta)}$. By the choice of $\delta$, ${{\mathcal D}}_df$ is equal to zero almost everywhere on the complement of $B(x,\eta)$. Since the function ${{\mathcal D}}_d f$ is continuous and since the measure $\mu$ has full support in $X$, ${{\mathcal D}}_df$ vanishes everywhere outside the closed ball $\bar B(x,\eta)$, which is included in $U$. \[fact:Dfnonzero\] If $f$ is a nonnegative continuous function on $X$, then ${{\mathcal D}}_df(x)>0$ for every $x\in X$ such that $f(x)>0$. It follows immediately from property ${{\mathcal P}(d)}$ that for every $x\in X$, there exists a probability measure $\mu{^{[d]}}_x$ on $X{^{[d]}}_*$ such that $$\frac 1{N^d}\sum_{0\leq n_1,\dots,n_d<N} \prod_{\substack{\epsilon\subset[d]\\ \epsilon\neq\emptyset}} f_\epsilon(T^{{\mathbf{n}}\cdot\epsilon }x) \to \int \prod_{\substack{\epsilon\subset[d]\\ \epsilon\neq\emptyset}} f_\epsilon(y_\epsilon)\,d\mu{^{[d]}}_x({\mathbf{y}}_*)$$ as $N\to+\infty$ for any continuous functions $f_\epsilon$, $\emptyset\neq\epsilon\subset[d]$, on $X$. By construction, the measure $\delta_x\times\mu{^{[d]}}_x$ is concentrated on the closed orbit ${{\bf K}}{^{[d]}}(x)$ of the point $x{^{[d]}}\in X{^{[d]}}$ under the group of face transformations ${{\mathcal F}}{^{[d]}}$, and is invariant under these transformations. Since $(X,T)$ is distal, the action of these transformations on ${{\bf K}}{^{[d]}}(x)$ is minimal and thus the topological support of the measure $\delta_x\times\mu{^{[d]}}_x$ is equal to ${{\bf K}}{^{[d]}}(x)$. Therefore, the point $x_*{^{[d]}}\in X{^{[d]}}_*$ belongs to the topological support of the measure $\mu{^{[d]}}_x$. If $f$ is a nonnegative continuous function on $X$ with $f(x)>0$ then, $${{\mathcal D}}_df(x)=\int \prod_{\substack{\epsilon\subset[d]\\ \epsilon\neq\emptyset}} f(y_\epsilon)\,d\mu{^{[d]}}_x({\mathbf{y}}_*)>0 \ ,$$ because the function in the integral is positive at the point $x_*{^{[d]}}$ which belongs to the support of the measure $\mu{^{[d]}}_x$. \[fact:Dddense\] The algebra of functions spanned by $\{{{\mathcal D}}_d f\colon f\in {{\mathcal C}}(X)\}$ is dense in ${{\mathcal C}}(X)$ under the uniform norm. By Lemmas \[fact:Ddzero2\] and \[fact:Dfnonzero\], for distinct $x,y\in X$, there exists a continuous function $f$ on $X$ with ${{\mathcal D}}_df(x)\neq{{\mathcal D}}_df(y)$. Recall that ${{\mathcal D}}_df$ is a continuous function on $X$. Noting that ${{\mathcal D}}_d1=1$, the statement follows from the Stone-Weierstrass Theorem. \[fact:factonil\] Let $p\colon (X,\mu,T)\to(X',\mu',T')$ be a measure theoretic factor map between inverse limits of $(d-1)$-step ergodic nilsystems. Then the factor map $p\colon X\to X'$ is equal almost everywhere to a topological factor map. Let ${{\mathcal A}}$ be a countable subset of ${{\mathcal C}}(X')$ that is dense under the uniform norm. By Lemmas \[fact:continuousDd\] and \[fact:Dddense\], $\{{{\mathcal D}}_df\colon f\in{{\mathcal A}}\}$ is included in ${{\mathcal C}}(X')$ and is dense in this algebra. By Lemma \[fact:dualfactor\], for every $f\in{{\mathcal A}}$ we have that ${{\mathcal D}}_df\circ p={{\mathcal D}}_d(f\circ p)$ almost everywhere. By Lemma \[fact:continuousDd\], ${{\mathcal D}}_d(f\circ p)$ is $\mu$-almost everywhere equal to a continuous function on $X$. Therefore, there exists $X_0\subset X$ of full measure such that for every $f\in{{\mathcal A}}$, the function $({{\mathcal D}}_df)\circ p$ coincides on $X_0$ with a continuous function on $X$. The same property holds for every function belonging to the algebra spanned by ${{\mathcal A}}$. Since $X_0$ is dense in $X$, by density the same property holds for every continuous function on $X$. This defines a homomorphism of algebras $\kappa\colon {{\mathcal C}}(X')\to{{\mathcal C}}(X)$ with $\kappa f(x)=f(p(x))$ for every $x\in X_0$ and every $f\in{{\mathcal C}}(X')$, and $\kappa$ commutes with the transformations $T$ and $T'$. Thus there exists a continuous map $p'\colon X\to X'$ such that $\kappa f=f\circ p'$ for all $f\in{{\mathcal C}}(X')$. \[fact:acionnil\] Let $(X,T,\mu)$ be an ergodic inverse limit of $(d-1)$-step nilsystems, $G$ be a Polish group, and $(g,x)\mapsto g\cdot x$ be a Borel action of $G$ on $X$ by measure preserving transformations commuting with $T$. There exists a continuous action $(g,x)\mapsto g*x$ of $G$ on $X$, commuting with $T$, such that for every $g\in G$, $g*x=g\cdot x$ for $\mu$-almost every $x\in X$. By hypothesis, the map $(g,x)\mapsto g\cdot x$ is Borel from $G\times X$ to $X$. The action of $G$ on $X$ we want must be such that the map $(g,x)\mapsto g*x$ is continuous from $G\times X$ to $X$. By Theorem \[fact:factonil\], for every $g\in G$ there exists a continuous map $x\mapsto g* x$, commuting with $T$ and preserving the measure $\mu$, such that $g* x=g\cdot x$ for $\mu$-almost every $x\in X$. For $g,h\in G$, we have that for $\mu$-almost every $x\in X$, $g* (h\cdot x)=gh* x$. By density, the same equality holds everywhere. Therefore, the map $(g,x)\mapsto g* x$ is an action of $G$ on $X$. We are left with showing that this map is jointly continuous. Let $f\in{{\mathcal C}}(X)$. For $g\in G$, write $f_g(x)=f(g* x)$. For each $g\in G$, the function $f_g$ is continuous and the map $x\mapsto g* x$ commutes with $T$. By Proposition \[fact:uniform\], ${{\mathcal D}}_df_g(x)={{\mathcal D}}_df(g* x)$ for every $x\in X$. For each $g\in G$, the functions $f_g$ and $x \mapsto g\cdot x$ are equal almost everywhere and represent the same element of $L^{2^d-1}(\mu)$. Since the action $(g,x)\mapsto g\cdot x$ of $G$ on $X$ is Borel and measure preserving, by [@BK] we have that the map $g\mapsto f_g$ is continuous from $G$ to $L^{2^d-1}(\mu)$. By Lemma \[fact:boundedDd\], the map $g\mapsto {{\mathcal D}}_df_g$ is continuous from $G$ to ${{\mathcal C}}(X)$, meaning that the function $(g,x)\mapsto{{\mathcal D}}_df_g(x)={{\mathcal D}}_df(g* x)$ is continuous on $G\times X$. By density (Lemma \[fact:Dddense\]), for every function $h\in{{\mathcal C}}(X)$, the function $(g,x)\mapsto h(g*x)$ is continuous on $G\times X$. We deduce that the map $(g,x)\mapsto g* x$ is continuous from $G\times X$ to $X$. [9999]{} J. Auslander. [*Minimal Flows and their Extensions*]{}. North-Holland Mathematics Studies [**153**]{} North-Holland Publishing Co., Amsterdam (1988). J. Auslander and E. Glasner. The distal order of a minimal flow. [*Israel J. Math.*]{} [**127**]{} (2002) 61–80. L. Auslander, L. Green and F. Hahn. Flows on homogeneous spaces. [*Ann. Math. Studies*]{} [**53**]{}, Princeton Univ. Press (1963). H. Becker and A. S.  Kechris. The descriptive theory of Polish group actions. [*London Math. Soc. Series*]{} 232, Cambridge Univ. Press (1996). V. Bergelson, B. Host and B. Kra, with an appendix by I.Z. Ruzsa. Multiple recurrence and nilsequences. [*Invent. Math.*]{}, [**160**]{} (2005) 261–303. R. Ellis. [*Lectures on topological dynamics*]{}. W. A. Benjamin, Inc., New York (1969). N. Frantzikinakis and M. Wierdl. A Hardy field extension of Szemerédi’s theorem. To appear, [*Adv. in Math.*]{} H. Furstenberg and B. Weiss. A mean ergodic theorem for $\frac{1}{N}\sum_{n=1}^N f(T^nx)g(T^{n^2}x)$. [*Convergence in Ergodic Theory and Probability*]{}, Eds: Bergelson, March, Rosenblatt, Walter de Gruyter & Co, Berlin, New York: 193–227, 1996. E. Glasner. Personal Communication. B. Green and T. Tao. An inverse theorem for the Gowers $U^3(G)$ norm. [*Proc. Edin. Math. Soc.*]{}, [**51**]{} (2008) 73–153. B. Green and T. Tao. Linear equations in primes. To appear [*Ann. of Math.*]{} B. Green and T. Tao. The Möbius function is strongly orthogonal to nilsequences. Preprint. B. Host and B. Kra. Convergence of Conze-Lesigne Averages. [*Erg. Th. & Dyn. Sys.*]{}, [**21**]{} (2001) 493–509. B. Host and B. Kra. Nonconventional averages and nilmanifolds. [*Ann. of Math.*]{}, [**161**]{} (2005) 398–488. B. Host and B. Kra. Parallelepipeds, nilpotent groups, and Gowers norms. [*Bull. Soc. Math. France*]{}, [**136**]{} (2008) 405–437. B. Host and B. Kra. Analysis of two step nilsequences. [*Ann. Inst. Fourier.*]{} [**58**]{} (2008) 1407–1453. B. Host and B. Kra. Uniformity norms on $\ell^\infty$ and applications. To appear, [*J. d’Analyse Mathématique*]{}. B. Host and A. Maass. Nilsystèmes d’ordre deux et parallélépipèdes. [*Bull. Soc. Math. France*]{}, [**135**]{} (2007) 367–405. A. Leibman. Pointwise convergence of ergodic averages for polynomial sequences of translations on a nilmanifold. [*Erg. Th. & Dyn. Sys.*]{} [**25**]{} (2005), no. 1, 201-213. W. Parry. Dynamical systems on nilmanifolds. [*Bull. London Math. Soc.*]{}, [**2**]{} (1970), 37–40. D.J. Rudolph. Eigenfunctions of $T\times S$ and the Conze-Lesigne algebra. [*Ergodic Theory and its Connections with Harmonic Analysis*]{}, Eds.: Petersen & Salama, Cambridge University Press, New York (1995), 369–432. [^1]: The first author was partially supported by the Institut Universitaire de France, the second author by NSF grant $0555250$, and the third author by the Millennium Nucleus Information and Randomness P04-069F, CMM-Fondap-Basal fund. This work was begun during the visit of the authors to MSRI and we thank the institute for its hospitality.
{ "pile_set_name": "ArXiv" }
--- abstract: '[*“The purpose of numerical models is not numbers but insight.”*]{} (Hamming) In the spirit of this adage, and of Don Cox’s approach to scientific speaking, we discuss the questions that the latest generation of numerical models of the interstellar medium raise, at least for us. The energy source for the interstellar turbulence is still under discussion. We review the argument for supernovae dominating in star forming regions. Magnetorotational instability has been suggested as a way of coupling disk shear to the turbulent flow. Models make evident that the unstable wavelengths are very long compared to thermally unstable wavelengths, with implications for star formation in the outer galaxy and low surface brightness disks. The perennial question of the factors determining the hot gas filling factor in a SN-driven medium remains open, in particular because of the unexpectedly strong turbulent mixing at the boundaries of hot cavities seen in the models. The formation of molecular clouds in the turbulent flow is also poorly understood. Dense regions suitable for cloud formation clearly form even in the absence of self-gravity, although their ultimate evolution remains to be computed.' author: - 'Mordecai-Mark Mac Low, Miguel A. de Avillez, & Maarit J. Korpi' --- Questions about Turbulence ========================== Numerical models often yield insight into the behavior of a physical system long before they can give quantitative results. In this contribution we review possible answers to three major questions about turbulence, relying on a combination of general energetic arguments and numerical models. The first question is, “What provides the energy to drive the turbulent flow?” Many sources have been proposed, but few have the required energy to counteract dissipation in the interstellar medium. Supernovae (SNe) seem likely to be the primary driver in parts of galaxies where star formation occurs, while the magnetorotational instability (MRI) may couple the gas to galactic rotational shear in other parts of galaxies. The second question is, “How does the driving shape the flow?” Most of the energy lies at the driving scale, so the large-scale structure is determined quite directly by the driving mechanism. Turbulent compression may be as important as thermodynamic phases in determining the pressure at any particular point in the ISM, as well as in determining the filling factor of the hot gas. The last question, of interest to understanding the rate of star formation from the ISM, is “How do molecular clouds form in this flow?” Turbulent compression and self-gravity both appear as possible mechanisms, but cannot yet be definitively distinguished. What Drives the Turbulence? =========================== Maintenance of observed motions appears to depend on continued driving of the turbulence, which has kinetic energy density $e = (1/2) \rho v_{\rm rms}^2$. Mac Low [@m99; @m02] estimates that the dissipation rate for isothermal, supersonic turbulence is $$\begin{aligned} \label{eqn:dissip} \dot{e} & \simeq & -(1/2)\rho v_{\rm rms}^3/L_{\rm d} \\ & = & -(3 \times 10^{-27} \,\mbox{erg}\,\mbox{cm}^{-3}\,\mbox{s}^{-1}) \left(\frac{n}{1\,\mbox{cm}^{-3}}\right) \left(\frac{v_{\rm rms}}{10\,\mbox{km}\,\mbox{s}^{-1}}\right)^3 \left(\frac{L_{\rm d}}{100 \,\mbox{pc}}\right)^{-1}, \nonumber\end{aligned}$$ where $L_{\rm d}$ is the driving scale, which we have somewhat arbitrarily taken to be 100$\,$pc (though it could well be smaller, or a broad range), and we have assumed a mean mass per particle $\mu= 2.11\times 10^{-24}$ g. The dissipation time for turbulent kinetic energy $\tau_{\rm d} = e / \dot{e} \simeq L_{\rm d}/v_{\rm rms},$ which is just the crossing time for the turbulent flow across the driving scale [@e00]. What then is the energy source for this driving? We here review the energy input rates for the most plausible mechanisms, feedback from massive stars, particularly SNe, and magnetorotational instabilities. A more extensive discussion covering a number of other possibilities as well is given by Mac Low & Klessen [@mk04]. An energy source for interstellar turbulence that has long been considered is shear from galactic rotation [@f81]. However, the question of how to couple from the large scales of galactic rotation to smaller scales remained open. Sellwood & Balbus [@sb99] suggested that the MRI [@bh91; @bh98] could couple the large and small scales efficiently. The instability generates Maxwell stresses (a positive correlation between radial $B_R$ and azimuthal $B_{\Phi}$ magnetic field components) that transfer energy from shear into turbulent motions at a rate $$\label{eqn:stress} \dot{e} = - T_{R\Phi} (d\Omega / d \ln R) = T_{R\Phi} \Omega,$$ where the last equality holds for a flat rotation curve [@sb99]. Numerical models suggest that the Maxwell stress tensor $T_{R\Phi} \simeq 0.6 B^2/(8\pi)$ [@h96]. For the Milky Way, the standard value of the rotation rate $\Omega = (220 \mbox{ Myr})^{-1} = 1.4 \times 10^{-16} \mbox{ rad s}^{-1}$, so the MRI can contribute energy at a rate $$\dot{e} = (3 \times 10^{-29}\,\mbox{erg}\,\mbox{cm}^{-3}\,\mbox{s}^{-1}) \left(\frac{B}{3 \mu\mbox{G}}\right)^2 \left(\frac{\Omega}{(220\,\mbox{Myr})^{-1}}\right).$$ For parameters appropriate to the H[i]{} disk of a sample small galaxy, NGC 1058, including $\rho = 10^{-24}\,$g$\,$cm$^{-3}$, Sellwood & Balbus [@sb99] find that the magnetic field required to produce the observed velocity dispersion of 6 km s$^{-1}$ is roughly 3 $\mu$G, a reasonable value for such a galaxy. This instability may provide a base value for the velocity dispersion below which no galaxy will fall. If that is sufficient to prevent collapse, little or no star formation will occur, producing something like a low surface brightness galaxy with large amounts of H[i]{} and few stars. This may also apply to the outer disk of our own Milky Way and other star-forming galaxies. In active star-forming galaxies, massive stars probably dominate the driving at radii where they form. They could do so through ionizing radiation and stellar winds from O stars, or clustered and field SN explosions, predominantly from B stars no longer associated with their parent gas. Mac Low & Klessen [@mk04] demonstrate that ionizing radiation is unlikely to dominate the kinetic energy budget, despite the large amount of energy going into heating and ionization. The total energy input from the line-driven stellar wind over the main-sequence lifetime of an early O star can equal the energy from its SN explosion, and the Wolf-Rayet wind can be even more powerful. However, the mass-loss rate from stellar winds drops as roughly the sixth power of the star’s luminosity, while the powerful Wolf-Rayet winds [@nl00] last only $10^5$ years or so, so only the very most massive stars contribute substantial energy from stellar winds. The energy from SN explosions, on the other hand, remains nearly constant down to the least massive star that can explode. As there are far more lower-mass stars than massive stars, SN explosions inevitably dominate over stellar winds after the first few million years of the lifetime of an OB association. To estimate the energy input rate from SNe, we begin with a SN rate for the Milky Way of (50 yr)$^{-1}$, which agrees well with the estimate in equation (A4) of McKee [@m89]). If we take the scale height of SNe $H_c \simeq 100$ pc and a star-forming radius $R_{sf} \simeq 15$ kpc, we can compute the energy input rate from SN explosions with energy $E_{SN} = 10^{51}\,$erg to be $$\begin{aligned} \dot{e} & = &\frac{\sigma_{SN} \eta_{SN} E_{SN}}{\pi R_{sf}^2 H_c} \\ & = & (3 \times 10^{-26} \mbox{ erg s$^{-1}$ cm}^{-3}) \left(\frac{\eta_{SN}}{0.1} \right) \left(\frac{\sigma_{SN}}{1 \mbox{ SNu}} \right) \left(\frac{H_c}{100 \mbox{ pc}} \right)^{-1} \times \nonumber \\ & \times & \left(\frac{R_{sf}}{15 \mbox{ kpc}} \right)^{-2} \left(\frac{E_{SN}}{10^{51} \mbox{ erg}} \right). \nonumber\end{aligned}$$ The efficiency of energy transfer from SN blast waves to the interstellar gas $\eta_{SN}$ depends on the strength of radiative cooling in the initial shock, which will be much stronger in the absence of a surrounding superbubble (e.g. [@h90]). Substantial amounts of energy can escape in the vertical direction in superbubbles as well, however. The scaling factor $\eta_{SN} \simeq 0.1$ used here was derived by Thornton et al. [@t98] from detailed, 1D, numerical simulations of SNe expanding in a uniform ISM. It can alternatively be drawn from momentum conservation arguments, comparing a typical expansion velocity of 100$\,$km$\,$s$^{-1}$ to typical interstellar turbulence velocity of 10$\,$km$\,$s$^{-1}$. Multi-dimensional models of the interactions of multiple SN remnants (e.g. [@a00]) are required to better determine the effective scaling factor. SN driving appears to be powerful enough to maintain the turbulence even with the dissipation rates estimated in Eq. (\[eqn:dissip\]). It provides a large-scale self-regulation mechanism for star formation in disks with sufficient gas density to collapse despite the velocity dispersion produced by the MRI. As star formation increases in such galaxies, the number of OB stars increases, ultimately increasing the SN rate and thus the velocity dispersion, which restrains further star formation. How Does Driving Shape the ISM? =============================== We now turn to the question of how these different driving mechanisms determine the structure of the ISM. Clearly, different mechanisms yield different results. To study the MRI, we used a parallel MHD code integrating $\ln \rho$ rather than density $\rho$ to handle strong density contrasts [@ck01], with shearing sheet horizontal boundary conditions implemented. The preliminary models shown here were run at $64 \times 64 \times 128$ zones on an $0.5 \times 0.5 \times 1$ kpc grid, with the ISM in vertical hydrostatic equilibrium with scale height $H = 250$ pc initially, and an initially vertical magnetic field with thermal to magnetic pressure ratio $\beta = 1000$. The initial wavelength of maximum instability was then 80 pc. Runs were extended to 10 orbits, or 2.5 Gyr. Radiative cooling was included based on an equilibrium ionization cooling curve including thermal instability below $10^4$ K, and heating proportional to density was chosen to exactly balance the cooling in the initial model. We ran models initially in thermally stable and unstable regimes. ![\[mri\] Log of density on vertical cuts through 3D shearing sheet models of MRI at times given. [*Top:*]{} thermally stable models, including a horizontal cut through the midplane of the final model; [*Bottom:*]{} thermally unstable models, showing the action of the MRI on the cooled clumps.](korpi-mri.ps){width="\textwidth"} In Figure \[mri\] we show the development of the MRI in these regimes. In the thermally stable regime, factor of 2–3 column density contrasts through the disk are created by the instability. In the thermally unstable regime, the thermal instability acts quickly to clump the gas, but after multiple orbits the MRI adds sufficient velocity dispersion to heat the gas and distribute it more uniformly. Rather more substantial column density contrasts still occur. Comparison with observed H[i]{} disks outside of the star-forming region should be revealing of whether this mechanism is in fact maintaining their velocity dispersion. Numerical models of the SN-driven ISM suggest that the hot gas filling factor $f$ is closer to the value $f \sim 0.2$ [@a00; @a04; @ab03] predicted by Slavin & Cox [@sc93] than to the values close to unity predicted by McKee & Ostriker [@mo77]. Why is this? McKee & Ostriker [@mo77] assumed a two-phase medium with cold, dense clouds embedded in a uniform density, warm, intercloud medium. Hot SN remnants then expanded into this medium. Was the cooling within the SN remnants underestimated because turbulent mixing was approximated with mass loading from the clumps overrun by the remnants, or was the effective external density underestimated by the two-phase model? ![\[sne\] Cuts through the midplane of the SN-driven model run at finest resolution of 1.25 pc. [*Left:*]{} Density; [*Right:*]{} Pressure.](av-den-prs.ps){width="50.00000%"} To study the SN-driven ISM we used an adaptive mesh refinement code described by Avillez & Mac Low [@am02], with a $1 \times 1 \times 20$ kpc grid set up as described in Avillez [@a00], with a SN rate equal to the galactic value. Figure \[sne\] shows that the pressures vary widely [@m04], so that no simple two-phase medium can actually form. Instead, the densities cover a broad range continuously, as shown in Figures \[sne\] and \[rho-pdf\]. ![\[rho-pdf\] Probability distribution function of density in the midplane of the SN-driven model. The two-phase medium assumed by McKee & Ostriker [@mo77] is shown schematically by the thick black rectangles. ](av-den-pdf.ps){width="60.00000%"} This continuous distribution of density may act to impede the expansion of SN remnants more effectively than the warm intercloud medium with cold embedded clouds shown schematically in Figure \[rho-pdf\]. On the other hand, Avillez & Mac Low [@am02] demonstrated using a tracer field that mixing occurs quite efficiently in the hot regions. In Figure \[sne\] widespread turbulent mixing at the edges of shells and supershells can be seen. This could substantially enhance the density in the hot interiors, thus enhancing the radiative cooling, which is proportional to $\rho^2$. However, a quantitative test of how well or poorly this turbulent mixing was modeled by the model of SN remnants overrunning conductively evaporating clouds used by McKee & Ostriker [@mo77] remains to be done. How Do Molecular Clouds Form in the Turbulent ISM? ================================================== Molecular clouds are high-density objects, with much of their mass at densities of $10^{3-5}$ cm$^{-3}$. With typical temperatures of order 10 K, their pressures are an order of magnitude or more above the average ISM pressure. It has usually been argued that these high pressures must be caused by self-gravity, since they would otherwise explode. However, turbulent ram pressure in a SN-driven ISM produces high-density, high-pressure regions even in the absence of self-gravity, as shown in Figure \[press\]. These may provide the sites for the formation of at least some molecular clouds, especially ones that do not show vigorous, efficient, star formation. ![\[press\] [*Left:*]{} Scatter plot of pressure vs.density in the midplane of the SN-driven model showing occupation of high-pressure, high-density region associated with molecular clouds. [*Right:*]{} Volume-weighted probability distribution function of pressure in the same model. Note that the small volumes occupied by high-density, cold gas have large mass. ](av-prs-pdf-scatter.ps){width="50.00000%"} M-MML is partly supported by NSF grants AST 99-85392 and AST 03-07854. This work made use of the NASA ADS Abstract Service. Avillez, M. A., 2000, MNRAS, 315, 479 Avillez, M. A., 2004, these proceedings Avillez, M. A., and D. Breitschwerdt, 2003, in [*Star Formation Through Time*]{}, edited by E. Pérez, R. M. González Delgado, and G. Tenorio-Tagle (Astronomical Society of the Pacific: San Francisco), in press (astro-ph/0303322) Avillez, M. A., and M.-M. Mac Low, 2002, ApJ, 581, 1047 Balbus, S. A., and J. F. Hawley, 1991, ApJ, 376, 214 Balbus, S. A., and J. F. Hawley, 1998, Rev. Mod.Phys., 70, 1 Caunt, S. E., & Korpi, M. J. 2001, A&A, 369, 706 Elmegreen, B. G., 2000, ApJ, 530, 277 Fleck, R. C., 1981, ApJ Lett, 246, L151 Heiles, C., 1990, ApJ, 354, 483 Hawley, J. F., C. F. Gammie, and S. A. Balbus, 1996, /apj, 464, 690 Mac Low, M.-M., 1999, ApJ, 524, 169 Mac Low, M.-M., 2002, in [*Turbulence and Magnetic Fields in Astrophysics*]{}, edited by E. Falgarone & T. Passot (Springer, Heidelberg), 182 Mac Low, M.-M., Balsara, D. S., Avillez, M. A., & Kim, J. 2004, ApJ, in revision (astro-ph/0106509) Mac Low, M.-M., & Klessen, R. S. 2004, Rev. Mod. Phys., in press (astro-ph/0301093) Matzner, C. D. 2002, ApJ, 566, 302 McKee, C. F., 1989, ApJ, 345, 782 McKee, C. F., and J. P. Ostriker, 1977, ApJ, 218, 148 Nugis, T., and H. J. G. L. M. Lamers, 2000, A&A, 360, 227 Sellwood, J. A., and S. A. Balbus, 1999, ApJ, 511, 660 Slavin, J. D., & Cox, D. P. 1993, ApJ, 417, 187 Thornton, K., M. Gaudlitz, H.-Th. Janka, and M. Steinmetz, 1998, ApJ, 500, 95 Discussion ========== [*Gaensler:*]{} We know from observations of scattering & scintillation that there is turbulence in the warm ionized phase of the ISM. Since we expect expanding SN remnants to sweep up neutral shells, can SN remnants still drive the turbulence seen in ionized gas? What sort of a contribution do ionization fronts make to turbulence in ionized gas? [*Mac Low:*]{} In our models, most of the turbulence comes from the interaction of multiple shells. Diffuse ionizing radiation will ionize some of that gas, producing diffuse, turbulent, ionized gas. H [ii]{} regions also contribute, of course. In Mac Low & Klessen [@mk04] we use results from Matzner [@cdm02] to argue that H [ii]{} region expansion is only a minor (&lt; 1%) contributor to ISM kinetic energy.\ [*Heiles:*]{} You emphasized the breadth of the pressure distribution. But it’s really not more than an order of magnitude, right? [*Mac Low:*]{} That is true for the volume-weighted FWHM. However, a mass-weighted view shows that a substantial fraction of the mass is at the high-density end.
{ "pile_set_name": "ArXiv" }
--- abstract: 'The representation of quantum states via phase-space functions constitutes an intuitive technique to characterize light. However, the reconstruction of such distributions is challenging as it demands specific types of detectors and detailed models thereof to account for their particular properties and imperfections. To overcome these obstacles, we derive and implement a measurement scheme that enables a reconstruction of phase-space distributions for arbitrary states whose functionality does not depend on the knowledge of the detectors, thus defining the notion of detector-agnostic phase-space distributions. Our theory presents a generalization of well-known phase-space quasiprobability distributions, such as the Wigner function. We implement our measurement protocol, using state-of-the-art transition-edge sensors without performing a detector characterization. Based on our approach, we reveal the characteristic features of heralded single- and two-photon states in phase space and certify their nonclassicality with high statistical significance.' author: - 'J. Sperling' - 'D. S. Phillips' - 'J. F. F Bulmer' - 'G. S. Thekkadath' - 'A. Eckstein' - 'T. A. W. Wolterink' - 'J. Lugani' - 'S. W. Nam' - 'A. Lita' - 'T. Gerrits' - 'W. Vogel' - 'G. S. Agarwal' - 'C. Silberhorn' - 'I. A. Walmsley' title: 'Detector-Agnostic Phase-Space Distributions' --- #### Introduction.| {#introduction. .unnumbered} The characterization of quantum light is a main challenge one encounters when implementing classically infeasible tasks, such as quantum communication protocols [@KMSUZ16; @BFV09; @RL09]. On a more fundamental level, studying the peculiarities of quantized radiation fields leads to a profound understanding of the role of quantum physics in nature in general, and how it is distinct from classical wave theories in particular. As in classical systems, quantum-optical phase-space distributions offer a versatile instrument to directly visualize unique features of nonclassical light, such as demonstrated for squeezing [@W83; @SBRF93; @BSM97]. Moreover, negativities in certain phase-space functions directly point at quantum properties of light; see, e.g., Refs. [@LHABMS01; @KVPZB08; @LKCGS10; @DEWKDKCM13; @HSRHMSS16; @BFS17]. For the above reasons, the representation of quantum light in phase space is one of the most frequently applied methods to characterize nonclassical light. However, the estimation of phase-space distributions from experimental data is a cumbersome task. Consequently, this reconstruction problem inspired a wide range of research [@WVO99; @S07; @LR09], leading to sophisticated analytical tools, such as solving inversion problems [@T97; @SSG09], employing diverging pattern functions [@R96; @LMKRR96], performing maximum-likelihood estimations [@H97; @L04; @KWR04], and using data pattern recognition [@RMH10; @MIMSRH13]. In addition, each family of detection devices has to be equipped with its own precise model to reliably extract information about phase-space functions [@WVO99; @S07; @LR09]. This treatment comprises a comprehensive analysis that assesses (i) how a detector responds to incident light [@KK64; @FW91], including, e.g., nonlinear detection responses [@JA69; @AML11], and (ii) how the light absorption is influenced by a number of possible imperfections, e.g., efficiencies [@PM07; @WCLMSPTW09]. Moreover, applying these methods can also require universally applicable, yet rather demanding theoretical and experimental techniques in practice, such as performing detector tomography and calibration [@K80; @LS99; @F01; @AMP04; @FLCEPW09; @CLPFSMSEPW09; @RFZMGLDFE12; @PHMH12; @BKSSV17]. Despite these challenges, phase-space distributions constitute a highly successful approach to revealing nonclassical properties of light [@LHABMS01; @KVPZB08; @LKCGS10; @DEWKDKCM13; @HSRHMSS16; @BFS17]. For example, $s$-parametrized quasiprobabilities [@CG69; @AW70], as well as their non-Gaussian generalizations [@K66; @KV10], can exhibit negativities that are incompatible with classical light. Even if a phase-space function does not exhibit negativities, observable patterns render it possible to identify quantum features, for instance, via the nonnegative Husimi function [@H40; @LGC18; @MS04] or through marginal distributions [@A93; @PLLSZZZKN17]. Because of its success, the concept of phase-space functions has been further extended to other physical scenarios; see Refs. [@SW18; @SV19]. To name a few, atomic ensembles [@A81; @DAS94; @MZHCV15; @LMKMIW96] and entanglement [@STV12; @SV09; @SMBBS18] have been successfully characterized using quasiprobability distributions. Nevertheless, there remains a dependency on well-defined detection schemes and reconstruction algorithms. In this contribution, we circumvent the reconstruction problem by devising a measurement protocol that results in detector-agnostic phase-space (DAPS) distributions, which can be directly estimated, encompass known quasiprobabilities, and apply to arbitrary quantum states of light. We demonstrate our scheme with transition-edge sensors (TESs), which have sophisticated physics underlying their operation, and analyze our data without relying on any specific detector models. Our DAPS functions reveal nonclassical features expected from our heralded multiphoton states with high statistical significance. Moreover, the measurement of vacuum alone enables us to predict the unique structures of DAPS distributions as demonstrated for our experimentally generated states. ![image](Overview.pdf){width="\textwidth"} #### Theory framework.| {#theory-framework. .unnumbered} Our measurement scheme is a combination of unbalanced homodyning [@WV96] and a multiplexed detection layout [@PTKJ96]; see Fig. \[fig:Layout\]. A signal light field, $\hat\rho$, is mixed with a local oscillator (LO), $|\beta\rangle$, on a beam splitter. One of the output states, represented through $\hat\rho(\beta)$, is injected into a multiplexing scheme that consists of $S$ steps. In each step, light is split into output fields with the same intensity, which then can be split again. The finally obtained $N=2^S$ output beams are individually measured with unknown detectors, which are not specified but assumed to operate in the same manner. Each detector returns one of the possible outcomes $\mathcal K=\{0,\ldots,K\}$. In Refs. [@SCEMRKNLGVAW17; @SECMRKNLGWAV17], we have shown for the multiplexing part that independently of the detector response, the probability to simultaneously measure $N_k$ times the outcome $k$ ($\forall k\in\mathcal K$) follows a quantum version of the multinomial distribution; its generalization to $\hat\rho(\beta)$ reads $$\begin{aligned} \label{eq:ClickStatistics} c_{N_0,\ldots,N_K}(\beta)=&\left\langle{:} \frac{N!}{N_0!\cdots N_K!}\hat\pi_0^{N_0}\cdots\hat\pi_{K}^{N_K} {:}\right\rangle_{\hat\rho(\beta)}, \end{aligned}$$ where ${:}\cdots{:}$ denotes the normal ordering and $\{\hat\pi\}_{k\in\mathcal K}$ is the unknown positive operator-valued measure of the detectors. The only assumptions made are a balanced splitting in the multiplexing and identical response functions for the $N$ detectors, including all imperfections. We can account for deviations from both assumptions by including a systematic error, directly estimated from asymmetries in the measured data; see the Supplemental Material (SM) for details [@supplement]. A probability distribution is entirely characterized through its generation function, which can be expressed as $$\begin{aligned} \label{eq:GeneratinFunctions} \begin{aligned} g_{z_0,\ldots,z_K}(\beta)=&\sum_{N_0,\ldots,N_K}z_0^{N_0}\cdots z_K^{N_K}c_{N_0,\ldots,N_K}(\beta) \\ =&\left\langle{:} \left(z_0\hat\pi_0+\cdots+z_K\hat\pi_K\right)^N {:}\right\rangle_{\hat\rho(\beta)}, \end{aligned} \end{aligned}$$ for $z_0,\ldots,z_K\in\mathbb R$. The second line is a result of the multinomial form of the statistics in Eq. . One salient feature is that classical light fields have a nonnegative generation function $g_{z_0,\ldots,z_K}$. To see this, first recall that a classical light field is described through a nonnegative Glauber-Sudarshan distribution [@G63; @S63], which is not affected by displacements and describes a state as a statistical mixture of coherent states. Furthermore, for all even $N$, we can define the operator $\hat f=\hat f^\dag=\left(z_0\hat\pi_0+\cdots+z_K\hat\pi_K\right)^{N/2}$. Since for any nonnegative Glauber-Sudarshan function $\langle {:}\hat f^\dag\hat f{:}\rangle\geq0$ holds true [@TG65; @M86; @VW06; @A12], we conclude $$\begin{aligned} \label{eq:Negativity} g_{z_0,\ldots,z_K}(\beta)\stackrel{\mathrm{cl.}}{\geq}0. \end{aligned}$$ A violation of this inequality certifies the nonclassicality of the signal light, $\hat \rho$. We can also define a special case of this generating function, $$\begin{aligned} \label{eq:GeneratinFunctionSpecial} G_z(\beta)=&g_{1,z,z^2,\ldots,z^K}(\beta). \end{aligned}$$ Similarly to the expression in Eq. , $G_z$ is straightforwardly estimated from the measured detector outcomes $c_{N_0,\ldots,N_K}(\beta)$ by setting $z_k=z^k$, and $G_z$ is nonnegative for classical light. As an example, we may consider photocounting [@KK64]. Although this model is not required for our approach and does not apply to our experiment (TESs have a finite photon-number resolution, a non-unit detection efficiency, and a nonlinear response function [@SECMRKNLGWAV17]), it demonstrates how $G_z$ generalizes the concept of well-known phase-space distributions. For photocounting, we find [@supplement; @WM64] $$\begin{aligned} \label{eq:Pfct} G_z(\beta)=&\langle{:}e^{ -[1-z]\eta\hat n }{:}\rangle_{\hat\rho(\beta)} =\frac{\pi(1-s)}{2}P\left(\frac{r}{t}\beta;s\right), \end{aligned}$$ with $\eta$ and $\hat n$ being the efficiency and the photon-number operator, respectively, and $s=1-2/[\eta|t|^2(1-z)]$ [@commentsParameter]. Thus, $G_z(\beta)$ resembles the $s$-parametrized distributions $P(r\beta/t;s)$ [@CG69; @AW70]. Beyond photoelectric detectors, we refer to $g_{z_0,\ldots,z_K}$ and $G_z$ as DAPS distributions as Eqs. and apply without any knowledge of the measurement operators $\{\hat\pi_k\}_{k\in\mathcal K}$. In this context, it is worth emphasizing that the first line in Eq. enables the estimation of our DAPS distributions as a result of the measured coincidence statistics $c_{N_0,\ldots,N_K}(\beta)$ alone. #### Implementation.| {#implementation. .unnumbered} By implementing a single multiplexing step, $N = 2^S= 2$ for $S=1$, we demonstrate how to apply our theoretical framework of DAPS distributions. To realize our protocol in Fig. \[fig:Layout\], we produce heralded photon states $\hat\rho$ and different LO amplitudes $\beta$. The detectors used for the multiplexing measurement and the heralding are TESs, which count photons up to a maximal number $K$. In the following, we describe the experimental setup [@supplement]. Femtosecond pulses with a $100\,\textrm{kHz}$ repetition rate from a titanium sapphire laser are coupled into two separate, periodically poled potassium titanyl phosphate (ppKTP) waveguides. Both pulses are filtered to a full-width at half-maximum of $\pm2\,\textrm{nm}$ using angle-tuned bandpass filters. With the first ppKTP waveguide, prepare the signal $\hat\rho$. When filtering the pump at $775\,\textrm{nm}$ the waveguide produces two-mode squeezed vacuum in approximately a single spatio-temporal mode via type-II parametric down conversion (PDC) [@ECMS11]. The signal mode at $1554\,\textrm{nm}$ and the herald mode at $1547\,\textrm{nm}$ are separated with a polarizing beam splitter, then filtered and coupled into single-mode optical fibers. The herald mode is then sent to a single TES detector. With the second ppKTP waveguide, we prepare the LO. In contrast to the signal state generation, we filter the pump at $783\,\textrm{nm}$ and stimulate the PDC process by seeding it with $2\,\mathrm{ns}$ pulses carved with an electro-optic modulator from a $1580\,\textrm{nm}$ continuous-wave laser. Because of the strong seed signal, this nonlinear mixing generates coherent light to an excellent approximation in the polarization mode orthogonal to the seed [@LS13]. We separate the LO from the seed with a polarizing beam splitter, then pass through a bandpass filter at $1554\,\textrm{nm}$. By pumping the two waveguides at different wavelengths, we are able to create an LO that is well mode-matched to the signal using a seed laser that is detuned from the heralding mode. This avoids a potential source of noise due to the seed laser passing through the filters for the heralding TES. The generated LO is attenuated to the single-photon level and coupled into single-mode optical fiber. Crucially, this process prepares an LO with Poissonian photon statistics with a measured second-order correlation function $g^{(2)}(0)$ of $1.005\pm0.002$. Finally, the LO $|\beta\rangle$ and signal $\hat\rho$ are combined on a $90:10$ fiber beam splitter. We consider the port that uses $|r|^2=10\%$ of the LO and transmits $|t|^2=90\%$ of the signal. The light from this port, $\hat\rho(\beta)$, is then impinged on a $50:50$ fiber beam splitter for realizing a multiplexing step; both outputs are then sent to two separate TESs. See Fig. \[fig:Layout\]. Our experiment uses three TES detectors that can have efficiencies above $\eta=90\%$ [@LMN08]. TESs are superconducting photon-number-resolving detectors that we operate in a dilution refrigerator at a temperature of around $80\,\mathrm{mK}$. Their response is amplified using an array of superconducting quantum interference devices [@WM91], followed by further amplification and filtering at room temperature. This electrical signal is read by an analogue-to-digital converter and processed using a matched filter technique [@FCMPSW00], which outputs a single value when triggered by a clock signal from the laser. We bin these values to assign a photon number. It should be noted that it is possible to extract slightly more accurate estimates of photon number, however, using more sophisticated signal processing techniques, yet without affecting the applicability of the DAPS distribution approach [@SECMRKNLGWAV17; @HMGHLNNDKW15]. We record the binned outcome at all three TESs for various LO amplitudes ($|\beta|^2$ from $0$ to ${\sim}28$ in steps of ${\sim}1$). The amplitude is controlled by varying the seed laser power. To obtain data for a specific heralded state $\hat\rho$, we consider the subset of trials with the appropriate detection outcome (i.e., heralding bin $k_h$) at the herald TES. #### Verification of nonclassicality.| {#verification-of-nonclassicality. .unnumbered} In a first step, we apply our DAPS distribution to uncover nonclassical features of our prepared states through the violation of condition . The optimal negativity we obtain from the DAPS function \[Eq. \] is given by the minimum $$\begin{aligned} \label{eq:MinimalPhaseSpace} g_{\min}=\min_{\beta}\min_{\substack{z_0,\ldots,z_K:\\|z_0|^2+\cdots+|z_K|^2\leq1}} g_{z_0,\ldots,z_K}(\beta). \end{aligned}$$ To assess the quality of this approach, we compared our verification of nonclassicality with other methods. In Ref. [@SCEMRKNLGVAW17], we demonstrated that a correlation matrix, $M$, obtained from the measured statistics in Eq. , is positive semidefinite for classical light, described through a nonnegative minimal eigenvalue $\mu_{\min}$ of $M$. The resulting notion of sub-multinomial light, $\mu_{\min}<0$, was shown to be a better figure of merit than other means of verifying nonclassicality [@SECMRKNLGWAV17], such as sub-Poisson light [@M79; @SM83] and sub-binomial light [@SVA12; @BDJDBW13]. The comparison of $g_{\min}$ and $\mu_{\min}$ for our data is shown in Table \[Tab:SomeFigures\] for different heralding bins $k_h$. For the heralded one-photon (two-photon) states, we confirm $g_{\min}<0$ with $9$ ($6$) standard deviations, while the sub-multinomial behavior is less significant, $5$ ($3$) standard deviations. For the vacuum state, i.e., $k_h=0$, both measures are consistent with the classical expectation, $g_{\min}=0=\mu_{\min}$. $k_h$ $\mu_{\min}$ $g_{\min}$ ------- ---------------------------- ---------------------------- 0 $(-0 \pm 9)\times 10^{-4}$ $(-0 \pm 2)\times 10^{-9}$ 1 $-0.15 \pm 0.03$ $-0.026 \pm 0.003$ 2 $-0.10 \pm 0.03$ $-0.017 \pm 0.003$ : For different heralding outcomes, $k_h$, we show the nonclassicality criteria $\mu_{\min}<0$ and $g_{\min}<0$. $g_{\min}$ is defined in Eq. . $\mu_\mathrm{min}$ is the minimal eigenvalue to the second-order correlation matrix $M$ defined in Eq. (6) of Ref. [@SCEMRKNLGVAW17]; see also the SM [@supplement]. “$-0$” indicates a slightly negative mean value that rounds to zero. []{data-label="Tab:SomeFigures"} ![image](NewFig.pdf){width="\textwidth"} #### Reconstructed distributions.| {#reconstructed-distributions. .unnumbered} From the data, we can directly estimate our DAPS distributions. The results of our extended analysis are shown in Fig. \[fig:NewFig\]. The estimation procedure is this: we run the experiment twice, once with the signal blocked and once with the signal unblocked. To have full detector-agnostic approach, we first define a detector-independent coherent amplitude, $$\begin{aligned} \label{eq:DefinitionAmplitude} |\beta^\mathrm{(DI)}|=\sqrt{\sum_{N_0,\ldots,N_K}[0N_0+\cdots +KN_K]c^\mathrm{(vac)}_{N_0,\ldots,N_K}(\beta)}, \end{aligned}$$ which is given by the statistics $c^\mathrm{(vac)}_{N_0,\ldots,N_K}(\beta)$ measured by blocking the signal [@supplement]. In case of photocounting, this gives $|\beta^\mathrm{(DI)}|=\sqrt\eta|r||\beta|$. As we do not record a phase, we consider full phase randomization. This does not affect the DAPS distributions of our heralded photon states. In Fig. \[fig:NewFig\], our DAPS distributions $G_z$ are shown as a function of the amplitude in Eq. , determined by means of the vacuum measurement. The same measurement renders it possible to theoretically predict the DAPS distribution of arbitrary states. Namely, a general DAPS distribution can be described as a convolution of the measured vacuum distribution $G_z^\mathrm{(vac)}$ and the Glauber-Sudarshan distribution $P(\beta';1)$ of the state under study [@supplement], $$\begin{aligned} \label{eq:VaccumConvolution} G_z(\beta)=\int d^2\beta'\, P(\beta';1) G^\mathrm{(vac)}_{z}\left(\beta-\frac{t}{r}\beta'\right). \end{aligned}$$ In our case, the Gaussian shape of $G_z^\mathrm{(vac)}$ implies that heralded single-photon (two-photon) states should follow a Gaussian distribution multiplied with a first-order (second-order) polynomial in $|\beta^\mathrm{(DI)}|^2$. In Fig. \[fig:NewFig\], this prediction (dashed lines) is confirmed as it correctly represents the DAPS distributions of the measured heralded photon states. The heralding to $k_h=1$ gives a characteristic dip at the origin $|\beta^\mathrm{(DI)}|=0$, and the two-photon case, $k_h=2$, leads to additional oscillations together with the appearance of a peak at the origin. We emphasize that the functional behavior $\beta\mapsto G_z(\beta)$ depends on the measurement operators, but the estimation of $G_z(\beta)$ is done without any specification of the detector operators, according to the first line in Eq. . Moreover, we are able to characterize defining features of other states without any other prior knowledge about the detectors from the data obtained using the vacuum state input \[Eq. \]. Based on our reconstruction, we were able to determine a number of other properties of the experimentally produced states [@supplement]. For instance, we can determine how well the DAPS distributions enable us to perform a quantum state discrimination task. The single- and two-photon states \[plots (b) and (c) in Fig. \[fig:NewFig\]\] can be distinguished from each other with more than $98\%$ certainty. Furthermore, we found that, for $z<-2.4$, the central dip of $G_z$ becomes negative, similar to the behavior of other phase-space quasiprobabilities. The negativity has the highest statistical significance for $z=-4.85$, where $G_{z}(0)=-0.51\pm0.08$ is more than 6 standard deviations below the classical threshold of zero. #### Summary and discussion.| {#summary-and-discussion. .unnumbered} We have developed a theory and realized an experiment to characterize quantum light by means of phase space that functions for any type of detector and without performing a prior detector characterization. Our framework is based on the generating function derived from the properties of a balanced linear optical network, enabling our DAPS distribution, to be directly estimated from measured correlations and that applies to arbitrary states. To demonstrate this concept, we showed that a single multiplexing step is already sufficient for applying our method. This renders it possible to verify the nonclassicality of multiphoton states based on DAPS distributions, which results in greater statistical significance than obtained with earlier approaches [@SCEMRKNLGVAW17; @SECMRKNLGWAV17], which themselves already outperformed previous quantifiers of nonclassicality. Moreover, our approach encompasses prominent phase-space quasiprobabilities and straightforwardly generalizes to multimode light. Our general theory also includes more recent phase-space functions based on on-off detectors [@LSV15], constituting the special case $K=1$ and being applicable to off-the-shelf detectors (e.g., avalanche photodiodes in Geiger mode and single-photon nanowire detectors); to prove this, see the corresponding experiment with $S=1,2,3$ multiplexing steps [@BTBSSV18]. Furthermore, recent advancements in detector technology (see, e.g., Refs. [@DetNew1; @DetNew2; @DetNew3; @DetNew4]) offer new photon counters to which our detector-agnostic framework is also readily applicable. With our approach, we are further able to predict defining phase-space features of any states by measuring vacuum as a reference. Thus, we can compare a target state with the actually reconstructed DAPS distribution, thus enabling us to estimate other quantum properties as well. As a practical example, a state discrimination task based on our DAPS distributions resulted in distinguishing one- and two-photon states with almost unit certainty, despite high losses in our setup. It is also worth noting that our DAPS distribution includes the full quantum information of the state that is accessible with the detectors used and does not require demanding reconstruction algorithms and detection models. Our experiment comprises state-of-the-art detectors combined with an advantageous method to create coherent states, well mode matched to our signal. As our method is detector agnostic, the detector efficiency need not be specified or even known; the number of data points merely has to be sufficient to produce statistically meaningful results. Also, our approach is not restricted to any specific states; currently, we are mainly limited by the available sources of nonclassical light. In the future, recording the LO’s phase would be beneficial for applying our scheme to phase-sensitive nonclassical states as well. Furthermore, generalizing other interferometric measurement schemes in a detector-agnostic manner is feasible, e.g., as done for on-off detectors [@SVA15]. In addition, we encounter the imperfections stemming from imbalances by assigning systematic errors. It may be possible to avoid this by using more sophisticated strategies [@LFPR16]. Generalized phase-space distributions are becoming increasingly important in identifying vastly different notions of quantumness; see Refs. [@SW18; @SV19] for thorough overviews. To date, however, such universally applicable techniques are also highly dependent on the particular response of the detectors. Our DAPS approach, however, sets a precedence that such limitations can be overcome in theory and experiment. In conclusion, our detector-agnostic framework provides a universally applicable approach to the robust characterization of quantum light in phase space under conditions where detailed knowledge of the measurement apparatus is not available, and forms a basis for future research. #### Acknowledgments.| {#acknowledgments. .unnumbered} The authors are grateful to William R. Clements for helpful discussions and Jelmer J. Renema for his assistance with the installation of the cryogenic infrastructure. The authors also thank Scott Glancy, Arik Avagyan, and Tim Bartley for valuable comments. The Integrated Quantum Optics group acknowledges financial support from the Gottfried Wilhelm Leibniz-Preis (Grant No. SI1115/3-1). This work received funding through the Networked Quantum Information Technologies (NQIT) hub (part of the UK National Quantum Technologies Programme) under Grant No. EP/N509711/1. G. S. T. acknowledges financial support from the Natural Sciences and Engineering Research Council of Canada and the Oxford Basil Reeve Graduate Scholarship. A. E. is supported by EPSRC (project EP/K034480/1 BLOQS). T. A. W. W. is supported by Fondation Wiener - Anspach. J. L. thanks the European Commission (H2020-FETPROACT-2014 grant QUCHIP). I. A. W. acknowledges ERC (Advanced Grant MOQUACINO). This work was supported by the Quantum Information Science Initiative (QISI). [99]{} M. Krenn, M. Malik, T. Scheidl, R. Ursin, and A. Zeilinger, *Quantum Communication with Photons*, [in *Optics in Our Time* (Springer, Cham, 2016), pp. 455–482](https://doi.org/10.1007/978-3-319-31903-2_18). J. L. O’Brien, A. Furusawa, and J. Vučković, *Photonic quantum technologies*, [Nat. Phot. **3**, 687 (2009)](https://doi.org/10.1038/nphoton.2009.229). T. C. Ralph and P. K. Lam, *A bright future for quantum communications*, [Nat. Phot. **3**, 671 (2009)](https://doi.org/10.1038/nphoton.2009.222). D. F. Walls, *Squeezed states of light*, [Nature (London) **306**, 141 (1983)](https://doi.org/10.1038/306141a0). D. T. Smithey, M. Beck, M. G. Raymer, and A. Faridani, *Measurement of the Wigner distribution and the density matrix of a light mode using optical homodyne tomography: Application to squeezed states and the vacuum*, [Phys. Rev. Lett. **70**, 1244 (1993)](https://doi.org/10.1103/PhysRevLett.70.1244). G. Breitenbach, S. Schiller, and J. Mlynek, *Measurement of the quantum states of squeezed light*, [Nature (London) **387**, 471 (1997)](https://doi.org/10.1038/387471a0). A. I. Lvovsky, H. Hansen, T. Aichele, O. Benson, J. Mlynek, and S. Schiller, *Quantum State Reconstruction of the Single-Photon Fock State*, [Phys. Rev. Lett. **87**, 050402 (2001)](https://doi.org/10.1103/PhysRevLett.87.050402). G. Harder, C. Silberhorn, J. Rehacek, Z. Hradil, L. Motka, B. Stoklasa, and L. L. Sánchez-Soto, *Local Sampling of the Wigner Function at Telecom Wavelength with Loss-Tolerant Detection of Photon Statistics*, [Phys. Rev. Lett. **116**, 133601 (2016)](https://doi.org/10.1103/PhysRevLett.116.133601). T. Kiesel, W. Vogel, V. Parigi, A. Zavatta, and M. Bellini, *Experimental determination of a nonclassical Glauber-Sudarshan P function*, [Phys. Rev. A 78, 021804(R) (2008)](https://doi.org/10.1103/PhysRevA.78.021804). K. Laiho, Katiúscia N. Cassemiro, D. Gross, and C. Silberhorn, *Probing the Negative Wigner Function of a Pulsed Single Photon Point by Point*, [Phys. Rev. Lett. **105**, 253603 (2010)](https://doi.org/10.1103/PhysRevLett.105.253603). T. Douce, A. Eckstein, S. P. Walborn, A. Z. Khoury, S. Ducci, A. Keller, T. Coudreau, and P. Milman, *Direct measurement of the biphoton Wigner function through two-photon interference*, [Sci. Rep. **3**, 3530 (2013)](https://doi.org/10.1038/srep03530). C. Baune, J. Fiurášek, and R. Schnabel, *Negative Wigner function at telecommunication wavelength from homodyne detection*, [Phys. Rev. A **95**, 061802(R) (2017)](https://doi.org/10.1103/PhysRevA.95.061802). D.-G. Welsch, W. Vogel, and T. Opatrný, *Homodyne Detection and Quantum-State Reconstruction*, [Prog. Opt. **39**, 63 (1999)](https://doi.org/10.1016/S0079-6638(08)70389-5). C. Silberhorn, *Detecting quantum light*, [Contemp. Phys. **48**, 143 (2007)](https://doi.org/10.1080/00107510701662538). A. I. Lvovsky and M. G. Raymer, *Continuous-variable optical quantum-state tomography*, [Rev. Mod. Phys. **81**, 299 (2009)](https://doi.org/10.1103/RevModPhys.81.299). S. M. Tan, *An inverse problem approach to optical homodyne tomography*, [J. Mod. Opt. **44**, 2233 (1997)](https://doi.org/10.1080/09500349708231881). V. N. Starkov, A. A. Semenov, and H. V. Gomonay, *Numerical reconstruction of photon-number statistics from photocounting statistics: Regularization of an ill-posed problem*, [Phys. Rev. A **80**, 013813 (2009)](https://doi.org/10.1103/PhysRevA.80.013813). T. Richter, *Pattern functions used in tomographic reconstruction of photon statistics revisited*, [Phys. Lett. A **211**, 327 (1996)](https://doi.org/10.1016/0375-9601(96)00029-1). U. Leonhard, M. Munroe, T. Kiss, T. Richter, and M. G. Raymer, *Sampling of photon statistics and density matrix using homodyne detection*, [Opt. Commun. **127**, 144 (1996)](https://doi.org/10.1016/0030-4018(96)00061-2). Z. Hradil, *Quantum-state estimation*, [Phys. Rev. A **55**, R1561(R) (1997)](https://doi.org/10.1103/PhysRevA.55.R1561). A. I. Lvovsky, *Iterative maximum-likelihood reconstruction in quantum homodyne tomography*, [J. Opt. B **6**, S556 (2004)](https://doi.org/10.1088/1464-4266/6/6/014). R. Kosut, I. A. Walmsley, and H. Rabitz, *Optimal Experiment Design for Quantum State and Process Tomography and Hamiltonian Parameter Estimation*, [arXiv:0411093 \[quant-ph\]](https://arxiv.org/abs/quant-ph/0411093). J. Řeháček, D. Mogilevtsev, and Z. Hradil, *Operational Tomography: Fitting of Data Patterns*, [Phys. Rev. Lett. **105**, 010402 (2010)](https://doi.org/10.1103/PhysRevLett.105.010402). D. Mogilevtsev, A. Ignatenko, A. Maloshtan, B. Stoklasa, J. Rehacek, and Z. Hradil, *Data pattern tomography: reconstruction with an unknown apparatus*, [New J. Phys. **15**, 025038 (2013)](https://doi.org/10.1088/1367-2630/15/2/025038). P. L. Kelley and W. H. Kleiner, *Theory of electromagnetic field measurement and photoelectron counting*, [Phys. Rev. **136**, A316 (1964)](https://doi.org/10.1103/PhysRev.136.A316). M. Fleischhauer and D. G. Welsch, *Nonperturbative approach to multimode photodetection*, [Phys. Rev. A **44**, 747 (1991)](https://doi.org/10.1103/PhysRevA.44.747). A. K. Jaiswal and G. S. Agarwal, *Photoelectric Detection with Two-Photon Absorption*, [J. Opt. Soc. Am. **59**, 1446 (1969)](https://doi.org/10.1364/JOSA.59.001446 ). M. K. Akhlaghi, A. H. Majedi, and J. S. Lundeen, *Nonlinearity in single photon detection: Modeling and quantum tomography*, [Opt. Express **19**, 21305 (2011)](https://doi.org/10.1364/OE.19.021305). S. V. Polyakov and A. L. Migdall, *High accuracy verification of a correlated-photon-based method for determining photon-counting detection efficiency*, [Opt. Express **15**, 1390 (2007)](https://doi.org/10.1364/OE.15.001390). A. P. Worsley, H. B. Coldenstrodt-Ronge, J. S. Lundeen, P. J. Mosley, B. J. Smith, G. Puentes, N. Thomas-Peter, and I. A. Walmsley, *Absolute efficiency estimation of photon-number-resolving detectors using twin beams*, [Opt. Express **17**, 4397 (2009)](https://doi.org/10.1364/OE.17.004397). D. N. Klyshko, *Use of two-photon light for absolute calibration of photoelectric detectors*, [Sov. J. Quantum Electron. **10**, 1112 (1980)](https://doi.org/10.1070/QE1980v010n09ABEH010660). A. Luis and L. L. Sánchez-Soto, *Complete Characterization of Arbitrary Quantum Measurement Processes*, [Phys. Rev. Lett. **83**, 3573 (1999)](https://doi.org/10.1103/PhysRevLett.83.3573). J. Fiurášek, *Maximum-likelihood estimation of quantum measurement*, [Phys. Rev. A **64**, 024102 (2001)](https://doi.org/10.1103/PhysRevA.64.024102). G. M. D’Ariano, L. Maccone, and P. Lo Presti, *Quantum Calibration of Measurement Instrumentation*, [Phys. Rev. Lett. **93**, 250407 (2004)](https://doi.org/10.1103/PhysRevLett.93.250407). A. Feito, J. S. Lundeen, H. Coldenstrodt-Ronge, J. Eisert, M. B. Plenio, and I. A. Walmsley, *Measuring measurement: Theory and practice*, [New J. Phys. **11**, 093038 (2009)](https://doi.org/10.1088/1367-2630/11/9/093038). H. B. Coldenstrodt-Ronge, J. S. Lundeen, K. L. Pregnell, A. Feito, B. J. Smith, W. Mauerer, C. Silberhorn, J. Eisert, M. B. Plenio, and I. A. Walmsley, *A proposed testbed for detector tomography*, [J. Mod. Opt. **56**, 432 (2009)](https://doi.org/10.1080/09500340802304929). J. J. Renema, G. Frucci, Z. Zhou, F. Mattioli, A. Gaggero, R. Leoni, M. J. A. de Dood, A. Fiore, and M. P. van Exter, *Modified detector tomography technique applied to a superconducting multiphoton nanodetector*, [Opt. Express **20**, 2806 (2012)](https://doi.org/10.1364/OE.20.002806). J. Peřina, Jr., O. Haderka, V. Michálek, and M. Hamar, *Absolute detector calibration using twin beams*, [Opt. Lett. **37**, 2475 (2012)](https://doi.org/10.1364/OL.37.002475). M. Bohmann, R. Kruse, J. Sperling, C. Silberhorn, and W. Vogel, *Direct calibration of click-counting detectors*, [Phys. Rev. A **95**, 033806 (2017)](https://doi.org/10.1103/PhysRevA.95.033806). K. E. Cahill and R. J. Glauber, *Density operators and quasiprobability distributions*, [Phys. Rev. **177**, 1882 (1969)](https://doi.org/10.1103/PhysRev.177.1882). G. S. Agarwal and E. Wolf, *Calculus for Functions of Noncommuting Operators and General Phase-Space Methods in Quantum Mechanics. II. Quantum Mechanics in Phase Space*, [Phys. Rev. D **2**, 2187 (1970)](https://doi.org/10.1103/PhysRevD.2.2187). J. R. Klauder, *Improved Version of Optical Equivalence Theorem*, [Phys. Rev. Lett. **16**, 534 (1966)](https://doi.org/10.1103/PhysRevLett.16.534). T. Kiesel and W. Vogel, *Nonclassicality filters and quasiprobabilities*, [Phys. Rev. A **82**, 032107 (2010)](https://doi.org/10.1103/PhysRevA.82.032107). K. Husimi, *Some formal properties of the density matrix*, [Proc. Phys. Math. Soc. Jpn. **22**, 264 (1940)](https://doi.org/10.11429/ppmsj1919.22.4_264). O. Landon-Cardinal, L. C. G. Govia, and A. A. Clerk, *Quantitative Tomography for Continuous Variable Quantum Systems*, [Phys. Rev. Lett. **120**, 090501 (2018)](https://doi.org/10.1103/PhysRevLett.120.090501). D. F. Mundarain and J. Stephany, *Husimi’s Q($\alpha$) function and quantum interference in phase space*, [J. Phys. A: Math. Gen. **37**, 3869 (2004)](https://doi.org/10.1088/0305-4470/37/12/010). G. S. Agarwal, *Nonclassical characteristics of the marginals for the radiation field*, [Opt. Commun. **95**, 109 (1993)](https://doi.org/10.1016/0030-4018(93)90059-E). J. Park, Y. Lu, J. Lee, Y. Shen, K. Zhang, S. Zhang, M. S. Zubairy, K. Kim, and H. Nha, *Revealing nonclassicality beyond Gaussian states via a single marginal distribution*, [Proc. Natl. Acad. Sci. U.S.A. **114**, 891 (2017)](https://doi.org/10.1073/pnas.1617621114). J. Sperling and I. A. Walmsley, *Quasiprobability representation of quantum coherence*, [Phys. Rev. A **97**, 062327 (2018)](https://doi.org/10.1103/PhysRevA.97.062327). J. Sperling and W. Vogel, *Quasiprobability distributions for quantum-optical coherence and beyond*, [arXiv:1907.12427](https://arxiv.org/abs/1907.12427). G. S. Agarwal, *Relation between atomic coherent-state representation, state multipoles, and generalized phase-space distributions*, [Phys. Rev. A **24**, 2889 (1981)](https://doi.org/10.1103/PhysRevA.24.2889). J. P. Dowling, G. S. Agarwal, and W. P. Schleich, *Wigner distribution of a general angular-momentum state: Applications to a collection of two-level atoms*, [Phys. Rev. A **49**, 4101 (1994)](https://doi.org/10.1103/PhysRevA.49.4101). R. McConnell, H. Zhang, J. Hu, S. Čuk, and V. Vuletić, *Entanglement with negative Wigner function of almost 3,000 atoms heralded by one photon*, [Nature (London) **519**, 439 (2015)](https://doi.org/10.1038/nature14293). D. Leibfried, D. M. Meekhof, B. E. King, C. Monroe, W. M. Itano, and D. J. Wineland, *Experimental Determination of the Motional Quantum State of a Trapped Atom*, [Phys. Rev. Lett. **77**, 4281 (1996)](https://doi.org/10.1103/PhysRevLett.77.4281). A. Sanpera, R. Tarrach, and G. Vidal, *Local description of quantum inseparability*, [Phys. Rev. A **58**, 826 (1998)](https://doi.org/10.1103/PhysRevA.58.826). J. Sperling and W. Vogel, *Representation of entanglement by negative quasiprobabilities*, [Phys. Rev. A **79**, 042337 (2009)](https://doi.org/10.1103/PhysRevA.79.042337). J. Sperling, E. Meyer-Scott, S. Barkhofen, B. Brecht, and C. Silberhorn, *Experimental Reconstruction of Entanglement Quasiprobabilities*, [Phys. Rev. Lett. **122**, 053602 (2019)](https://doi.org/10.1103/PhysRevLett.122.053602). S. Wallentowitz and W. Vogel, *Unbalanced homodyning for quantum state measurements*, [Phys. Rev. A **53**, 4528 (1996)](https://doi.org/10.1103/PhysRevA.53.4528). H. Paul, P. Törmä, T. Kiss, and I. Jex, *Photon Chopping: New Way to Measure the Quantum State of Light*, [Phys. Rev. Lett. **76**, 2464 (1996)](https://doi.org/10.1103/PhysRevLett.76.2464). J. Sperling, W. R. Clements, A. Eckstein, M. Moore, J. J. Renema, W. S. Kolthammer, S. W. Nam, A. Lita, T. Gerrits, W. Vogel, G. S. Agarwal, and I. A. Walmsley, *Detector-Independent Verification of Quantum Light*, [Phys. Rev. Lett. **118**, 163602 (2017)](https://doi.org/10.1103/PhysRevLett.118.163602). J. Sperling, A. Eckstein, W. R. Clements, M. Moore, J. J. Renema, W. S. Kolthammer, S. W. Nam, A. Lita, T. Gerrits, I. A. Walmsley, G. S. Agarwal, and W. Vogel, *Identification of nonclassical properties of light with multiplexing layouts*, [Phys. Rev. A **96**, 013804 (2017)](https://doi.org/10.1103/PhysRevA.96.013804). See the Supplemental Material, which includes the Refs. [@SCEMRKNLGVAW17; @SECMRKNLGWAV17; @WV96; @WM64; @VW06; @SVA14], for technical details about the experiment and theory, the data and error analysis, and additional results. R. J. Glauber, *Coherent and incoherent states of the radiation field*, [Phys. Rev. **131**, 2766 (1963)](https://doi.org/10.1103/PhysRev.131.2766). E. C. G. Sudarshan, *Equivalence of Semiclassical and Quantum Mechanical Descriptions of Statistical Light Beams*, [Phys. Rev. Lett. **10**, 277 (1963)](https://doi.org/10.1103/PhysRevLett.10.277). U. M. Titulaer and R. J. Glauber, *Correlation functions for coherent fields*, [Phys. Rev. **140**, B676 (1965)](https://doi.org/10.1103/PhysRev.140.B676). L. Mandel, *Non-classical states of the electromagnetic field*, [Phys. Scr. **T12**, 34 (1986)](https://doi.org/10.1088/0031-8949/1986/T12/005). W. Vogel and D.-G. Welsch, *Quantum Optics* ([Wiley-VCH, Weinheim, 2006](https://doi.org/10.1002/3527608524)). G. S. Agarwal, *Quantum Optics* ([Cambridge University Press, 2012](https://doi.org/10.1017/CBO9781139035170)). E. Wolf and C. L. Mehta, *Determination of the Statistical Properties of Light from Photoelectric Measurements*, [Phys. Rev. Lett. **13**, 705 (1964)](https://doi.org/10.1103/PhysRevLett.13.705). Note that in our case, the $z$ parameter is not limited to values that correspond to $-1\leq s\leq 1$ because of $\mathcal K$ being a finite set and, thus, guaranteeing convergence of $G_z$ for any $z$. A. Eckstein, A. Christ, P. J. Mosley, and C. Silberhorn, *Highly Efficient Single-Pass Source of Pulsed Single-Mode Twin Beams of Light*, [Phys. Rev. Lett. **106**, 013603 (2011)](https://doi.org/10.1103/PhysRevLett.106.013603). M. Liscidini and J. E. Sipe, *Stimulated Emission Tomography*, [Phys. Rev. Lett. **111**, 193602 (2013)](https://doi.org/10.1103/PhysRevLett.111.193602). A. E. Lita, A. J. Miller, and S.W. Nam, *Counting near infrared single-photons with 95% efficiency*, [Opt. Express **16**, 3032 (2008)](https://doi.org/10.1364/OE.16.003032). R. P. Welty and J. M. Martinis, *A series array of DC SQUIDs*, [IEEE Trans. Magn. **27**, 2924 (1991)](https://doi.org/10.1109/20.133821). E. Figueroa-Feliciano, B. Cabrera, A. J. Miller, S. F. Powell, T. Saab, and A. B. C. Walker, *Optimal filter analysis of energy-dependent pulse shapes and its application to TES detectors*, [Nucl. Instrum. Methods Phys. Res. A **444**, 453 (2000)](https://doi.org/10.1016/S0168-9002(99)01434-5). P. C. Humphreys, B. J. Metcalf, T. Gerrits, T. Hiemstra, A. E. Lita, J. Nunn, S. W. Nam, A. Datta, W. S. Kolthammer, and I. A. Walmsley, *Tomography of photon-number resolving continuous-output detectors*, [New J. Phys. **17**, 103044 (2015)](https://doi.org/10.1088/1367-2630/17/10/103044). L. Mandel, *Sub-Poissonian photon statistics in resonance fluorescence*, [Opt. Lett. **4**, 205 (1979)](https://doi.org/10.1364/OL.4.000205). R. Short and L. Mandel, *Observation of Sub-Poissonian Photon Statistics*, [Phys. Rev. Lett. **51**, 384 (1983)](https://doi.org/10.1103/PhysRevLett.51.384). J. Sperling, W. Vogel, and G. S. Agarwal, *Sub-Binomial Light*, [Phys. Rev. Lett. **109**, 093601 (2012)](https://doi.org/10.1103/PhysRevLett.109.093601). T. J. Bartley, G. Donati, X.-M. Jin, A. Datta, M. Barbieri, and I. A. Walmsley, *Direct Observation of Sub-Binomial Light*, [Phys. Rev. Lett. **110**, 173602 (2013)](https://doi.org/10.1103/PhysRevLett.110.173602). A. Luis, J. Sperling, and W. Vogel, *Nonclassicality Phase-Space Functions: More Insight with Fewer Detectors*, [Phys. Rev. Lett. **114**, 103602 (2015)](https://doi.org/10.1103/PhysRevLett.114.103602). M. Bohmann, J. Tiedau, T. Bartley, J. Sperling, C. Silberhorn, and W. Vogel, *Incomplete Detection of Nonclassical Phase-Space Distributions*, [Phys. Rev. Lett. **120**, 063607 (2018)](https://doi.org/10.1103/PhysRevLett.120.063607). C. Cahall, K. L. Nicolich, N. T. Islam, G. P. Lafyatis, A. J. Miller, D. J. Gauthier, and J. Kim, *Multi-photon detection using a conventional superconducting nanowire single-photon detector*, [Optica **4**, 1534 (2017)](https://doi.org/10.1364/OPTICA.4.001534). W. Guo, X. Liu, Y. Wang, Q. Wei, L. F. Wei, J. Hubmayr, J. Fowler, J. Ullom, L. Vale, M. R. Vissers, and J. Gao, *Counting near infrared photons with microwave kinetic inductance detectors*, [Appl. Phys. Lett. **110**, 212601 (2017)](https://doi.org/10.1063/1.4984134). K. L. Nicolich, C. Cahall, N. T. Islam, G. P. Lafyatis, J. Kim, A. J. Miller, and D. J. Gauthier, *Universal Model for the Turn-On Dynamics of Superconducting Nanowire Single-Photon Detectors*, [Phys. Rev. Appl. **12**, 034020 (2019)](https://doi.org/10.1103/PhysRevApplied.12.034020). D. Zhu, M. Colangelo, C. Chen, B. A. Korzh, F. N. C. Wong, M. D. Shaw, and K. K. Berggren, *Resolving photon numbers using a superconducting tapered nanowire detector*, [arXiv:1911.09485](https://arxiv.org/abs/1911.09485). J. Sperling, W. Vogel, and G. S. Agarwal, *Balanced homodyne detection with on-off detector systems: Observable nonclassicality criteria*, [Europhys. Lett. **109**, 34001 (2015)](https://doi.org/10.1209/0295-5075/109/34001). C. Lee, S. Ferrari, W. H. P. Pernice, and C. Rockstuhl, *Sub-Poisson-binomial light*, [Phys. Rev. A **94**, 053844 (2016)](https://doi.org/10.1103/PhysRevA.94.053844). J. Sperling, W. Vogel, and G. S. Agarwal, *Quantum state engineering by click counting*, [Phys. Rev. A **89**, 043829 (2014)](https://doi.org/10.1103/PhysRevA.89.043829). Supplemental Material {#sec:Appendices .unnumbered} ===================== Details on the experiment {#sec:Experiment} ========================= The experimental setup is shown in Fig. \[fig:Setup\]. Our pump laser is a titanium sapphire (Ti:Saph) regenerative amplifier that generates femtosecond pulses (center wavelength $780\,\mathrm{nm}$, full width at half maximum \[FWHM\] $15\,\mathrm{nm}$) at a rate of $100\,\mathrm{kHz}$. This rate is chosen to accommodate the thermal relaxation time (${\sim}10\,\mathrm{\mu s}$) of the transition-edge sensors (TESs). We split the pulses into two paths, each pumping a periodically poled potassium titanyl phosphate (ppKTP) waveguide. We pump the first waveguide (right in Fig. \[fig:Setup\]) using filtered (center $775\,\mathrm{nm}$, FWHM $2\,\mathrm{nm}$) pulses from the Ti:Saph. Pumping the ppKTP waveguide generates two-mode squeezed vacuum via type-II parametric down-conversion. The pump is then discarded using a longpass filter. The two down-converted modes (signal $1554\,\mathrm{nm}$, idler $1547\,\mathrm{nm}$) are orthogonally polarized and separated by a polarisation beam splitter. Each mode is sent through a bandpass filter (FWHM $10\,\mathrm{nm}$). The idler mode is sent to a TES detector to herald photon-number states in the signal mode (heralding efficiency ${\sim}40\%$) by postselecting to a specific outcome $k_h$. In the second waveguide (left in Fig. \[fig:Setup\]), we prepare the local oscillator (LO). Since the first and second waveguides have slightly different phase-matching properties, a different pump spectrum (center $783\,\mathrm{nm}$, FWHM $2\,\mathrm{nm}$) is used. This pump spectrum is chosen to maximize the spectral overlap between the LO and signal (SI). We also carve $2\,\mathrm{ns}$ square seed pulses from a continuous-wave laser (center $1580\,\mathrm{nm}$) using an electro-optic modulator. The pump and seed pulses are temporally overlapped and coupled into the second ppKTP waveguide. Through difference frequency generation, the LO is generated in the polarization orthogonal to the seed. The LO is separated from the seed using a polarisation beam splitter. As before, we discard the pump by a longpass filter. The LO’s polarization is adjusted with a half-wave plate to match the SI’s polarization. Then, the LO is sent through a bandpass filter (FWHM $10\,\mathrm{nm}$) to further eliminate seed light as well as increase the spectral overlap with the signal. Finally, neutral-density filters attenuate the LO to the single-photon level. The SI and LO are combined on a 90:10 beam splitter. The resulting light field of one output then enters the multiplexing step (50:50 beam splitter) and the then resulting beams are measured with two TESs. The recorded coincidences give the detection events $E(k_1,k_2)$, which we use for our analysis. ![ Outline of the setup; see Sec. \[sec:Experiment\] for the full description. BP: bandpass filter, BS: beam splitter, CW: continuous-wave, HWP: half-wave plate, ppKTP: periodically poled potassium titanyl phosphate, LP: longpass filter, ND: neutral-density filter, PBS: polarizing beam splitter, TES: transition-edge sensor. []{data-label="fig:Setup"}](arXiv_Setup.pdf){width="\columnwidth"} In addition, we characterized the mode overlap of the signal and LO by combining the two on a 50:50 beam splitter. We consider the specific case of a single photon ($k_h=1$) and a weak LO ($|\beta|\ll 1$). By scanning the delay between the SI and LO, we expect to measure a Hong-Ou-Mandel-type dip in two-fold coincidences at the output of the beam splitter. We measured a dip of ${\sim}80\%$ visibility, suggesting that the mode overlap is at least $80\%$. By blocking the signal, this setup constitutes a Hanbury Brown-Twiss interferometer that allows us to measure the LO’s $g^{(2)}(0)$. We measured $g^{(2)}(0) = 1.005 \pm 0.002$, which is consistent with the expected Poisson distribution for the LO’s photon statistics. Details on the theory {#sec:Theory} ===================== General approach ---------------- Let us formulate some additional details on the theory. As we can expand any SI state in the Glauber-Sudarshan decomposition, $\hat\rho=\int d^2\alpha\,P(\alpha)|\alpha\rangle\langle \alpha|$, it is sufficient to consider the propagation of coherent states $|\alpha\rangle$. Our detection scheme consists of a combination of the SI with the LO state $|\beta\rangle$ on a beam splitter, the multiplexing, and the detection. Applying a beam splitter transformation, we map an input, consisting of SI and LO, as follows: $|\alpha\rangle\otimes|\beta\rangle\mapsto |t\alpha-r\beta\rangle\otimes|r^\ast\alpha+t^\ast\beta\rangle$, where $t$ and $r$ define the trasmissivity and reflectivity ($|t|^2+|r|^2=1$). When tracing over the second mode, we obtain the state that enters the multiplexing stage, $$\begin{aligned} \label{eq:DisplacedState} \hat\rho(\beta)=\int d^2\alpha\, P(\alpha)|t\alpha-r\beta\rangle\langle t\alpha-r\beta|. \end{aligned}$$ Further, the multiplexing distributes the coherent state components in Eq. among the $N=2^S$ output beams, where $S$ is the depth of the multiplexing scheme, $|\gamma\rangle\mapsto |\gamma/\sqrt N\rangle^{\otimes N}$, resulting in $\int d^2\alpha\, P(\alpha)[|(t\alpha-r\beta)/\sqrt N\rangle\langle (t\alpha-r\beta)/\sqrt N|]^{\otimes N}$. Using some combinatorics (see Ref. [@SECMRKNLGWAV17] for details), we find $$\begin{aligned} \begin{aligned} c_{N_0,\ldots,N_K}(\beta)=&\int d^2\alpha\,P(\alpha)\frac{N!}{N_0!\cdots N_K!} \\&\times \prod_{k=0}^K{\underbrace{\left\langle \frac{t\alpha-r\beta}{\sqrt N}\right|\hat\pi_k\left|\frac{t\alpha-r\beta}{\sqrt N}\right\rangle}_{\stackrel{\mathrm{def.}}{=}p_k\left(\frac{t\alpha-r\beta}{\sqrt N}\right)}}^{N_k}, \end{aligned} \end{aligned}$$ where $\{\hat\pi_k\}_{k=0,\ldots,K}$ is an unknown positive operator-valued measure (POVM) that describes the detector. Consequently, the two types of generating functions under consideration read $$\begin{aligned} \label{eq:GeneratingFunction} g_{z_0,\ldots,z_K}(\beta)=&\int d^2\alpha\,P(\alpha)\left[ \sum_{k=0}^K z_kp_k\left(\frac{t\alpha-r\beta}{\sqrt N}\right) \right]^N \intertext{and} \label{eq:GeneratingFunction2} G_z(\beta)=&\int d^2\alpha\,P(\alpha)\left[ \sum_{k=0}^K z^kp_k\left(\frac{t\alpha-r\beta}{\sqrt N}\right) \right]^N, \end{aligned}$$ using the Glauber-Sudarshan $P$ function. As long as $N$ is even and $P\geq0$, both expressions are necessarily nonnegative. Photoelectric counting with loss -------------------------------- Let us analyze our scheme for the special case of photoelectric counting. A simple photoelectric detection is described through POVM elements $\hat\pi_k={:}e^{-\eta\hat n}(\eta\hat n)^k/k!{:}$ for $k=0,1,\ldots$ ($K=\infty$), where $\eta$ is the quantum efficiency and $\hat n$ is the photon-number operator. In this scenario, the generating function in Eq. can be further evaluated [@WM64] and reads $$\begin{aligned} \label{eq:Photoelectric} \begin{aligned} G_z(\beta)=&\int d^2\alpha\,P(\alpha)\exp\left( -[1-z]|t|^2\eta\left|\alpha-\frac{r}{t}\beta\right|^2 \right) \\ =&\left\langle{:} \exp\left[ -(1-z)|t|^2\eta\hat n\left(\frac{r}{t}\beta\right) \right] {:}\right\rangle_{\hat\rho}, \end{aligned} \end{aligned}$$ where $\hat n(\gamma)$ is the displaced photon-number operator. Since we have $P(\gamma;s)=2(\pi[1-s])^{-1}\langle{:}e^{-2\hat n(\gamma)/[1-s]}{:}\rangle_{\hat\rho}$ [@WV96], the above expression can be related to $s$-paramterized distributions. Also note that according to the characterization performed in Ref. [@SECMRKNLGWAV17] (Sec. III), our TESs are more precisely described through POVMs of the form $\hat\pi_k={:}e^{-\Gamma(\hat n)}\Gamma(\hat n)^k/k!{:}$ for $k=0,\ldots, K-1$ and $\hat\pi_K=\hat 1-\sum_{k=0}^{K-1}\hat\pi_k$. Therein, $K<\infty$ reflects the finite photon-number resolution, and the response function $\Gamma$ has a nonlinear form, $\Gamma(\hat n)\approx\eta\hat n+\eta^{(2)}\hat n^2$, where the quantum efficiency $\eta$ is not one ($\eta<1$) and the nonlinear contribution does not vanish ($\eta^{(2)}\neq0$). As $\eta^{(2)}$ is small, the nonlinear behavior mainly affects higher LO and SI intensities. Coincidences and systematic errors {#sec:Coincidence} ================================== A single multiplexing step, $N=2$, was implemented. Thus, it is convenient to formulate the data processing in terms of measured coincidences. For this purpose, we denote with $E(k_1,k_2)$ the number of coincidence events for the measurement outcomes $k_1$ and $k_2$ ($k_1,k_2\in\{0,\ldots,K\}$), resembling the detection bins of the TESs $1$ and $2$, respectively. $E=\sum_{k_1,k_2}E(k_1,k_2)$ defines the total number of events. The coincidences are directly related to the desired quantum version of a multinomial distribution, $c_{N_0,\ldots,N_K}$, cf. Eq. (1) in the main text. Since $N_0+\cdots+N_K=N=2$, we have $$\begin{aligned} c_{N_0,\ldots,N_K} =\left\{\begin{array}{ll} c_{0,\ldots,0,N_i=2,0,\ldots,0} & \text{ for }0\leq i\leq K, \\ c_{0,\ldots,0,N_i=1,0,\ldots,0,N_j=1,0,\ldots,0} & \text{ for }0\leq i<j\leq K, \\ 0 & \text{ otherwise}, \end{array}\right. \end{aligned}$$ where we can identify $c_{0,\ldots,0,N_i=2,0,\ldots,0}=E(i,i)/E$ and $c_{0,\ldots,0,N_i=1,0,\ldots,0,N_j=1,0,\ldots,0}=(E(i,j)+E(j,i))/E$. To estimate the value $\overline f$ of a function $f_{N_0,\ldots,N_K}$, we can recast the standard sampling formula as follows: $$\begin{aligned} \nonumber \overline{f} =&\sum_{\substack{N_0,\ldots,N_K:\\N_0+\cdots+N_K=N}}f_{N_0,\ldots,N_K}\,c_{N_0,\ldots,N_K} \\\nonumber =&\sum_{0\leq i\leq K}\frac{E(i,i)}{E} \overbrace{f_{0,\ldots,0,N_i=2,0,\ldots,0}}^{\stackrel{\mathrm{def}}{=}f(i,i)} \\\nonumber &+\sum_{0\leq i<j\leq K}\frac{E(i,j)+E(j,i)}{E} \underbrace{f_{0,\ldots,0,N_i=1,0,\ldots,0,N_j=1,0,\ldots,0}}_{\stackrel{\mathrm{def}}{=}f(i,j)=f(j,i)} \\ =&\frac{1}{E}\sum_{i,j=0}^K f(i,j)E(i,j). \end{aligned}$$ See the Supplemental Material to Ref. [@SCEMRKNLGVAW17] for the generalization to $N>2$. In order to apply the multinomial framework described in the main text, one has to satisfy the premise that the coincidence statistics is symmetric, $E(k_1,k_2)=E(k_2,k_1)$. However, in reality, this is only true to a limited extent since the beam splitters in the multiplexing might not be perfectly balanced, and the detectors after the multiplexing might have slightly different responses. In Ref. [@SCEMRKNLGVAW17], we provided a rough systematic error estimate to account for such imperfections, which is further refined in the following. The premises mentioned above state that the coincidences are symmetric. The actual measurements $E(k_1,k_2)$ naturally exhibit a certain amount of asymmetry; if not, no systematic error needs to be assigned. We can decompose the coincidences as follows: $$\begin{aligned} \label{eq:CoincidenceDecomposition} E(k_1,k_2)= \overbrace{\frac{E(k_1,k_2)+E(k_2,k_1)}{2}}^{\text{(symmetric part)}} +\overbrace{\frac{E(k_1,k_2)-E(k_2,k_1)}{2}}^{\text{(asymmetric part)}}. \end{aligned}$$ Furthermore, assume we estimate a function $f(k_1,k_2)$ to obtain the mean $\overline{f}=\sum_{k_1,k_2} f(k_1,k_2)E(k_1,k_2)/E$. Inserting the above decomposition and denoting with $\overline{f}^\mathrm{(sym)}$ the value obtained from the symmetric part in Eq. , we apply the triangle inequality and find $$\begin{aligned} \label{eq:SystematicError} \left|\overline f-\overline{f}^\mathrm{(sym)}\right|\leq \sum_{k_1,k_2}|f(k_1,k_2)|\left|\frac{E(k_1,k_2)-E(k_2,k_1)}{2E}\right|=\epsilon_f, \end{aligned}$$ which is the systematic error resulting from the asymmetry in the measured data. As we use the typical quadratic error propagation—rather than the linear form used for the above derivation—, we replace the right-hand-side expression in Eq. with $\epsilon_f^2=\sum_{k_1,k_2}|f(k_1,k_2)|^2\left|[E(k_1,k_2)-E(k_2,k_1)]/[2E]\right|^2$. Recall that the general relation between linear and quadratic error expansion for a function $F(x_1,x_2,\ldots)$ is given by $\Delta^\mathrm{(lin)} F=\sum_j |\partial F/\partial x_j|\Delta x_j$ and $\Delta^\mathrm{(quad)} F=(\sum_j |\partial F/\partial x_j|^2[\Delta x_j]^2)^{1/2}$. In addition, let us remind ourselves that the random error reads $\sigma_f=[(\overline{f^2}-\overline{f}^2)/(E-1)]^{1/2}$, which is combined with the systematic error to give the overall uncertainty, $\Delta f=\sqrt{\epsilon_f^2+\sigma_f^2}$. Sub-multinomial light {#sec:SubMultinomial} ===================== We assess our results with our previously derived nonclassicality criteria [@SCEMRKNLGVAW17; @SECMRKNLGWAV17]. Let us briefly recapitulate this approach and its implementation for a self-consistent reading. The previously devised method is based on the observation that a correlation matrix, $M$, is positive semidefinite for classical light, i.e., $M=(M_{i,j})_{i,j=0,\ldots, K}\geq0$, where $$\begin{aligned} \begin{aligned} M_{i,j} =&N\overline{N_i(N_j+\delta_{i,j})}-(N-1)\overline{N_i}\,\overline{N_j} \\ =&N^2(N-1)\left( \langle{:}\hat\pi_i\hat\pi_j{:}\rangle-\langle{:}\hat\pi_i{:}\rangle\langle{:}\hat\pi_j{:}\rangle \right), \end{aligned} \end{aligned}$$ with $\delta$ denoting the Kronecker symbol. It was shown that the required first- and second-order moments can be obtained from coincidences as [@SCEMRKNLGVAW17] $$\begin{aligned} \overline{N_i} =&N\langle{:}\hat\pi_i{:}\rangle =\sum_{k_1,k_2}\left(\delta_{k_1,i}+\delta_{k_2,i}\right)\frac{E(k_1,k_2)}{E}, \\\nonumber \overline{N_i(N_j+\delta_{i,j})}=&N(N-1)\langle{:}\hat\pi_i\hat\pi_j{:}\rangle \\ =&\sum_{k_1,k_2}\left(\delta_{k_1,i}\delta_{k_2,j}+\delta_{k_1,j}\delta_{k_2,i}\right)\frac{E(k_1,k_2)}{E}. \end{aligned}$$ Finally, the minimal eigenvalue $\mu_{\min}$ of the correspondingly reconstructed matrix $M$ is computed to probe for positive semidefiniteness. The heralding with a TES enables us to generate higher-order photon-number states. In Table \[tab:SubMultinomial\], we listed the nonclassicality in terms of the criteria $\mu_{\min}<0$ for data with a coherent amplitude zero. The observed nonclassicality in Table \[tab:SubMultinomial\] for heralding bins larger than two is no longer significant within a three-standard-deviation error margin as the total number of events $E$ is too small in those cases. For this reason, we restrict our considerations to heralding bins $k_h\in\{0,1,2\}$. $k_h$ sub-multinomial ------- -- -------------------------------- 0 $-0.000\,0 \pm 0.000\,9$ 1 $-0.15\phantom{0\,0} \pm 0.03$ 2 $-0.10\phantom{0\,0} \pm 0.03$ 3 $-0.17\phantom{0\,0} \pm 0.07$ 4 $-0.3\phantom{00\,0} \pm 0.2$ : For the available heralding bins, $k_h$, the minimal eigenvalues $\mu_{\min}$ of the matrix $M$ are shown. Significant negativities defines the notion of nonclassical sub-multinomial light [@SCEMRKNLGVAW17; @SECMRKNLGWAV17]. []{data-label="tab:SubMultinomial"} Local oscillator amplitudes {#sec:LocalOsci} =========================== From the first derivative of Eq. for the photoelectric model, we can infer the dimensionless and displaced intensity, $$\begin{aligned} I(\beta)=&\left.\frac{\partial G_z(\beta)}{\partial z}\right|_{z=1} =|t|^2\eta\langle{:}\hat n\left(\frac{r}{t}\beta\right){:}\rangle_{\hat\rho} \\\nonumber =&\sum_{\substack{N_0,N_1,\ldots\geq0:\\N_0+N_1+\cdots=N}}(0N_0+1N_1+\cdots)c_{N_0,\ldots,N_K}(\beta), \end{aligned}$$ where the latter expression results from the definition $G_z(\beta)=\sum_{N_0,N_1,\ldots}z^{0N_0+1N_1+\cdots}c_{N_0,\ldots,N_K}(\beta)$. Most importantly, in the case that the SI is vacuum, we find $\langle{:}\hat n(\gamma){:}\rangle_{|0\rangle\langle 0|}=|\gamma|^2$. We abstract this observation and define for the general detection scenario a detector-independent amplitude $|\beta^\mathrm{(DI)}|=\sqrt{I(\beta)}=\sqrt{\partial G_z(\beta)/\partial z|_{z=1}}$ for general POVMs and the SI $\hat\rho=|0\rangle\langle 0|$. This also explains the following Eq. as well as Eq. (7) in the main text. We block the SI to infer the amplitude of the LO. The intensity, here represented through the dimensionless quantity $$\begin{aligned} \label{eq:EstimateAmplitude} \left|\beta^\mathrm{(DI)}\right|^2=\sum_{i=0}^K i\overline{N_i}\in[0,NK], \end{aligned}$$ has been varied, realized via $n=0,\ldots,28$ power settings of the seed laser. These settings have been also applied when the SI is not blocked. The settings are chosen such that an equidistant intensity grid is generated. This is confirmed through the linear fit in Fig. \[fig:Calibration\]. ![ The estimated intensity \[cf. Eq. \] as a function of the used setting number $n$. The slope of the linear fit (orange line) is $0.1$. This confirms the intended difference of $1$ photon between two settings when correcting for the $90:10$ splitting that uses $10\%$ of the LO intensity. []{data-label="fig:Calibration"}](arXiv_LOcalibration.pdf){width=".9\columnwidth"} Additional analysis and results {#sec:Additional} =============================== Typical data and error estimates {#sec:SamplingError} -------------------------------- In order to further assess the impact of uncertainties, let us consider a typical example. In Fig. \[fig:RawData\], we depict a typical data set, there for a single photon (i.e., heralding to bin $k_h=1$) and a vanishing LO amplitude (i.e., the setting $n=0$). ![ Raw coincidence data for $k_h=1$ and $\beta=0$. We recorded $E=184\,426$ events in $K+1=5$ bins. []{data-label="fig:RawData"}](arXiv_E_1phot_LO-1.png){width=".8\columnwidth"} The resulting (systematic and random) observational errors for the generating function are depicted in Fig. \[fig:Uncertainty\]. The expected trend of monotonicity and the diverging behavior of the uncertainties as a function of $|z|$ are clearly visible. As $G_{z=1}(\beta)=1$ corresponds to the total probability, which is not subject to fluctuations, the random error vanishes for $z=1$ (Fig. \[fig:Uncertainty\], top-left panel). By construction, the systematic errors are symmetric with respect to $z=0$ (top-right plot in Fig. \[fig:Uncertainty\]). See also the following discussion in Sec. \[sec:Quasiprobability\]. For probing the nonclassicality through the generating function, obtained as $g_{z_0,\ldots,z_K}(\beta)=\sum_{k_1,k_2} z_{k_1}z_{k_2} E(k_1,k_2)/E$, we compute the eigenvector $Z=[z_0,\ldots,z_K]$ to the normalized coincidence matrix $[E(k_1,k_2)/E]_{k_1,k_2=0,\ldots,K}$ that corresponds to the minimal eigenvalue. For the example under study, we get $Z=[0.228, -0.973, -0.033, -0.001, -0.000]$. With this information, we can now estimate the general generating function and get $g_{z_0,\ldots,z_K}(\beta)=-0.026\pm0.003$ as the optimal negativity, here for $\beta=0$. The minimum over all measured LO amplitudes then yields $g_{\min}$, cf. Eq. (6) in the main text. ![ Error composition for the sampling of $G_z(\beta=0)$ as a function of $z$ for the data shown in Fig. \[fig:RawData\] \[top-left: random error, $\sigma_{G_z(0)}$; top-right: systematic error, $\epsilon_{G_z(0)}$; bottom: combination of both random and systematic uncertainties, $\Delta{G_z(0)}$\]. []{data-label="fig:Uncertainty"}](arXiv_Uncertainty.pdf){width="\columnwidth"} Optimal quasiprobability distribution and error estimates {#sec:Quasiprobability} --------------------------------------------------------- We found that the parameter $z=-4.85$ is optimal in the sense that $G_z(\beta)$ has the most statistically significant negativity at the origin (see Fig. \[fig:Quasiprobability\]), i.e., $-G_z(0)/\Delta G_z(0)$ is maximized. In the plotted scenario, the error estimates $\Delta G_z(\beta)$ are rapidly increasing for increasing $|\beta|$, which we discuss in the following based on the sampling formula $$\begin{aligned} G_z(\beta)=\sum_{N_0+\cdots+N_K=N} z^{0N_0+\cdots+KN_K} c_{N_0,\ldots,N_K}(\beta). \end{aligned}$$ For increasing LO amplitudes, the components of $c_{N_0,\ldots,N_K}(\beta)$ that relate to a higher power of $z$ have a higher contribution to the estimate of this function. Similarly, $|z|> 1$ also leads to a most relevant term that corresponds to a higher power of $z$. Recall that the exponent $0N_0+\cdots+KN_K$ relates to the overall intensity \[cf. Eq. \]. Consequently, both a large LO amplitude and $|z|> 1$ result in the fact that the contribution for $z^p$ for larger $p$ becomes the most relevant one. Standard error propagation then implies a relative error scaled by the large factor $p$, for increasing LO amplitudes and increasing $|z|$ values, which also explains why the increase of observational uncertainties in those scenarios is expected. ![ Radial component of the phase-space distribution of the heralded single-photon state ($k_h=1$) for the optimal value $z=-4.85$. For almost all data points with $|\beta^\mathrm{(DI)}|>\sqrt{0.5}$, the error estimates exceeds the plot range while allowing for consistency with expected mean values close to zero, cf. the discussion in Sec. \[sec:Quasiprobability\]. []{data-label="fig:Quasiprobability"}](arXiv_Quasiprobability.pdf){width="\columnwidth"} Quantum state discrimination {#sec:Discrimination} ============================ As our distributions can, in principle, take arbitrary forms for arbitrary detectors, let us formulate the statistical model to discriminate states based on the reconstructed phase-space functions alone. The probability that two distributions, described through multivariate random variables $X$ and $X'$, are indistinguishable within a $\delta$-uncertainty can be expressed as $$\begin{aligned} \label{eq:Discrimination} \mathrm{Prob}(|X-X'|\leq \delta)=\int\limits_{-\delta}^{+\delta} du \int\limits_{-\infty}^{+\infty} dz\,p(X=z)p(X'=z+u), \end{aligned}$$ where $p(X)$ and $p(X')$ are the probability densities of the uncertainties for the two random variables. We identify $X=[G^{(k_h)}_z(\beta_0),\ldots,G^{(k_h)}_z(\beta_{28})]$ and $X'=[G^{(k'_h)}_z(\beta_n)]_{n=0,\ldots,28}$ for different heralded states and LO settings and use a Gaussian error model with a mean and variance that corresponds to the reconstructed distributions for each measured setting $n$. Consequently, we get from Eq. the following probability for the discrimination: $$\begin{aligned} \nonumber &\mathrm{Prob}\left(G^{(k_h)}_z\neq G^{(k'_h)}_z\right)=1-\mathrm{Prob}(\forall n:|X_n-X_n'|\leq \delta_n) \\ =&1-\prod_{n} \underbrace{ \int\limits_{-\delta_n}^{+\delta_n}du_n\, \frac{ \exp\left[-\frac{(u_n-[\mu_n-\mu_n'])^2}{2\Delta_n^2}\right] }{\sqrt{2\pi\Delta_n^2}} }_{ =\mathrm{Err}\left[\frac{|\mu_n-\mu'_n|}{\Delta_n}+3\right] -\mathrm{Err}\left[\frac{|\mu_n-\mu'_n|}{\Delta_n}-3\right] } , \end{aligned}$$ where we set the vector $\delta$ to correspond to three combined standard deviations for each setting, $\delta_n=3\Delta_n$ and $\Delta_n=[\Delta G^{(k_h)}_z(\beta_n)^2+\Delta G^{(k_h)}_z(\beta_n)^2]^{1/2}$, and using the mean values $\mu_n=\overline{G^{(k_h)}_z(\beta_n)}$ and $\mu'_n=\overline{G^{(k'_h)}_z(\beta_n)}$. Note that $\mathrm{Err}[z]=\int^z d\xi e^{-\xi^2/2}/\sqrt{2\pi}$ denotes the error function. For instance, we find for our measured data that the likelihood to discriminate the phase-space distributions for $k_h$ from the one for $k'_h$ for $z=-1.5$ is given by the matrix $$\begin{aligned} \!\!\!\!\left[\!\mathrm{Prob}\!\!\left(\!\! G^{(k_h)}_z{\neq} G^{(k'_h)}_z \!\!\right)\!\!\right]_{\!\! k_h,k'_h{=}0,1,2}\!\! {=}\!\!\!\begin{bmatrix} 7.5\% & 100.\% & 100.\% \\ 100.\% & 7.5\% & 98.9\% \\ 100.\% & 98.9\% & 7.5\% \end{bmatrix}\!\!\!, \end{aligned}$$ where “$100.\%$” corresponds to a value which is $100\%$ within the used numerical precision. Note that the diagonal elements are nonzero as identical distributions could still represent different states when considering a finite error margin. Fit model from vacuum measurements {#sec:Gauge} ================================== From the measurement in which the SI is blocked (i.e., vacuum SI), we can extrapolate the general shape of the phase-space distribution for photon states, without relying on any particular detector model. This approach is also used to fit the reconstructed distributions for the heralding to $k_h$. ![ Phase-space distribution for vacuum, $G_z(\beta)=G_z^\mathrm{(vac)}(\beta)=G^{(0)}_z(\beta)$, with $z=-1.5$. The dashed line corresponds to a fit function $f_{0}\exp[-b|\beta^\mathrm{(DI)}|^2]$ for real-valued constants $f_0$ and $b$. []{data-label="fig:Vacuum"}](arXiv_Vacuum.pdf){width=".8\columnwidth"} Using the data where the signal is blocked, we find that a Gaussian distribution describes the reconstructed phase-space distribution for vacuum quite well; see Fig. \[fig:Vacuum\]. This information can be used to predict the phase-space distributions of $m$-photon states as well. Because of Eq. and the known representation $P^{(m)}(\alpha)=\sum_{j=0}^m\binom{m}{j}j!^{-1}\partial_{\alpha}^j\partial_{\alpha^\ast}^jP^{(0)}(\alpha)$ [@VW06], where $P^{(0)}(\alpha)$ describes the delta distribution centered at the origin, we find that the $m$-th photon state is given by $$\begin{aligned} \label{eq:FockStates} G^{(m)}_z(\beta)=\sum_{j=0}^m\binom{m}{j}\frac{1}{j!}\left[\frac{|t|^2}{|r|^2}\right]^j\partial_{\beta}^j\partial_{\beta^\ast}^jG^{(0)}_z(\beta), \end{aligned}$$ where $G^{(0)}_z(\beta)$ is experimentally obtained by blocking the signal (Fig. \[fig:Vacuum\]) and which is determined without relying on any detection models. For deriving Eq. , note that the argument of the vacuum function implies that a partial integration of Eq. with derivatives of delta distributions results in $\left.\partial^j_\alpha\partial^{j}_{\alpha^\ast} f(\beta-t\alpha/r)\right|_{\alpha=0}=(-t/r)^j(-t^\ast/r^\ast)^{j}\partial^j_\beta\partial^{j}_{\beta^\ast} f(\beta)$. From Eq. and the fit obtained from $G^{(0)}_z(\beta)$, we can therefore predict the phase-space distribution of an $m$-photon state. In our case, this means that $G^{(m)}_z(\beta)$ is a Gaussian function multiplied with a fixed $m$th-order polynomial in $|\beta|^2$. In this context, also recall the linear relation between the actual intensity (via the setting number $n$) and the detector-independent intensity in Fig. \[fig:Calibration\]. In addition, it is known (see, e.g., Ref. [@SVA14]) that imperfect heralding for the kind of photon source used leads to additional noise contributions. For this reason, our fit for an heralding to the $k_h$th bin is described through $G^{(k_h)}_z(\beta)=\sum_{j=0}^{k_h} f_j|\beta^\mathrm{(DI)}|^{2j}\exp[-b|\beta^\mathrm{(DI)}|^2]$, which constitutes the generalized fit function used in the main text \[Figs. 2(a)–(c)\] and is determined from the vacuum measurements alone and without relying on any detection models. As a final remark, it is worth mentioning that the above treatment can be straightforwardly generalized to predict $G_z(\beta)$ for arbitrary states, resulting in Eq. (8) in the main text. This is based on the fact that the $P$ function of an arbitrary state can be written as a convolution, $P(\alpha)=\int d^2\alpha'\,P(\alpha-\alpha')P^{(0)}(\alpha')$, recalling that the vacuum state is described by a delta distribution, $P^{(0)}$. Thus, Eq. implies that $G_z$ of an arbitrary state, represented through the Glauber-Sudarshan distribution $P$, is predicted to resemble the convolution of the already measured vacuum state’s $G_z^{(0)}$ and $P$. Even more generally, we can write $$\begin{aligned} g_{z_0,\ldots,z_K}(\beta)=\int d^2\alpha\, P(\alpha) g^\mathrm{(0)}_{z_0,\ldots,z_K}\left(\beta-\frac{t}{r}\alpha\right), \end{aligned}$$ where $g^\mathrm{(0)}_{z_0,\ldots,z_K}(\beta)=\left[\sum_{k=0}^K z_kp_k\left(-r\beta/\sqrt N\right)\right]^N$, cf. Eq. , to predict the phase-space distribution for a state, theoretically described through $P(\alpha)$, via the measured $g^\mathrm{(0)}_{z_0,\ldots,z_K}$.
{ "pile_set_name": "ArXiv" }
--- bibliography: - 'biblioChaGarPouTon.bib' --- **B-urns[^1]** [Brigitte Chauvin $ ^a$, Danièle Gardy $ ^b$, Nicolas Pouyanne $ ^a$ and Dai-Hai Ton-That $ ^b$ ]{} $ ^a$ Laboratoire de Mathématiques de Versailles, CNRS UMR 8100.\ $ ^b$ Laboratoire PRiSM, CNRS UMR 8144.\ Université de Versailles - St-Quentin,\ 45, avenue des Etats-Unis, 78035 Versailles Cedex, France. July 22nd, 2015\ 1truecm [**Abstract.**]{} The fringe of a B-tree with parameter $m$ is considered as a particular Pólya urn with $m$ colors. More precisely, the asymptotic behaviour of this fringe, when the number of stored keys tends to infinity, is studied through the composition vector of the fringe nodes. We establish its typical behaviour together with the fluctuations around it. The well known phase transition in Pólya urns has the following effect on B-trees: for $m\leq 59$, the fluctuations are asymptotically Gaussian, though for $m\geq 60$, the composition vector is oscillating; after scaling, the fluctuations of such an urn strongly converge to a random variable $W$. This limit is $\g C$-valued and it does not seem to follow any classical law. Several properties of $W$ are shown: existence of exponential moments, characterization of its distribution as the solution of a smoothing equation, existence of a density relatively to the Lebesgue measure on $\g C$, support of $W$. Moreover, a few representations of the composition vector for various values of $m$ illustrate the different kinds of convergence. Introduction {#intro} ============ B-trees are a fundamental structure in computer science, they have been introduced in the early seventies by Bayer and McCreight [@Bay; @BayMcC], to store large quantities of data. These particular search trees are conceived in order to have all their leaves at the same level. The nodes at the deepest level are called the *fringe nodes*. A precise description can be found in Section \[sec-algo\] where are presented two classical algorithms giving a B-tree. The actual writing of these algorithms can be found for example in Cormen et al. [@CorLeiRiv] for one of them (the so-called *prudent* algorithm in the sequel), in Kruse and Ryba [@KruRyb] for the other one (called the *optimistic* algorithm in the sequel). The fringe analysis of B-trees goes back to Yao [@Yao] and has been developped by many authors (see for example the Baeza-Yates’ survey [@BaeSurvey]), both for B-trees and B$^{+}$-trees (where all the keys are stored in the fringe nodes). In Yao’s paper [@Yao] appears the Pólya urn model, which we develop in this article. Indeed, the fringe of a B-tree with parameter $m$ (where $m$ is a positive integer) can be considered as a particular Pólya urn with $m$ colors, so that a lot of information can be obtained concerning the asymptotic behaviour of this fringe, when the number of stored keys tends to infinity. Let us describe a Pólya urn process as follows. Consider an urn that contains balls of, say, $d$ different colors. Start with a finite number of different color balls as initial composition (possibly monochromatic). At each discrete time $n$, draw a ball at random, check its color, put it back into the urn and add balls according to the following rule: if the drawn ball is of color $i$, add $a_{i,j}$ balls of color $j$, where the $a_{i,j}$ are integer-valued. Thus, the replacement rule is described by the so-called *replacement matrix*, which is a dimension $d$ matrix, whose coefficients are the $a_{i,j}$, for $i$ and $j$ in $\{ 1, \dots , d\}$. Usually, the integers $a_{i,j}$ are assumed to be nonnegative for $i\not= j$ and the integers $a_{i,i}$ are nonpositive or nonnegative. A negative coefficient $a_{i,i}$ means that, if a ball of color $i$ is drawn, then $a_{i,i}$ balls of color $i$ are removed from the urn. In this case, we have to ensure that at least $a_{i,i}$ balls of color $i$ exist in the urn. This quality is called the *tenability* of the urn. To ensure that an urn with a negative coefficient $a_{i,i}$ is tenable, it is necessary and sufficient to have the following arithmetical condition (this can be easily proved by induction on $n$). Fix an initial composition $(\alpha_1, \dots , \alpha_d)$, meaning that there are $\alpha_j$ balls of colour $j$ at time zero in the urn, then the tenability condition can be written as $$\label{tenable} - a_{i,i} \hbox{ divides } \alpha_i, a_{1,i}, \dots , a_{d,i} .$$ Moreover, in the present paper, the urn is assumed to be [*balanced*]{}, which means that the total number of balls added at each step is a constant: there exists an integer $S$ such that, for any $i$ in $\{ 1, \dots , d\}$, $\displaystyle\sum_{j=1}^d a_{i,j} = S$. Let us emphasize that “drawing a ball at random” means choosing *uniformly* among the balls contained in the urn. That is why this model is related to many situations in mathematics, algorithmics or theoretical physics where a uniform choice among objects determines the evolution of a process. See Johnson and Kotz’s book [@JK], Mahmoud’s book [@Mah08] or Flajolet et al. [@FlaDumPuy] for many examples. For a general probabilistic treatment of Pólya urns, see [@Pou08], Janson [@Jan] or Mailler [@Mailler]. In Yao’s paper [@Yao], the focus is on the average number of nodes in the B-tree. Nevertheless, the main ideas are already there, namely the dynamics transforming a tree of size $n$ into a tree of size $n+1$, which is the same dynamics as in a Pólya urn process. The recent progresses in Pólya urn processes and their asymptotic behaviour ([@Pou08; @ChaPouSah; @ChaMaiPou], Janson [@Jan], Mailler [@Mailler]) lead to a more complete landscape for the B-trees. Our aim in this article is to present in a hopefully concise form a collection of results about the asymptotic behaviour of the fringe nodes in a B-tree, namely their typical behaviour and the fluctuations around it. Our main interest is focused on these fluctuations, which happen to have a phase transition: for $m\leq 59$, the fluctuations are of order $\sqrt n$ and have a Gaussian limit in distribution. But for $m\geq 60$, the fluctuations are of order $n^{\sigma}$, where $\sigma$ is larger than $1/2$ and increases to $1$ when $m$ tends to infinity. Moreover, an oscillating and significative phenomenon occurs in the fluctuation term. After scaling, the fluctuations strongly converge (meaning almost surely) to a random limit, here called $W$. The random variable $W$ is $\g C$-valued and does not seem to follow any classical law. The paper is organized as follows. In Section \[sec-algo\] are presented two classical algorithms allowing to construct a B-tree. In that section is also precised how the insertion dynamics is that of a suitable Pólya urn. In Section \[sec-ProcessusDiscrets\] are introduced the random vectors which describe the fringe of a B-tree. In Section \[sec-transition\] is established the phase transition, and we get the precise asymptotic behaviour of the fringe nodes, in Corollary \[asymptoticsBtreeL\] of Theorem \[asymptoticsDTL\]. For $m\geq 60$, the fluctuations around the drift are expressed via a random variable $W$, which is studied in the last sections. Thanks to an embedding into continuous time (Section \[sec-embedding\]), a multitype branching process is put forward. Properties of the continuous-time limit process can be translated to the discrete-time process, via an explicit connection. Several properties of $W$ are proved in Section \[sec-limitlaw\]: $W$ admits a density on the whole complex plane; it has exponential moments; it is the unique solution of a certain “smoothing equation” in a convenient probability distribution space. Finally, in Section \[sec-simulations\], a few pictures provide a synthetic and concrete illustration of the different kinds of convergence, depending on whether $m\leq 59$ or $m\geq 60$. B-tree algorithms {#sec-algo} ================= Description of a B-tree ----------------------- For a positive integer $m\geq 2$, a B-tree with parameter $m$ is a search[^2] tree, where the keys are stored into the internal nodes and the leaves[^3] represent insertion possibilities (we call them *gaps*), they do not contain any key; furthermore all the leaves are at the same depth. A *fringe node* is an internal node whose only descendants are leaves. In the literature, these fringe nodes are sometimes called *final internal nodes* or *leaf-nodes*, or *internal leaves*. We try to be non-ambiguous in the following, and use the terms fringe nodes and fringe node process. In the figures below, internal nodes are represented by ellipses and leaves by squares. As is the case for the leaves, the fringe nodes of a B-tree are at the same depth. Moreover, each internal node (fringe or otherwise) has a capacity; the root contains between $1$ and $C(m)$ keys, and the other internal nodes between $c(m)$ and $C(m)$ keys. When a node contains $C(m)$ keys, we say that the node is *saturated*. The minimal – $c(m)$ – and maximal – $C(m)$ – values depend both on the parameter $m$ and on the precise definition of the B-tree, which is itself closely related to the exact insertion algorithm, of which we present two versions below. Let us just state that $c(m) = m-1$ and $C(m) = 2m-1$ for the first algorithm, and $c(m) = m$ and $C(m) = 2m$ for the second one. In both cases we want to insert a new key into a tree of size $n$, i.e. having already $n$ keys in its internal nodes, and consequently $n+1$ leaves, or insertion possibilities. =\[ellipse,draw,dashed, fill=gray!20\] =\[ellipse,draw\] =\[ellipse,fill=pink,draw=pink\] =\[ellipse,fill=red,draw=red\] =\[rectangle,draw,inner sep=2pt\] =\[rectangle,draw,fill=pink,inner sep=2pt\] =\[rectangle,draw,fill=green,inner sep=2pt\] =\[rectangle,draw,fill=red,inner sep=2pt\] =\[&gt;=latex,&lt;-,thick,level distance = 8mm\] \[level 1/.style=[sibling distance=22mm]{}, level 2/.style=[sibling distance=4mm]{}, level 3/.style=[sibling distance=4mm]{}, noedge/.style=[edge from parent/.style=]{}, nonode/.style= \] at (3,0) [$\bullet$ $\bullet$]{} child [node \[entree\] [ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [$\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [$\bullet$ $\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} ; The prudent algorithm --------------------- In what we call here the *prudent algorithm* for insertion into a B-tree with parameter $m$, the nodes contain between $m-1$ and $2m-1$ keys. An insertion of a new key concerns a given leaf, so that a branch (the nodes between the root and this leaf) is determined for this insertion. The algorithm proceeds by *going down* from the root to the leaf, along this branch. We begin by checking the root: if it is saturated, it is split, a new root is created with a single key which is the median of the keys of the old root (remember that it has an odd number of keys, hence the median is defined without any ambiguity) and two sons, and the height of the tree increases by $1$. If the root is not saturated, we do not modify it. We then proceed along the branch to the insertion gap. When we meet a (non-root) saturated node, the median key of that node moves to the parent node (which is not saturated – if it initially was, we have already taken care of it) and the saturated node is split. Then, when we finally arrive at a fringe node, we split it when necessary, and the insertion of the new key always takes place into a non-saturated fringe node: the saturated nodes are dealt with *before* we find the node in which the insertion of the new key will take place. This algorithm, which can be presented both recursively and iteratively (there being only a descent from the root to a leaf), is found, e.g., in the book of Cormen et al [@CorLeiRiv]. If we consider the fringe nodes, insertion on a saturated node (with $2m-1$ keys) gives rise to $2$ new fringe nodes with respectively $m$ and $m-1$ keys. See Figure \[fig:Btree-standardNC\]. =\[ellipse,draw,dashed, fill=gray!20\] =\[ellipse,draw\] =\[ellipse,fill=pink,draw=pink\] =\[ellipse,fill=red,draw=red\] =\[rectangle,draw,inner sep=2pt\] =\[rectangle,draw,fill=pink,inner sep=2pt\] =\[rectangle,draw,fill=green,inner sep=2pt\] =\[rectangle,draw,fill=red,inner sep=2pt\] =\[&gt;=latex,&lt;-,thick,level distance = 8mm\] \[level 1/.style=[sibling distance=22mm]{}, level 2/.style=[sibling distance=4mm]{}, level 3/.style=[sibling distance=4mm]{}, noedge/.style=[edge from parent/.style=]{}, nonode/.style= \] at (3,0) [$\bullet$ $\bullet$]{} child [node \[entree\] [ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [$\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [$\bullet$ ${\color{red}\bullet}$ $\bullet$]{} child[node \[gapvert\] child\[pointeur\][node \[nonode\]]{} child\[noedge\][node \[nonode\]]{} ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} ; at (12,0) [$\bullet$ $\bullet$ ${\color{red}\bullet}$]{} child [node \[entree\] [$\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [$\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [${\color{green}\bullet}$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [$\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} ; -20pt The optimistic algorithm ------------------------ In what we call here the *optimistic algorithm*, for insertion into a B-tree with parameter $m-1$, the nodes contain between $m-1$ and $2m-2$ keys. Here the saturated nodes are dealt with *after* we have found the place for insertion. An insertion of a new key concerns a given leaf. If the corresponding fringe node is not saturated, the insertion occurs in this node; if it is saturated, the algorithm has to create a non-saturated node into which we can insert the new key. It needs to find what would be the place of the new key among the (already sorted) $2m-2$ keys; the middle key among these $2m-2 +1 = 2m-1$ keys moves to the parent node, and the saturated node is split into $2$ new fringe nodes with $m-1$ keys. If the parent node is saturated, a key is pushed up into the grandparent node, etc... all the way up to the root if necessary; if the root is saturated, it is split as well and the height of the tree increases by $1$. This algorithm proceeds by going down from the root to the gap of insertion, and then up to (some node on) the branch from that leaf to the root, and is possibly best understood recursively. Figure \[fig:BtreeNC\] illustrates an insertion on a saturated node for a B-tree with parameter $m=3$. =\[ellipse,draw,dashed, fill=gray!20\] =\[ellipse,draw\] =\[ellipse,fill=pink,draw=pink\] =\[ellipse,fill=red,draw=red\] =\[rectangle,draw,inner sep=2pt\] =\[rectangle,draw,fill=pink,inner sep=2pt\] =\[rectangle,draw,fill=green,inner sep=2pt\] =\[rectangle,draw,fill=red,inner sep=2pt\] =\[&gt;=latex,&lt;-,thick,level distance = 8mm\] \[level 1/.style=[sibling distance=22mm]{}, level 2/.style=[sibling distance=4mm]{}, level 3/.style=[sibling distance=4mm]{}, noedge/.style=[edge from parent/.style=]{}, nonode/.style= \] at (3,0) [$\bullet$ $\bullet$]{} child [node \[entree\] [$\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [$\bullet$ $\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [$\bullet$ ${\color{red}\bullet}$ $\bullet$ $\bullet$]{} child[node \[gapvert\] child\[pointeur\][node \[nonode\]]{} child\[noedge\][node \[nonode\]]{} ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} ; at (12,0) [$\bullet$ $\bullet$ ${\color{red}\bullet}$]{} child [node \[entree\] [$\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [$\bullet$ $\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [${\color{green}\bullet}$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [$\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} ; -20pt Insertion as the evolution of a Pólya urn ----------------------------------------- For both the prudent and the optimistic algorithms, let us define different types of fringe nodes: we say that a fringe node is *of type $k$*, when it contains $m+k-2$ keys and has thus $m+k-1$ gaps. For the prudent algorithm, $k$ varies between $1$ and $m+1$; for the optimistic algorithm, $k$ varies between $1$ and $m$ . We analyse the fringe of the tree through the so-called *composition vector* $L_n$, which counts the number of fringe nodes of each type in a B-tree of parameter $m$, at time $n$, i.e., assuming that we start from an empty tree and add the keys one by one, when the tree contains $n$ keys. Thus, $L_n^{(k)}$, the $k$-th coordinate of $L_n$, is the number of fringe nodes of type $k$. For the prudent algorithm, $L_n$ is a vector of dimension $m+1$, whereas it is a vector of dimension $m$ for the optimistic algorithm . For both algorithms, we define $G_n$ as the composition vector of *gaps* at time $n$. We say that a gap is *of type $k$*, when it is attached to a fringe node of type $k$. Thus $G_n^{(k)}$, the $k$-th coordinate of $G_n$, is the number of gaps of type $k$. In other words: $$\label{gap-node} (m+k-1) L_n^{(k)} = G_n^{(k)}.$$ For both algorithms, the process $(G_n)_{n\in\g N}$ is a Pólya urn process, as defined in the Introduction, where the balls are the gaps and the colors are the different types. Indeed, when the keys are randomly chosen under the so-called random permutation model, meaning that the keys are independently identically distributed (i.i.d.), then the insertion of a new key in a B-tree of size $n$ occurs *uniformly* on any of the $n+1$ gaps of the tree. $\bullet$ In the prudent algorithm, the number of keys in a fringe node ranges from $m-1$ to $2m-1$, there are $m+1$ types, and the vector $L_n$ is of dimension $m+1$. The replacement matrix of the gap process is of dimension $m+1$ and equal to[^4] $$r_m= \left( \begin{array}{ccccc} -m&(m+1)&&&\\ &-(m+1)&(m+2)&&\\ &&\ddots&\ddots&\\ &&&\ddots&2m\\ m&(m+1)&&&-2m \end{array} \right).$$ Figure \[fig:Btree-standard\] illustrates the same insertion as in Figure \[fig:Btree-standardNC\], taking into account the different types (colors) of the fringe nodes. =\[ellipse,draw,dashed, fill=gray!20\] =\[ellipse,draw\] =\[ellipse,fill=pink,draw=pink\] =\[ellipse,fill=red,draw=red\] =\[rectangle,draw,inner sep=2pt\] =\[rectangle,draw,fill=pink,inner sep=2pt\] =\[rectangle,draw,fill=red,inner sep=2pt\] =\[&gt;=latex,&lt;-,thick,level distance = 8mm\] \[level 1/.style=[sibling distance=22mm]{}, level 2/.style=[sibling distance=4mm]{}, level 3/.style=[sibling distance=4mm]{}, noedge/.style=[edge from parent/.style=]{}, nonode/.style= \] at (3,0) [$\bullet$ $\bullet$]{} child [node \[entree\] [ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entreerose\] [$\bullet$ $\bullet$]{} child[node \[gaprose\] ]{} child[node \[gaprose\] ]{} child[node \[gaprose\] ]{} ]{} child [node \[entreerouge\] [$\bullet$ $\bullet$ $\bullet$]{} child[node \[gaprouge\] child\[pointeur\][node \[nonode\]]{} child\[noedge\][node \[nonode\]]{} ]{} child[node \[gaprouge\] ]{} child[node \[gaprouge\] ]{} child[node \[gaprouge\] ]{} ]{} ; at (12,0) [$\bullet$ $\bullet$ $\bullet$]{} child [node \[entree\] [$\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entreerose\] [$\bullet$ $\bullet$]{} child[node \[gaprose\] ]{} child[node \[gaprose\] ]{} child[node \[gaprose\] ]{} ]{} child [node \[entreerose\] [$\bullet$ $\bullet$]{} child[node \[gaprose\] ]{} child[node \[gaprose\] ]{} child[node \[gaprose\] ]{} ]{} child [node \[entree\] [$\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} ; -20pt $\bullet$ In the optimistic algorithm, the number of keys in a fringe node ranges from $m-1$ to $2m-2$, there are $m$ types, and the vector $L_n$ is of dimension $m$. The replacement matrix of the gap process is of dimension $m$ and equal to $$R_m=\begin{pmatrix} -m&m+1&&&\\ &-(m+1)&m+2&&\\ &&\ddots&&\\ &&&-(2m-2)&2m-1\\ 2m&&&&-(2m-1) \end{pmatrix}.$$ Figure \[fig:Btree\] illustrates the same insertion as in Figure \[fig:BtreeNC\], taking into account the different types (colors) of the fringe nodes. =\[ellipse,draw,dashed, fill=gray!20\] =\[ellipse,draw\] =\[ellipse,fill=pink,draw=pink\] =\[ellipse,fill=red,draw=red\] =\[rectangle,draw,inner sep=2pt\] =\[rectangle,draw,fill=pink,inner sep=2pt\] =\[rectangle,draw,fill=red,inner sep=2pt\] =\[&gt;=latex,&lt;-,thick,level distance = 8mm\] \[ level 1/.style=[sibling distance=22mm]{}, level 2/.style=[sibling distance=4mm]{}, level 3/.style=[sibling distance=4mm]{}, noedge/.style=[edge from parent/.style=]{}, nonode/.style= \] at (3,0) [$\bullet$ $\bullet$]{} child [node \[entree\] [$\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entreerose\] [$\bullet$ $\bullet$ $\bullet$]{} child[node \[gaprose\] ]{} child[node \[gaprose\] ]{} child[node \[gaprose\] ]{} child[node \[gaprose\] ]{} ]{} child [node \[entreerouge\] [$\bullet$ $\bullet$ $\bullet$ $\bullet$]{} child[node \[gaprouge\] child\[pointeur\][node \[nonode\]]{} child\[noedge\][node \[nonode\]]{} ]{} child[node \[gaprouge\] ]{} child[node \[gaprouge\] ]{} child[node \[gaprouge\] ]{} child[node \[gaprouge\] ]{} ]{} ; at (12,0) [$\bullet$ $\bullet$ $\bullet$]{} child [node \[entree\] [$\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entreerose\] [$\bullet$ $\bullet$ $\bullet$]{} child[node \[gaprose\] ]{} child[node \[gaprose\] ]{} child[node \[gaprose\] ]{} child[node \[gaprose\] ]{} ]{} child [node \[entree\] [$\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} child [node \[entree\] [$\bullet$ $\bullet$]{} child[node \[gap\] ]{} child[node \[gap\] ]{} child[node \[gap\] ]{} ]{} ; -30pt Observe that both replacement matrices $r_m$ and $R_m$ are balanced (any row sums to $1$), which is an immediate consequence of the dynamics, since one key (one ball) is added at each unit of time. All the results in this paper hold for both algorithms, including the phase transition depending on whether $m\leq 59$ or $m\geq 60$. However, the proofs and results for the optimistic algorithm being somewhat simpler than those for the prudent algorithm, we choose to present them and to leave the other case to the reader: from now on, *we consider a B-tree constructed by the optimistic algorithm.* Gaps of a B-tree as a Pólya urn {#sec-ProcessusDiscrets} =============================== Let us remind from Section \[sec-algo\] that a fringe node of a B-tree contains from $m-1$ to $2m-2$ keys and from $m$ to $2m-1$ gaps. For any $k\in\{ 1,\dots ,m\}$, a fringe node that contains $m+k-2$ keys is called *of type $k$*. We are interested in the *fringe node composition vector* $L_n$ of a B-tree at time $n$, whose $k$-th coordinate counts the number of fringe nodes of type $k$. Let $\left( e_k\right) _{1\leq k\leq m}$ be the canonical basis of $\g R^m$. Denote by $w_1,\dots ,w_m$ the vectors defined by $$\label{increments} \left\{ \begin{array}{l} w_k=-e_k+e_{k+1},~~1\leq k\leq m-1\\ w_m=2e_1-e_{m}. \end{array} \right.$$ The $w_k$’s are the increment vectors of the fringe node dynamics: when a key is inserted in a fringe node of type $k\in\{ 1,\dots ,m-1\}$, the fringe node is replaced by a fringe node of type $k+1$ (addition of vector $w_k$) and when a key is inserted in a fringe node of type $m$, the fringe node is replaced by two fringe nodes of type $1$ (addition of vector $w_m$). When the keys are randomly drawn under the permutation model, the insertion is *uniform* on the gaps and the fringe node composition process of a B-tree is modelized by the $\g R^m$-valued Markov chain $\left( L_n\right) _{n\in\g N}$ defined as follows by its transition probabilities. For any $k\in\{ 1,\dots ,m\}$, $$\label{markovLeaves} \Proba\left( L_{n+1}=L_n+w_k\Big| L_n\right) =\frac{(m+k-1)\langle L_n,e_k\rangle}{K_n}$$ where the scalar product $\langle L_n,e_k\rangle = L_n^{(k)}$ is the $k$-th coordinate of $L_n$ and where $K_n$ denotes the total number of gaps at time $n$. Of course, when the process starts initially with $N_0$ keys at time $0$, then $K_n=1+N_0+n$. Note that, considering the B-tree, the number of gaps in a fringe node of type $k$ is $m+k-1$ whereas the total number of gaps in the tree at time $n$ is exactly $K_n$ so that Formulae  completely define a Markov process; it reflects the uniform insertions of the keys in the gaps. Alternatively, one can consider the gap process. A gap is called of type $k$ when it is contained in a fringe node of type $k$. Note once more that a fringe node of type $k$ contains $m+k-1$ gaps. Expressed in terms of gaps, the dynamics of key insertion in the B-tree is the following: when the key is inserted in a gap of type $k\in\{ 1,\dots ,m-1\}$, then $m+k-1$ gaps of type $k$ disappear and are replaced by $m+k$ gaps of type $k+1$; when the key is inserted in a gap of type $m$, then $2m-1$ gaps of type $m$ disappear and are replaced by $2m$ gaps of type $1$. Moreover, under the random permutation model, *the gaps are drawn uniformly*. In other words, the gap composition process of a B-tree is modelized by the following $m$-color Pólya urn process $\left( G_n\right) _{n\in\g N}$, having $R_m$ as replacement matrix and $(m,0,\dots ,0)$ as initial composition. Denote by $\left( G_n\right) _{n\in\g N}$ the $m$-color Pólya urn process defined by the $m$-dimensional replacement matrix $$R_m=\begin{pmatrix} -m&m+1&&&\\ &-(m+1)&m+2&&\\ &&\ddots&&\\ &&&-(2m-2)&2m-1\\ 2m&&&&-(2m-1) \end{pmatrix}.$$ Its balance, namely its common row sum, equals $1$. Note that the diagonal entries are negative. Nevertheless, the urn is tenable because, in any column, all entries are multiple of the diagonal coefficient: when a ball of color $2$ is drawn, $m+1$ extra balls must be withdrawn from the urn which is always possible because balls of type $2$ are put $m+1$ by $m+1$ in the urn as can be seen on $R_m$’s second column. The same phenomenon occurs for any color. Of course, these negative diagonal entries imply that one must necessarily take an initial composition that satisfies such divisibility conditions as well: for any $k\in\{ 1,\dots ,m\}$, the initial number of balls of type $k$ satisfies $$m+k-1~\hbox{ divides }~\langle L_0,e_k\rangle.$$ The symbol $\langle .,. \rangle$ denotes here the standard scalar product on $\g R^m$. Thus, the condition is fulfilled. Both Markov processes $\left( L_n\right) _{n}$ and $\left( G_n\right) _{n}$ are related by Relation , stated hereunder. Let $P$ be the $m$-dimensional diagonal invertible matrix $$\label{defP} P=\Diag\left( m,m+1,\dots ,2m-1\right).$$ When $V\in\g N^m\setminus\{ 0\}$, denote by $\left( L_n^V\right) _{n\geq 0}$ the fringe node process starting with $L_0=V$ and by $\left( G_n^V\right) _{n\geq 0}$ the gap process starting from $G_0=V$. Then, for any $V\in\g N^m\setminus\{ 0\}$, one has[^5] immediatly from $$\label{lienLeafGapDiscret} \left( G_n^{PV}\right) _{n\in\g N}\egalLoi \left( PL_n^V\right) _{n\in\g N}.$$ In particular, denoting by $|V|$ the sum of $V$’s coordinates, the total number of gaps in the B-tree at time $n$ is $$K_n=n+|PV| =\sum _{k=1}^m(m+k-1)\langle L_n^{V},e_k\rangle =\sum _{k=1}^m\langle G_n^{PV},e_k\rangle .$$ In the following, when no confusion is possible, we lighten the notation $G_n^{PV}$ into $G_n$, and $L_n^{V}$ into $L_n$, like in Theorems \[asymptoticsDTG\] and \[asymptoticsDTL\] below. Phase transition: *small* and *large* B-trees {#sec-transition} ============================================= With regard to the asymptotics of their composition vector, Pólya urns are subject to a well known phase transition. See the above references on Pólya urns for a general treatment. Let us translate the phenomenon for our B-urns, looking at the spectral properties of the replacement matrix. Spectral decomposition of the ambient space ------------------------------------------- In this section, we state notations relative to the spectral decomposition of the matrix $R_m$. These notations are used all along the paper. The (unitary) characteristic polynomial of $R_m$ turns out to be $$\label{polycar} \chi _m(X)= \prod _{k=m}^{2m-1}(X+k)-\frac{(2m)!}{m!}. $$ Its complex roots are all simple, the one having the largest real part one being $1$. Furthermore, two distinct eigenvalues that have the same real part are conjugated. We denote by $$\label{deflambda2} \lambda _2=\sigma _2+i\tau _2$$ the eigenvalue of $R_m$ having the second largest real part $\sigma _2$ and a positive imaginary part $\tau _2$. We adopt also the following notations: $$\label{notationsSpectrales} \left\{ \begin{array}{l} H_{m+1}(X)=\displaystyle \sum _{k=1}^m\frac{1}{X+k} \\ \\ v(\lambda )=\displaystyle \frac{1}{(m+\lambda )H_{m+1}(m+\lambda -1)}\times \\ \\ \hskip 20pt\displaystyle \left( 1,\frac{m+1}{m+1+\lambda}, \frac{(m+1)(m+2)}{(m+1+\lambda)(m+2+\lambda)}, \dots ,\frac{(m+1)\dots(2m-1)}{(m+1+\lambda)\dots (2m-1+\lambda)} \right) \\ \\ \langle v(\lambda ),e_k\rangle = \displaystyle \frac{1}{(m+\lambda )H_{m+1}(m+\lambda -1)} \prod_{j=1}^{k-1}\frac{m+j}{m+j+\lambda} \\ \\ u(\lambda )\left( x_1,\dots ,x_m\right) =\displaystyle \sum _{k=1}^m\left( \prod _{j=0}^{k-2}\frac{\lambda+m+j}{1+m+j}\right) x_k \\ \\ \hskip 90pt =x_1 + \displaystyle \frac{\lambda+m}{1+m}\ x_2 +\frac{(\lambda+m)(\lambda+m+1)}{(1+m)(2+m)}\ x_3 + \cdots \\ \\ \hskip 100pt + \displaystyle\frac{(\lambda+m)\dots(\lambda+2m-2)}{(1+m)\dots(2m-1)}\ x_m. \end{array} \right.$$ When $\lambda$ is an eigenvalue of $R_m$, the vector $v(\lambda )$ is an eigenvector of $\transp{R_m}$ associated with $\lambda$. The linear form $u(\lambda )$ is an eigenform of $\transp{R_m}$ associated with $\lambda$, which means that for any (column) vector $V$, $u(\lambda )\left(\transp{R_m}V\right) =\lambda u(\lambda )(V)$. Moreover, if $\lambda$ and $\mu$ are eigenvalues of $R_m$, then $u(\lambda)\left[ v(\mu )\right]=\delta _{\lambda ,\mu}$ (Kronecker). In other words, $\left( v(\lambda )\right) _{\lambda\in\Sp\left( R_m\right)}$ and $\left( u(\lambda )\right) _{\lambda\in\Sp\left( R_m\right)}$ are dual basis of respectively eigenvectors and eigenforms of $\transp{R_m}$. In the sequel, for more simplicity, we denote $$\label{notations12} \left\{ \begin{array}{l} v_1=v(1)=\displaystyle \frac 1{H_{m+1}(m)}\left( \frac 1{m+1}, \frac 1{m+2}, \dots , \frac 1{2m} \right) \\ \\ u_1\left( x_1,\dots ,x_m\right) =u(1)\left( x_1,\dots ,x_m\right) =\displaystyle \sum _{k=1}^mx_k \\ \\ v_2=v(\lambda _2) {\rm ~~and~~} u_2=u(\lambda _2). \end{array} \right.$$ The complex vector space $\g C^m$ admits the decomposition as direct sum of $\transp{R_m}$-stable lines $$\g C^m=\bigoplus _{\lambda\in\Sp\left( R_m\right)}\g Cv(\lambda )$$ and the corresponding projection on any line $\g Cv(\lambda )$ is $u(\lambda )v(\lambda)$. In the real field, we use the decomposition $$\label{defRondV1} \g R^m=\g Rv_1\oplus\rond V_1$$ where $\rond V_1$ is the only subspace which is simultaneously $\transp{R_m}$-stable and supplementary to $\g Rv_1$. It is generated by the vectors respectively constituted by the real parts and the imaginary parts of the coordinates of the complex vectors $v(\lambda )$, $\lambda\in\Sp\left( R_m\right)\setminus\{ 1\}$. In the same vein, we denote by $\rond V_2$ the only $\transp{R_m}$-stable subspace of $\g R^m$ that satisfies $$\g R^m=\g Rv_1\oplus\g R\Re\left( v_2\right)\oplus\g R\Im\left( v_2\right)\oplus\rond V_2.$$ Phase transition for gaps and fringe nodes ------------------------------------------ The phase transition on urns is expressed on the gap process $(G_n)_n$ in the following result. Note the two very different convergence modes: a weak one for small phases *vs* a strong one with periodic phenomena for large phases. See simulations in Section \[sec-simulations\] for an illustration. \[asymptoticsDTG\] Let $m\geq 2$. Let $V\in\g N^m$ be a non zero vector and let $\left( G_n\right) _{n\geq 0}$ be the (discrete-time) gap process starting with the initial condition $G_0=PV$. Then, with notations  and , \(i) (Small phases) if $m\leq 59$, as $n$ tends to infinity, $\displaystyle\frac{G_n-nv_1}{\sqrt n}$ converges in distribution to a centered Gaussian vector; \(ii) (Large phases) if $m\geq 60$, as $n$ tends to infinity, $$\label{expansionDTG} G_n=nv_1+2\Re \left( n^{\lambda _2}W^{DT}v_2\right) +o\left( n^{\sigma _2}\right),$$ almost surely and in any ${\rm L}^p$, $p\geq 1$, where $W^{DT}$ is a complex-valued random variable with expectation $\displaystyle\frac{\Gamma\left( |PV|\right)}{\Gamma\left( |PV|+\lambda _2\right)}u_2(PV)$ ($\Gamma$ denotes Euler Gamma function). [[Proof.]{} ]{} These results come from the general theory of balanced Pólya urn processes. See Janson [@Jan] or [@Pou08]. The numerical values of $\sigma_2$ leading to the phase transition are given in the Appendix. $\hfill{\sqcap\!\!\!\!\sqcup}$ In Theorem \[asymptoticsDTG\](ii), for any $p\in ]0,1[$, the asymptotics is also valid in the (not locally convex) complete metric space ${\rm L}^p$ defined by the usual quasi-norm. This is true for Theorem \[asymptoticsDTL\], Corollary \[asymptoticsBtreeL\], Theorem \[asymptoticsCTG\], Theorem \[asymptoticsCTL\] and Remark \[rem-expansionBtreeCT\] as well. When $V$ is a complex vector, $\Re (V)$ denotes the vector made of real parts of $V$’s coordinates. The random variable $W^{DT}$, which is more deeply studied below, appears as a martingale limit in the field of urn theory. It can be also described as the almost sure limit of $G_n$ after normalisation and projection along the principal direction defined by $v_2$: $$W^{DT}=\lim _{n\to\infty}\frac{1}{n ^{\lambda _2}}u_2\left( G_n\right).$$ Of course, the phase transition and the asymptotics can be straightforwardly translated on the fringe node process $\left( L_n\right) _n$ *via* the diagonal matrix $P$ defined in . The random variable $W^{DT}$ that appears for large phases in Theorem \[asymptoticsDTL\] has the same law as the one of the gap process in Theorem \[asymptoticsDTG\]. \[asymptoticsDTL\] Let $m\geq 2$. Let $V\in\g N^m$ be a non zero vector and let $\left( L_n\right) _{n\geq 0}$ be the (discrete-time) fringe node process starting with the initial condition $L_0=V$. Then, with notations , and , \(i) (Small phases) if $m\leq 59$, as $n$ tends to infinity, $\displaystyle\frac{L_n-nP^{-1}v_1}{\sqrt n}$ converges in distribution to a centered Gaussian vector; \(ii) (Large phases) if $m\geq 60$, as $n$ tends to infinity, $$\label{expansionDTL} L_n=nP^{-1}v_1+2\Re \left( n^{\lambda _2}W^{DT}P ^{-1}v_2\right) +o\left( n^{\sigma _2}\right),$$ almost surely and in any ${\rm L}^p$, $p\geq 1$, where $W^{DT}$ is a complex-valued random variable which has the same distribution as in the variable named the same way in Theorem \[asymptoticsDTG\]. Geometrically speaking, expansion (and expansion as well) can be understood as follows. Notice that an analogous explanation holds for expansions and in Theorem \[asymptoticsCTG\] and Theorem \[asymptoticsCTL\] respectively. \[spirale\] (420,190) (217,-118)[![image](spirale-eps-converted-to.pdf){height="400pt" width="250pt"}]{} (350,-10)[$\Re (v_2)$]{} (380,25)[$\Im (v_2)$]{} (350,145)[$v_1$]{} (0,80) Let us denote by $\varphi$ any argument of the complex number $W^{DT}$. T he trajectory of the random vector $G_n$, projected in the $3$-dimensional real vector space spanned by the vectors $(\Re (v_2),\Im (v_2),v_1)$ is almost surely asymptotic to the (random) spiral $$\left\{ \begin{array}{l} x_n=2|W|n^{\sigma _2}\cos (\tau _2 \log n+\varphi), \\ y_n=-2|W|n^{\sigma _2}\sin (\tau _2 \log n+\varphi), \\ z_n=n, \end{array} \right.$$ drawn on the (random) revolution surface $$4|W|^2z^{2\sigma _2}=x^2+y^2,$$ when $n$ tends to infinity. As is well known in the field of Pólya urn processes, the phase transition is due to the number $\sigma _2$. When $\sigma _2<1/2$, the Pólya urn is *small* and admits a weak Gaussian asymptotics. On the contrary, when $\sigma _2>1/2$, the urn is *large* and has a strong and oscillating ($\lambda _2$ is nonreal) asymptotic behaviour. Considering the replacement matrix $R_m$, it turns out that $\sigma _2$ is an increasing function of $m$ and that: $\bullet$ when $m=59$, $\lambda _2=(0.49534...)+(9.10305...)i$ while $\bullet$ when $m=60$, $\lambda _2=(0.50378...)+(9.10270...)i$. These numerical values have been computed by a Newton approximation algorithm, which can be found in the Appendix. The monotonicity of $\sigma _2$ as a function of $m$ (it increases to $1$ when $m$ tends to infinity) has been evocated by Hennequin [@Hen] in a figure. Qualitatively, let us emphasize the fact that for large values of $m$ (which is the actual use in computer science, since $m$ amounts to several hundreds), the fluctuation term with $W^{DT}$ is highly significant. We deduce from these theorems the asymptotic behaviour of the composition vector of the fringe nodes of different types in a B-tree, which is a particular case of Theorem \[asymptoticsDTL\] with the initial condition $V=(1,0,\dots ,0)$. We stated the theorems above for arbitrary initial conditions because of the further study of the limit law $W^{DT}$ that requires these wider statements. \[asymptoticsBtreeL\] Let $m\geq 2$. Let $\rond L_n$ be the composition vector at time $n$ of the fringe nodes of different types in a B-tree with minimum degree $m$. Then, as $n$ goes of to infinity, with notations  and , \(i) when $m\leq 59$, $\displaystyle\frac{\rond L_n-nP^{-1}v_1}{\sqrt n}$ converges in distribution to a centered gaussian vector; \(ii) when $m\geq 60$, $\rond L_n=nP^{-1}v_1+2\Re \left( n^{\lambda _2}W_m^{\rm B-tree}P ^{-1}v_2\right) +o\left( n^{\sigma _2}\right)$, almost surely and in any ${\rm L}^p$, $p\geq 1$, where $W_m^{\rm B-tree}$ is a complex-valued random variable with expectation $\displaystyle\frac{m!}{\Gamma\left( m+\lambda _2\right)}$. Embeddings into continuous time {#sec-embedding} =============================== Going further and obtaining significant properties of the random limit $W^{DT}$ is not so easy. As can be seen in this section, the classical method of embedding in continuous time turns out to be very fruitful: this idea of embedding discrete urn models in continuous-time branching processes goes back at least to Athreya and Karlin [@AK]. A description is given in Athreya and Ney’s book [@AN Section 9]. The method has been recently revisited and developed by Janson [@Jan], it is the core of recent results on Pólya urns in [@ChaPouSah; @ChaMaiPou]. Definition of the continuous-time fringe node process ----------------------------------------------------- Denote by $\left( L(t)\right) _{t\in\g R_{\geq 0}}$ the $\g N^m\setminus\{ 0\}$-valued continuous time Markov process having $\rond G$ as infinitesimal generator, where $\rond G$ is defined, for any function $f:\g N^m\setminus\{ 0\}\to \rond V$ ($\rond V$ is any real or complex vector space) and for any nonzero $X=\left( x_1,\dots ,x_m\right)\in\g N^m$, by $$\rond G(f)(X)= \sum _{k=1}^m(m+k-1)x_k\Big[ f\left( X+w_k\right)- f\left( X\right)\Big]$$ where the increment vectors $w_k$ have already be defined by . This process is a multitype branching process, embedding of the Markov chain $\left( L_n\right) _n$ into continuous time, as classically done (see for example Bertoin [@Bertoin]). One can think of it the following way. At each (real) time $t\geq 0$, one gets particles of $m$ different types named $1,2,\dots ,m$. Each particle is equipped with a clock that rings at random times. The clock of any particle of type $k$ is exponentially distributed, with parameter $m+k-1$ and all the clocks are independent. The dynamics of the process is the same as in discrete time: for any $k\in\{ 1,\dots ,m-1\}$, when the clock of a particle of type $k$ rings, the particle disappears and is replaced by a particle of type $k+1$; when the clock of a particle of type $m$ rings, the particle disappears and is replaced by two particles of type $1$. Having the same dynamics, the distributions of the processes $\left( L_n\right) _{n\in\g N}$ and $\left( L(t)\right) _{t\in\g R_{\geq 0}}$ are as usual related by the finite-time connection $$\label{finiteTimeConnectionL} (L_n)_{n\in\g N}\egalLoi \left(L\left(\tau _{(n)}\right)\right)_{n\in\g N}$$ where $\tau _{(n)}$ denotes the $n$-th splitting time (the $n$-th ringing time). This relation allows us to transfer results on one process to the other. In particular, the results below strongly rely on the fact that $(e^{- (\transp{R_m}) t}L(t))_{t\in\g R_{\geq 0}}$ is a vector-valued martingale (see Janson [@Jan] or Athreya and Ney [@AN] for this). Definition of the continuous-time gap process --------------------------------------------- Define the vector-valued continuous-time Markov process $\left( G(t)\right) _{t\in\g R_{\geq 0}}$ as being the embedding into continuous time of the discrete-time urn process $\left( G_n\right) _{n\in\g N}$. It takes its values in the set of vectors of the form $PV$ where $V\in\g N^m\setminus\{ 0\}$. With notations as above, its infinitesimal generator is given by $$\rond H(f)(X)= \sum _{k=1}^mx_k\Big[ f\left( X+Pw_k\right)- f\left( X\right)\Big] ,$$ the increment vectors $Pw_k$ being the rows of the urn replacement matrix $R_m$. One can think of this process the following way. Take an urn that contains clocks of $m$ different colors named $1,\dots ,m$. Each clock rings at a random time, exponentially distributed with parameter $1$ and all the clocks are independent. As soon as a clock rings, the following replacement mechanism occurs: if the ringing clock has color $k\in\{ 1,\dots ,m-1\}$, then it disappears together with $m+k-2$ other clocks of color $k$ and $m+k$ clocks of color $k+1$ arise in the urn; if the ringing clock has color $m$, then it disappears together with $2m-2$ other clocks of color $k$ and $2m$ clocks of color $1$ arise in the urn. The fact that the ringing times are exponentially distributed allows to think as if all clocks were restarted as soon as one of them rings. Note that the fact that many clocks disappear at the same time prevents $\left( G(t)\right) _{t\in\g R_{\geq 0}}$ from being a multitype branching process. As in the preceding case, the processes $\left( G_n\right) _{n\in\g N}$ and $\left( G(t)\right) _{t\in\g R_{\geq 0}}$ have the same dynamics, so that $$\label{finiteTimeConnectionG} \left(G_n\right)_{n\in\g N}\egalLoi \left(G\left(\tau '_{(n)}\right)\right)_{n\in\g N},$$ where $\tau '_{(n)}$ denotes the $n$-th ringing time. When $V\in\g N^m\setminus\{ 0\}$, denote by $\left( L(t)^V\right) _{t\geq 0}$ the fringe node process starting with $L(0)=V$ and by $\left( G(t)^V\right) _{t\geq 0}$ the gap process starting from $G(0)=V$. Then, as in the discrete-time case in , for any $V\in\g N^m\setminus\{ 0\}$, $$\label{lienLeafGapContinu} \left( G(t)^{PV}\right) _{t\geq 0} \egalLoi \left( P L(t)^V\right) _{t\geq 0},$$ where $P$ is the diagonal matrix defined in . Asymptotics of both continuous-time processes --------------------------------------------- The asymptotics of the continuous-time processes admit the same kind of phase transition as in discrete time. We state this asymptotics for both continuous-time processes in Theorems \[asymptoticsCTL\] and \[asymptoticsCTG\]. Since $G$ is the image of $L$ by $P$, any of these theorem implies the other one. Nevertheless, as explained below, we prove both of them together using results on branching processes and results on Pólya urns. \[asymptoticsCTG\] Let $m\geq 2$. Let $V\in\g N^m$ be a non zero vector and let $\left( G(t)\right) _{t\in\g R_{\geq 0}}$ be the continous-time gap process that satisfies $G(0)=PV$. Then, with notations , \(i) (Small phases) when $m\leq 59$, as $t$ tends to infinity, $e^{-t}G(t)$ converges almost surely and in any ${\rm L}^p$, $p\geq 1$, to $\xi v_1$ where $\xi$ is a positive random variable which is Gamma-distributed with parameter $|PV|$. Furthermore, if one writes $G(t)=G_1(t)+G'_1(t)$ where the random vector $G_1(t)$ is proportional to $v_1$ and where $G'_1(t)$ is $\rond V_1$-valued (see), then $e^{-t}G_1(t)$ converges almost surely and in any ${\rm L}^p$ to $\xi v_1$ while $e^{-t/2}G'_1(t)$ converges in distribution to $\sqrt\xi N$ where $N$ is a centered $\rond V_1$-valued gaussian vector independant of $\xi$. \(ii) (Large phases) when $m\geq 60$, as $t$ tends to infinity, $$\label{expansionCTG} G(t)=e^t\xi v_1 \left( 1+o(1)\right) +2\Re\left( e^{\lambda _2t}W^{CT}v_2\right)\left( 1+o(1)\right) +o\left( e^{\sigma _2t}\right)$$ almost surely and in any ${\rm L}^p$, $p\geq 1$, where $W^{CT}$ is a complex-valued random variable with expectation $u_2(PV)$ and $\xi$ a positive random variable that is Gamma distributed with parameter $|PV|$. The almost sure remainder $o\left( e^{\sigma t}\right)$ is a $\rond V_2$-valued random vector. \[asymptoticsCTL\] Let $m\geq 2$. Let $V\in\g N^m$ be a non zero vector and let $\left( L(t)\right) _{t\in\g R_{\geq 0}}$ be the continous-time fringe node process that satisfies $L(0)=V$. Then, with notations  and , \(i) (Small phases) when $m\leq 59$, as $t$ tends to infinity, $e^{-t}L(t)$ converges almost surely and in any ${\rm L}^p$, $p\geq 1$, to $\xi P^{-1}v_1$ where $\xi$ is a positive random variable which is Gamma-distributed with parameter $|PV|$. Furthermore, if one writes $L(t)=L_1(t)+L'_1(t)$ where the random vector $L_1(t)$ is proportional to $P^{-1}v_1$ and where $L'_1(t)$ is $P^{-1}\rond V_1$-valued (see), then $e^{-t}L_1(t)$ converges almost surely and in any ${\rm L}^p$ to $\xi P^{-1}v_1$ while $e^{-t/2}L'_1(t)$ converges in distribution to $\sqrt\xi N'$ where $N'$ is a centered $P^{-1}\rond V_1$-valued gaussian vector independant of $\xi$. \(ii) (Large phases) when $m\geq 60$, as $t$ tends to infinity, $$\label{expansionCTL} L(t)=e^t\xi P^{-1}v_1 \left( 1+o(1)\right) +2\Re\left( e^{\lambda _2t}W^{CT}P^{-1}v_2\right)\left( 1+o(1)\right) +o\left( e^{\sigma _2t}\right)$$ almost surely and in any ${\rm L}^p$, $p\geq 1$, where $W^{CT}$ is a complex-valued random variable with expectation $u_2(PV)$ and $\xi$ a positive random variable that is Gamma distributed with parameter $|PV|$. The almost sure remainder $o\left( e^{\sigma t}\right)$ is a $P^{-1}\rond V_2$-valued random vector. Note that the random variables $\xi$ and $W^{CT}$ that appear in both theorems have been denoted the same way because their distributions are the same in both cases. This comes immediately from . . Despite the fact that similar results can be found in Janson [@Jan] and Mailler [@Mailler], the particular case of our processes is not properly contained in their statement. The proofs are essentially made the same way as in both papers [@ChaLiuPouContinu] and [@ChaLiuPouDiscret]. We give hereunder the general scheme of the argumentation. The first tool comes from the fact that the normalised projection $\left( e^{-t}u_1(G(t))\right) _{t\geq 0}$ is always a convergent positive martingale. The random variable $\xi$ is its limit. \(i) Small phases. The process $\left( L(t)\right) _{t\in\g R_{\geq 0}}$ is a multitype branching process so that [*(i)*]{} in Theorem \[asymptoticsCTL\] is covered by [@AN] and [@Jan]. Relation  thus implies [*(i)*]{} in Theorem \[asymptoticsCTG\]. \(ii) Large phases. As for the first projection, $\left( e^{-\lambda _2t}u_2(G(t))\right) _{t\geq 0}$ is a martingale, which is convergent if, and only if $\sigma _2>1/2$, *i.e.* when $m\geq 60$. The complex-valued random variable $W^{CT}$ is its limit. The oscillating term $\Re\left( e^{\lambda _2t}W^{CT}P^{-1}v_2\right)$ in Theorem \[asymptoticsCTL\] is a consequence of [@AN] and [@Jan]’s results. In order to establish the almost sure remainders $o\left( e^{\sigma _2t}\right)$, we use results on discrete-time Pólya urns shown in [@Pou08]. The work is done on the gap process $(G(t))_t$ viewed as an embedded urn into continuous time. For any $t\geq 0$, decompose $G(t)$ as the sum $G(t)=G_1(t)+G_2(t)+G_\ell (t)+G_s(t)$ of its respective following projections on the described supplementary subspaces: $\bullet$ $G_1(t)$ is the projection on $\g Rv_1$ as before; $\bullet$ $G_2(t)$ is the projection on the real plane generated by the real part and the imaginary part of $v_2$; $\bullet$ $G_\ell (t)$ is the projection on the subspace of $\g R^m$ generated by the real and imaginary parts of the eigenvectors $v(\lambda )$ for all eigenvalues $\lambda$ different from $1$ and $\lambda _2$ such that $\Re\left(\lambda\right) >1/2$ (*large* projections); $\bullet$ finally, $G_s(t)$ is the projection on the subspace of $\g R^m$ generated by the real and imaginary parts of the eigenvectors $v(\lambda )$ for all eigenvalues $\lambda$ such that $\Re\left(\lambda\right)\leq 1/2$ (*small* projections). As seen before, $e^{-t}G_1(t)$ converges to $\xi v_1$ almost surely and in ${\rm L}^p$, $p\geq 1$, by martingale techniques; this gives rise to the first term $e^t\xi v_1$ in the asymptotics of $G(t)$. Since $G_2(t)=2\Re\left[ u_2\left( G_2(t)\right)v_2\right]$ and because of the convergence in ${\rm L}^p$, $p\geq 1$, of the complex martingale $\left( e^{-\lambda _2t}u_2(G(t))\right) _{t\geq 0}$ mentioned above, one gets the second term $\Re\left( e^{\lambda _2t}W^{CT}v_2\right)$ of $G(t)$’s asymptotics. The remainder $o\left( e^{\sigma _2t}\right)$ is obtained from $G_\ell$ and $G_s$ asymptotics. As for $G_2$, if $\lambda$ is an eigenvalue of $R_m$ such that $\Re (\lambda )>1/2$, by martingale arguments, the complex projection of $G(t)$ on any eigenline $\g Cv(\lambda )$ is equivalent to $e^{\lambda t}W_\lambda$ almost surely and in any ${\rm L}^p$ where $W_\lambda$ is a complex-valued random variable. In particular, the whole projection $G_\ell (t)$ is $o\left( e^{\sigma _2t}\right)$, almost surely and in any ${\rm L}^p$. To make the proof complete, it remains to show that $G_s(t)$ is $o\left( e^{\sigma _2t}\right)$ as well. To prove this fact, we use the technique detailed in [@ChaLiuPouContinu] (Theorem 4.1 and Lemma 4.2). It consists in considering the same projection for the discrete-time urn process $(G_n)_n$, in using the moment bounds proven in [@Pou08] for small projections of discrete-time Pólya urns and in coming back to continuous time by Relation . By this means, one shows after some probabilistic arguments that for any $\eta >0$, the whole projection $G_s$ satisfies that $e^{-\left(\eta +\frac12\right)t}G_s(t)$ is bounded, almost surely and in ${\rm L}^p$, $p\geq 1$, implying the expected result on $G(t)$. The corresponding asymptotics of $L(t)$ is obtained by taking the image of $G(t)$ by $P^{-1}$. $\hfill{\sqcap\!\!\!\!\sqcup}$ \[rem-expansionBtreeCT\] For $m\geq 60$, we deduce from these theorems the asymptotic behaviour of the continuous-time fringe node process, denoted by $(\rond L(t))_t$, starting from the B-tree initial condition $V=(1, 0, \dots , 0)$: $$\label{expansionBtreeCTL} \rond L(t)=e^t\xi P^{-1}v_1 \left( 1+o(1)\right) +2\Re\left( e^{\lambda _2t}\rond W^{CT}P^{-1}v_2\right)\left( 1+o(1)\right) +o\left( e^{\sigma _2t}\right)$$ almost surely and in any ${\rm L}^p$, $p\geq 1$, where $\rond W^{CT}$ is a complex-valued random variable with expectation $m$ and $\xi$ a positive random variable that is Gamma distributed with parameter $m$. The almost sure remainder $o\left( e^{\sigma t}\right)$ is a $P^{-1}\rond V_2$-valued random vector. For large phases, the finite time connections  or  lead to a relation between the random variables $W$ in discrete and continuous times. This relation, commonly named *martingale connection* will be stated and used below in the article. We indicate hereafter how one can get it. Take for instance Relation  concerning the gap processes $(G_n)_n$ and $(G(t))_t$ starting with the same initial condition $G_0=G(0)=PV$. Using Theorems \[asymptoticsDTG\] and \[asymptoticsCTG\], since $\tau _{(n)}$ tends almost surely to $+\infty$ as $n$ goes of to infinity, one gets successively $\xi =\lim _{t\to\infty}e^{-t}u_1\left( G(t)\right) =\lim _{n\to\infty}e^{-\tau _{(n)}}u_1(G_n) =\lim _{n\to\infty}ne^{-\tau _{(n)}}$ on one hand. On the other hand, $W^{CT}=\lim _{t\to\infty}e^{-\lambda _2t}u_2\left( G(t)\right) =\lim _{n\to\infty}e^{-\lambda _2\tau _{(n)}}u_2(G_n) =\lim _{n\to\infty}\left[ ne^{-\tau _{(n)}}\right]^{\lambda _2}\left[ n^{-\lambda _2}u_2(G_n)\right]$. This entails the martingale connection $$\label{martingaleConnection} W^{CT}\egalLoi\xi ^{\lambda _2}W^{DT}.$$ We just recall here that the random variable $\xi$ is Gamma-distributed with expectation $|PV|$. Limit law of large B-trees {#sec-limitlaw} ========================== In this section appear the benefits of the embedding in continuous time. Indeed, the branching property applied to the fringe node process $(L(t))_t$, together with the asymptotics proved in Theorem \[asymptoticsCTL\], allow us to see the limit $W^{CT}$ as a solution of a distributional equation. This is detailed in Section \[sec-dislocation\]. It is the starting point to deduce several properties of $W^{CT}$: its distribution is the unique solution of such an equation in a convenient space of probability distributions (Theorem \[th-contraction\] in Section \[sec-smoothing\]); it admits exponential moments in a neighborhood of $0$ (Theorem \[Expmoments\] in Section \[sec-cascade\]); up to a change of function, its Laplace transform is a solution of the quite simple (but unsolvable!) differential equation $y^{(m)}=y^2$ (Theorem \[equaDiffLaplaceCT\] in Section \[sec-laplace\]); it admits a density relatively to Lebesgue measure on $\g C$ and its support is the whole complex plane (Theorem \[densiteSupportWCT\] in Section \[sec-density\]). Thanks to connection between $W^{CT}$ and $W^{DT}$, corresponding results are true for $W^{DT}$ and consequently for $W_m^{\rm B-tree}$. Dislocation equations in continuous time {#sec-dislocation} ---------------------------------------- In this section, using the branching property of the continuous-time process $\left( L(t)\right)_t$, we show that the complex-valued random variable $W^{CT}$ is solution of a very simple distributional equation. In order to simplify the notations, for any $k\in\{ 1,\dots ,m\}$, denote by $W_k$ the limit random variable $W^{CT}$ (or its distribution) of the continuous-time fringe node process $\left( L(t)^{e_k}\right)_t$ that starts with one particle of type $k$, which means that its initial composition $L(0)$ is the $k$-th vector $e_k$ of $\g R^m$ canonical basis. Denote also by $\tau _k$ the *first* splitting time of the process $\left( L(t)^{e_k}\right)_t$; its is exponentially distributed, with parameter $m+k-1$. Because of the branching property of the process $\left( L(t)\right)_t$, for any time $t\geq \tau _1$, the processes $\left( L(t)^{e_1}\right)_{t\geq 0}$ and $\left( L(t)^{e_2}\right)_{t\geq 0}$ are related by the distributional equation $$L(t)^{e_1}\egalLoi L(t-\tau _1)^{e_2}.$$ In the asymptotic form given by Theorem \[asymptoticsCTL\], consider the second order term on both sides of the equality, which consists in projecting, normalizing and letting $t$ tend to infinity. This leads to the distributional equality $$W_1=e^{-\lambda _2\tau _1}W_2,$$ the random variables $W_2$ and $\tau _1$ being independent. Doing the same for all values of $k\in\{1,\dots ,m\}$ leads to the distributional system: $$\label{systLoiWCT} \left\{ \begin{array}{ccl} W_1&\egalLoi&e^{-\lambda _2\tau _1}W_2 \\ W_2&\egalLoi&e^{-\lambda _2\tau _2}W_3 \\ &\vdots& \\ W_{m-1}&\egalLoi&e^{-\lambda _2\tau _{m-1}}W_m \\ W_m&\egalLoi&e^{-\lambda _2\tau _m}\left( W_1^{(1)}+W_1^{(2)}\right) \end{array} \right.$$ where $\bullet$ for any $k\in\{ 1,\dots ,m-1\}$, the random variables $\tau _k$ and $W_{k+1}$ of the $k$-th equation’s right-hand sides are independent; $\bullet$ in the right-hand side of the last equation, the random variables $W_1^{(1)}$ and $W_1^{(2)}$ are independent copies of $W_1$, both being independent of $\tau _m$ as well. We recall that for any $k\in\{ 1,\dots ,m\}$, the random variable $\tau _k$ is exponentially distributed, with parameter $m+k-1$ (see Section \[sec-embedding\]). In particular, $W_1$ is a solution of the following distributional equation, sometimes called fixed point equation or smoothing equation in some branching processes contexts (see Liu [@Liu98] or Biggins and Kyprianou [@BigKyp05]). $$\label{eqLoiWCT} W_1\egalLoi B^{\lambda _2}\left( W_1^{(1)}+W_1^{(2)}\right)$$ where $\bullet$ the random variables $W_1^{(1)}$ and $W_1^{(2)}$ are independent copies of $W_1$; $\bullet$ $B$ is a random variable, independent of $W_1^{(1)}$ and $W_1^{(2)}$, Beta distributed with parameters $(m,m)$ which means that it admits $t^{m-1}(1-t)^{m-1}{\leavevmode\hbox{\rm \small1\kern-0.35em\normalsize1}}_{[0,1]}(t)$ as a density (${\leavevmode\hbox{\rm \small1\kern-0.35em\normalsize1}}_A$ denotes the indicatrix function of the set $A$). The distribution of $B$ is computed the following way. By immediate computation from System , one sees that $B=e^{-\left( \tau _1+\tau _2+\dots +\tau _m\right)}$, the variables $\tau _k$ being mutually independent. To recognize the Beta$(m,m)$ law, one can make a direct computation of its density or compute its moments (a Beta distribution is characterized by its moments because its support is compact). Smoothing equation in discrete time ----------------------------------- In a general setting of $m$-color Pólya urns, including the case of negative entries on the diagonal of the replacement matrix, Mailler [@Mailler] proves that $W^{DT}$ is a solution of a distributional equation which turns to be in our case $$\label{eqLoiWDT} W \ \egalenloi \ B_1^{ \lambda_2 }W^{(1)} + B_2^{ \lambda_2 }W^{(2)} ,$$ where $\bullet$ the random variables $W^{(1)}$ and $W^{(2)}$ are independent copies of $W$; $\bullet$ $(B_1,B_2)$ is a random vector, independent of $W^{(1)}$ and $W^{(2)}$, Dirichlet distributed with parameters $(m,m)$, which means that $B_1+B_2 = 1$ and that $B_1$ and $B_2$ are Beta distributed with parameters $(m,m)$. One proof of this result in [@Mailler] uses the tree structure of the urn. Nevertheless, we do not actually understand what kind of “divide-and-conquer” type argument, applied to B-trees, could lead to this equation. Indeed, in other cousin models, like $m$-ary search trees (see Fill and Kapur [@FillKapur]), such a backward decomposition leads to a finite time decomposition equation and passing to the limit, it gives the distributional equation. Contraction methods {#sec-smoothing} ------------------- The question of existence and unicity of solutions of equations like or is classically solved using the Banach fixed point theorem. One point of view, frequent in analysis of algorithms, consists in starting from a decomposition property of the algorithm at finite time, deduce a distributional equation on a cost variable, and pass to the limit to get a smoothing equation on the limit random variable. See Knape and Neininger [@KnaNei] for Pólya urns, and also the general paper by Neininger and Rüschendorf [@NeiRusaap] or their survey [@NeiRusSurvey] for many examples of this so-called contraction method. Another point of view (in this article) consists in taking advantage of the dynamics of the algorithm and exhibiting a martingale limit, solution of a smoothing equation. Thus, the existence is automatically achieved. In both points of view, to get the unicity, the contraction property has to be established, in a convenient space of probability distributions, classically equipped with a Wasserstein distance to get a complete metric space of measures. We do not prove here the theorem below, since it is done in a general frame by Mailler [@Mailler]. The same kind of results can be found in Janson [@Jan proof of Th 3.9 (iii)] and in Knape and Neininger [@KnaNei] even if the only case $a_{i,i}\geq -1$ is considered there. See also [@ChaLiuPouContinu; @ChaLiuPouDiscret]. \[th-contraction\] When $A$ is a complex number, let $\rond M_2\left( A\right)$ be the space of probability distributions on $\g C$ that have $A$ as expectation and a finite second moment, endowed with a complete metric space structure by the Wasserstein distance. Let $\lambda\in\C$ be any root of the characteristic polynomial (\[polycar\]) such that $\Re(\lambda) >\frac 12$. Then, - Each of the two equations $$W \ \egalenloi \ B_1^{ \lambda }W^{(1)} + B_2^{ \lambda }W^{(2)}$$ where $W^{(1)}$ and $W^{(2)}$ are independent copies of $W$, and where $(B_1,B_2)$ is a random vector, independent of $W^{(1)}$ and $W^{(2)}$, Dirichlet distributed with parameters $(m,m)$, and $$W\egalLoi B^{\lambda }\left( W^{(1)}+W^{(2)}\right)$$ where $W^{(1)}$ and $W^{(2)}$ are independent copies of $W$, and where $B$ is independent of $W^{(1)}$ and $W^{(2)}$, Beta distributed with parameters $(m,m)$, have a unique solution in $\rond M_2\left( A\right)$. - For $m\geq 60$, the variable $W_m^{\rm B-tree}$, defined in Corollary \[asymptoticsBtreeL\], is the unique solution of having $\displaystyle\frac{m!}{\Gamma\left( m+\lambda _2\right)}$ as expectation and a finite second moment. - For $m\geq 60$, the variable $\rond W^{CT}$, defined in , is the unique solution of having $m$ as expectation and a finite second moment. Cascades and exponential moments {#sec-cascade} -------------------------------- Let $\lambda\in\C$ be any root of the characteristic polynomial (\[polycar\]) such that $\Re(\lambda) >\frac 12$ and let $B$ be a Beta distribution with parameters $(m,m)$. A simple computation leads to $2\g E\left(B^{\lambda}\right)=1$. This is coherent with equation $$\label{eqLoiWCTgeneral} W\egalLoi B^{\lambda }\left( W^{(1)}+W^{(2)}\right).$$ Moreover, for any positive real $s$, $$2\g E\left( B^s\right) = \frac{(2m)\dots (m+1)}{(2m-1+s) \dots (m+s)}<1\Longleftrightarrow s>1.$$ Consequently, when $2\Re (\lambda) >1$, one has $$\label{coeffcontraction} 2\g E\left( |B^{\lambda}|^2\right) <1.$$ Theorem \[Expmoments\] below states that any solution $W$ of Equation admits exponential moments in a neighbourhood of $0$, so that the moment exponential generating series of $W$ defines an analytic function in a neighbourhood of the origin. Another consequence is that the law of $W$ is determined by its moments. The proof relies on a Mandelbrot’s cascade here defined in a complex setting (see Barral et al. [@Ba10] for complex Mandelbrot’s cascades). To lighten the notations, denote for a while $A:=B^{\lambda}$ and let $A_u, u\in U$ be independent copies of $A$, indexed by all finite sequences of $0$ and $1$: $$u = u_1\dots u_n \in U:= \bigcup_{n\geq 1} \{ 0, 1\}^n.$$ Let $Y_0=m$, $Y_1 = 2mA$ and for $n\geq 2$, $$Y_n = \sum_{u_1\dots u_{n-1} \in \{0,1\}^{n-1}} 2mAA_{u_1} A_{u_1u_2} \dots A_{u_1 \dots u_{n-1}}.$$ By the branching property, and using $2\g EA=1$, it is easy to see that $(Y_n)_n$ is a martingale with expectation $m$. This martingale has been studied by many authors in the real-valued random variable case, especially in the context of Mandelbrot’s cascades, see for example Liu $\cite{Liu01}$ and the references therein. It can be easily seen that $$\label{EqYn} Y_{n+1} = B^{\lambda} \left( Y_n^{(1)} + Y_n^{(2)} \right)$$ where $Y_n^{(1)}$ and $Y_n^{(2)}$ are independent of each other and independent of $B^{\lambda}$ and each has the same distribution as $Y_n$. Therefore for $n\geq 1$, $Y_n$ is square-integrable and $$\Var Y_{n+1} = 2\g E|B^{\lambda}|^2 \Var Y_n + (4\g E |B^{\lambda}|^2 - 1)$$ where $\Var X = \g E\left( |X-\g EX|^2\right)$ denotes the variance of $X$. Since $2\g E|B^{\lambda}|^2 < 1$, the martingale $(Y_n)_n$ is bounded in $L^2$, so that the following result holds. \[YnL2\] Let $\lambda\in\C$ be any root of the characteristic polynomial (\[polycar\]) such that $\Re(\lambda) >\frac 12$ and let $B$ be a Beta distribution with parameters $(m,m)$. When $n\rightarrow + \infty$, $$Y_n \rightarrow Y_\infty \mbox{ a.s. and in } L^2,$$ where $Y_\infty$ is a (complex-valued) random variable with variance $$\Var (Y_\infty) = \frac {4\g E |B^{\lambda}|^2 - 1}{1 - 2\g E|B^{\lambda}|^2 }.$$ Notice that, passing to the limit in (\[EqYn\]) gives a new proof of the existence of a solution $W$ of Eq. (\[eqLoiWCTgeneral\]) with a given expectation and finite second moment whenever $\Re (\lambda) >1/2$. From Section \[sec-smoothing\], we have the uniqueness of solution of this equation so that Theorem \[Expmoments\] below will be proved as soon as it holds for $Y_\infty$. \[expmomentsY\] There exist some constants $C>0$ and $\varepsilon >0$ such that for all $t \in \C$ with $|t| \leq \varepsilon $, we have $$\label{bound-phi-inf} \g E e^{\langle t, Y_\infty \rangle } \leq e^{m\Re (t) + C |t|^2 }.$$ [[Proof.]{} ]{}By Fatou lemma, it is sufficient to prove the existence of $C>0$ and $\varepsilon >0$ such that for all $t \in \C$ with $|t| \leq \varepsilon $, and for every integer $n$, $$\label{bound-phi-n} \g E e^{\langle t, Y_n \rangle } \leq e^{m\Re (t) + C |t|^2 }.$$ Denote $\varphi_n(t):=\g E e^{\langle t, Y_n \rangle }$ and notice that $\varphi_{n+1}(t) = \g E\left( \varphi_n^2(tB^{\overline{\lambda}})\right)$ thanks to Equation , allowing to prove by recursion on $n\geq 0$. For $n=0$, $$\varphi_0(t):=\g E e^{\langle t, Y_0 \rangle } = e^{m\Re(t)}$$ and by the recursion assumption, $$\varphi_{n+1}(t)\leq \g E\left( e^{2C|t|^2 |B^{\lambda}|^2 + 2m\Re\left( tB^{\overline{\lambda}}\right)}\right)= e^{m\Re (t) + C |t|^2 }f(t_1,t_2)$$ where for any $t \in \C$, written $t = t_1+it_2$ with $t_1,t_2\in\g R$, $$f(t_1,t_2) = \g E\left( e^{C|t|^2 (2|B^{\lambda}|^2-1) + 2m\Re\left( tB^{\overline{\lambda}}\right) -m\Re(t)}\right),$$ so that it is sufficient to prove that $(0,0)$ is a local maximum of $f$. Writing $\lambda = \sigma + i\tau$, with $\sigma, \tau\in\g R$, $$f(t_1,t_2) = \g E\exp\left[ C(t_1^2 + t_2^2)(2B^{2\sigma}-1) + 2m B^{\sigma}(t_1\cos(\tau) + t_2\sin(\tau)) -mt_1\right].$$ Remembering $2\g E\left(B^{\lambda}\right)=1$, which means $2\g E\left( B^{\sigma}\cos(\tau)\right)=1$ and $\g E\left( B^{\sigma}\sin(\tau)\right)=0$, we get that the first derivatives vanish at $(0,0)$ which is a critical point. Moreover, the calculation of the second partial derivatives gives $$\begin{aligned} \frac{\partial^2 f}{\partial t_1} (0,0) &=& \g E \left[ \left(2mB^{\sigma}\cos(\tau)-m\right)^2+ 2C \left(2B^{2\sigma}-1 \right)\right], \\ \frac{\partial^2 f}{\partial t_2} (0,0)&= &\g E \left[ \left(2mB^{\sigma}\sin(\tau)\right)^2+ 2C \left(2B^{2\sigma}-1 \right)\right] ,\\ \frac{\partial^2 f}{\partial t_1 \partial t_2} (0,0) &=& \g E \left(2mB^{\sigma}\cos(\tau)-m\right)\left(2mB^{\sigma}\sin(\tau)\right).\end{aligned}$$ By , $\g E\left(2B^{2\sigma}-1 \right) <0$, so that the Hessian matrix at $(0,0)$ is definite negative for $C >0$ large enough which implies that $(0,0)$ is a local maximum of $f$. ${\sqcap\!\!\!\!\sqcup}$ The following theorem is a direct consequence of Lemma \[expmomentsY\], like in [@ChaLiuPouContinu]. \[Expmoments\] Let $\lambda \in \g C$ be a root of the characteristic polynomial (\[polycar\]) with $\Re (\lambda) >1/2$ and let $W$ be a solution of Eq. (\[eqLoiWCTgeneral\]). There exist some constants $C>0$ and $\varepsilon >0$ such that for all $t\in \C$ with $ |t| \leq \varepsilon$, $$\label{bdExpmoments} \g Ee^{ \langle t,W\rangle } \leq e^{ m\Re (t) + C |t|^2 } \; \mbox{ and } \; \g Ee^{ |tW | } \leq 4 e^{ m|t| + 2C |t|^2 } .$$ Laplace transform {#sec-laplace} ----------------- Theorem \[Expmoments\] above concerning $W^{CT}$ and Theorem 7 (ii) in Mailler [@Mailler] (which establishes that the Laplace series of $W^{DT}$ has an infinite radius of convergence) answer the question of the convergence of the Laplace series of $W^{DT}$ and $W^{CT}$. Nevertheless, a natural investigation consists in searching more information about these Laplace transforms coming from the smoothing equations. Indeed, the dislocation equations lead to a system of differential equations on the Laplace transforms $$\label{defLaplaceWCT} \forall k = 1, 2, \dots , m, \hskip 1cm \varphi_k(z):= \g E\left( e^{<z,W_k>}\right),$$ where we recall that $W_k$ is the limit random variable $W^{CT}$ of the continuous-time fringe node process $\left( L(t)^{e_k}\right)_t$ that starts with one particle of type $k$. Using the independence between the splitting times $\tau_k$ (which are exponentially distributed, with parameter $m+k-1$) and the $W_k$, for $k = 1, 2, \dots , m-1$, $$\varphi_k(z) = \int_0^{+\infty} \varphi_{k+1}\left(ze^{-\overline{\lambda_2} t}\right)(m+k-1)e^{-(m+k-1)t}dt,$$ and after a change of variable, and derivation, for $k = 1, 2, \dots , m-1$, $$\frac{m+k-1}{\overline{\lambda_2}}\varphi_k(z) +z\varphi_k'(z) = \varphi_{k+1}(z),$$ and for $k=m$, $$\frac{2m-1}{\overline{\lambda_2}}\varphi_m(z) +z\varphi_m'(z) = \varphi_{1}^2(z).$$ Thanks to a convenient change of function, a simple calculation gives the following theorem about the Laplace transforms of the $W_k$, for $k = 1, 2, \dots , m$. \[equaDiffLaplaceCT\] For $k = 1, 2, \dots , m$, let $\varphi_k$ be the Laplace transform of $W_k$ defined in , and let $$\psi_k(z):= \left( -\overline{\lambda_2} \right)^{m+k-1} \ \frac{\varphi_k\left( z^{ -\overline{\lambda_2} }\right)}{z^{m+k-1}},$$ (for any determination of the logarithm). Then the functions $\psi_k$ satisfy the simple differential system $$\left\{ \begin{array}{l} \displaystyle \psi'_k=\psi_{k+1}, \hskip 1cm \forall k\in\{1,\dots ,m-1\}, \\ \displaystyle \psi'_{m}=\psi_1^2. \end{array} \right.$$ In particular, $\psi_1$ is a solution of the differential equation $$y^{(m)}=y^2.$$ Density and support {#sec-density} ------------------- Liu’s method has been developped in [@Liu99] and [@Liu01] for positive real-valued random variables solution of smoothing equations of the same type as or . Adapting this method to $\g C$-valued random variables, Mailler [@Mailler] gets the support and the existence of a density for the limit law of a $d$-color Pólya urn. The theorem below is a particular case. \[densiteSupportWCT\] Let $m\geq 2$. Let $V\in\g N^m$ be a non zero vector and let $W^{DT}$ and $W^{CT}$ be the $W$-distributions of the respective discrete-time and continuous-time fringe node processes having $V$ as initial composition (see (ii) in Theorem \[asymptoticsDTL\] and Theorem \[asymptoticsCTL\]). Then \(i) the supports of $W^{DT}$ and $W^{CT}$ are both the whole complex plane $\g C$; \(ii) $W^{DT}$ and $W^{CT}$ are absolutely continuous relatively to Lebesgue’s measure on $\g C$; \(iii) as $|t| \rightarrow \infty$, $ \g Ee^{i\langle t,W^{DT}\rangle } = O(|t|^{-a})$ for each $a\in\left] 0, \displaystyle\frac m{\Re (\lambda_2)}\right[$, and the same is true for the Fourier transform of $W^{CT}$ as well. Perspectives {#sec-perspectives} ------------ Some open questions remain about $W^{DT}$ and $W^{CT}$, let us say $W$: - can the $W$ distribution be expressed by means of usual distributions? Same question for $|W|$ and $\Arg(W)$? - how heavy are the tails of $W$? - what is the order of magnitude of $W$’s $p$-th moment as $p$ tends to $+\infty$? Simulations {#sec-simulations} =========== Let us here summarize and illustrate the asymptotic results concerning the fringe of a random B-trees. We only show the behaviour of the gap process $(G_n)_n$ and illustrate Theorem \[asymptoticsDTG\]; nevertheless, the simulations would be analogous for the fringe node process $(\rond L_n)_n$ to illustrate Corollary \[asymptoticsBtreeL\]. In all the simulations below, sequences of $10^7$ random keys have been inserted in a B-tree for different value of the parameter $m$. Notice that in “real life” computer science implementations, $m$ is most of the time taken around $100$ or more. Simulations of $G_n$ {#sec-simulationsGn} -------------------- Figures \[fig:mpetit\] and \[fig:mgrand\] represent the trajectories of three coordinates of the random vector $G_n$: for any given value $m=10$, $30$, $55$, $65$, $100$ or $237$ of the parameter, we make one random drawing of a sequence of $10^7$ keys and insert them in a B-tree. On the pictures, the $x$-axis represents the time $n\in\{ 0,\dots ,10^7\}$ while the $y$-axis represents the number $G_n^{(k)}$ of gaps of type $k$ for $k=1$, $\lfloor m/2\rfloor$ and $m$. In each case, the picture illustrates the almost sure asymptotics $G_n\sim nv_1$ when $n$ tends to infinity (remember that $v_1$ is a non random $m$-dimensional vector). In Figure \[fig:mpetit\], $m$ is *small* ($m=10$, $30$, $55$). One can already catch sight of the gaussian fluctuations around the deterministic vector $nv_1$. Notice that the variance of the gaussian limit increases with $m$, so that the amplitude of the fluctuation becomes more visible for $m=30$ and even more for $m=55$. (600,150) (-10,0)[![\[fig:mpetit\] Simulations for $3$ coordinates of the gap process $\left( G_n\right) _n$ for *small* $m$.](gnm10-eps-converted-to.pdf "fig:"){height="45"}]{} (30,0)[$m=10$]{} (145,0)[![\[fig:mpetit\] Simulations for $3$ coordinates of the gap process $\left( G_n\right) _n$ for *small* $m$.](gnm30-eps-converted-to.pdf "fig:"){height="45"}]{} (185,0)[$m=30$]{} (300,0)[![\[fig:mpetit\] Simulations for $3$ coordinates of the gap process $\left( G_n\right) _n$ for *small* $m$.](gnm55-eps-converted-to.pdf "fig:"){height="45"}]{} (340,0)[$m=55$]{} In Figure \[fig:mgrand\], $m$ is *large* ($m=65$, $100$, $237$). On can see the almost sure oscillations around $nv_1$ appear and become more visible when $m$ grows. Notice that they are particularly clear for $m=237$, which is the threshold value when the *third* largest real part of the roots of $\chi_m$ becomes larger than $\frac 12$. See the Appendix for more details. (600,150) (-10,0)[![\[fig:mgrand\] Simulations for $3$ coordinates of the gap process $\left( G_n\right) _n$ for *large* $m$.](gnm65-eps-converted-to.pdf "fig:"){height="45"}]{} (30,0)[$m=65$]{} (145,0)[![\[fig:mgrand\] Simulations for $3$ coordinates of the gap process $\left( G_n\right) _n$ for *large* $m$.](gnm100-eps-converted-to.pdf "fig:"){height="45"}]{} (185,0)[$m=100$]{} (300,0)[![\[fig:mgrand\] Simulations for $3$ coordinates of the gap process $\left( G_n\right) _n$ for *large* $m$.](gnm237-eps-converted-to.pdf "fig:"){height="45"}]{} (340,0)[$m=237$]{} Of course, one can make similar graphs for trajectories of the vector $\frac{G_n}n$ which converges to the deterministic vector $v_1$. This is done in Figure \[fig:deriveGn\] where the convergence can be seen on the three drawn coordinates. Once more, the fluctuations around the limit $v_1$ are of different nature depending on $m\leq 59$ or $m\geq 60$, which is also illustrated on this figure. In particular, on can see the “$\cos\log n$” almost sure oscillations arise when $m\geq 60$ and become more evident when $m$ increases. (600,310) (-10,160)[![\[fig:deriveGn\] Simulations of $3$ coordinates of $\frac{G_n}n$ for *small* and *large* values of $m$.](gnsurnm10-eps-converted-to.pdf "fig:"){height="45"}]{} (30,160)[$m=10$]{} (145,160)[![\[fig:deriveGn\] Simulations of $3$ coordinates of $\frac{G_n}n$ for *small* and *large* values of $m$.](gnsurnm30-eps-converted-to.pdf "fig:"){height="45"}]{} (185,160)[$m=30$]{} (300,160)[![\[fig:deriveGn\] Simulations of $3$ coordinates of $\frac{G_n}n$ for *small* and *large* values of $m$.](gnsurnm55-eps-converted-to.pdf "fig:"){height="45"}]{} (340,160)[$m=55$]{} (-10,0)[![\[fig:deriveGn\] Simulations of $3$ coordinates of $\frac{G_n}n$ for *small* and *large* values of $m$.](gnsurnm65-eps-converted-to.pdf "fig:"){height="45"}]{} (30,0)[$m=65$]{} (145,0)[![\[fig:deriveGn\] Simulations of $3$ coordinates of $\frac{G_n}n$ for *small* and *large* values of $m$.](gnsurnm100-eps-converted-to.pdf "fig:"){height="45"}]{} (185,0)[$m=100$]{} (300,0)[![\[fig:deriveGn\] Simulations of $3$ coordinates of $\frac{G_n}n$ for *small* and *large* values of $m$.](gnsurnm237-eps-converted-to.pdf "fig:"){height="45"}]{} (340,0)[$m=237$]{} Simulations of $G_n$ after scaling {#sec-simulationsGnScaling} ---------------------------------- A second kind of simulations focus on the possible scalings of the centered gap process $\left( G_n-nv_1\right) _n$. In order to get convergence, according to Theorem \[asymptoticsDTG\], one has to divide $G_n-nv_1$ by $\sqrt n$ when $m\leq 59$ and by $n^{\sigma _2}$ when $m\geq 60$. Figures \[fig:mpetitScaling\] and \[fig:mgrandScaling\] represent trajectories of the median coordinate (the $\lfloor m/2\rfloor$-th) of the normalized vector process. Hereunder, $X_n$ denotes this median coordinate $X_n=G_n^{\lfloor m/2\rfloor}$. Figure \[fig:mpetitScaling\] deals with *small* values of $m$, namely $m=10$, $30$, $55$ again. On the $x$-axis, time $n\in\{ 0,\dots , 10^7\}$ ; on the $y$-axis, the normalized coordinate $\displaystyle\frac{X_n-nv_1^{\lfloor m/2\rfloor}}{\sqrt n}$ which converges in distribution to a normal law. Note that even if the random vector $\displaystyle\frac{G_n-nv_1}{\sqrt n}$ converges in distribution, it almost surely diverges, which is illustrated by its brownian-like trajectory. One can refer to Gouet [@Gouet93] for more details on this continuous type process limit. (600,140) (-10,0)[![\[fig:mpetitScaling\] Simulations of one coordinate of $G_n$ after normalisation, for *small* values of $m$.](gnNorm10-eps-converted-to.pdf "fig:"){height="45"}]{} (30,-5)[$m=10$]{} (145,0)[![\[fig:mpetitScaling\] Simulations of one coordinate of $G_n$ after normalisation, for *small* values of $m$.](gnNorm30-eps-converted-to.pdf "fig:"){height="45"}]{} (185,-5)[$m=30$]{} (300,0)[![\[fig:mpetitScaling\] Simulations of one coordinate of $G_n$ after normalisation, for *small* values of $m$.](gnNorm55-eps-converted-to.pdf "fig:"){height="45"}]{} (340,-5)[$m=55$]{} Figure \[fig:mgrandScaling\] deals with $m=65$, $100$, $237$ which are *large* values of $m$. On the $y$-axis: the normalized coordinate $\displaystyle\frac{X_n-nv_1^{\lfloor m/2\rfloor}}{n^{\sigma _2}}$, which is almost surely equivalent to some $\rho\cos\left( \tau _2\log n+\varphi\right)$ when $n$ tends to infinity, where $\rho$ is a positive random variable (random amplitude), $\varphi$ a $[0,2\pi [$-valued random variable (random phase) and $\tau _2$ the imaginary part of the complex eigenvalue $\lambda _2=\sigma _2+i\tau _2$. The random variables $\rho$ and $\varphi$ are proportional to the module and the argument of the complex-valued random variable $W_m$ in Theorem \[asymptoticsDTG\]. (600,140) (-10,0)[![\[fig:mgrandScaling\] Simulations of one coordinate of $G_n$ after normalization, for *large* values of $m$.](gnNorm65-eps-converted-to.pdf "fig:"){height="45"}]{} (30,0)[$m=65$]{} (145,0)[![\[fig:mgrandScaling\] Simulations of one coordinate of $G_n$ after normalization, for *large* values of $m$.](gnNorm100-eps-converted-to.pdf "fig:"){height="45"}]{} (185,0)[$m=100$]{} (300,0)[![\[fig:mgrandScaling\] Simulations of one coordinate of $G_n$ after normalization, for *large* values of $m$.](gnNorm237-eps-converted-to.pdf "fig:"){height="45"}]{} (340,-5)[$m=237$]{} Appendix. The phase transition and $\sigma _2(m)$ {#appendix} ================================================= The phase transition that occurs for B-trees with parameter $m$ relies on the roots of the characteristic polynomial $$\chi _m(X)= \prod _{k=m}^{2m-1}(X+k)-\frac{(2m)!}{m!}. $$ Denote by $\lambda _2=\lambda _2(m)$ the root of $\chi _m$ having the second largest real part and a positive imaginary part. Denote also $\sigma _2=\sigma _2(m)$ the real part of $\lambda _2(m)$. As shown in Section \[sec-transition\], the B-tree admits a Gaussian central limit theorem when $m\leq 59$ (small régime) whereas it admits an almost sure nonnormal fluctuation term of order $n^{\sigma _2(m)}$ around the drift when $m\geq 60$ (large régime). Coming from Pólya urn theory, this asymptotic behaviour depends on whether $\sigma _2(m)<1/2$ (small régime) or $\sigma _2(m)>1/2$ (large régime). Let $F$ the two-variable meromorphic function defined by $$F(x,y)= \frac {\Gamma\left(x+2y\right)\Gamma\left( 1+y\right)} {\Gamma\left( 1+2y\right)\Gamma\left( x+y\right)}$$ where $\Gamma$ denotes Euler’s Gamma function. For a given $m\geq 2$, $\lambda _2=\lambda _2(m)=\sigma _2+i\tau _2$ is the root of equation $F\left( X,m\right) =1$ having the the second largest real part $\sigma _2$ (the first one being reached by the evident root $1$) and a positive imaginary part $\tau _2$. Denote by $\psi$ the classical Digamma function, the logarithmic derivative of Euler’s Gamma. Since $\frac{\partial}{\partial x}F(x,1/y)=\psi\left( x+2/y\right) -\psi\left(x+1/y\right) =\log 2+O(y)$ as $y$ tends to $0$, the analytic implicit function theorem shows that $\lambda _2(m)$ is an analytic function of $1/m$ as $m$ tends to $+\infty$. Using the expansion $$\log\Gamma (z)=z\log z-z-\frac 12\log z+\frac 12\log 2\pi+O\left(\frac 1z\right) ~{\rm mod}~2i\pi$$ as $|z|$ tends to infinity (the *mod* coming from the determination of the logarithm), writing $\lambda _2(m)$ as a power series in $1/m$ and putting the first terms of this expansion of in the equation $\log F\left(\lambda _2(m),m\right)=0~{\rm mod~2i\pi}$, one gets by identification $$\left\{ \begin{array}{l} \displaystyle \sigma _2(m)=1-\frac{\pi ^2}{\log ^32}\times\frac 1m+O\left(\frac 1{m^2}\right) \\ \\ \displaystyle \tau _2(m)=\frac{2\pi}{\log 2}+\frac{\pi}{2\log ^22}\times\frac 1m+O\left(\frac 1{m^2}\right) \end{array} \right.$$ as $m$ tends to infinity. The graph of the function $m\mapsto\sigma _2(m)$ is given in Figure Ê\[fig-sigma\]. Numerical values of the expansions give $\sigma _2\approx 1-29.63/m+\dots$ while $\tau _2\approx 9.06+3.27/m+\dots$ ![\[fig-sigma\] The graph of the sequence $m\mapsto\sigma _2(m)$.](courbeSigma2Debut.pdf "fig:"){height="210pt"} ![\[fig-sigma\] The graph of the sequence $m\mapsto\sigma _2(m)$.](courbeSigma2.pdf "fig:"){height="210pt"} Moreover, the numerical values of $\sigma _2$ around $m=60$ are the following ones, showing more accurately that $\sigma _2(m)<1/2$ if, and only if $m\leq 59$. These numerical values have been computed applying the Newton method to the $\chi _m$ function starting from the point $0.5+9.0i$, as suggested by the above expansions of $\sigma _2$ and $\tau _2$. $m$ $\sigma _2(m)$ ----- ---------------- 57 0.4775726941 58 0.4866133472 59 0.4953467200 60 0.5037882018 61 0.5119521623 62 0.5198520971 In order to justify the choice of $m=237$ in our drawings, denote by $\sigma _3(m)$ the third largest real part of the roots of $\chi _m$. The threshold value when $\sigma _3(m)$ becomes larger than $\frac 12$ is $m=237$. Using general statements on Pólya urns, this shows that a second almost sure phenomenon with magnitude $n^{\lambda _3}$ is added to the one we describe in Theorem \[asymptoticsDTG\] as soon as $m\geq 238$. That is the reason why the above figures have been selected for $m=237$; indeed, for $m$ increasing from $60$ until $237$, the asymptotic expansion of $G_n$ contains the oscillating term of amplitude $n^{\sigma _2(m)}$ more and more visible compared to brownian terms in $n^{\frac 12}$. For $m>237$, the second oscillating term of amplitude $n^{\sigma _3(m)}$ appears, making the $n^{\sigma _2(m)}$ oscillation less visible. The numerical values of $\sigma _3(m)$ around $m=237$ are the following ones. $m$ $\sigma _3(m)$ ----- ---------------- 236 0.4971039325 237 0\. 4992277960 238 0.5013338161 239 0.5034221856 The authors kindly thank Karine Zeitouni for valuable discussions on B-trees and Andrea Sportiello for insightful comments. [^1]: [*2000 Mathematics Subject Classification.*]{} Primary: 60J80. Secondary: 68W40, 68Q87. [*Key words and phrases.*]{} B-tree. Fringe analysis. Pólya urn. Urn model. Martingale. Multitype branching process. Smoothing transforms. Contraction method. [^2]: A search tree is a tree where internal nodes contain sorted keys and where a node containing $k$ keys $x_1, \dots , x_k$ defines $k+1$ intervals such that, for $j=1, \dots , k+1$, the keys in the $j$-th subtree belong to the $j$-th interval. [^3]: The leaves, sometimes called external nodes, are the nodes without any descendant. [^4]: An empty entry stands for a zero in all the matrices of this article. [^5]: When no confusion is possible, we denote by PV the product of the square matrix $P$ by the vector $V\in\g R^m$ instead of the correct form $P\transp V$.
{ "pile_set_name": "ArXiv" }
![image](NCRA_logo.png){width="2cm" height="2.2cm"} **National Centre for Radio Astrophysics**\ **Tata Institute of Fundamental Research** Internal Technical Report: R1401 [ **The Polarization Convention of the uGMRT in Band 4\ ** ]{} [Barnali Das, Sanjay Kudale, Poonam Chandra, Bhaswati Bhattacharya, Jayanta Roy, Yashwant Gupta]{}\ *Corresponding Author’s email id:*\ April 18, 2020\ [**Objective: To check the polarization configuration of the uGMRT in band 4**]{}\ Introduction {#sec:intro} ============ In astronomy, the polarization of a celestial signal is expressed in terms of the four Stokes parameters: $I$, $Q$, $U$ and $V$; where $I$ is the total intensity, $Q$ and $U$ represent linearly polarized power and $V$ quantifies the amount of circular polarization. These four quantities are derived from the voltages (coming from the celestial object) recorded by two receivers in the telescope corresponding to two orthogonal polarizations: $X$ and $Y$, for telescopes with dipolar feeds; or $R$ and $L$, for telescopes with circular feeds. Because of the degeneracy involved in forming a set of orthogonal polarizations, a convention is adopted by the IAU to define a set of universal $X$ and $Y$: in this convention, $X$ and $Y$ point towards the north and east respectively and the $+Z$ axis is the direction along which the radiation propagates from the source towards the observer [e.g. @hamaker1996]. In this convention, $R$ and $L$ are related to $X$ and $Y$ as $R=(X+iY)/\sqrt{2}$, $L=(X-iY)\sqrt{2}$, $i=\sqrt{-1}$. The four Stokes parameters are then obtained using: $$\begin{aligned} I&=XX^*+YY^*\equiv RR^*+LL^*\label{eq:I}\\ Q&=XX^*-YY^*\equiv RL^*+R^*L\label{eq:Q}\\ U&=XY^*+X^*Y\equiv -i(RL^*-R^*L)\label{eq:U}\\ V&=-i(XY^*-X^*Y)\equiv RR^*-LL^*\label{eq:V}\end{aligned}$$ The asterisk implies complex conjugate. Note that these four equations are w.r.t. the incoming radiation (and not what is received by the feed since the polarization state ‘seen’ by the feed can be different from the intrinsic state if the feed is at the primary or tertiary focus of the antenna dish). Out of these four equations, there is a lot of confusion about the fourth equation that defines Stokes $V$. After defining $X$ and $Y$ according to the IAU convention and taking $R=(X+iY)/\sqrt{2}$, $L=(X-iY)\sqrt{2}$; $RR^*$ and $LL^*$ correspond to right and left hand circular polarization respectively. A right hand circularly polarized radiation is the one where for the incoming radiation, the electric field vector rotates counter-clockwise. Similarly, for left hand circularly polarized radiation, the electric field vector rotates clockwise for the incoming radiation [@robishaw2018]. This way of defining right and left hand circular polarization corresponds to the IEEE convention, also adopted by the IAU. Stokes $V$ is then positive for right hand circularly polarized radiation under the “IAU/IEEE” convention. However the alternate way of defining Stokes V, i.e. $V=LL^*-RR^*$ is also prevalent, especially among the pulsar community. This convention, where Stokes $V$ is positive for left hand circularly polarized radiation, is known as the “PSR/IEEE” convention. In this report, we aim to find out the polarization convention adopted by the upgraded Giant Metrewave Radio Telescope [uGMRT, @gupta2017] in band 4 (550–900 MHz). The feeds in this band are dipolar; however these are converted to circular polarizations and the final products that an observer can get are $RR^*$, $LL^*$ and the cross terms of $R$ and $L$. The Stokes parameters can then be obtained using Eq.s \[eq:I\] to \[eq:V\]. This report is structured as follows: in §\[sec:strategy\], we describe the strategy adopted to perform the experiment; this is followed by observation and data analysis (§\[sec:data\_analysis\]). We present our results in §\[sec:results\], and then compare those with the conventions adopted for the Very Large Array (VLA) telescope in §\[sec:vla\_comparison\]. We present our conclusion in §\[sec:conc\]. Strategy {#sec:strategy} ======== The strategy that we adopted was to observe a source with known polarizations with the uGMRT and then compare the result with its known values. We choose the pulsar B1702–19 (RA: $17^h05^m36^s.099$ Dec: $-19^d06^m38^s.60$) as a test source to achieve this goal. This pulsar was chosen as it shows high circular polarization ($\approx 40\%$) according to the EPN database (<http://www.epta.eu.org/epndb/#gl98/J1705-1906/gl98_610.epn>) at 610 MHz. The Stokes $V$ sign is predominantly negative. This property is crucial if we want to detect circular polarization by imaging as it involves averaging over time. The convention used for Stokes $V$ was $V=LL^*-RR^*$ [obtained by @gould1998 using the Lovell telescope]. Comparison of the 1369 MHz Stokes V profile for PSR B1702–19 obtained by @johnston2017 using the Parkes radio telescope (explicitly mentioning the PSR/IEEE convention) with that obtained by @gould1998 using the Lovell telescope at 1408 MHz indicate that convention adopted by Lovell telescope is the PSR/IEEE version. This implies that at 610 MHz, the pulsar is $\approx 40\%$ right circularly polarized according to the IEEE convention. In Figure \[fig:epn\_profile\], we show the 610 MHz $I$, $Q$, $U$, $V$ profiles for this pulsar, generated from the data available in the EPN database [@gould1998]. To make it consistent with the rest of the paper, we have plotted $V=RR^*-LL^*$, in stead of the original convention of $V=LL^*-RR^*$. Observation and data analysis {#sec:data_analysis} ============================= We observed this pulsar in band 4 of the uGMRT on March 19, 2020 with 200 MHz bandwidth, divided into 2048 channles. 3C286 was used as the flux and bandpass calibrator, and J1822-096 (RA: $18^h22^m28.71^s$, Dec: $-09^d38^m56^s.84$) was used as the phase calibrator. The on-source time was around 40 minutes. The data were recorded both in interferometric and pulsar phased array mode. In the pulsar mode, the array was first phased using the nearby calibrator J1822-096. The data were then recorded on target in GMRT format (co- and cross-polar voltage products, viz. $RR^*$, $LL^*$, and real and imaginary parts of $RL^*$) with a time resolution of 327.68 microseconds ($\approx 0.33$ ms, the rotation period of the pulsar is $\approx 280$ ms). These data were analysed using the full polarization data analysis pipeline of the GMRT described in @kudale2008. $R$ and $L$ powers are separately scaled by its average bandshape to take out the bandpass effect. The final outputs were Stokes $I$, $Q$, $U$ and $V=RR^*-LL^*$, which are normalized by the maximum in total power (i.e. Stokes-I). Note that no polarization calibration was applied. The interferometric data were flagged and calibrated using standard tasks in $\textsc{casa}$ [e.g. @das2019]. The $RR^*$ and $LL^*$ data of the target were separately self-calibrated to make the final images. ![The 610 MHz profiles for the pulsar B1702–19 for Stokes $I$ (magenta curve), $Q$ (green curve), $U$ (blue curve) and $V$ (orange curve) obtained by @gould1998 using the Lovell telescope. Also shown are the angle of polarization (PA) on top panel and linear polarization $L=\sqrt{Q^2+U^2}$ in yellow in the bottom panel. Note that the convention for Stokes $V$, here, is $V=RR^*-LL^*$ even though the convention used in the original paper [@gould1998] was $V=LL^*-RR^*$. This is done to avoid confusion since GMRT convention for defining $V$ is $RR^*-LL^*$. For details, refer to §\[sec:strategy\].\[fig:epn\_profile\]](b1702_19_epn_profile_IQUV_PA_2.png){width="45.00000%"} Results and discussions {#sec:results} ======================= ![image](compare_ll_rr.png){width="40.00000%"} We present the results obtained from the data taken in interferometric mode and pulsar (beamformer) mode in the following two subsections. Results from interferometric mode {#subsec:interferometry_mode} --------------------------------- In the left panel of Figure \[fig:band4\_ll\_rr\], we show the result of the imaging exercise, where we plot the flux denities in $LL^*$ against the flux densities in $RR^*$ for the target (marked in red square) as well as a few other sources (marked in blue circles) in the field of view. This was done to make sure that the inferred difference in flux densities is not due to any systematics. We clearly see that for these test sources, the flux densities nearly follow the $y=x$ line, meaning that they have equal flux densities in $RR^*$ and $LL^*$. But the target is significantly offset from the $y=x$ line. Its $RR^*$ flux density came out to be $15.8\pm 0.1$ mJy, whereas the $LL^*$ flux density came out to be $23.9\pm 0.2$ mJy which give percentage circular polarization $V/I=-20\%$, while using $V=(RR^*-LL^*)/2$ and $I=(RR^*+LL^*)/2$. Thus, according to the uGMRT data, the pulsar is left hand circularly polarized at 610 MHz, whereas it is right hand circularly polarized under the IEEE convention. Therefore, effectively the uGMRT $RR^*$ is actually $LL^*$ in IEEE convention and vice-versa in band 4. Result from pulsar mode {#subsec:pulsar_mode} ----------------------- In the right panel of Figure \[fig:band4\_ll\_rr\], we show the pulse profiles for Stokes $I$, $Q$, $U$ and $V\,(=RR^*-LL^*)$. The maximum circular polarization observed is $-40\%$. By comparing these profiles with those in Figure \[fig:epn\_profile\], we find that: 1. The signs of the Stokes $Q$ profiles in the two figures do not match, but the Stokes $U$ profiles agree with each other. 2. The sign of the uGMRT Stokes $V$ profile does not match with that of the EPN Stokes $V$ profile. 3. The sweeps of the polarization angle (PA) are opposite in the two figures. The third point is a consequence of the first point. The second point suggests that the uGMRT $RR^*$ and $LL^*$ in band 4 are in opposite conventions to that defined by the IEEE (also suggested by the data taken in interferometric mode). However simply swapping $R$ and $L$ cannot explain why the sign of $Q$ is opposite to what should have been obtained according to the IEEE convention, whereas signs of $U$ match. To explain this behaviour, we refer to Eq. \[eq:Q\] and Eq.\[eq:U\]: $$\begin{aligned} Q=RL^*+R^*L,\quad U=-i(RL^*-R^*L)\nonumber\end{aligned}$$ Thus, swapping of $R$ and $L$ will give us opposite sign of $U$ and not $Q$. To explain the reversal of the sign of $Q$, we next consider the possibility that $X$ and $Y$ are swapped (§\[sec:intro\]). For clarity, let us denote uGMRT $R$ and $L$ as $\tilde{R}$ and $\tilde{L}$ respectively. If $X$ and $Y$ are swapped, we will get: $$\begin{aligned} \tilde{R}&=\frac{Y+iX}{\sqrt{2}}=i\frac{X-iY}{\sqrt{2}}=iL\\ \tilde{L}&=\frac{Y-iX}{\sqrt{2}}=-i\frac{X+iY}{\sqrt{2}}=-iR\end{aligned}$$ In that case, we will get $\tilde{V}=\tilde{R}\tilde{R^*}-\tilde{L}\tilde{L^*}=LL^*-RR^*$, so that the Stokes $V$ profile obtained with the uGMRT will match the one from EPN even though they have different conventions. For $Q$ and $U$, we will get: $$\begin{aligned} \tilde{Q}&=\tilde{R}\tilde{L}^*+\tilde{R}^*\tilde{L}=(iL)(iR^*)+(-iL^*)(-iR)=-(RL^*+R^*L)=-Q\\ \tilde{U}&=-i(\tilde{R}\tilde{L}^*-\tilde{R}^*\tilde{L})=-i\{(iL)(iR^*)-(-iL^*)(-iR)\}=-i(RL^*-R^*L)=U\end{aligned}$$ Thus, we infer that indeed the $X$ and $Y$ dipoles are swapped so that the resulting $RR^*$ and $LL^*$ are not in accordance with the IEEE convention in band 4 of the uGMRT. Comparison with the Very Large Array (VLA) telescope {#sec:vla_comparison} ==================================================== To compare the definitions of circular polarization for the uGMRT and the VLA, we again used the same pulsar. According to the EPN database, the pulsar is predominantly right hand circularly polarized with $\approx 32\%$ circular polarization at 1.4 GHz [@gould1998]. We analyzed archival VLA data for this pulsar acquired on February 9, 2002 at 1.5 GHz (project ID: AC629). The on-source time was around 11 minutes. No flux calibrator was observed. The phase calibrator was J1733-130 (J2000 RA: $17^h33^m02.70^s$, Dec: $-13^d04^m49^s.55$). Despite the lack of a flux calibrator, we did the analysis assuming a flat spectrum for the phase calibrator (with flux density of 1 Jy) over the observing band which was of width 22 MHz (fractional bandwidth is $1.5\%$). This implies that the flux densities that we get for the target could be incorrect, however the sign of Stokes $V$ will not be affected, and so is the case for the percentage circular polarization ($V/I$). The percentage circular polarization came out to be $+12\%$, while using $V=(RR^*-LL^*)/2$. We analyzed one more VLA archival dataset for this pulsar, obtained on February 5, 1990 (project ID: AR218) at 1.4 GHz with 100 MHz bandwidth (fractional bandwidth $\approx 7\%$). The on-source time was around 2 hours and 37 minutes. 3C286 was used as the flux and bandpass calibrator, and J1743-038 (J2000 RA: $17^h43^m58^s.86$, Dec: $-03^d50^m04^s.62$) was used as the phase calibrator. We obtained the flux densities of the target to be $7.7\pm 0.4$ mJy in $LL$ and $10.0\pm 0.4$ mJy in $RR$, which again gives $V/I\approx +13\%$. Thus, according to the VLA convention also, the pulsar is right circularly polarized, implying that the $RR^*$ and $LL^*$ for the VLA are defined in accordance with the IEEE convention. Note that the above result need not apply to the upgraded VLA, which is the Karl G. Jansky VLA or the JVLA. However we have indirect evidence in support of the assumption that the old VLA and the JVLA obey the same convention for defining $RR^*$ and $LL^*$. This comes from the observation of the magnetic star CUVir. This is the first main-sequence star which was discovered to be a persistent emitter of electron cyclotron maser emission (ECME) at 1.4 GHz [@trigilio2000]. This result came out from an observation performed with the VLA in the year 1998. One characteristic of ECME is that it produces very highly circularly polarized pulsed emission. For CUVir, the ECME pulses were found to be right circularly polarized [@trigilio2000]. The star was again observed with the VLA in the year 2010 at 1.45 GHz, and the ECME pulses were found to be right circularly polarized [@trigilio2011]. Most recently, we observed the star with the JVLA in the year 2019, and we also observed that the pulses are right circularly polarized in both L (1–2 GHz) and S (2–4 GHz) bands (Das et al. in prep). This suggests that the JVLA convention for defining $RR^*$ and $LL^*$ is same as that for the old VLA, and hence in accordance with the IEEE convention. Conclusion {#sec:conc} ========== The primary conclusion drawn from this work is the following: 1. The $X$ and $Y$ dipoles of the uGMRT in band 4 are swapped (w.r.t. the IAU convention). 2. If we have to compare uGMRT band 4 polarization results with that from the VLA, Lovell or Parkes telescope, we must interchange $RR^*$ and $LL^*$ for the uGMRT data; and also change the sign of Stokes $Q$. We would like to mention that the observatory is likely to take measures to make the convention (for defining $X$ and $Y$) consistent with the standard convention. These new measures, once implemented and fully tested, will be notified in a future technical report. Until then, the above conclusions are valid. Acknowledgements {#acknowledgements .unnumbered} ================ We thank the staff of the GMRT that made these observations possible. The GMRT is run by the National Centre for Radio Astrophysics of the Tata Institute of Fundamental Research. BD thanks Minhajur Rahaman for suggesting the pulsar B1702–19 to use as the polarization standard. BD thanks Biny Sebastian and Jayaram N. Chengalur for useful discussions. SK thanks Kadaladi Pavankumar for helping in beam polarization data. PC thanks Dipanjan Mitra for independently suggesting the polarization standard. This research has used NASA’s Astrophysics Data system. [99]{} Das, B., Chandra, P., Shultz, M. E., et al. 2019, [ApJ]{}, 877, 123 Gould, D. M., & Lyne, A. G. 1998, [MNRAS]{}, 301, 235 Gupta, Y., Ajithkumar, B., Kale, H. S., et al. 2017, Current Science, 113, 707 Hamaker, J. P., & Bregman, J. D. 1996, [A&AS]{}, 117, 161 Johnston, S., & Kerr, M. 2018, [MNRAS]{}, 474, 4629 Kudale, S.  2008, M.Sc. Thesis, Technical Report NCRA-T1400 Robishaw, T., & Heiles, C. 2018, arXiv e-prints, arXiv:1806.07391 Trigilio, C., Leto, P., Leone, F., Umana, G., & Buemi, C. 2000, [A&A]{}, 362, 281 Trigilio, C., Leto, P., Umana, G., et al. 2011, [ApJ]{}, 739, L10 van Straten, W., Manchester, R. N., Johnston, S., et al. 2010, [Publ. Astron. Soc. Australia]{}, 27, 104
{ "pile_set_name": "ArXiv" }
--- abstract: 'In this paper, we propose an end-to-end graph learning framework, namely **D**eep **I**terative and **A**daptive **L**earning for **G**raph **N**eural **N**etworks ([<span style="font-variant:small-caps;">DIAL-GNN</span>]{}), for jointly learning the graph structure and graph embeddings simultaneously. We first cast the graph structure learning problem as a similarity metric learning problem and leverage an adapted graph regularization for controlling smoothness, connectivity and sparsity of the generated graph. We further propose a novel iterative method for searching for a hidden graph structure that augments the initial graph structure. Our iterative method dynamically stops when the learned graph structure approaches close enough to the optimal graph. Our extensive experiments demonstrate that the proposed [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}model can consistently outperform or match state-of-the-art baselines in terms of both downstream task performance and computational time. The proposed approach can cope with both transductive learning and inductive learning.' author: - | Yu Chen^1^, Lingfei Wu^2^[^1], Mohammed J. Zaki^1^\ \ ^1^Rensselaer Polytechnic Institute\ ^2^IBM Research\ [email protected], [email protected], [email protected] bibliography: - 'dgl\_aaai20.bib' title: Deep Iterative and Adaptive Learning for Graph Neural Networks --- Introduction ============ Recent years have seen a growing amount of interest in graph neural networks (GNNs) [@kipf2016semi; @li2016gated; @GraphSage:hamilton2017inductive], with successful applications in broad areas such as computer vision [@norcliffe2018learning], natural language processing [@xu2018graph2seq; @xu2018exploiting; @xu2018sql] and healthcare informatics [@gao2019dyngraph2seq]. Unfortunately, GNNs can only be used when graph-structured data is available. Many real-world applications naturally admit graph-structured data like social networks. However, it is questionable if these intrinsic graph-structures are optimal for the downstream tasks. More importantly, many applications such as those in natural language processing may only have non-graph structured data or even just the original feature matrix, requiring additional graph construction from the original data matrix to formulate graph data. In the field of graph signal processing, researchers have explored various ways of learning graphs from data, but without considering the downstream tasks [@dong2016learning; @kalofolias2016learn; @kalofolias2017large; @egilmez2017graph]. Independently, there has been an increasing amount of work studying the dynamic model of interacting systems utilizing implicit interaction models [@sukhbaatar2016learning; @hoshen2017vain; @van2018relational; @kipf2018neural]. However, these methods cannot be directly applicable to jointly learning the graph structure and graph representations when the graph is noisy or even not available. Recently, researchers have explored methods to automatically construct a graph [@choi2019graph; @li2018adaptive; @liu2018contextualized; @chen2019graphflow; @chen2019reinforcement] when applying GNNs to non-graph structured data. However, these methods merely optimize the graphs towards the downstream tasks without utilizing the techniques which have proven to be useful in graph signal processing. More recently, [@franceschi2019learning] presented a new approach for jointly learning the graph and the parameters of GNNs by approximately solving a bilevel program. However, this approach has severe scalability issue since it needs to learn $N^2$ number of (Bernoulli) random variables to model joint probability distribution on the edges of the graph consisting of $N$ number of vertices. More importantly, it can only be used for transductive setting, which means this method cannot consider new nodes during the testing. To address these limitations, in this paper, we propose a **D**eep **I**terative and **A**daptive **L**earning for **G**raph **N**eural **N**etworks ([<span style="font-variant:small-caps;">DIAL-GNN</span>]{}) framework for jointly learning the graph structure and the GNN parameters that are optimized towards some prediction task. In particular, we present a graph learning neural network that casts a graph learning problem as a data-driven similarity metric learning task for constructing a graph. We then adapt techniques for learning graphs from smooth signals [@kalofolias2016learn] to serve as graph regularization. More importantly, we propose a novel iterative method to search for a hidden graph structure that augments the initial graph structure towards an optimal graph for the (semi-)supervised prediction tasks. The proposed approach can cope with both transductive learning and inductive learning. Our extensive experiments demonstrate that our model can consistently outperform or match state-of-the-art baselines on various datasets. Approach ======== ![image](images/arch.png) With this paper we address the challenging problem of automatic graph structure learning for GNNs. We are given a set of $n$ objects $V$ associated with a feature matrix $\vec{X} \in \mathbb{R}^{d\times n}$ encoding the feature descriptions of the objects. The goal is to automatically learn the graph structure $\mathcal{G}$, typically in the form of an adjacency matrix $\vec{A} \in \mathbb{R}^{n\times n}$, underlying the set of objects, which will be consumed by a GNN-based model for a downstream prediction task. Unlike most existing methods that construct graphs based on hand-crafted rules or features during preprocessing, our proposed [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}framework formulates the problem as an iterative learning problem that jointly learns the graph structure and the GNN parameters iteratively in an end-to-end manner. The overall model architecture is shown in Fig. \[fig:overall\_arch\]. Graph Learning as Similarity Metric learning -------------------------------------------- A common strategy of graph construction is to first compute the similarity between pairs of nodes based on some metric, and then consume the constructed graph in a downstream task. Unlike these methods, in this work, we design a learnable metric function for graph structure learning, which will be jointly trained with a task-dependent prediction model. ### Similarity Metric Learning After preliminary experiments, we design a multi-head weighted cosine similarity, $$\label{eq:similarity_metric_learning} \begin{aligned} s_{ij}^k &= \text{cos}(\vec{w}_k \odot \vec{v}_i, \vec{w}_k \odot \vec{v}_j)\\ s_{ij} &= \frac{1}{m}\sum_{k=1}^{m}{s_{ij}^k} \end{aligned}$$ where $\odot$ denotes the Hadamard product. Specifically, we use $m$ weight vectors (each has the same dimension as the input vectors and represents one perspective) to compute $m$ cosine similarity matrices independently and take their average as the final similarity $\vec{S}$. Intuitively, $s_{ij}^k$ computes the cosine similarity between the two input vectors $\vec{v}_i$ and $\vec{v}_j$, for the $k$-th perspective where each perspective considers one part of the semantics captured in the vectors. This idea of multi-head similarity is similar to those in multi-head attention [@vaswani2017attention; @velivckovic2017graph]. ### Graph Sparsification via $\varepsilon$-neighborhood An adjacency matrix (same for a metric) is supposed to be non-negative while $s_{ij}$ ranges between $[-1, 1]$. In addition, many underlying graph structures are much more sparse than a fully connected graph, which is not only computationally expensive but also makes little sense for most applications. We hence proceed to extract a symmetric sparse adjacency matrix $\vec{A}$ from $\vec{S}$ by considering only the $\varepsilon$-neighborhood for each node. Specifically, we mask off those elements in $\vec{S}$ which are smaller than certain non-negative threshold $\varepsilon$. $$\label{eq:epsilon_neigh} \begin{aligned} \vec{A}_{ij} = \left\{ \begin{array}{ll} s_{ij} & \quad s_{ij} > \varepsilon \\ 0 & \quad \text{otherwise} \end{array} \right. \end{aligned}$$ Graph Regularization -------------------- In graph signal processing [@shuman2013emerging], a widely adopted assumption for graph signals is that values change smoothly across adjacent nodes. Given an undirected graph with symmetric weighted adjacency matrix $A$, the smoothness of a set of $n$ graph signals $\vec{x}_1, \dots, \vec{x}_n \in \mathbb{R}^{d}$ is usually measured by the Dirichlet energy [@belkin2002laplacian], $$\label{eq:smoothness_loss} \begin{aligned} \Omega(\vec{A}, \vec{X})=\frac{1}{2n^2}\sum_{i,j}\vec{A}_{ij}||\vec{x}_i - \vec{x}_j||^2=\frac{1}{n^2}\text{tr}(\vec{X}^T \vec{L} \vec{X}) \end{aligned}$$ where $\text{tr}(\cdot)$ denotes the trace of a matrix, $\vec{L}=\vec{D}-\vec{A}$ is the graph Laplacian, and $\vec{D}=\sum_{j}\vec{A}_{ij}$ is the degree matrix. As can be seen, minimizing $\Omega(\vec{A}, \vec{X})$ forces adjacent nodes to have similar features, thus enforces smoothness of the graph signals on the graph associated to $\vec{A}$. However, solely minimizing the above smoothness loss will result in the trivial solution $\vec{A}=0$. Also, it is desirable to have control of how sparse the resulting graph is. Following [@kalofolias2016learn], we impose additional constraints to the learned graph, $$\label{eq:degree_sparsity_loss} \begin{aligned} f(\vec{A})= \frac{-\beta}{n} \vec{1}^T \text{log}(\vec{A}\vec{1}) + \frac{\gamma}{n^2} ||\vec{A}||_F^2 \end{aligned}$$ where $||\cdot||_F$ denotes the Frobenius norm. As we can see, the first term penalizes the formation of disconnected graphs via the logarithmic barrier, and the second term controls sparsity by penalizing large degrees due to the first term. In this work, we borrow the above techniques, and apply them as regularization to the graph learned by \[eq:similarity\_metric\_learning,eq:epsilon\_neigh\]. The overall graph regularization loss is defined as the sum of the above losses, which is able to control the smoothness, connectivity and sparsity of the resulting graph where $\alpha$, $\beta$ and $\gamma$ are all non-negative hyperparameters. $$\label{eq:graph_loss} \begin{aligned} \mathcal{L}_{\mathcal{G}} = \alpha \Omega(\vec{A}, \vec{X}) + f(\vec{A}) \end{aligned}$$ An Iterative Graph Learning Method ---------------------------------- ### Joint Graph Structure and Representation Learning We expect the graph structure underlying a set of objects to serve two purposes: i) it should respect the semantic relations among the objects, which is enforced by the metric function (\[eq:similarity\_metric\_learning\]) and the smoothness loss (\[eq:smoothness\_loss\]); ii), it should suit the needs of the downstream prediction task. Compared to previous works which directly optimize the adjacency matrix based on either some graph regularization loss [@kalofolias2017large], or some task-dependent prediction loss [@franceschi2019learning], we propose to learn by minimizing a joint loss function combining both the task prediction loss and the graph regularization loss, namely, $\mathcal{L} = \mathcal{L}_{\text{pred}} + \mathcal{L}_{\mathcal{G}}$. Note that our graph learning framework is agnostic to various GNNs and prediction tasks. In this paper, we adopt a two-layered GCN [@kipf2016semi] where the first layer maps the node features to the node embedding space (\[eq:update\_node\_vec\]), and the second layer further maps the intermediate node embeddings to the output space (\[eq:prediction\]). $$\begin{aligned} & \vec{Z} = \text{ReLU}(\widetilde{\vec{A}} \vec{X} \vec{W}_1) \label{eq:update_node_vec}\\ & \widehat{\vec{y}} = \sigma(\widetilde{\vec{A}} \vec{Z} \vec{W}_2) \label{eq:prediction}\\ & \mathcal{L}_{\text{pred}} = \ell(\widehat{\vec{y}}, \vec{y}) \label{eq:pred_loss}\end{aligned}$$ where $\widetilde{\vec{A}}$ is the normalized adjacency matrix, $\sigma(\cdot)$ is a task-dependent output function, and $\ell(\cdot)$ is a task-dependent loss function. For instance, for node classification problem, $\sigma(\cdot)$ is a softmax function for predicting a probability distribution over a set of classes, and $\ell(\cdot)$ is a cross-entropy function for computing the prediction loss. We now discuss how to obtain the normalized adjacency matrix $\widetilde{\vec{A}}$. Our preliminary experiments showed that it is harmful to totally discard the initial graph structure when it is available. Previous works [@velivckovic2017graph; @jiang2019semi] inject the initial graph structure into the graph learning mechanism by performing masked attention, which might limits its graph learning ability. This is because there is no way for their methods to learn weights for those edges that do not exist in the initial graph, but carry useful topological information. With the assumption that the optimal graph structure is potentially a small shift from the initial graph structure, we combine the learned graph structure with the initial graph structure as follows, $$\label{eq:adj_norm} \begin{aligned} \widetilde{\vec{A}} = \lambda \vec{L}_0 + (1 - \lambda) \frac{\vec{A}_{ij}}{\sum_j{\vec{A}_{ij}}} \end{aligned}$$ where $\vec{L}_0$ is the normalized adjacency matrix of the initial graph, defined as, $\vec{L}_0 = \vec{D}_0^{-1/2}\vec{A}_0\vec{D}_0^{-1/2}$, and $\vec{D}_0$ is its degree matrix. The adjacency matrix learned by \[eq:similarity\_metric\_learning,eq:epsilon\_neigh\] is row normalized such that each row sums to 1. A hyperparameter $\lambda$ is used to balance the trade-off between the learned graph structure and the initial graph structure. If such an initial graph structure is not available, we instead use a kNN graph constructed based on cosine similarity. $[\vec{A}_0 \leftarrow \text{kNN}( \vec{X}, k)]$ $\vec{A}^{(0)}, \widetilde{\vec{A}}^{(0)} \leftarrow \{\vec{X}, \vec{A}_0\}$ using  \[eq:similarity\_metric\_learning,eq:epsilon\_neigh,eq:adj\_norm\] $\vec{Z}^{(0)} \leftarrow \{\widetilde{\vec{A}}^{(0)}, \vec{X}\}$ using \[eq:update\_node\_vec\] $\mathcal{L}_{\text{pred}}^{(0)} \leftarrow \{\widetilde{\vec{A}}^{(0)}, \vec{Z}^{(0)}, \vec{y}\}$ using \[eq:prediction,eq:pred\_loss\] $\mathcal{L}_{\mathcal{G}}^{(0)} \leftarrow \{\vec{A}^{(0)}, \vec{X}\}$ using \[eq:smoothness\_loss,eq:degree\_sparsity\_loss,eq:graph\_loss\] $\mathcal{L}^{(0)} \leftarrow \mathcal{L}_{\text{pred}}^{(0)} + \mathcal{L}_{\mathcal{G}}^{(0)}$ $t \leftarrow 0$ $\mathcal{L} \leftarrow \mathcal{L}^{(0)} + \sum_{i=1}^t{\mathcal{L}^{(i)}} / t$ \[alg\_line:BP\_end\] ### Iterative Method for Graph Learning Some previous works [@velivckovic2017graph] rely solely on raw node features to learn the graph structure based on some attention mechanism, which we think have some limitations since raw node features might not contain enough information to learn good graph structures. Our preliminary experiments showed that simply applying some attention function upon these raw node features does not help learn meaningful graphs (i.e., attention scores are kind of uniform). Even though we train the model jointly using the task-dependent prediction loss, we are limited by the fact that the similarity metric is computed based on the potentially inadequate raw node features. To address the above limitation, we propose a Deep Iterative and Adaptive Learning framework for Graph Neural Networks ([<span style="font-variant:small-caps;">DIAL-GNN</span>]{}). A sketch of the [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}framework is presented in \[alg:DIAL-GNN\]. Inputs and operations in squared brackets are optional. Specifically, besides computing the node similarity based on their raw features, we further introduce another learnable similarity metric function ( \[eq:similarity\_metric\_learning\]) that is rather computed based on the intermediate node embeddings, as demonstrated in \[alg\_line:refine\_adj\]. The aim is that the metric function defined on this node embedding space is able to learn topological information supplementary to the one learned solely based on the raw node features. In order to combine the advantages of both the raw node features and the node embeddings, we make the final learned graph structure as a linear combination of them, $$\label{eq:combine_adj_norm_t_0} \begin{aligned} \widebar{\vec{A}}^{(t)} = \eta \widetilde{\vec{A}}^{(t)} + (1 - \eta) \widetilde{\vec{A}}^{(0)} \end{aligned}$$ where $\widetilde{\vec{A}}^{(t)}$ and $\widetilde{\vec{A}}^{(0)}$ are the two normalized adjacency matrices learned by \[eq:adj\_norm\] at the $t$-th iteration and the initialization step before the iterative loop, respectively. Furthermore, as we can see from \[alg\_line:refine\_adj\] to \[alg\_line:refine\_node\_vec\], the algorithm repeatedly refines the adjacency matrix $\widetilde{\vec{A}}^{(t)}$ with the updated node embeddings $\vec{Z}^{(t-1)}$, and in the meanwhile, refines the node embeddings $\vec{Z}^{(t)}$ with the updated adjacency matrix $\widetilde{\vec{A}}^{(t)}$. The iterative procedure dynamically stops when the learned adjacency matrix converges (with certain threshold $\delta$) or the maximal number of iterations is reached ( \[alg\_line:stopping\_criterion\]). Compared to using a fixed number of iterations globally, the advantage of applying this dynamical stopping strategy becomes more clear when we are doing mini-batch training since we can adjust when to stop dynamically for each example graph in the mini-batch. At each iteration, a joint loss combining both the task-dependent prediction loss and the graph regularization loss is computed ( \[alg\_line:joint\_loss\_t\]). After all iterations, the overall loss will be back-propagated through all previous iterations to update the model parameters (\[alg\_line:BP\]). Formal Analysis {#sec:formal_analysis} =============== Convergence of the Iterative Learning Procedure {#sec:convergence_analysis} ----------------------------------------------- While it is challenging to theoretically prove the convergence of the proposed iterative learning procedure due to the arbitrary complexity of the involved learning model, here we want to conceptually understand why it works in practice. \[fig:info\_flow\] shows the information flow of the learned adjacency matrix $\vec{A}$ and the intermediate node embedding matrix $\vec{Z}$ during the iterative procedure. For the sake of simplicity, we omit some other variables such as $\widetilde{\vec{A}}$. As we can see, at $t$-th iteration, $\vec{A}^{(t)}$ is computed based on $\vec{Z}^{(t-1)}$ (\[alg\_line:refine\_adj\]), and $\vec{Z}^{(t)}$ is computed based on $\widetilde{\vec{A}}^{(t)}$ (\[alg\_line:refine\_node\_vec\]) which is computed based on $\vec{A}^{(t)}$ (\[eq:adj\_norm\]). We further denote the difference between the adjacency matrices at the $t$-th iteration and the previous iteration by $\delta_A^{(t)}$. Similarly, we denote the difference between the node embedding matrices at the $t$-th iteration and the previous iteration by $\delta_Z^{(t)}$. If we assume that $\delta_Z^{(1)} < \delta_Z^{(0)}$, then we can expect that $\delta_A^{(2)} < \delta_A^{(1)}$ because conceptually more similar node embedding matrix (i.e., smaller $\delta_Z$) is supposed to produce more similar adjacency matrix (i.e., smaller $\delta_A$) given the fact that model parameters keep the same through iterations. Similarly, given that $\delta_A^{(2)} < \delta_A^{(1)}$, we can expect that $\delta_Z^{(2)} < \delta_Z^{(1)}$. Following this chain of reasoning, we can easily extend it to later iterations. In order to see why the assumption $\delta_Z^{(1)} < \delta_Z^{(0)}$ makes sense in practice, we need to recall the fact that $\delta_Z^{(0)}$ measures the difference between $\vec{Z}^{(0)}$ and $\vec{X}$, which is usually larger than the difference between $\vec{Z}^{(1)}$ and $\vec{Z}^{(0)}$, namely $\delta_Z^{(1)}$. We will empirically examine the convergence property of the iterative learning procedure in the experimental section. ![Information flow of the proposed iterative learning procedure.[]{data-label="fig:info_flow"}](images/info_flow.png) Model Complexity ---------------- The cost of learning an adjacency matrix is $\mathcal{O}(n^2h)$ for $n$ nodes and data in $\mathbb{R}^h$, while computing node embeddings costs $\mathcal{O}(n^2d + ndh)$, computing task output costs $\mathcal{O}(n^2h)$, and computing the total loss costs $\mathcal{O}(n^2d)$. We set the maximal number of iterations to $T$, hence the overall complexity is $\mathcal{O}(Tn(nh+nd+hd))$. If we assume that $d \approx h$ and $n \gg d$, the overall complexity is $\mathcal{O}(Tdn^2)$. Experiments =========== Methods 20News MRD ------------------------------------------------------------ -- ---------------- ---------------- BiLSTM 80.0 (0.4) 53.1 (1.4) kNN-GCN 81.3 (0.6) 60.1 (1.5) [<span style="font-variant:small-caps;">DIAL-GNN</span>]{} **83.6 (0.4)** **63.7 (1.8)** : Test scores ($\pm$ standard deviation) in percentage on classification (accuracy) and regression ($R^2$) datasets in the inductive setting.[]{data-label="table:inductive_results"} Methods Cora Wine 20News ------------------------------------------------------------ -- ---------------- ---------------- ---------------- -- -- [<span style="font-variant:small-caps;">DIAL-GNN</span>]{} **84.5 (0.3)** **97.8 (0.6)** **83.6 (0.4)** w/o graph reg. 84.3 (0.4) 97.3 (0.8) 83.4 (0.5) w/o IL 83.5 (0.6) 97.2 (0.8) 83.0 (0.4) : Ablation study on various classification datasets.[]{data-label="table:ablation_results"} In this section, we conducted a series of experiments to verify the effectiveness of the proposed model and assess the impact of different model components. The implementation of the model will be made publicly available at <https://github.com/hugochan/IDGL> soon. The details on model settings are provided in the appendix. Datasets and Setup ------------------ The benchmarks used in our experiments include two network benchmarks, three data point benchmarks and two text benchmarks. Cora and Citeseer are two commonly used network benchmarks for evaluating graph-based learning algorithms [@sen2008collective]. The input features are bag of words and the task is node classification. In addition to Cora and Citeseer where the graph topology is available, we evaluate [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}on three data point benchmarks (i.e., Wine, Breast Cancer (Cancer) and Digits from the UCI machine learning repository [@Dua:2019]). The task is also node classification. Finally, to demonstrate the effectiveness of [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}on inductive learning problems, we conduct document classification and regression tasks on the 20Newsgroups data (20News) and the movie review data (MRD) [@pang2004sentimental], respectively. In this setting, we regard each document as a graph containing each word as a node. For Cora and Citeseer, we follow the experimental setup of previous works [@kipf2016semi; @velivckovic2017graph; @franceschi2019learning]. For Wine, Cancer and Digits, we follow the experimental setup of [@franceschi2019learning]. For 20News, we randomly select 30% examples from the training data as the development set. For MRD, we split the data to train/dev/test sets using a 60%/20%/20% split. The reported results are averaged over 5 runs with different random seeds. Please refer to the appendix for data statistics. Baselines --------- Our main baseline in the transductive setting is LDS. Similar to our work, LDS also jointly learns the graph structure and the parameters of GNNs. However, LDS is incapable of handling inductive learning problems since it aims at directly optimizing the discrete probability distribution on the edges of the underlying graph, which makes it unable to handle unseen nodes/graphs in the testing phase. The experimental results of several semi-supervised embedding (SemiEmb) [@weston2012deep]) and supervised learning (support vector machines (RBF SVM)) baselines are reported in the LDS paper. For the sake of completeness, we directly copy their results here. For ease of comparison, we also copy the reported results of LDS even though we rerun the experiments of LDS using the official code released by the authors. In addition, for Cora and Citeseer, we include GCN [@kipf2016semi] and GAT [@velivckovic2017graph] as baselines. In order to evaluate the robustness of [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}to noisy graphs, we also compare [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}with GCN on graphs with edge deletions or additions. For data point benchmarks where the graph topology is not available, we conceive a kNN-GCN baseline where a kNN affinity graph on the data set is first constructed as a preprocessing step before applying a GCN. For 20News and MRD in the inductive setting, we compare [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}with a BiLSTM [@hochreiter1997long] baseline and kNN-GCN. Results and Analysis -------------------- The results of transductive and inductive experiments are shown in \[table:transductive\_results\] and \[table:inductive\_results\]. First of all, we can see that [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}outperforms all baseline methods in 6 out of 7 benchmarks, which demonstrates the effectiveness of [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}. We can clearly see that [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}can greatly help the node classification task even when the graph topology is given. When the graph topology is not given, compared to kNN-GCN, [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}consistently achieves much better results on all datasets, which shows the power of jointly learning graph structures and GNN parameters. Compared to LDS, [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}achieves better performance in 4 out of 5 benchmarks. The good performance on 20News and MRD verifies the capability of [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}on inductive learning problems. We perform an ablation study to assess the impact of different model components. As shown in \[table:ablation\_results\], we can see a significant performance drop consistently on all datasets (e.g., 3.1% on Citeseer) by turning off the iterative learning component, which demonstrates the effectiveness of the proposed iterative learning framework for the graph learning problem. We can also see the benefits of jointly training the model with the graph regularization loss. Benchmarks Cora Citeseer Wine Cancer Digits ------------------------------------------------------------------ -- ---------- ----------- --------- --------- --------- GCN 3 (1) 5 (1) GAT 26 (5) 28 (5) LDS 390 (82) 585 (181) 33 (15) 25 (6) 72 (35) [<span style="font-variant:small-caps;">DIAL-GNN</span>]{} 237 (21) 563 (100) 20 (7) 21 (11) 65 (12) [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}w/o IL 49 (8) 61 (15) 3 (2) 3 (1) 2 (1) To evaluate the robustness of [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}on noisy graphs, we construct graphs with random edge deletions or additions. Specifically, we randomly remove or add 25%, 50% and 75% of the edges in the original graphs. The results on the edge deletion graphs and edge addition graphs are shown in \[fig:missing\_edges\_plot\] and \[fig:added\_edges\_plot\], respectively. As we can clearly see, compared to GCN, [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}achieves better results in all scenarios and is much more robust to noisy graphs. While GCN completely fails in the edge addition scenario, [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}is still able to perform reasonably well. We conjecture this is because \[eq:adj\_norm\] is formulated in a form of skip-connection, by lowering the value of $\lambda$, we enforce the model to rely less on the initial noisy graph that contains too much additive random noise. ![Test accuracy ($\pm$ standard deviation) in percentage for the edge deletion scenario on Cora.[]{data-label="fig:missing_edges_plot"}](images/cora_missing_edges_aaai.png) ![Test accuracy ($\pm$ standard deviation) in percentage for the edge addition scenario on Cora.[]{data-label="fig:added_edges_plot"}](images/cora_added_edges_aaai.png) In \[fig:convergence\], we show the evolution of the learned adjacency matrix and accuracy through iterations in the iterative learning procedure in the testing phase. We compute the difference between adjacency matrices at consecutive iterations as $\delta_A^{(t)} = ||\vec{A}^{(t)} - \vec{A}^{(t-1)}||_F^2 / ||\vec{A}^{(t)}||_F^2$ which typically ranges from 0 to 1. As we can see, both the adjacency matrix and accuracy converge quickly through iterations. This empirically verifies the analysis we made on the convergence property of the iterative learning procedure. ![Evolution of the learned adjacency matrix and test accuracy (in %) through iterations in the iterative learning procedure.[]{data-label="fig:convergence"}](images/cora_convergence_new.png) There are two natural ways of designing the stopping strategy for iterative learning methods. We can either use a fixed number of iterations, or dynamically determine if the learning procedure already converges or not based on some stopping criterion. In \[fig:effect\_stopping\], we empirically compare the effectiveness of the above two strategies. We run [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}on Cora (left) and Citeseer (right) using different stopping strategies with 5 runs, and report the average accuracy. As we can see, dynamically adjusting the number of iterations using the stopping criterion works better in practice. ![Performance comparison (i.e., test accuracy in %) of two different stopping strategies: i) using a fixed number of iterations (blue line), and ii) using a stopping criterion to dynamically determine the convergence (red line).[]{data-label="fig:effect_stopping"}](images/cora_stopping_criterion.png) Timing ------ Finally, we compare the training efficiency of [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}, LDS and other classic GNNs (e.g., GCN and GAT) on various benchmarks. All experiments are conducted on the same machine which has an Intel i7-2700K CPU, an Nvidia Titan Xp GPU and 16GB RAM, and are repeated 5 times with different random seeds. Results are shown in Table \[table:training\_time\]. As we can see, both [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}and LDS are slower than GCN and GAT, which is as expected since GCN and GAT do not need to learn graph structures simultaneously. [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}is consistently faster than LDS, but in general, they are comparable. We also find that the iterative learning part is the most time consuming in [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}. Conclusion ========== In this paper, we proposed a Deep Iterative and Adaptive Learning framework for Graph Neural Networks ([<span style="font-variant:small-caps;">DIAL-GNN</span>]{}) for jointly learning the graph structure and graph embeddings by optimizing a joint loss combining both task prediction loss and graph regularization loss. The proposed method is able to iteratively search for hidden graph structures that better help the downstream prediction task. Our extensive experiments demonstrate that the proposed [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}model can consistently outperform or match state-of-the-art baselines on various datasets. We leave how to design more effective and scalable metric functions as future work. ### Acknowledgments {#acknowledgments .unnumbered} This work is supported by IBM Research AI through the IBM AI Horizons Network. We thank the anonymous reviewers for their helpful feedback. Data Statistics {#sec:data_statistics} =============== The benchmarks used in our experiments include two network benchmarks, three data point benchmarks and two text benchmarks. Below we show the brief data statistics. Model Settings {#sec:model_settings} ============== In all our experiments, we apply a dropout ratio of 0.5 after GCN layers except for the output GCN layer. During the iterative learning procedure, we also apply a dropout ratio of 0.5 after the intermediate GCN layer, except for Citeseer (no dropout) and Digits (0.3 dropout). For experiments on text benchmarks, we keep and fix the 300-dim GloVe vectors for words that appear more than 10 times in the dataset. For long documents, for the sake of efficiency, we cut the text length to maximal 1,000 words. We apply a dropout ratio of 0.5 after word embedding layers and BiLSTM layers. The batch size is set to 16. And the hidden size is set to 128 and 64 for 20News and MRD, respectively. For all other benchmarks, the hidden size is set to 16 to follow the original GCN paper. We use Adam [@kingma2014adam] as the optimizer. For the text benchmarks, we set the learning rate to 1e-3. For all other benchmarks, we set the learning rate to 0.01 and apply L2 norm regularization with weight decay set to 5e-4. \[table:hyperparam\] shows the hyperparameters associated to [<span style="font-variant:small-caps;">DIAL-GNN</span>]{}for all benchmarks. All hyperparameters are tuned on the development set. [^1]: Corresponding author.
{ "pile_set_name": "ArXiv" }
--- abstract: 'A 2.5-dimensional magnetohydrodynamics simulation analysis of the energy release for three different reconnection regimes is presented. The system under investigation consists in a current-sheet located in a medium with a strong density variation along the current layer: such system is modeled as it were located in the high chromosphere/low solar corona as in the case of pre- flare and coronal mass ejection (CME) configurations or in the aftermath of such explosive phenomena. By triggering different magnetic-reconnection dynamics, that is from a laminar slow evolution to a spontaneous non-steady turbulent reconnection [@lapenta08; @bettarini09; @skender09], we observe a rather different efficiency and temporal behavior with regard to the energy fluxes associated with each of these reconnection-driven evolutions. These discrepancies are fundamental key-properties to create realistic models of the triggering mechanisms and initial evolution of all those phenomena requiring fast (and high power) magnetic reconnection events within the solar environment.' author: - Lapo Bettarini - Giovanni lapenta title: Analysis of the energy release for different magnetic reconnection regimes within the solar environment --- [ address=[Centrum voor Plasma-Astrofysica, Departement Wiskunde, Katholieke Universiteit Leuven, Celestijnenlaan 200B, B-3001 Leuven, Belgium]{} ]{} [ address=[Centrum voor Plasma-Astrofysica, Departement Wiskunde, Katholieke Universiteit Leuven, Celestijnenlaan 200B, B-3001 Leuven, Belgium]{} ]{} Introduction ============ Magnetic reconnection is essential to model partially or completely most of the phenomena occurring within the solar atmosphere, being the sun’s magnetic field the huge reservoir of energy driving the dynamical evolution of such structures. A detailed and realistic model of magnetohydrodynamics (MHD) reconnection is an unescapable step in order to achieve a deep understanding of the highly dynamic solar atmosphere as well as of the interaction of solar plasma structures with the interplanetary medium. In fact, the MHD approach is adequate to describe the initial evolution and the following nonlinear stages of such phenomena, even though the detailed physics of their initiation depends primarily on kinetic effects. In general, the starting point is either the well-known Sweet-Parker’s model of a slow and laminar reconnection [@sweet58; @parker57] in a harris or force-free current-sheet, or the fast process described by Petschek’s model [@petschek64] and usually obtained by the introduction in the MHD system of a localized enhancement of the plasma resistivity [@biskamp00; @yokoyama01]. Yet, a new high-power and non-steady turbulent reconnection process is possible in MHD without invoking the presence of any *ad-hoc* term resembling kinetic effects in MHD equations. This has been recently shown by @lapenta08 and applied to the solar environment by @bettarini09. Such mechanism includes a spontaneous transition from a slow to a fast reconnection with time-scales of the order of Alfvén times and no factitious anomalous effects are considered. During its evolution, current-sheets develop a chaotic structure of multiple small-scale interacting reconnection sites determining a macroscopic turbulent state. As in @bettarini09, the system under investigation in the present work is a current-sheet embedded in the high chromosphere/low solar corona region. It consists in a current-sheet determined by a force-free magnetic field in the presence of a strong density variation along the current layer provided by a proper step function. We vary the global resistivity of our numerical domain (by means of the Lundquist number) such that different dynamical regimes of the current-sheet are set, from a slow and diffusive reconnection to a fast and turbulent process and an in-between evolution. Here, we present an energy-release analysis of such three different processes in order to (a) provide useful hints of the best-promising mechanism to be considered for more and more realistic models of solar (explosive) phenomena, and (b) to determine the key-features of events that could be observed by current and near-future solar and space missions. Experiment Settings =================== We solve numerically the viscous-resistive compressible $2.5$D one-fluid MHD equations parametrized by the global Lundquist and kinetic Reynolds number, [**S**]{} and [**R**]{}$_\nu$ respectively. We define a vertical or stream-wise direction ($z$) oriented away from the sun, wherein open boundary conditions are set at $z = 0$ and $L_z$, and a horizontal or cross-stream direction ($x$) where we set reflecting boundary conditions. We consider an ideal equation of state with polytropic index $\gamma$ such that $p = \rho \, (\gamma-1) \, I$ where $p$, $\rho$ and $I$ are respectively the kinetic pressure, the density and the enthalpy of the plasma. We use the Lagrangian code, FLIP3D-MHD [@brackbill91], and we consider $600$ (in $x$) $\times$ $960$ (in $z$) Lagrangian markers arrayed initially in a $3 \times 3$ uniform formation in each of the $200 \times 320$ cells of our numerical grid. [lccc]{} & & &\ S & $10^2$ & $10^3$ & $10^4$\ \ \ \ As already pointed out in the introduction, our system consists in a current-sheet configuration determined by a force-free magnetic field with $B_{z}(x) = \tanh x$, $B_{x}(x) = 0$ and the out-of-plane component of the magnetic field $B_{y}(x) = -{{\rm sech\,}}x$ as guide field. With this configuration, we obtain high-resolution simulations where the current-sheet width is resolved by $30$ Lagrangian markers. The density is modeled as a step function where the ratio between the density in the region $0 < z < L_z/8$ (modeling the high chromosphere) and that one in $L_z/8 < 0 < L_z$ (the low corona) is about $10^5$ as considered in @yokoyama01 [@bettarini09]. We consider an initial GEM-like perturbation [@birn01] but exponentially localized at the center of the numerical box (i.e. within the solar corona) with a perturbation amplitude of $\epsilon = 0.5$. Physical Parameters ------------------- The MHD equations are normalized according to reasonable values for the high chromosphere/low solar corona: the number density $\rho/m \approx 10^9 {\ensuremath{\,\mbox{cm}^{-3}}}$; the length scale $L \approx 1000-2000 {\ensuremath{\,\mbox{km}}}$, where $L$ is the current-sheet width and $L = L_x/20 = L_z/80$; the Alfvén velocity $v_A \approx 400-600 {\ensuremath{\,\mbox{km} \, \mbox{s}^{-1}}}$ (approximately three times the sound speed). So, we have a time scale of $t_A = L/v_A \approx 1.5-5 {\ensuremath{\,\mbox{s}}}$. See Tab. \[table1\] for more details on simulations’ setting and parameters. Results ======= ![Top panel: Measured reconnected flux as a function of time where the dashed line corresponds to Run A, dotted line to Run B and solid and dotted-dashed lines correspond both to Run C (see Tab. \[table1\]). Bottom panel: Reconnection rate as a function of time for the three simulations.[]{data-label="figure1"}](figure1.eps){height=".25\textheight"} In the top panel of Fig. \[figure1\], the reconnected magnetic flux (in dimensionless units) as a function of time for three different values of the Lundquist number (so, of the global resistivity) is shown. This quantity is measured by computing the maximum and minimum values of the out-of-plane magnetic potential, $A_y$ at the mid axis ($x = 0$): their difference provides the amount of flux in the closed field lines, caused by reconnection as described in @biskamp00. A low value of [**S**]{}, [**S**]{}$ = 10^2$ results in a laminar and slow dynamics of the current-sheet. After an initial transient time, a nearly-constant reconnected flux is observed (Run A, dashed line). By setting [**S**]{} to $10^3$, the system presents again an initial nearly-constant increase in the reconnected flux resembling a slow reconnecting evolution, although it is more and more evident a smooth transition to a faster evolution as the time goes by (Run B, dotted line). By increasing [**S**]{} further, [**S**]{} $=10^4$, smaller scales can be reached and the high resolution allows us to observe a completely different regime [@lapenta08; @bettarini09]. Because of a tearing instability of the initial Sweet-Parker configuration [@furth63; @skender09], the system experiences an abrupt switch from an initial slow and laminar reconnection regime to a fast reconnection process leading to a macroscopic turbulent configuration of the initial system. ![image](figure2.eps){height=".3\textheight"} The different behavior of the reconnection process in the three simulations is pointed out also in the evolution of the reconnection rate as a function of time. This is shown in the bottom panel of Fig. \[figure1\]. Different rates but the same general behavior are shown during the first few decades of time steps where the system is driven by a slow reconnection. At about $t = 40$ (measured in terms of alfvén times), it is evident the rapid increase of the reconnection rate in Run C. With regard to this case, the rate is measured both by computing the slope of the corresponding reconnected flux (solid line in Fig. \[figure1\]) as for the other runs and also by means of the out-of-plane electric field variation through the reconnection diffusion region (dotted-dashed line). With the latter method, the curve of the rate is more noisy, yet the two methods give two manifestly consistent results. Note that also the rate of Run B (dotted line) is becoming more and more consistent with a fast evolution as the time goes by, whereas the Run A’s rate is slowly decreasing. As described in @bettarini09, during such three different regimes, the slow reconnection evolution, the spontaneous and abrupt variation from a slow to a fast regime and the intermediate dynamics, the current-layer develops asymmetrically because of the strong density variation resembling the transition from the high chromosphere and the low solar corona. Moreover, characteristic features in magnetic field topology of the aftermath of solar explosive phenomena are observed. In order to understand the importance of such reconnection mechanisms within the solar atmosphere, a fundamental step is to analyze such processes from the point of view of their efficiency in releasing energy and particles in the upper atmosphere. In Fig. \[figure2\], the flux of the Poynting vector, the enthalpy flux, the kinetic flux, and the mass flux through the upper boundary of the numerical box ($z = L_z$) are shown as functions of time. In the figure also the normalization values for all the shown quantities are reported. As the perturbation is approximately in the middle of our numerical box ($z \sim L_z/2 = 40$, that is in the solar corona) and there the alfvén velocity is $1$, we start to measure the effects of reconnection at the upper boundary from $t \sim 40$. Of course, we neglect the initial transient time caused by a perturbation that is not exactly an eigenstate of the system. In the Run A, the current-sheet evolves accordingly a diffusive reconnection process that determines a pile-up of the enthalpy flux of the system (top-left panel of Fig. \[figure2\]) as well as an increase of the mass flux (top-right panel) to the upper layers of the corona because of an acceleration process as shown by the kinetic flux’s profile (bottom-right panel of Fig. \[figure2\]). The flux through the upper boundary of the stream-wise component of the Poynting vector is almost null throughout the simulation. As previously pointed-out, Run A’s reconnection rate sensitively decreases in the last phase and after a plateau of a few time steps, we observe a corresponding decrease of all fluxes. The slow-to-fast reconnection evolution of the current-sheet observed in Run C passes through a chaotic state during which the formation of several smaller-scale current-sheets and reconnection sites interacting each other is observed [@bettarini09]. So, it is the interaction of such structures continuously moving backward and forward that prevents the system to release a large amount of energy and particles to the upper atmosphere. Yet, as the process goes on and the system reaches a macroscopic turbulent configuration, an impulsive event is observed and a release of mass as well as an increase of the temperature is rapidly determined in the upper corona. A different consideration has to be done with regard to Poynting vector’s flux. In fact, it increases constantly and sensitively during all the fast regime, but it does not characterize the impulsive phase of the system. As already pointed out, an intermediate evolution is observed during the Run B. The system switches to a fast reconnection regime according to a slow rather than an abrupt process. From the energy- and mass-flux point of view, the profiles shown in \[figure2\] stay in-between the behavior described for the previous cases. An initial slow but constant increase of the fluxes is followed by a plateau of few time-steps that does not lead to a final decrease of the fluxes as in the Run A but to a impulsive phase as in the Run C, though being the peak values sensitively lower. Conclusions =========== Here, we observe the energetic and mass release of a force-free current-sheet evolving in a small region of the high/chromosphere/low solar corona. According to the new mechanism proposed by @lapenta08 and applied by @bettarini09 within the solar environment, it is possible that this system follows a magnetic field-line reconnection dynamics and that spontaneously evolves from a low- and constant-rate phase to a fast and high power regime within the pure-resistive MHD-theory framework (and so without invoking the presence of any anomalous resistivity). Hence, it becomes fundamental the understanding whether this process can provide us with evolution-related observables that can improve our models of solar (explosive) phenomena and of MHD structures in general. By performing 2.5D MHD high-resolution simulations and by varying the characteristic global Lundquist number, we switch three different magnetic reconnection regimes: a Sweet-Parker evolution, the new spontaneously slow-to-fast reconnection and an intermediate dynamics. By means of the Sweet-Parker regime, we can obtain an quantitatively efficient but slow conversion of the initial magnetic energy available in the basic configuration, but it does not provide some important features observed for example in solar phenomena like flares or CMEs, that is a spontaneous impulsive event that evolves on time-scales of tens/hundreds of seconds (see again Fig. \[figure2\] and the normalization values there reported) and releases large quantities of mass, enthalpy and high-velocity structures that eventually can power and trigger secondary events in the upper solar atmosphere. Even an intermediate evolution combining an initial laminar reconnection dynamics and evolving to a faster and faster regime is not able to recover such dynamical features, even if a less rapid and power impulsive process is observed. Here, we show that the spontaneous non-steady reconnection process can provide time-scales, reconnection rates and energy- and mass-flux values that, on the one hand, are consistent with the theoretical framework of pure resistive-MHD and, on the other hand, they allow to use such mechanism as starting model of the complex triggering event and initial evolution of solar phenomena. Future studies will include a detailed analysis both from the dynamical and energetic point of view of the smaller-scale structures forming during the chaotic evolution of the current-sheet by means of adaptive-mesh refinement numerical tools. Furthermore, at this stage the introduction of a more realistic model of the high chromosphere/low solar corona in a three-dimensional configuration is mandatory. The research leading to these results has received funding from the European Commission’s Seventh Framework Programme (FP7/2007-2013) under the grant agreement n$^\circ$ 218816 (SOTERIA project, [www.soteria-space.eu](www.soteria-space.eu). The simulations shown were conducted using processors on the VIC cluster of the Vlaams Supercomputer Centrum at K.U. Leuven (Belgium). [11]{} natexlab\#1[\#1]{}\[1\][“\#1”]{} url \#1[`#1`]{}urlprefix\[2\]\[\][[\#2](#2)]{} G. Lapenta, *Phys. Rev. Lett.* **100**, 235001 (2008). L. Bettarini, and G. Lapenta, *ApJ* **Submitted** (2009). M. Skender, and G. Lapenta, *Phys. Plasmas* **submitted** (2009). P. A. Sweet, “The Neutral Point Theory of Solar Flares,” in *Electromagnetic Phenomena in Cosmical Physics*, edited by B. Lehnert, Cambridge University Press, Cambridge, England, 1958, vol. 6 of *IAU Symposium*, pp. 123–+. E. N. Parker, *J. Geophys. Res.* **62**, 509–520 (1957). H. E. Petschek, “Magnetic field annihilation,” in *The Physics of Solar Flares*, edited by W. N. Hess, 1964, pp. 425–+. D. Biskamp, *Magnetic reconnection in plasmas*, Cambridge Monograph on Plasma Physics, Cambridge University Press, Cambridge, England, 2000. T. Yokoyama, and K. Shibata, *ApJ* **549**, 1160–1174 (2001). J. U. Brackbill, *J. Comp. Phys.* **96**, 163–192 (1991), <http://adsabs.harvard.edu/abs/1991JCoPh..96..163B>. J. Birn, J. F. Drake, M. A. Shay, B. N. Rogers, R. E. Denton, M. Hesse, M. Kuznetsova, Z. W. Ma, A. Bhattacharjee, A. Otto, and P. L. Pritchett, *J. Geophys. Res.* **106**, 3715–3720 (2001). H. P. [Furth]{}, J. [Killeen]{}, and M. N. [Rosenbluth]{}, *Phys. Fluids* **6**, 459–484 (1963).
{ "pile_set_name": "ArXiv" }
--- abstract: 'We prove that any derivation of the $*$-algebra $LS(\mathcal{M})$ of all locally measurable operators affiliated with a properly infinite von Neumann algebra $\mathcal{M}$ is continuous with respect to the local measure topology $t(\mathcal{M})$. Building an extension of a derivation $\delta:\mathcal{M}\longrightarrow LS(\mathcal{M})$ up to a derivation from $LS(\mathcal{M})$ into $LS(\mathcal{M})$, it is further established that any derivation from $\mathcal{M}$ into $LS(\mathcal{M})$ is $t(\mathcal{M})$-continuous.' address: - 'Department of Mathematics,National University of Uzbekistan, Vuzgorodok, 100174, Tashkent, Uzbekistan' - 'Department of Mathematics, National University of Uzbekistan, Vuzgorodok, 100174, Tashkent, Uzbekistan' - 'School of Mathematics and Statistics, University of New South Wales, Sydney, NSW 2052, Australia ' author: - 'A. F. Ber' - 'V. I. Chilin' - 'F. A. Sukochev' title: Continuity of derivations in algebras of locally measurable operators --- Introduction ============ The theory of derivations of various classes of Banach $*$-algebras (e.g. $C^*$, $AW^*$ and $W^*$-algebras) is very well developed (see, for example, [@Brat; @Sak; @Sak1]). It is well known that every derivation of a $C^*$-algebra is norm continuous and every derivation of a $AW^*$-algebra (in particular, of a $W^*$-algebra) is inner [@Olesen; @Sak]. The development of the theory of noncommutative integration, initiated by I. Segal’s paper [@Seg] prompted the introduction of numerous non-trivial $*$-algebras of unbounded operators, which, in a certain sense, are close to $AW^*$ and $W^*$-algebras. The main interest here is represented by the $*$-algebra $LS(\mathcal{M})$ (respectively, $S(\mathcal{M})$) of all locally measurable (respectively, measurable) operators, affiliated with a $W^*$-algebra (or with a $AW^*$-algebra) $\mathcal{M}$ and also by the $*$-algebra $S(\mathcal{M},\tau)$ of all $\tau$-measurable operators from $S(\mathcal{M})$, where $\tau$ is a faithful normal semifinite trace on $\mathcal{M}$ [@Ne; @San]. The importance of the algebra $LS(\mathcal{M})$ for the theory of unbounded derivations on von Neumann algebras may be seen from the following classical example. Consider the algebra $\mathcal{M}= L_\infty(0,\infty)$ equipped with the semifinite trace given by Lebesgue integration and consider (a partially defined) derivation $\delta=d/dt$ on $\mathcal{M}$. A simple argument shows that the algebra $LS(\mathcal{M})$, which in this case coincides with the space of all measurable complex functions on $(0,\infty)$ is the only natural receptacle of $\delta$. Similar examples can be produced in much more sophisticated circumstances and clearly indicate that the algebra $LS(\mathcal{M})$ is the most suitable object for studying unbounded derivations on a given von Neumann algebra $\mathcal{M}$. However, the study of derivations in the setting of $LS(\mathcal{M})$ has been greatly impeded by the fact that it is not a Banach algebra (it is not even a Frechet algebra or locally convex algebra when endowed with its natural topology). An additional difficulty (especially, in comparison with rather well studied algebras $S(\mathcal{M},\tau)$) is represented by the lack of developed analytical techniques in $LS(\mathcal{M})$. Only recently, in the series of papers [@AAK; @AK; @BdPS; @BCS; @Ber] meaningful attempts have been made to study the structure of derivations on such algebras. Of particular interest is the problem of identifying the class of von Neumann algebras, for which any derivation of the $*$-algebra $LS(\mathcal{M})$ is inner. In the setting of commutative $W^*$-algebras (respectively, commutative $AW^*$-algebras) this problem is fully resolved in [@BCS] (respectively, in [@Kusraev]). In the setting of von Neumann algebras of type $I$, a thorough treatment of this problem may be found in [@AAK] and [@BdPS]. The papers [@AAK; @BCS] contain examples of non-inner derivations of the $*$-algebra $LS(\mathcal{M})$, which are not continuous with respect to the topology $t(\mathcal{M})$ of local convergence in measure on $LS(\mathcal{M})$. The latter topology is the only topology considered on algebras $LS(\mathcal{M})$ to date, it may be also viewed as a noncommutative generalization of the classical topology of convergence in measure on the sets of finite measure in the case when $\mathcal{M}$ is given by the algebra $L^\infty(\Omega,\Sigma,\mu)$, where $(\Omega,\Sigma,\mu)$ is a $\sigma$-finite measure space (in this case the algebra $LS(\mathcal{M})$ coincides with the algebra of all measurable complex functions on $\Omega$). See details in Section 2 below. On the other hand, it is shown in [@AAK] that in the special case when $\mathcal{M}$ is a properly infinite von Neumann algebra of type $I$, every derivation of $LS(\mathcal{M})$ is continuous with respect to the local measure topology $t(\mathcal{M})$. Moreover, all such derivations are inner. Using a completely different technique, a similar result was also obtained in [@BdPS] under the additional assumption that the predual space $\mathcal{M}_*$ to $\mathcal{M}$ is separable. It is of interest to observe that an analogue of this result (that is the continuity of an arbitrary derivation of $(LS(\mathcal{M}),t(\mathcal{M}))$) also holds for any von Neumann algebra $\mathcal{M}$ of type $III$ [@AK]. In [@AK] the following problem is formulated (Problem 3): Let $\mathcal{M}$ be a von Neumann algebra of type $II$ and let $\tau$ be a faithful, normal, semifinite trace on $\mathcal{M}$. Is any derivation of a $*$-algebra $S(\mathcal{M},\tau)$ equipped with the (classical) measure topology generated by the trace necessarily continuous? In [@Ber] this problem is solved affirmatively for a properly infinite algebra $\mathcal{M}$. In view of the example we mentioned above, a natural problem (analogous to Problem 3 from [@AK]) is whether any derivation in a $*$-algebra $LS(\mathcal{M})$ is necessarily continuous with respect to the topology $t(\mathcal{M})$, where $\mathcal{M}$ is a properly infinite von Neumann algebra of type $II$. The main results of this paper provide an affirmative solution to this problem. In fact, we establish a much stronger result that any derivation $\delta: \mathcal{A}\longrightarrow LS(\mathcal{M})$, where $\mathcal{A}$ is any subalgebra in $LS(\mathcal{M})$ containing the algebra $\mathcal{M}$, is necessarily continuous with respect to the topology $t(\mathcal{M})$. The proof proceeds in two stages. Firstly, we establish the $t(\mathcal{M})$-continuity of any derivation $\delta: LS(\mathcal{M})\longrightarrow LS(\mathcal{M})$ for a properly infinite von Neumann algebra $\mathcal{M}$ (section 3). Then, in Section 4, a special construction of extension of a derivation $\delta: \mathcal{M}\longrightarrow LS(\mathcal{M})$ up to a derivation defined on the whole algebra $LS(\mathcal{M})$ is given (here $\mathcal{M}$ is actually an arbitrary von Neumann algebra). We also hope that our approach to unbounded derivations on $\mathcal{M}$ as well as techniques developed for dealing with locally measurable operators and the topology of local convergence in measure in this paper are of interest in their own right and may be used elsewhere. We use terminology and notations from von Neumann algebra theory [@Sak; @Tak] and theory of locally measurable operators from [@MCh; @San; @Yead]. Preliminaries ============= Let $H$ be a Hilbert space, let $B(H)$ be the $*$-algebra of all bounded linear operators on $H$, and let $\mathbf{1}$ be the identity operator on $H$. Given a von Neumann algebra $\mathcal{M}$ acting on $H$, denote by $\mathcal{Z}(\mathcal{M})$ the center of $\mathcal{M}$ and by $\mathcal{P}(\mathcal{M})=\{p\in\mathcal{M}:\ p=p^2=p^*\}$ the lattice of all projections in $\mathcal{M}$. Let $P_{fin}(\mathcal{M})$ be the set of all finite projections in $\mathcal{M}$. Denote by $\tau_{so}$ the strong operator topology on $B(H)$, that is the locally convex topology generated by the family of seminorms $p_\xi(x)=\|x\xi\|_H,\xi\in H$, where $\|\cdot\|_H$ is the Hilbert norm on $H$. A linear operator $x:\mathfrak{D}\left( x\right) \rightarrow H $, where the domain $\mathfrak{D}\left( x\right) $ of $x$ is a linear subspace of $H$, is said to be [*affiliated*]{} with $\mathcal{M}$ if $yx\subseteq xy$ for all $y$ from the commutant $\mathcal{M}^{\prime }$ of algebra $\mathcal{M}$. A densely-defined closed linear operator $x$ (possibly unbounded) affiliated with $\mathcal{M}$ is said to be *measurable* with respect to $\mathcal{M}$ if there exists a sequence $\{p_n\}_{n=1}^\infty\subset \mathcal{P}(\mathcal{M})$ such that $p_n\uparrow \mathbf{1},\ p_n(H)\subset \mathfrak{D}(x)$ and $p_n^\bot=\mathbf{1}-p_n\in P_{fin}(\mathcal{M})$ for every $n\in\mathbb{N}$, where $\mathbb{N}$ is the set of all natural numbers. Let us denote by $S(\mathcal{M})$ the set of all measurable operators. Let $x,y\in S(\mathcal{M})$. It is well known that $x+y,\ xy$ and $x^*$ are densely-defined and preclosed operators. Moreover, the closures $\overline{x+y}$ (strong sum), $\overline{xy}$ (strong product) and $x^*$ are also measurable, and equipped with this operations (see [@Seg]) $S(\mathcal{M})$ is a unital $*$-algebra over the field $\mathbb{C}$ of complex numbers. It is clear that $\mathcal{M}$ is a $*$-subalgebra of $S(\mathcal{M})$. A densely-defined linear operator $x$ affiliated with $\mathcal{M}$ is called *locally measurable* with respect to $\mathcal{M}$ if there is a sequence $\{z_n\}_{n=1}^\infty$ of central projections in $\mathcal{M}$ such that $z_n \uparrow \mathbf{1}$ and $z_nx\in S(\mathcal{M})$ for all $n\in\mathbb{N}$. The set $LS(\mathcal{M})$ of all locally measurable operators (with respect to $\mathcal{M}$) is a unital $*$-algebra over the field $\mathbb{C}$ with respect to the same algebraic operations as in $S(\mathcal{M})$ [@Yead] and $S(\mathcal{M})$ is a $*$-subalgebra of $LS(\mathcal{M})$. If $\mathcal{M}$ is finite, or if $\dim(\mathcal{Z}(\mathcal{M}))<\infty$, the algebras $S(\mathcal{M})$ and $LS(\mathcal{M})$ coincide [@MCh Corollary 2.3.5 and Theorem 2.3.16]. If von Neumann algebra $\mathcal{M}$ is of type $III$ and $\dim(\mathcal{Z}(\mathcal{M}))=\infty$, then $S(\mathcal{M})=\mathcal{M}$ and $LS(\mathcal{M})\neq \mathcal{M}$ [@MCh Theorem 2.2.19, Corollary 2.3.15,]. For every subset $E\subset LS(\mathcal{M})$, the sets of all self-adjoint (resp., positive) operators in $E$ will be denoted by $E_h$ (resp. $E_+$). The partial order in $LS(\mathcal{M})$ is defined by its cone $LS_+(\mathcal{M})$ and is denoted by $\leq$. We shall need the following important property of the $*$-algebra $LS(\mathcal{M})$. Let $\{z_i\}_{i\in I}$ be a family of pairwise orthogonal non-zero central projections from $\mathcal{M}$ with $\sup_{i\in I}z_i=\mathbf{1}$, where $I$ is an arbitrary set of indices (in this case, the family $\{z_i\}_{i\in I}$ is called a central decomposition of the unity $\mathbf{1}$). Consider the $*$-algebra $\prod_{i\in I}LS(z_i\mathcal{M})$ with the coordinate-wise operations and involution and set $$\phi:\ LS(\mathcal{M})\rightarrow \prod_{i\in I} LS(z_i\mathcal{M}),\ \phi(x):=\{z_ix\}_{i\in I}.$$ [@MCh],[@Saito]. \[Saito\] The mapping $\phi$ is a $*$-isomorphism from $LS(\mathcal{M})$ onto $\prod_{i\in I}LS(z_i\mathcal{M})$. Observe that the analogue of Proposition \[Saito\] for the $*$-algebra $S(\mathcal{M})$ does not hold in general [@MCh §2.3]. Proposition \[Saito\] implies that given any central decomposition $\{z_i\}_{i\in I}$ of the unity and any family of elements $\{x_i\}_{i\in I}$ in $LS(\mathcal{M})$, there exists a unique element $x\in LS(\mathcal{M})$ such that $z_ix=z_ix_i$ for all $i\in I$. This element is denoted by $x=\sum_{i\in I}z_ix_i$. It is shown in [@MCh2] that if $\mathcal{M}$ is of type $I$ or $III$, then for any $x\in LS(\mathcal{M})$ there exists a countable central decomposition of unity $\{z_n\}_{n=1}^\infty$, such that $x=\sum_{n=1}^\infty z_n x$ and $z_nx\in \mathcal{M}$ for all $n\in \mathbb{N}$. Let $x$ be a closed operator with dense domain $\mathfrak{D}(x)$ in $H$, let $x=u|x|$ be the polar decomposition of the operator $x$, where $|x|=(x^*x)^{\frac{1}{2}}$ and $u$ is a partial isometry in $B(H)$ such that $u^*u$ is the right support $r(x)$ of $x$. It is known that $x\in LS(\mathcal{M})$ (respectively, $x\in S(\mathcal{M})$) if and only if $|x|\in LS(\mathcal{M})$ (respectively, $|x|\in S(\mathcal{M})$) and $u\in \mathcal{M}$ [@MCh §§2.2,2.3]. If $x$ is a self-adjoint operator affiliated with $\mathcal{M}$, then the spectral family of projections $\{E_\lambda(x)\}_{\lambda\in \mathbf{R}}$ for $x$ belongs to $\mathcal{M}$ [@MCh §2.1]. A locally measurable operator $x$ is measurable if and only if $E_\lambda^\bot(|x|)\in \mathcal{P}_{fin}(\mathcal{M})$ for some $\lambda>0$ [@MCh §2.2]. Let us now recall the definition of the local measure topology. First let $\mathcal{M}$ be a commutative von Neumann algebra. Then $\mathcal{M}$ is $*$-isomorphic to the $*$-algebra $L^\infty(\Omega,\Sigma,\mu)$ of all essentially bounded measurable complex-valued functions defined on a measure space $(\Omega,\Sigma,\mu)$ with the measure $\mu$ satisfying the direct sum property (we identify functions that are equal almost everywhere) (see e.g. [@Tak Ch. III, §1]). The direct sum property of a measure $\mu$ means that the Boolean algebra of all projections of the $*$-algebra $L^\infty(\Omega,\Sigma,\mu)$ is order complete, and for any non-zero $ p\in \mathcal{P}(\mathcal{M})$ there exists a non-zero projection $q\leq p$ such that $\mu(q)<\infty$. The direct sum property of a measure $\mu$ is equivalent to the fact that the functional $\tau(f):=\int_\Omega f\,d\mu$ is a semi-finite normal faithful trace on the algebra $L^\infty(\Omega,\sigma,\mu)$. Consider the $*$-algebra $LS(\mathcal{M})=S(\mathcal{M})=L^0(\Omega,\Sigma,\mu)$ of all measurable almost everywhere finite complex-valued functions defined on $(\Omega,\Sigma,\mu)$ (functions that are equal almost everywhere are identified). On $L^0(\Omega,\Sigma,\mu)$, define the local measure topology $t(L^\infty(\Omega))$, that is, the Hausdorff vector topology, whose base of neighborhoods of zero is given by $$W(B,\varepsilon,\delta):= \{f\in\ L^0(\Omega,\, \Sigma,\, \mu) \colon \ \hbox{there exists a set} \ E\in \Sigma\ \mbox{such that}$$ $$E\subseteq B, \ \mu(B\setminus E)\leq\delta, \ f\chi_E \in L^\infty(\Omega,\Sigma,\mu), \ \|f\chi_E\|_{{L^\infty}(\Omega,\Sigma,\mu)}\leq\varepsilon\},$$ where $\varepsilon, \ \delta >0$, $B\in\Sigma$, $\mu(B)<\infty$, and $$\chi(\omega)=\left\{\begin{array}{rcl} 1&,& \ \ \omega\in E, \\ 0&,& \ \ \ \omega \ \not\in \ E. \end{array}\right.$$ Convergence of a net $\{f_\alpha\}$ to $f$ in the topology $t(L^\infty(\Omega))$, denoted by $f_\alpha \stackrel{t(L^\infty(\Omega))}{\longrightarrow}f$, means that $f_\alpha \chi_B \longrightarrow f\chi_B$ in measure $\mu$ for any $B\in \Sigma$ with $\mu(B)<\infty$. Note, that the topology $t(L^\infty(\Omega))$ does not change if the measure $\mu$ is replaced with an equivalent measure [@Yead]. Now let $\mathcal{M}$ be an arbitrary von Neumann algebra and let $\varphi$ be a $*$-isomorphism from $\mathcal{Z}(\mathcal{M})$ onto the $*$-algebra $L^\infty(\Omega,\Sigma,\mu)$, where $\mu$ is a measure satisfying the direct sum property. Denote by $L^+(\Omega,\, \Sigma,\, m)$ the set of all measurable real-valued functions defined on $(\Omega,\Sigma,\mu)$ and taking values in the extended half-line $[0,\, \infty]$ (functions that are equal almost everywhere are identified). It was shown in [@Seg] that there exists a mapping $$\mathcal{D}\colon \mathcal{P}(\mathcal{M})\to L^+(\Omega,\Sigma,\mu)$$ that possesses the following properties: - $\mathcal{D}(p)\in L_+^0(\Omega,\Sigma,\mu)\Longleftrightarrow p\in \mathcal{P}_{fin}(\mathcal{M})$; - $\mathcal{D}(p\vee q)=\mathcal{D}(p)+\mathcal{D}(q)$ if $pq=0$; - $\mathcal{D}(u^*u)=\mathcal{D}(uu^*)$ for any partial isometry $u\in \mathcal{M}$; - $\mathcal{D}(zp)=\varphi(z)\mathcal{D}(p)$ for any $z\in \mathcal{P}(\mathcal{Z}(\mathcal{M}))$ and $p\in \mathcal{P}(\mathcal{M})$; - if $p_\alpha, p\in \mathcal{P}(\mathcal{M})$, $\alpha\in A$ and $p_\alpha\uparrow p$, then $\mathcal{D}(p)=\sup\limits_{\alpha\in A}\mathcal{D}(p_\alpha)$. A mapping $\mathcal{D}\colon \mathcal{P}(\mathcal{M})\to L^+(\Omega,\Sigma,\mu)$ that satisfies properties (D1)—(D5) is called a *dimension function* on $\mathcal{P}(\mathcal{M})$. A dimension function $\mathcal{D}$ also has the following properties [@Seg]: - if $p_n\in\mathcal{P}(\mathcal{M})$, $n\in\mathbb{N}$, then $\mathcal{D}(\sup_{n\geq 1} p_n)\leq\sum_{n=1}^\infty\mathcal{D}(p_n)$, in addition, when $p_np_m=0$, $n\neq m$, the equality holds; - if $p_n\in\mathcal{P}_{fin}(\mathcal{M})$, $n\in\mathbb{N}$, $p_n\downarrow 0$, then $\mathcal{D}(p_n)\rightarrow 0$ almost everywhere. For arbitrary scalars $\varepsilon , \delta >0$ and a set $B\in \Sigma$, $\mu(B)<\infty$, we set $$V(B,\varepsilon, \delta ) := \{x\in LS(\mathcal{M})\colon \ \mbox{there exist} \ p\in \mathcal{P}(\mathcal{M}),\ z\in \mathcal{P}(\mathcal{Z}(\mathcal{M})),$$ $$\mbox{such that} \ xp\in \mathcal{M}, \|xp\|_{\mathcal{M}}\leq\varepsilon, \ \varphi(z^\bot) \in W(B,\varepsilon,\delta), \ \mathcal{D}(zp^\bot)\leq\varepsilon \varphi(z)\},$$ where $\|\cdot\|_{\mathcal{M}}$ is the $C^*$-norm on $\mathcal{M}$. It was shown in [@Yead] that the system of sets $$%\label{eq_xV} \{x+V(B,\,\varepsilon,\,\delta)\colon \ x \in LS(\mathcal{M}),\ \varepsilon, \ \delta >0,\ B\in\Sigma,\ \mu(B)<\infty\}$$ defines a Hausdorff vector topology $t(\mathcal{M})$ on $LS(\mathcal{M})$ such that the sets $\{x+V(B,\,\varepsilon,\,\delta)\}$, $\varepsilon, \ \delta >0$, $B\in \Sigma$, $\mu(B)<\infty$ form a neighborhood base of an operator $x\in LS(\mathcal{M})$. It is known that $(LS(\mathcal{M}), t(\mathcal{M}))$ is a complete topological $*$-algebra, and the topology $t(\mathcal{M})$ does not depend on a choice of dimension function $\mathcal{D}$  and on the choice of $*$-isomorphism $\varphi$ (see e.g. [@MCh §3.5], [@Yead]). The topology $t(\mathcal{M})$ on $LS(\mathcal{M})$ is called the *local measure topology* (or the *topology of convergence locally in measure*). Note, that in case when $\mathcal{M}=B(H)$ the equality $LS(\mathcal{M})=\mathcal{M}$ holds [@MCh §2.3] and the topology $t(\mathcal{M})$ coincides with the uniform topology, generated by the $C^*$-norm $\|\cdot\|_{B(H)}$. We will need the following criterion for convergence of nets from $LS(\mathcal{M})$ with respect to this topology. \[plm-spk1\] (i). A net $\{p_\alpha\}_{\alpha\in A}\subset \mathcal{P}(\mathcal{M})$ converges to zero with respect to the topology $t(\mathcal{M})$ if and only if there is a net $\{z_\alpha\}_{\alpha\in A}\subset \mathcal{P}(\mathcal{Z}(\mathcal{M}))$ such that $z_\alpha p_\alpha\in \mathcal{P}_{fin}(\mathcal{M})$ for all $\alpha\in A$, $\varphi(z^\bot_\alpha) \stackrel{t(L^\infty(\Omega))}{\longrightarrow} 0$, and $\mathcal{D}(z_\alpha p_\alpha)\stackrel{t(L^\infty(\Omega))}{\longrightarrow} 0$, where $t(L^\infty(\Omega))$ is the local measure topology on $L^0(\Omega, \Sigma, \mu)$, and $\varphi$ is a $*$-isomorphism of $\mathcal{Z}(\mathcal{M})$ onto $L^\infty(\Omega,\Sigma,\mu)$. (ii). A net $\{x_\alpha\}_{\alpha\in A} \subset LS(\mathcal{M})$ converges to zero with respect to the topology $t(\mathcal{M})$ if and only if $E^\bot_\lambda(|x_\alpha|) \stackrel{t(\mathcal{M})}{\longrightarrow} 0$ for every $\lambda>0$, where $\{E_\lambda(|x_\alpha|)\}$ is the spectral projection family for the operator $|x_\alpha|$. \[rem\_plm-spk1\] It follows from Proposition \[plm-spk1\] that, if $q_\alpha,p_\alpha\in \mathcal{P(M)},q_\alpha\leq p_\alpha$ and $p_\alpha \stackrel{t(\mathcal{M})}{\longrightarrow} 0$, then $q_\alpha \stackrel{t(\mathcal{M})}{\longrightarrow} 0$. Since the involution is continuous in the topology $t(\mathcal{M})$, the set $LS_h(\mathcal{M})$ is closed in $(LS(\mathcal{M}),t(\mathcal{M}))$. The cone $LS_+(\mathcal{M})$ of positive elements is also closed in $(LS(\mathcal{M}),t(\mathcal{M}))$ [@Yead]. Proposition \[plm-spk1\] will be used in the proof of the following convergence criterion. \[p3\] If $x_\alpha\in LS(\mathcal{M}),\ 0\neq z \in \mathcal{P}(\mathcal{Z}(\mathcal{M}))$, then $$zx_\alpha\stackrel{t(\mathcal{M})}{\longrightarrow} 0\Longleftrightarrow zx_\alpha\stackrel{t(z\mathcal{M})}{\longrightarrow} 0.$$ Fix a $*$-isomorphism $\varphi: \mathcal{Z}(\mathcal{M})\to L^\infty(\Omega,\Sigma,\mu)$ and $0\neq z\in \mathcal{P}(\mathcal{Z}(\mathcal{M}))$. Let $E\in \Sigma$ be such that $\varphi(z)=\chi_E$. Define the mapping $$\psi:\mathcal{Z}(z\mathcal{M})=z\mathcal{Z}(\mathcal{M})\to L^\infty(E,\Sigma_E, \mu|_E)$$ by setting $$\psi(za)=\varphi(za)|_E, \ \hbox{for} \ a\in \mathcal{Z}(\mathcal{M}).$$ Here, $\Sigma_E:=\{A\cap E: A\in \Sigma\}$ and $\mu|_E$ is the restriction of $\mu$ to $\Sigma_E$. It is clear that $\psi$ is a $*$-isomorphism. Now define $\mathcal{D}_z:\mathcal{P}(z\mathcal{M})\rightarrow L_+(E,\Sigma_E, \mu|_E)$ by setting $\mathcal{D}_z(q)=\mathcal{D}(q)|_E$ for $q\in \mathcal{P}(z\mathcal{M})$. It is straightforward that $\mathcal{D}_z$ is a dimension function on $\mathcal{P}(z\mathcal{M})$. Let $\{q_\alpha\}_{\alpha\in A}\subset \mathcal{P}(z\mathcal{M})$. We claim $$q_\alpha \stackrel{t(\mathcal{M})}{\longrightarrow} 0\Longleftrightarrow q_\alpha \stackrel{t(z\mathcal{M})}{\longrightarrow}0.$$ To see the claim, assume that the first convergence holds and observe that by Proposition \[plm-spk1\]$(i)$, there exists a net $\{z_\alpha\}_{\alpha\in A}\subset \mathcal{P}(\mathcal{Z}(\mathcal{M}))$ such that $z_\alpha q_\alpha \in \mathcal{P}_{fin}(\mathcal{M})$ for any $\alpha\in A$, $\varphi(z_\alpha^\bot)\stackrel{t(L^\infty(\Omega))}{\longrightarrow}0$, and $\mathcal{D}(z_\alpha q_\alpha)\stackrel{t(L^\infty(\Omega))}{\longrightarrow}0$. The projection $r_\alpha=zz_\alpha$ belongs to the center $\mathcal{Z}(z\mathcal{M})$ of the von Neumann algebra $z\mathcal{M}$, and $r_\alpha q_\alpha=z_\alpha q_\alpha$ is a finite projection in $z\mathcal{M}$ for each $\alpha\in A$. Also $$\psi(z-r_\alpha)=\psi(z(\mathbf{1}-z_\alpha))=\varphi(z z_\alpha^\bot)|_E=\varphi(z)\varphi( z_\alpha^\bot)|_E\stackrel{t(L^\infty(E))}{\longrightarrow}0,$$ where $t(L^\infty(E))$ is the local measure topology on $L^0(E,\Sigma_E,\mu|_E)$, and $$\mathcal{D}_z(r_\alpha q_\alpha)=\mathcal{D}_z(z_\alpha q_\alpha)=\mathcal{D}(z_\alpha q_\alpha)|_E\stackrel{t(L^\infty(E))}{\longrightarrow}0.$$ Hence, by Proposition \[plm-spk1\](i) we get that $q_\alpha \stackrel{t(z\mathcal{M})}{\longrightarrow}0$. We will show now that the convergence $q_\alpha \stackrel{t(z\mathcal{M})}{\longrightarrow}0$ for $\{q_\alpha\}_{\alpha\in A}\subset \mathcal{P}(z\mathcal{M})$ implies the convergence $q_\alpha\stackrel{t(\mathcal{M})}{\longrightarrow}0$. Let $\{r_\alpha\}_{\alpha\in A}$ be a net in $\mathcal{P}(\mathcal{Z}(z\mathcal{M}))$ such that $r_\alpha q_\alpha\in \mathcal{P}_{fin}(z\mathcal{M})$ for every $\alpha\in A$, $$\psi(z-r_\alpha)\stackrel{t(L^\infty(E))}{\longrightarrow}0$$ and $$\mathcal{D}_z(r_\alpha q_\alpha)\stackrel{t(L^\infty(E))}{\longrightarrow}0.$$ Put $z_\alpha=z^\bot+r_\alpha$. Then $z_\alpha\in \mathcal{P}(\mathcal{Z}(\mathcal{M}))$ and $z_\alpha q_\alpha= r_\alpha q_\alpha\in\mathcal{P}_{fin}(\mathcal{M}) $. Since $z^\bot_\alpha=z(\mathbf{1}-r_\alpha)$, we have $\varphi(z_\alpha^\bot)=\chi_E\varphi(z_\alpha^\bot)$ and $$\varphi(z^\bot_\alpha)|_E=\chi_E\varphi(z(\mathbf{1}-r_\alpha))|_E=\chi_E\psi(z-r_\alpha)\stackrel{t(L^\infty(E))}{\longrightarrow}0.$$ Also $$\mathcal{D}(z_\alpha q_\alpha)=\mathcal{D}(zr_\alpha q_\alpha)=\chi_E\mathcal{D}(r_\alpha q_\alpha),$$ and so $\mathcal{D}(z_\alpha q_\alpha)\stackrel{t(L^\infty(\Omega))}{\longrightarrow}0$, since $\mathcal{D}(r_\alpha q_\alpha)|_E=\mathcal{D}_z(r_\alpha q_\alpha)\stackrel{t(L^\infty(E))}{\longrightarrow}0$. Again appealing to Proposition \[plm-spk1\](i), we conclude that $q_\alpha\stackrel{t(\mathcal{M})}{\longrightarrow}0$. Now let $\{x_\alpha\}\subset LS(z\mathcal{M})$ and $x_\alpha \stackrel{t(\mathcal{M})}{\longrightarrow}0$. By Proposition \[plm-spk1\](ii), we have that $E^\bot_\lambda(|x_\alpha|) \stackrel{t(\mathcal{M})}{\longrightarrow}0$ for any $\lambda>0$, where $\{E_\lambda(|x_\alpha|)\}$ is the spectral family for $|x_\alpha|$. Denote by $\{E^z_\lambda(|x_\alpha|)\}$ the family of spectral projections for $|x_\alpha|$ in $LS(z\mathcal{M})$, $\lambda>0$. It is clear that $E_\lambda(|x_\alpha|)=z^\bot+E^z_\lambda(|x_\alpha|)$ and $E^\bot_\lambda(|x_\alpha|)=z-E^z_\lambda(|x_\alpha|)$ for all $\lambda>0$. It follows from above that $z-E^z_\lambda(|x_\alpha|) \stackrel{t(z\mathcal{M})}{\longrightarrow}0$ for all $\lambda>0$. Hence, by Proposition \[plm-spk1\](ii), it follows that $x_\alpha \stackrel{t(z\mathcal{M})}{\longrightarrow}0$. The proof of the implication $x_\alpha \stackrel{t(z\mathcal{M})}{\longrightarrow}0\Longrightarrow x_\alpha \stackrel{t(\mathcal{M})}{\longrightarrow}0$ is similar and therefore omitted. The lattice $\mathcal{P}(\mathcal{M})$ is said to have a countable type, if every family of non-zero pairwise orthogonal projections in $\mathcal{P}(\mathcal{M})$ is, at most, countable. A von Neumann algebra is said to be $\sigma$-finite, if the lattice $\mathcal{P(M)}$ has a countable type. It is shown in [@Seg Lemma 1.1] that a finite von Neumann algebra $\mathcal{M}$ is $\sigma$-finite, provided that the lattice $\mathcal{P}(\mathcal{Z}(\mathcal{M}))$ of central projections has a countable type. If $\mathcal{M}$ is a commutative von Neumann algebra and $\mathcal{P}(\mathcal{M})$ has a countable type, then $\mathcal{M}$ is $*$-isomorphic to a $*$-algebra $L^\infty(\Omega,\Sigma,\mu)$ with $\mu(\Omega)<\infty$. In this case, the topology $t(L^\infty(\Omega))$ is metrizable and has a base of neighborhoods of $0$ consisting of the sets $W(\Omega,1/n,1/n),\ n\in\mathbb{N}$. In addition, $f_n \stackrel{t(L^\infty(\Omega))}{\longrightarrow}0\Leftrightarrow f_n\longrightarrow 0$ in measure $\mu$, where $f_n,f\in L^0(\Omega,\Sigma,\mu)=LS(\mathcal{M})$. Let $\mathcal{M}$ be a commutative von Neumann algebra such that $\mathcal{P}(\mathcal{M})$ does not have a countable type. Denote by $\varphi$ a $*$-isomorphism from $\mathcal{M}$ on $L^\infty(\Omega,\Sigma,\mu)$, where $\mu$ is a measure with the direct sum property. Due to the latter property, there exists a family $\{z_i\}_{i\in I}$ of non-zero pairwise orthogonal projections from $\mathcal{P}(\mathcal{M})$, such that $\sup_{i\in I} z_i=\mathbf{1}$ and $\mu(\varphi(z_i))<\infty$ for all $i\in I$, in particular, $\mathcal{P}(z_i\mathcal{Z}(\mathcal{M}))$ has a countable type. Select $A_i\in\Sigma$ so that $\varphi(z_i)=\chi_{A_i}$ and set $$\Sigma_{A_i}=\{A\cap A_i:\ A\in\Sigma\},\ \mu_i(A\cap A_i)=\mu(A\cap A_i),\ i\in I.$$ Let $t(L^\infty(A_i))$ be the local measure topology on $L^0(A_i,\Sigma_{A_i},\mu_i)$. Since $\mu_i(A_i)<\infty$, we see that the topology $t(L^\infty(A_i))$ coincides with the topology of convergence in measure $\mu_i$ in $L^0(A_i,\Sigma_{A_i},\mu_i)$. \[p4\] For a net $\{f_\alpha\}_{\alpha\in A}$ and $f$ from $L^0(\Omega,\Sigma,\mu)$ the following conditions are equivalent: (i). $f_\alpha\stackrel{t(L^\infty(\Omega))}{\longrightarrow} f$ ; (ii). $f_\alpha\chi_{A_i}\stackrel{t(L^\infty(A_i))}{\longrightarrow} f\chi_{A_i}$ for all $i\in I$. The implication $(i)\Rightarrow (ii)$ follows from the definitions of topologies $t(L^\infty(\Omega))$ and $t(L^\infty(A_i))$. $(ii)\Rightarrow (i)$. It is sufficient to consider the case when $f=0$. Consider the set $\Gamma$ of all finite subsets $\gamma$ from $I$ and order it with respect to inclusion. Consider an increasing net $\chi_{D_\gamma}\uparrow \chi_\Omega$ in $L^0_h(\Omega,\Sigma,\mu)$, where $D_\gamma=\bigcup_{i\in\gamma} A_i,\ \gamma\in\Gamma$. Take an arbitrary neighborhood of zero $U$ (in the topology $t(L^\infty(\Omega))$ ) and select $W(B,\varepsilon,\delta)$ in such a way that $W(B,\varepsilon,\delta)+W(B,\varepsilon,\delta)\subset U$. Since $\mu(B\cap D_\gamma)\uparrow \mu(B)<\infty$, we can locate such $\gamma_0\in\Gamma$ that $\mu(B\setminus D_{\gamma_0})\leq \delta$. Hence, $f_\alpha\chi_{\Omega\setminus D_{\gamma_0}}\in W(B,\varepsilon,\delta)$ for all $\alpha\in A$. Since $f_\alpha\chi_{A_i}\stackrel{t(L^\infty(A_i))}{\longrightarrow} 0$ for all $i\in\gamma_0$ and $\gamma_0$ is a finite set, it follows $f_\alpha\chi_{D_{\gamma_0}}=\sum_{i\in\gamma_0}f_\alpha\chi_{A_i}\stackrel{t(L^\infty(\Omega))}{\longrightarrow} 0$. Thus, there exists such $\alpha_0\in A$ that $f_\alpha\chi_{D_{\gamma_0}}\in W(B,\varepsilon,\delta)$ for all $\alpha\geq \alpha_0$. In particular, $$f_\alpha=f_\alpha\chi_{D_{\gamma_0}}+f_\alpha\chi_{\Omega\setminus D_{\gamma_0}}\in W(B,\varepsilon,\delta)+W(B,\varepsilon,\delta)\subset U,\ \forall\alpha\geq\alpha_0,$$ which implies the convergence $f_\alpha\stackrel{t(L^\infty(\Omega))}{\longrightarrow} 0$. Let us now establish a variant of Proposition \[p4\] for an arbitrary von Neumann algebra $\mathcal{M}$. Let $\varphi$ be a $*$-isomorphism from $\mathcal{Z}(\mathcal{M})$ onto $L^\infty(\Omega,\Sigma,\mu)$ and let $\{z_i\}_{i\in I}$ be a central decomposition of the unity. As before, we denote $\Gamma$ the directed set of all finite subsets from $I$. For every $\gamma\in\Gamma$ we set $z^{(\gamma)}=\sum_{i\in\gamma} z_i$. Since $\varphi(z_i)=\chi_{A_i}$ for some $A_i\in\Sigma$, we see that $\varphi(z^{(\gamma)})=\chi_{D_\gamma}$, where $D_\gamma=\bigcup_{i\in\gamma}A_i$, and, in addition, $z^{(\gamma)} \uparrow \mathbf{1}$ which implies $z^{(\gamma)}\stackrel{t(\mathcal{M})}{\longrightarrow} \mathbf{1}$ (see Proposition \[plm-spk1\] (i) for $p_\alpha=z_\alpha^\bot$ ). As it the proof of Proposition \[p4\] for a given $V(B,\varepsilon,\delta)$ we choose $\gamma_0\in\Gamma$ such that $x(\mathbf{1}-z^{(\gamma_0)})\in V(B,\varepsilon,\delta)$ for every $x\in LS(\mathcal{M})$. If $x_\alpha\in LS(\mathcal{M})$ and $x_\alpha z_i\stackrel{t(z_i\mathcal{M})}{\longrightarrow} 0$ for all $i\in I$, then by Proposition \[p3\], we have $x_\alpha z_i \stackrel{t(\mathcal{M})}{\longrightarrow} 0$ for all $i\in I$, and so $x_\alpha z^{(\gamma_0)}=\sum_{i\in\gamma_0} x_\alpha z_i\stackrel{t(\mathcal{M})}{\longrightarrow} 0$. Hence, there exists such $\alpha_0\in A$ that $x_\alpha z^{(\gamma_0)}\in V(B,\varepsilon,\delta)$ for all $\alpha\geq\alpha_0$. This means that $$x_\alpha=x_\alpha z^{(\gamma_0)}+x_\alpha (\mathbf{1}-z^{(\gamma_0)})\in V(B,\varepsilon,\delta)+V(B,\varepsilon,\delta)\subset V(B,2\varepsilon,2\delta).$$ The argument above justifies the following result. \[p5\] Let $\mathcal{M}$ be an arbitrary von Neumann algebra, $x_\alpha,x\in LS(\mathcal{M})$, $0\neq z_i\in \mathcal{P}(\mathcal{Z}(\mathcal{M})),\ z_iz_j=0$ when $i\neq j$, $\sup_{i\in I} z_i=\mathbf{1}$. The following conditions are equivalent: (i). $x_\alpha\stackrel{t(\mathcal{M})}{\longrightarrow} x$; (ii). $z_ix_\alpha\stackrel{t(z_i\mathcal{M})}{\longrightarrow} z_ix$ for any $i\in I$. \[r1\] From Propositions \[Saito\] and \[p5\] it follows that the topology $t(\mathcal{M})$ coincides with the Tikhonov product of topologies $t(z_i\mathcal{M}),\ i\in I$. In particular, the isomorphism $\phi$ from Proposition \[Saito\] is a topological $*$-isomorphism from $(LS(\mathcal{M}),t(\mathcal{M}))$ onto $\prod_{i\in I}(LS(z_i\mathcal{M}),t(z_i\mathcal{M}))$. Let $\{z_i\}_{i\in I}$ be the same as in the assumption of Proposition \[p5\], let $T: LS(\mathcal{M})\rightarrow LS(\mathcal{M})$ be a linear operator such that $T(z_ix)=z_iT(x)$ for all $x\in LS(\mathcal{M}),\ i\in I$. It is clear that $T_{z_i}(y):=T(y)$, $y\in LS(z_i\mathcal{M})$ is a linear operator acting in $LS(z_i\mathcal{M})$. Due to Proposition \[p5\], the next result follows immediately. \[c1\] Let $\mathcal{M}$ and let $\{z_i\}_{i\in I}$ satisfy the same assumptions of Proposition \[p5\], and let $T: LS(\mathcal{M})\rightarrow LS(\mathcal{M})$ be a linear operator such that $T(z_ix)=z_iT(x)$ for all $x\in LS(\mathcal{M}),\ i\in I$. The following conditions are equivalent: (i). The mapping $T: (LS(\mathcal{M}),t(\mathcal{M}))\rightarrow (LS(\mathcal{M}),t(\mathcal{M}))$ is continuous; (ii). The mapping $T_{z_i}: (LS(z_i\mathcal{M}),t(z_i\mathcal{M}))\rightarrow (LS(z_i\mathcal{M}),t(z_i\mathcal{M}))$ is continuous for every $i\in I$. Continuity of derivations in $*$-algebra $LS(\mathcal{M})$ ========================================================== Let $\mathcal{M}$ be an arbitrary von Neumann algebra, let $\mathcal{A}$ be a subalgebra in $LS(\mathcal{M})$. A linear mapping $\delta: \mathcal{A} \rightarrow LS(\mathcal{M})$ is called a *derivation* on $\mathcal{A}$ with values in $LS(\mathcal{M})$, if $\delta(xy)=\delta(x)y+x\delta(y)$ for all $x,y\in \mathcal{A}$. Each element $a\in \mathcal{A}$ defines a derivation $\delta_a(x):=ax-xa$ on $\mathcal{A}$ with values in $\mathcal{A}$. Derivations $\delta_a,a\in\mathcal{A}$ are said to be *inner derivations* on $\mathcal{A}$. Since the operation of multiplication is continuous with respect to the topology $t(\mathcal{M})$, it immediately follows that any inner derivation of $\mathcal{A}$ is continuous with respect to the topology $t(\mathcal{M})$. Now, we list a few properties of derivations on $\mathcal{A}$ which we shall need below. \[l1\] If $\mathcal{P}(\mathcal{Z}(\mathcal{M}))\subset \mathcal{A}$, $\delta$ is a derivation on $\mathcal{A}$ and $z\in\mathcal{P}(\mathcal{Z}(\mathcal{M}))$, then $\delta(z)=0$ and $\delta(zx)=z\delta(x)$ for all $x\in \mathcal{A}$. We have that $\delta(z)=\delta(z^2)=\delta(z)z+z\delta(z)=2z\delta(z)$. Hence, $z\delta(z)=z(2z\delta(z))=2z\delta(z)$, that is $z\delta(z)=0$. Therefore, we have $\delta(z)=0$. In particular, $\delta(zx)=\delta(z)x+z\delta(x)=z\delta(x)$. Let $\mathcal{A}$ be an $*$-subalgebra in $LS(\mathcal{M})$, let $\delta$ be a derivation on $\mathcal{A}$ with values in $LS(\mathcal{M})$. Let us define a mapping $$\delta^*: \mathcal{A}\rightarrow LS(\mathcal{M}),$$ by setting $\delta^*(x)=(\delta(x^*))^*$, $x\in \mathcal{A}$. A direct verification shows that $\delta^*$ is also a derivation on $\mathcal{A}$. A derivation $\delta$ on $\mathcal{A}$ is said to be *self-adjoint*, if $\delta=\delta^*$. Every derivation $\delta$ on $\mathcal{A}$ can be represented in the form $\delta= Re(\delta)+ i Im(\delta)$, where $Re(\delta)=(\delta+\delta^*)/2,\ Im(\delta)=(\delta-\delta^*)/2i$ are self-adjoint derivations on $\mathcal{A}$. Since $(LS(\mathcal{M}),t(\mathcal{M}))$ is a topological $*$-algebra, the following result holds. \[l2\] A derivation $\delta: \mathcal{A}\rightarrow LS(\mathcal{M})$ is continuous with respect to the topology $t(\mathcal{M})$ if and only if the self-adjoint derivations $Re(\delta)$ and $Im(\delta)$ are continuous with respect to that topology. As we already stated in the introduction, in the special case, when $\mathcal{M}$ is a properly infinite von Neumann algebra of type $I$ or von Neumann algebra of type $III$, any derivation of the algebra $LS(\mathcal{M})$ is continuous with respect to the topology $t(\mathcal{M})$ [@AK]. The next theorem extends this result to an arbitrary properly infinite von Neumann algebra. \[main\] If $\mathcal{M}$ properly infinite von Neumann algebras, then any derivation $\delta: LS(\mathcal{M})\rightarrow LS(\mathcal{M})$ is continuous with respect to the topology $t(\mathcal{M})$ of local convergence in measure. By Lemma \[l2\], we may assume that $\delta^*=\delta$. Since $\mathcal{Z}(\mathcal{M})$ is a commutative von Neumann algebra, there exists a system $\{z_i\},\ i\in I$ of non-zero pairwise orthogonal projections from $\mathcal{Z}(\mathcal{M})$ such that $\sup_{i\in I}z_i=\mathbf{1}$ and the Boolean algebra $\mathcal{P}(z_i\mathcal{Z}(\mathcal{M}))$ has a countable type for all $i\in I$. By Lemma \[l1\] we have that $\delta(z_ix)=z_i\delta(x)$ for all $x\in LS(\mathcal{M}), i\in I$. Therefore, by Corollary \[c1\], it is sufficient to prove that each derivation $\delta_{z_i}$ is $t(z_i\mathcal{M})$-continuous, $i\in I$. Thus, we may assume without loss of generality that the Boolean algebra $\mathcal{P}(\mathcal{Z}(\mathcal{M}))$ has a countable type. In this case the topology $t(\mathcal{M})$ is metrizable, and the sets $V(\Omega,1/n,1/n)$, $n\in\mathbb{N}$ form a countable base of neighborhoods of $0$; in particular, $(LS(\mathcal{M}),t(\mathcal{M}))$ is an $F$-space. Therefore it is sufficient to show that the graph of the linear operator $\delta$ is a closed set. Arguing by a contradiction, let us assume that the graph of $\delta$ is not closed. This means that there exists a sequence $\{x_n\}\subset LS(\mathcal{M})$, such that $x_n\stackrel{t(\mathcal{M})}{\longrightarrow} 0$ and $\delta(x_n)\stackrel{t(\mathcal{M})}{\longrightarrow} x\neq 0$. Recalling that $(LS(\mathcal{M}),t(\mathcal{M}))$ is a topological $*$-algebra and that $\delta=\delta^*$, we may assume that $x=x^*,\ x_n=x_n^*$ for all $n\in\mathbb{N}$. In this case, $x=x_+-x_-$, where $x_+,x_-\in LS_+(\mathcal{M})$ are respectively the positive and negative parts of $x$. Without loss of generality, we shall also assume that $x_+\neq 0$, otherwise, instead of the sequence $\{x_n\}$ we consider the sequence $\{-x_n\}$. Let us select scalars $0<\lambda_1<\lambda_2$ so that the projection $$p:=E_{\lambda_2}(x)-E_{\lambda_1}(x)$$ does not vanish. We have that $0<\lambda_1 p\leq pxp=px\leq\lambda_2 p$ and $\|px\|_\mathcal{M}\leq \lambda_2$. Replacing, if necessary, $x_n$ on $x_n/\lambda_1$, we may assume that $$\label{pxp} pxp\geq p.$$ By the assumption, $\mathcal{M}$ is a properly infinite von Neumann algebra and therefore, there exist pairwise orthogonal projections $\{p_m^{(1)}\}_{m=1}^\infty\subset \mathcal{P}(\mathcal{M})$, such that $\sup_{m\geq 1} p_m^{(1)}=\mathbf{1}$ and $p_m^{(1)}\sim \mathbf{1}$ for all $m\in\mathbb{N}$, in particular, $p\preceq p_m^{(1)}$. Here, the notation $p\sim q$ denotes the equivalence of projections $p,q\in\mathcal{P}(\mathcal{M})$, and the notation $p\preceq q$ means that there exists a projection $e\leq q$ such that $p\sim e$. In course of the proof of our main result we shall frequently use the following well-known fact: if $p\sim q$ and $z\in\mathcal{P}(\mathcal{Z}(\mathcal{M}))$ then $pz\sim qz$. For every $m\in\mathbb{N}$ we select a projection $p_m\leq p_m^{(1)}$, for which $p_m\sim p$ and denote by $v_m$ a partial isometry from $\mathcal{M}$ such that $v_m^*v_m=p,\ v_mv_m^*=p_m$. Clearly, we have $p_mp_k=0$ whenever $m\neq k$ and the projection $$\label{p_0} p_0:=\sup_{m\geq 1}p_m$$ is infinite as the supremum of pairwise orthogonal and equivalent projections. Taking into account that $$p_m=v_mpv_m^*\stackrel{(\ref{pxp})}\leq v_mpxpv_m^*=v_mxv_m^*\in p_m\mathcal{M}p_m,$$ and $$\|v_mxv_m^*\|_{\mathcal{M}}=\|v_mpxpv_m^*\|_{\mathcal{M}}\leq\|pxp\|_{\mathcal{M}}\leq \lambda_2,$$ we see that the series $\sum_{m=1}^\infty v_mxv_m^*$ converges with respect to the topology $\tau_{so}$ to some operator $y\in\mathcal{M}$ satisfying $$\label{yp_0} \|y\|_{\mathcal{M}}=\sup_{m\geq 1}\|v_mxv_m^*\|_{\mathcal{M}}\leq \|pxp\|_{\mathcal{M}},\ \text{and}\ y\geq p_0.$$ In what follows, we shall assume that the central support $c(p_0)$ of the projection $p_0$ is equal to $\mathbf{1}$ (otherwise, we replace the algebra $\mathcal{M}$ with the algebra $c(p_0)\mathcal{M}$). Let $\varphi$ be a $*$-isomorphism from $\mathcal{Z}(\mathcal{M})$ onto $L^\infty(\Omega,\Sigma,\mu)$. By the assumption, the Boolean algebra $\mathcal{P}(\mathcal{Z}(\mathcal{M}))$ has a countable type, and so we may assume that $\mu(\Omega)=\int_\Omega \mathbf{1}_{L^\infty(\Omega)} \, d\mu =1$, where $\mathbf{1}_{L^\infty(\Omega)}$ is the identity of the $*$-algebra $L^\infty(\Omega,\Sigma,\mu)$. In this case, the countable base of neighborhoods of $0$ in the topology $t(\mathcal{M})$ is formed by the sets $V(\Omega,1/n,1/n),\ n\in\mathbb{N}$. Recalling that we have $x_n\stackrel{t(\mathcal{M})}{\longrightarrow} 0$ and $\delta(x_n)\stackrel{t(\mathcal{M})}{\longrightarrow} x$, we obtain $$v_mx_nv_m^*\stackrel{t(\mathcal{M})}{\longrightarrow} 0,\ \delta(v_m)x_nv_m^*\stackrel{t(\mathcal{M})}{\longrightarrow} 0,\ v_m\delta(x_n)v_m^*\stackrel{t(\mathcal{M})}{\longrightarrow} v_mxv_m^*$$ when $n\rightarrow\infty$ for every fixed $m\in\mathbb{N}$. Fix $k\in\mathbb{N}$, and using the convergence $v_mx_nv_m^*\stackrel{t(\mathcal{M})}\longrightarrow 0$ for $n\longrightarrow 0$, for each $m\in \mathbb{N}$ select an index $n_1(m,k)$ and projections $q^{(1)}_{m,n}\in\mathcal{P}(\mathcal{M})$, $z^{(1)}_{m,n}\in \mathcal{P}(\mathcal{Z}(\mathcal{M}))$, such that $$\|v_mx_nv_m^*q^{(1)}_{m,n}\|_{\mathcal{M}}\leq 2^{-m}(k+1)^{-1};$$ $$\int_\Omega\varphi(\mathbf{1}-z^{(1)}_{m,n})d\mu\leq 3^{-1}2^{-m-k-1}$$ and $$\mathcal{D}(z^{(1)}_{m,n}(\mathbf{1}-q^{(1)}_{m,n}))\leq 3^{-1}2^{-m-k-1}\varphi(z^{(1)}_{m,n})$$ for all $n\geq n_1(m,k)$. Similarly, using the convergence $\delta(v_m)x_nv_m^*\stackrel{t(\mathcal{M})}\longrightarrow 0$ (respectively, $v_m\delta(x_n)v_m^*\stackrel{t(\mathcal{M})}\longrightarrow v_mxv_m^*$) for $n\longrightarrow\infty$, for each $m\in\mathbb{N}$ select indexes $n_2(m,k)$ and $n_3(m,k)$ and projections $q^{(2)}_{m,n},q^{(3)}_{m,n}\in\mathcal{P}(\mathcal{M})$, $z^{(2)}_{m,n},z^{(3)}_{m,n}\in \mathcal{P}(\mathcal{Z}(\mathcal{M}))$, such that $$\|\delta(v_m)x_nv_m^*q^{(2)}_{m,n}\|_{\mathcal{M}}\leq (3(k+1)2^m)^{-1}$$ (respectively, $\|(v_m\delta(x_n)v_m^*-v_mxv_m^*)q^{(3)}_{m,n}\|_{\mathcal{M}}\leq (3(k+1)2^m)^{-1}$); $$\int_\Omega\varphi(\mathbf{1}-z^{(i)}_{m,n})d\mu\leq 3^{-1}2^{-m-k-1}$$ and $\mathcal{D}(\mathbf{1}-q^{(i)}_{m,n})\leq3^{-1}2^{-m-k-1}\varphi(z^{(i)}_{m,n})$, $i=2,3$, for all $n\geq n_2(m,k)$ (respectively, $n\geq n_3(m,k)$). Set $n(m,k)=\max_{i=1,2,3} n_i(m,k)$, $z_m=\inf_{i=1,2,3} z^{(i)}_{m,n(m,k)}$, $q_m=\inf_{i=1,2,3} q^{(i)}_{m,n(m,k)}$. Due to the selection of projections $q_m\in\mathcal{P}(\mathcal{M})$, $z_m\in\mathcal{P}(\mathcal{Z}(\mathcal{M}))$ and indexes $n(m,k)$, we have that for each $m\in\mathbb{N}$ inequalities hold - $\|v_mx_{n(m,k)}v_m^*q_m\|_{\mathcal{M}}\leq 2^{-m}(k+1)^{-1}$; - $\|\delta(v_m)x_{n(m,k)}v_m^*q_m\|_{\mathcal{M}}\leq (3(k+1)2^{m})^{-1}$; - $\|q_m(v_m\delta(x_{n(m,k)})v_m^*-v_mxv_m^*)\|_{\mathcal{M}}\leq (3(k+1)2^{m})^{-1}$; - $\mathcal{D}(z_m(\mathbf{1}-q_m))\stackrel{(D6)}\leq \sum_{i=1}^3 \mathcal{D}(z_m(\mathbf{1}-q^{(i)}_{m,n(m,k)})) \leq 2^{-m-k-1}\varphi(z_m)$; - $1-\int_\Omega\varphi(z_m)d\mu=\int_\Omega\varphi(\mathbf{1}-z_m)d\mu\leq$ $\leq\sum_{i=1}^3\int_ \Omega\varphi(\mathbf{1}-z^{(i)}_{m,n(m,k)})d\mu\leq 2^{-m-k-1}.$ Fix $m_1,m_2\in\mathbb{N}$ with $m_1<m_2$ and set $$q_{m_1,m_2}:=\inf_{m_1<m\leq m_2}q_m,\ z_{m_1,m_2}:=\inf_{m_1<m\leq m_2}z_m.$$ Since $(\mathbf{1}-z_{m_1,m_2})=\sup\limits_{m_1<m\leq m_2}(\mathbf{1}-z_m)$ and $(\mathbf{1}-q_{m_1,m_2})=\sup\limits_{m_1<m\leq m_2}(\mathbf{1}-q_m)$, it follows that $\varphi(\mathbf{1}-z_{m_1,m_2})=\sup\limits_{m_1<m\leq m_2}\varphi(\mathbf{1}-z_m)$ and $\varphi(\mathbf{1}-q_{m_1,m_2})=\sup\limits_{m_1<m\leq m_2}\varphi(\mathbf{1}-q_m)$, and therefore $$\label{m1m21} 1-\int_\Omega\varphi(z_{m_1,m_2})d\mu=\int_\Omega\varphi(\mathbf{1}-z_{m_1,m_2})d\mu\leq \sum_{m=m_1+1}^{m_2}\int_\Omega\varphi(\mathbf{1}-z_m)d\mu\stackrel{(A5)} \leq 2^{-m_1-k-1};$$ $$\label{m1m22} \mathcal{D}(z_{m_1,m_2}(\mathbf{1}-q_{m_1,m_2}))\stackrel{(D6)}\leq\sum_{m=m_1+1}^{m_2}\mathcal{D}(z_{m_1,m_2}(\mathbf{1}-q_m)) \stackrel{(A4)}\leq 2^{-m_1-k-1}\varphi(z_{m_1,m_2})$$ and $$\label{m1m23} \|\sum_{m=m_1+1}^{m_2}(v_mx_{n(m,k)}v_m^*)q_{m_1,m_2}\|_{\mathcal{M}}\leq \sum_{m=m_1+1}^{m_2}\|v_mx_{n(m,k)}v_m^*q_m\|_{\mathcal{M}}\stackrel{(A1)}\leq 2^{m_1}(k+1)^{-1}.$$ Inequalities (\[m1m21\])-(\[m1m23\]) mean that the sequence $$S_{l,k}=\sum_{m=1}^lv_mx_{n(m,k)}v_m^*,\ l\geq 1$$ is a Cauchy sequence in the $F$-space $(LS(\mathcal{M},t(\mathcal{M})))$ for each fixed $k\in\mathbb{N}$. Consequently, there exists $y_k\in LS(\mathcal{M})$ such that $S_{l,k}\stackrel{t(\mathcal{M})}\longrightarrow y_k$ for $l\longrightarrow\infty$, i.e. the series $$\begin{gathered} \label{y_k} y_k=\sum_{m=1}^\infty v_mx_{n(m,k)}v_m^*\end{gathered}$$ converges in $LS(\mathcal{M})$ with respect to the topology $t(\mathcal{M})$. Since the involution is continuous in topology $t(\mathcal{M})$ and $S^*_{l,k}=S_{l,k}$, we conclude $y_k=y_k^*$. Setting $$\label{r_m} r_m:=p_m\wedge q_m, \ m\in \mathbb{N},$$ and using the relation $z_m(p_m-p_m\wedge q_m)\sim z_m(p_m\vee q_m-q_m)$ ( see e.g. [@Tak ch. 5, Proposition 1.6]) we have $$\label{Dp_mq_m} \begin{split} \mathcal{D}(z_m(p_m-r_m))&=\mathcal{D}(z_m(p_m-p_m\wedge q_m))\stackrel{(D3)}=\mathcal{D}(z_m(p_m\vee q_m - q_m))\cr \\&\leq \mathcal{D}(z_m(\mathbf{1}-q_m))\stackrel{(A4)}\leq 2^{-m-k-1}\varphi(z_m). \end{split}$$ Setting $$\label{q0z0} q_0^{(k)}:=\sup_{m\geq 1} r_m,\ z_0^{(k)}:=\inf_{m\geq 1}z_m,$$ we have (see (\[p\_0\]), (\[yp\_0\]) and (\[r\_m\])) $$\label{p0q0} y \geq p_0 \geq q_0^{(k)},\ k\in\mathbb{N}.$$ From (\[m1m21\]) it follows that $$\label{z0} 1-\int_\Omega\varphi(z_0^{(k)})d\mu=\int_\Omega\varphi(\mathbf{1}-z_0^{(k)})d\mu\leq 2^{-k-1}.$$ Since $p_mp_j=0,\ m\neq j$, and $r_m\leq p_m$ (see (\[r\_m\])) we obtain $p_0-q_0^{(k)}=\sup_{m\geq 1}(p_m-r_m)$ and hence, by (\[Dp\_mq\_m\]), $$\label{Dz_kp0q0k} \mathcal{D}(z_0^{(k)}(p_0-q_0^{(k)}))\stackrel{(D6)}=\sum_{m=1}^\infty \mathcal{D}(z_0^{(k)}(p_m-r_m))\stackrel{(\ref{Dp_mq_m})}\leq 2^{-k-1}\varphi(z_0^{(k)}).$$ Due to (\[r\_m\]), we have $p_mq^{(k)}_0=r_mq^{(k)}_0=r_m=r_mq_m$ for all $m\in \mathbb{N}$. Hence, $$v_mx_{n(m,k)}v_m^* q_0^{(k)}= v_mx_{n(m,k)}v_m^* p_mq_0^{(k)}= v_mx_{n(m,k)}v_m^* r_m$$ and $$\begin{gathered} \label{yq}\|y_k q_0^{(k)}\|_{\mathcal{M}} = \|(\sum_{m=1}^\infty v_mx_{n(m,k)}v_m^*)q_0^{(k)}\|_{\mathcal{M}}= \notag\\ = \|\sum_{m=1}^\infty v_mx_{n(m,k)}v_m^*q_0^{(k)}\|_{\mathcal{M}}\leq \sup_{m\geq 1}\| v_mx_{n(m,k)}v_m^* r_m\|_{\mathcal{M}}\leq \notag\\ \leq \sup_{m\geq 1}\| v_mx_{n(m,k)}v_m^* q_m\|_{\mathcal{M}}\stackrel{(A1)}\leq (k+1)^{-1}.\end{gathered}$$ Using the properties of the derivation $\delta$ and equalities $p_nv_n=v_n,\ v_n^*=v_n^*p_n$ and (\[r\_m\]), (\[q0z0\]), we have $$\begin{split} q_0^{(k)} & \delta(v_mx_{n(m,k)}v_m^*)q_0^{(k)}\cr &= q_0^{(k)}((\delta(v_mx_{n(m,k)}v_m^*)-v_mxv_m^*)+v_mxv_m^*)q_0^{(k)}\cr & = (q_0^{(k)}\delta(v_m)x_{n(m,k)}v_m^*q_0^{(k)}+q_0^{(k)}v_mx_{n(m,k)}\delta(v_m^*)q_0^{(k)})\cr & + q_0^{(k)}(v_m\delta(x_{n(m,k)})v_m^*-v_mxv_m^*)q_0^{(k)}+q_0^{(k)}(v_mxv_m^*)q_0^{(k)}\cr & = q_0^{(k)}\delta(v_m)x_{n(m,k)}v_m^*q_mr_m+r_mq_mv_mx_{n(m,k)}\delta(v_m^*)q_0^{(k)}\cr &+ r_mq_m(v_m\delta(x_{n(m,k)})v_m^*-v_mxv_m^*)q_mr_m+q_0^{(k)}(v_mxv_m^*)q_0^{(k)}. \end{split}$$ Consider the following formal series suggested by the preceding $$\label{s1} \sum_{m=1}^\infty q_0^{(k)}\delta(v_m)x_{n(m,k)}v_m^*q_mr_m;$$ $$\label{s2} \sum_{m=1}^\infty r_mq_mv_mx_{n(m,k)}\delta(v_m^*)q_0^{(k)};$$ $$\label{s3} \sum_{m=1}^\infty r_mq_m(v_m\delta(x_{n(m,k)})v_m^*-v_mxv_m^*)q_mr_m;$$ $$\label{s4} \sum_{m=1}^\infty q_0^{(k)}(v_mx_{n(m,k)}v_m^*)q_0^{(k)}.$$ By the condition (A2) the first series (\[s1\]) and the second series (\[s2\]) converge with respect to the norm $\|.\|_{\mathcal{M}}$ to some elements $a,b\in\mathcal{M}$ respectively and $\|a\|_{\mathcal{M}}\leq (3(k+1))^{-1}$ and $\|b\|_{\mathcal{M}}\leq (3(k+1))^{-1}$. Similarly, by the condition (A3), the third series (\[s3\]) also converges with respect to the norm $\|.\|_{\mathcal{M}}$ to some element $c\in\mathcal{M}$, satisfying $\|c\|_{\mathcal{M}}\leq (3(k+1))^{-1}$. Finally, since $y=\sum_{m=1}^\infty v_mxv_m^*$ (the convergence of the latter series is taken in the $\tau_{so}$ topology), we see that the fourth series (\[s4\]) converges with respect to the topology $\tau_{so}$ to some element $q_0^{(k)}yq_0^{(k)}$. Hence, the series $$\label{sq0deltaq0} \sum_{m=1}^\infty q_0^{(k)}\delta(v_mx_{n(m,k)}v_m^*)q_0^{(k)}$$ converges with respect to the topology $\tau_{so}$ to some element $a_k\in\mathcal{M}$, and, in addition, we have $$\label{aq} \|a_k-q_0^{(k)}yq_0^{(k)}\|_{\mathcal{M}}\leq (k+1)^{-1}.$$ We shall show that $$\label{qd} a_k=q_0^{(k)}\delta(y_k)q_0^{(k)},$$ where $y_k=\sum_{m=1}^\infty v_mx_{n(m,k)}v_m^*$ (the convergence of the latter series is taken in the $t(\mathcal{M})$-topology (see (\[y\_k\])). Using (\[q0z0\]) for any $m_1,m_2\in\mathbb{N}$ we have $$\begin{split} r_{m_1}q_0^{(k)}\delta(y_k)q_0^{(k)}r_{m_2}&= \delta(r_{m_1}q_0^{(k)}y_k)q_0^{(k)}r_{m_2}-\delta(r_{m_1}q_0^{(k)})y_kq_0^{(k)}r_{m_2}\cr & =\delta(r_{m_1}v_{m_1}x_{n(m_1,k)}v_{m_1}^*)r_{m_2}-\delta(r_{m_1})v_{m_2}x_{n(m_2,k)}v_{m_2}^*r_{m_2}. \end{split}$$ Since the series $\sum_{m=1}^\infty q_0^{(k)}\delta(v_mx_{n(m,k)}v_m^*)q_0^{(k)}$ converges with respect to the topology $\tau_{so}$ (see \[sq0deltaq0\]), it follows that the series $$\sum_{m=1}^\infty r_{m_1}(q_0^{(k)}\delta(v_mx_{n(m,k)}v_m^*)q_0^{(k)})r_{m_2}$$ also converges with respect to this topology ([@R-S ch. VI]), in addition, the following equalities hold $$\begin{split} r_{m_1}a_kr_{m_2}&=\sum_{m=1}^\infty r_{m_1}(q_0^{(k)}\delta(v_mx_{n(m,k)}v_m^*)q_0^{(k)})r_{m_2}\cr &\stackrel{(\ref{q0z0})}=\sum_{m=1}^\infty r_{m_1}\delta(v_mx_{n(m,k)}v_m^*)r_{m_2}\cr &= \sum_{m=1}^\infty (\delta(r_{m_1}v_mx_{n(m,k)}v_m^*)r_{m_2}-\delta(r_{m_1})v_mx_{n(m,k)}v_m^*r_{m_2})\cr &\stackrel{(\ref{r_m})}= \delta(r_{m_1}v_{m_1}x_{n(m_1,k)}v_{m_1}^*)r_{m_2}-\delta(r_{m_1})v_{m_2}x_{n(m_2,k)}v_{m_2}^*r_{m_2}, \end{split}$$ which guarantees $$\label{qk} r_{m_1}q_0^{(k)}\delta(y_k)q_0^{(k)}r_{m_2}=r_{m_1}a_kr_{m_2}.$$ Since $$r_{m_1}(\delta(y_k)-a_k)r_m\stackrel{(\ref{qk})}=0,$$ we see that for the right support $r(r_{m_1}(\delta(y_k)-a_k))$ of the operator $r_{m_1}(\delta(y_k)-a_k)$ satisfies the inequality $$r(r_{m_1}(\delta(y_k)-a_k))\leq \mathbf{1}-r_m, m\in\mathbb{N},$$ and therefore $$r(r_{m_1}(\delta(y_k)-a_k))\leq \inf_{m\geq 1}(\mathbf{1}-r_m)\stackrel{(\ref{q0z0})}=\mathbf{1}-q_0^{(k)}.$$ Consequently, $r_{m_1}(\delta(y_k)-a_k)q_0^{(k)}=0$ for all $m_1\in\mathbb{N}$. Similarly, using the left support of the operator $(\delta(y_k)-a_k)q_0^{(k)}$, we claim that $q_0^{(k)}(\delta(y_k)-a_k)q_0^{(k)}=0$. Since $q_0^{(k)}a_kq_0^{(k)}=a_k$, the equality (\[qd\]) holds. Thus, the inequality (\[aq\]) can be restated as follows $$\label{aqk} \|q_0^{(k)}(\delta(y_k)-y)q_0^{(k)}\|_{\mathcal{M}}\leq (k+1)^{-1}.$$ It follows from the inequalities (\[yq\]) and (\[aqk\]), that $$\label{ky} \|(k+1)q_0^{(k)}y_k\|_{\mathcal{M}}=\|(k+1)y_kq_0^{(k)}\|_{\mathcal{M}}\leq 1$$ and $$\label{yky} \|q_0^{(k)}\delta((k+1)y_k)q_0^{(k)}-(k+1)q_0^{(k)}yq_0^{(k)}\|_{\mathcal{M}}\leq 1.$$ Due to (\[yky\]), and taking into account (\[p0q0\]), we obtain $$\begin{split} (k+1)q_0^{(k)}-q_0^{(k)}\delta((k+1)y_k)q_0^{(k)}&\leq (k+1)q_0^{(k)}yq_0^{(k)}-q_0^{(k)}\delta((k+1)y_k)q_0^{(k)}\cr &\leq q_0^{(k)}, \end{split}$$ that is $$\label{yqky} kq_0^{(k)}\leq q_0^{(k)}\delta((k+1)y_k)q_0^{(k)}.$$ Let us now consider the projections $$\label{q0z01}q_0:=\inf_{k\geq 1} q_0^{(k)},\ z_0:=\inf_{k\geq 1} z_0^{(k)}.$$ Using (\[p0q0\]), (\[q0z01\]) we have that $p_0-q_0=\sup_{k\geq 1} (p_0-q_0^{(k)})$. Therefore, combining (\[Dz\_kp0q0k\]) and (\[q0z01\]), we obtain $$\label{Dq0z01}\mathcal{D}(z_0(p_0-q_0))=\mathcal{D}(\sup\limits_{k\geq 1}(z_0(p_0-q_0^{(k)})))\stackrel{(D6)} \leq \sum_{k=1}^\infty \mathcal{D}(z_0(p_0-q_0^{(k)}))\stackrel{(\ref{Dz_kp0q0k})}\leq \varphi(z_0),$$ that is the projection $z_0(p_0-q_0)$ is finite (see (D1)). Moreover, due to inequalities (\[ky\]) (respectively, (\[yqky\])), we have $$\label{kyq}\|(k+1)q_0y_k\|_{\mathcal{M}}=\|(k+1)y_kq_0\|_{\mathcal{M}}\leq 1,\quad k\in\mathbb{N}$$ (respectively, $$\label{kqy}kq_0\leq q_0\delta((k+1)y_k)q_0,\quad k\in\mathbb{N}.)$$ Since $\varphi$ is a $*$-isomorphism from $\mathcal{Z(M)}$ onto $L^\infty(\Omega,\Sigma,\mu)$, by (\[z0\]), we have that $$\int_\Omega \varphi(\mathbf{1}-z_0)d\mu=\int_\Omega \sup_{k\geq 1}\varphi(\mathbf{1}-z_0^{(k)})d\mu\leq\sum_{k=1}^\infty \int_\Omega \varphi(\mathbf{1}-z_0^{(k)})d\mu \stackrel{(\ref{z0})}\leq 2^{-1},$$ in particular, $z_0\neq 0$. Since $\mathbf{1}=c(p_0)$ and $c(p_0z_0)=c(p_0)z_0=z_0\neq 0$, we have $z_0p_0\neq 0$, and therefore there exists such $n\in\mathbb{N}$ that $z_0p_n\neq 0$ (see (\[p\_0\])). Since $z_0p_n\sim z_0p_m$, we have $z_0p_m\neq 0$ for all $m\in\mathbb{N}$. Hence, $z_0p_0$ is an infinite projection. Since the projection $z_0(p_0-q_0)$ is finite (see (\[Dq0z01\])), we see that the projection $z_0q_0$ must be infinite. By [@KR Proposition 6.3.7], there exists a central projection $$0\neq e_0\in\mathcal{P}(\mathcal{Z}(\mathcal{M})),\ e_0\leq z_0,$$ such that $e_0q_0$ is properly infinite, in particular, there exist pairwise orthogonal projections $$\label{e0}e_n\leq e_0q_0,\ e_n\sim e_0q_0$$ for all $n\in\mathbb{N}$ (see, for example, [@Sak Proposition 2.2.4]). In addition, $$\label{int_e0q0} \int_\Omega\varphi(c(q_0)e_0)\,d\mu\neq 0.$$ For every $n\in\mathbb{N}$ the operator $$b_n:=\delta(e_n)e_n$$ is locally measurable, and therefore there exists such a sequence $\{z_m^{(n)}\} \subset\mathcal{P}(\mathcal{Z}(\mathcal{M}))$ that $z_m^{(n)}\uparrow \mathbf{1}$ when $m\rightarrow\infty$ and $z_m^{(n)}b_n\in S(\mathcal{M})$ for all $m\in\mathbb{N}$. Since $\varphi(z_m^{(n)})\uparrow \varphi(\mathbf{1})=\mathbf{1}_{L^\infty(\Omega)}$ it follows that $\int_\Omega\varphi(z_m^{(n)})d\mu \uparrow \mu(\mathbf{1}_{L^\infty(\Omega)})=1$ when $m\rightarrow\infty$, and therefore, by (\[int\_e0q0\]), for every $n\in\mathbb{N}$ there exists such a projection $z^{(n)}\in\mathcal{P}(\mathcal{Z}(\mathcal{M}))$, that $z^{(n)}b_n\in S(\mathcal{M})$ and $$\label{int}1-2^{-n-1}\int_\Omega\varphi(c(q_0)e_0)d\mu<\int_\Omega\varphi(z^{(n)})d\mu.$$ Consider the central projection $$g_0:=\inf_{n\geq 1}z^{(n)}.$$ Since $z^{(n)}b_n\in S(\mathcal{M}), g_0=g_0z^{(n)}$ we have that $g_0b_n\in S(\mathcal{M})$ for all $n\in\mathbb{N}$. Due to (\[int\]) we have $$\begin{gathered} \begin{split}1&-\int_\Omega\varphi(g_0)d\mu=\int_\Omega\varphi(\mathbf{1}-g_0)d\mu=\int_\Omega\sup\varphi(\mathbf{1}-z^{(n)})d\mu\leq\\& \sum_{n=1}^\infty\int_\Omega\varphi(\mathbf{1}-z^{(n)})d\mu=\sum_{n=1}^\infty(1-\int_\Omega\varphi(z^{(n)})d\mu)\leq 2^{-1}\int_\Omega\varphi(c(q_0)e_0)d\mu. \end{split}\end{gathered}$$ Consequently, $1-2^{-1}\int_\Omega\varphi(c(q_0)e_0)d\mu\leq \int_\Omega\varphi(g_0)d\mu$, and therefore $$\label{phi}1+2^{-1}\int_\Omega\varphi(c(q_0)e_0)d\mu\leq \int_\Omega\varphi(g_0)d\mu+\int_\Omega\varphi(c(q_0)e_0)d\mu.$$ From (\[int\_e0q0\]) and inequality (\[phi\]), it follows that $\int_\Omega \varphi(g_0c(q_0)e_0)d\mu>0$, i.e. $g_0c(q_0)e_0\neq 0$ and so $g_0e_0q_0\neq 0$. Since $e_0q_0$ is a properly infinite projection it follows that $g_0e_0q_0$ is a properly infinite projection. From the relationship $g_0e_n\stackrel{(\ref{e0})}\sim g_0e_0q_0$, we see that the projection $g_0e_n$ is also properly infinite for all $n\in\mathbb{N}$. Since $$c(g_0e_n)=g_0c(e_n)\stackrel{(\ref{e0})}\leq q_0c(q_0e_0)=g_0c(q_0)e_0,$$ it follows that $ze_n$ is also properly infinite projection for every $0\neq z \in \mathcal{P}(\mathcal{Z}(\mathcal{M}))$ with $z\leq g_0c(q_0)e_0$. Indeed, if $z'\in\mathcal{P(Z(M))}$ and $z'ze_n\neq 0$, then $0\neq z'ze_n=(z'zc(q_0)e_0)g_0e_n$, and therefore, since the projection $g_0e_n$ is properly infinite, we have $(z'zc(q_0)e_0)g_0e_n\notin\mathcal{P}_{fin}(\mathcal{M})$. Consequently, the projection $ze_n$ is also properly infinite. Passing, if necessary to the algebra $g_0c(q_0)e_0\mathcal{M}$, we may assume that $g_0c(q_0)e_0=\mathbf{1}$. In this case, we also may assume that $b_n\in S(\mathcal{M}),\ e_n\sim q_0$, $c(e_n)=\mathbf{1}$ and $ze_n$ is a properly infinite projection for every non-zero $z\in\mathcal{P}(\mathcal{Z}(\mathcal{M}))$. The assumption $b_n\in S(\mathcal{M})$ means that for every fixed $n\in\mathbb{N}$ there exists such a sequence $\{p^{(n)}_m\}_{m=1}^\infty\subset \mathcal{P}_{fin}(\mathcal{M})$, that $p^{(n)}_m\downarrow 0$ when $m\rightarrow\infty$ and $b_n(\mathbf{1}-p^{(n)}_m)\in\mathcal{M}$ for all $m\in\mathbb{N}$. Since $\mathcal{D}(p^{(n)}_m)\in L^0(\Omega,\Sigma,\mu)$ and $\mathcal{D}(p^{(n)}_m)\downarrow 0$ (see (D7)), it follows that $\{\mathcal{D}(p_m^{(n)})\}_{n=1}^\infty$ converges in measure $\mu$ to zero. Consequently, we may select a central projection $f_n$ and a finite projection $s_n=p_{m_n}^{(n)}\in \mathcal{P}_{fin}(\mathcal{M})$ as to guarantee $\mathcal{D}(f_ns_n)<2^{-n}\varphi(f_n)$, $1-2^{-n-1}<\int\varphi(f_n)d\mu$ and $$\label{f_nb_n} f_nb_n(\mathbf{1}-s_n)\in\mathcal{M}$$ for all $n\in\mathbb{N}$. Setting $$f:=\inf_{n\geq 1} f_n,\ s:=\sup_{n\geq 1}s_n,$$ we have that $$1/2<\int\varphi(f)d\mu,\quad \mathcal{D}(fs)\stackrel{(D6)}\leq\sum_{n=1}^\infty\mathcal{D}(fs_n)\leq \varphi(f).$$ This means that $f\neq 0$ and $fs\in \mathcal{P}_{fin}(\mathcal{M})$ (see (D1)). In addition, since $f\leq f_n, (\mathbf{1}-s)\leq (\mathbf{1}-s_n)$ from (\[f\_nb\_n\]) it follows that $fb_n(\mathbf{1}-s)\in\mathcal{M}$ for all $n\in\mathbb{N}$. Consider the projections $t=f(\mathbf{1}-s)$ and $g_n=f(e_n\wedge(\mathbf{1}-s)),\ n\in\mathbb{N}$. Clearly (see (\[e0\])), $$\label{gze} g_n\leq fe_n\leq q_0,\ b_ng_n\in\mathcal{M},\ g_n\leq t$$ for all $n\in\mathbb{N}$, and also $$fe_n-g_n=f(e_n-e_n\wedge(\mathbf{1}-s))\sim f(e_n\vee(\mathbf{1}-s)-(\mathbf{1}-s))\leq fs,$$ that is $fe_n-g_n\in \mathcal{P}_{fin}(\mathcal{M})$. Hence, for every non-zero central projection $z\leq f$, we have that the projection $ze_n-zg_n$ is finite. Since the projection $ze_n$ is infinite, the projection $zg_n$ is also infinite, i.e. $$\label{zgn} zg_n\notin\mathcal{P}_{fin}(\mathcal{M})$$ for any $0\neq z\in \mathcal{P}(\mathcal{Z}(\mathcal{M}))$ and $n\in\mathbb{N}$. Since $b_nt=fb_n(\mathbf{1}-s) \in \mathcal{M}$, we see that there exists such an increasing sequence $\{l_n\}\subset\mathbb{N}$ that $l_n>n+2\|b_nt\|_{\mathcal{M}}$ for all $n\in\mathbb{N}$. Appealing to the inequalities (\[kyq\]), (\[gze\]) and taking into account the equality $b_n=\delta(e_n)e_n$, we deduce $$\begin{split} \|g_n({l_n}+1)y_{l_n}\delta(e_n)e_ng_n\|_{\mathcal{M}}&\leq \|g_n({l_n}+1)y_{l_n}\|_{\mathcal{M}}\|\delta(e_n)e_ng_n\|_{\mathcal{M}}\cr &\leq \|q_0({l_n}+1)y_{l_n}\|_{\mathcal{M}}\|\delta(e_n)e_nt\|_{\mathcal{M}}\cr &< (l_n-n)/2. \end{split}$$ Hence, $$\label{ineq_l_n-n} \|g_ne_n\delta(e_n)({l_n}+1)y_{l_n}g_n+g_n({l_n}+1)y_{l_n}\delta(e_n)e_ng_n\|_{\mathcal{M}}\leq l_n-n.$$ For every $x=x^*\in\mathcal{M}$ the inequalities $-\|x\|_\mathcal{M}\mathbf{1}\leq x\leq \|x\|_\mathcal{M}\mathbf{1}$ holds, in particular, $-g_n\|x\|_\mathcal{M}\leq q_nxq_n\leq g_n\|x\|_\mathcal{M}$. Hence, inequality (\[ineq\_l\_n-n\]) implies that $$\label{ge}g_ne_n\delta(e_n)({l_n}+1)y_{l_n}g_n+g_n({l_n}+1)y_{l_n}\delta(e_n)e_ng_n\geq (n-l_n)g_n.$$ Since $e_ne_m=0$ whenever $n\neq m$, we see (due to inequalities (\[kyq\]) and (\[gze\])) that the series $\sum_{n=1}^\infty e_n({l_n}+1)y_{l_n}e_n$ converges with respect to the topology $\tau_{so}$ to a self-adjoint operator $h_0\in\mathcal{M}$, satisfying $$\|h_0\|_{\mathcal{M}}\leq \sup_{n\geq 1}\|e_n({l_n}+1)y_{l_n}e_n\|_{\mathcal{M}}\leq 1.$$ Again appealing to the inequalities (\[kqy\]), (\[gze\]) and (\[ge\]), we infer that $$\begin{split} n g_n & =l_ng_n+(n-l_n)g_n \cr &\leq g_n({l_n}+1)\delta(y_{l_n})g_n+g_ne_n\delta(e_n)({l_n}+1)y_{l_n}g_n+g_n({l_n}+1)y_{l_n}\delta(e_n)e_ng_n\cr &= ({l_n}+1)(g_n\delta(y_{l_n})g_n+g_ne_n\delta(e_n)y_{l_n}g_n+g_ny_{l_n}\delta(e_n)e_ng_n)\cr &= ({l_n}+1)g_n\delta(e_ny_{l_n}e_n)g_n\cr &= \delta(g_ne_n({l_n}+1)y_{l_n}e_n)g_n-\delta(g_n)e_n({l_n}+1)y_{l_n}e_ng_n\cr &= \delta(g_nh_0)g_n-\delta(g_n)h_0g_n= g_n\delta(h_0)g_n. \end{split}$$ Thus, $$\label{ng_n} n g_n\leq g_n\delta(h_0)g_n$$ for every $n\in\mathbb{N}$. Set $g_n^{(0)}=g_n\wedge E_{n-1}(\delta(h_0)),\ n\in\mathbb{N}$, where $\{E_\lambda(\delta(h_0))\}$ is the spectral family of projections for self-adjoint operator $\delta(h_0)$. For every $n\in\mathbb{N}$ we have $$\begin{gathered} \begin{split} ng_n^{(0)} & =ng_n^{(0)}g_ng_n^{(0)}\stackrel{(\ref{ng_n})}\leq g_n^{(0)}(g_n\delta(h_0)g_n)g_n^{(0)}\cr &= g_n^{(0)}\delta(h_0)g_n^{(0)}=g_n^{(0)}E_{n-1}(\delta(h_0))\delta(h_0)g_n^{(0)}\cr & \leq g_n^{(0)}(n-1)E_{n-1}(\delta(h_0))g_n^{(0)}=(n-1)g_n^{(0)}. \end{split}\end{gathered}$$ Hence, $g_n\wedge E_{n-1}(\delta(h_0))=g_n^{(0)}=0$ which implies $$g_n=g_n-g_n\wedge E_{n-1}(\delta(h_0))\sim g_n\vee E_{n-1}(\delta(h_0))-E_{n-1}(\delta(h_0))\leq \mathbf{1}-E_{n-1}(\delta(h_0)),$$ i.e. $g_n\preceq \mathbf{1}-E_{n-1}(\delta(h_0))$. Then $g_n\stackrel{(\ref{gze})}\leq fg_n\preceq f(\mathbf{1}-E_{n-1}(\delta(h_0)))$, and therefore $$\label{Dgn} \mathcal{D}(g_n)\stackrel{(D3)}\leq \mathcal{D}(f(\mathbf{1}-E_{n-1}(\delta(h_0))))$$ for all $n\in\mathbb{N}$. Since $|f\delta(h_0)|\in LS(\mathcal{M})$, we see that there exists such a non-zero central projection $f_0\leq f$, that $|f_0\delta(h_0)|\in S_h(\mathcal{M})$. Hence, we may find such $\lambda_0>0$, that $(f_0-E_{\lambda}(|f_0\delta(h_0)|))\in \mathcal{P}_{fin}(\mathcal{M})$ for all $\lambda\geq \lambda_0$ ([@MCh §2.2]), that is $\mathcal{D}(f_0(\mathbf{1}-E_{\lambda}(|f_0\delta(h_0)|)))\in L_+^0(\Omega,\Sigma,\mu)$ when $\lambda>\lambda_0$. Since $f_0(\mathbf{1}-E_{\lambda}(|f_0\delta(h_0)|))= f_0(\mathbf{1}-E_{\lambda}(|\delta(h_0)|))$, we infer from (\[Dgn\]) that $$\mathcal{D}(f_0g_n)\in L_+^0(\Omega,\Sigma,\mu)$$ for all $n\geq \lambda_0+1$ which contradicts with the property (D1) in the definition of the dimension function $\mathcal{D}$, since $f_0g_n$ is an infinite projection (see (\[zgn\])). Hence, our assumption that the derivation $\delta$ fails to be continuous in $(LS(\mathcal{M}),t(\mathcal{M}))$ has led to a contradiction. Observe that in the special case of properly infinite von Neumann algebras of type $I$ or $III$ , Theorem \[main\] gives a new proof of the results concerning the continuity of a derivation of $(LS(\mathcal{M}),t(\mathcal{M}))$ established earlier in [@AAK; @AK; @BdPS]. Extension of a derivation $\delta : \mathcal{M}\rightarrow LS(\mathcal{M})$ up to a derivation on $LS(\mathcal{M})$ =================================================================================================================== In this section the construction of extension of any derivation, acting on a von Neumann algebra $\mathcal{M}$ with values in $LS(\mathcal{M})$, up to a derivation from $LS(\mathcal{M})$ into $LS(\mathcal{M})$ is given. Using this extension and Theorem \[main\] it is established that in case the of a properly infinite von Neumann algebra $\mathcal{M}$, any derivation $\delta: \mathcal{A}\longrightarrow LS(\mathcal{M})$ from a subalgebra $\mathcal{A}$ satisfying $\mathcal{M}\subset\mathcal{A}\subset LS(\mathcal{M})$ is continuous with respect to the local measure topology. Let $\mathcal{\mathcal{M}}$ be an arbitrary von Neumann algebra and let $\{z_n\}_{n=1}^\infty$ be a sequence of central projections from $\mathcal{\mathcal{M}}$, such that $z_n\uparrow \mathbf{1}$. A sequence $\{x_n\}_{n=1}^\infty$ is called *consistent* with the sequence $\{z_n\}_{n=1}^\infty$, if for any $n,m\in\mathbb{\mathbb{N}}$ the equality $x_mz_n=x_nz_n$ holds for $n<m$. \[p8\] Let $\{x_n\}_{n=1}^\infty\subset LS(\mathcal{M})$ (respectively, $\{y_n\}_{n=1}^\infty\subset LS(\mathcal{M})$) be a sequence consistent with the sequence $\{z_n\}_{n=1}^\infty\subset\mathcal{P}(\mathcal{Z}(\mathcal{M}))$ (respectively, with the sequence $\{z'_n\}_{n=1}^\infty\subset\mathcal{P}(\mathcal{Z}(\mathcal{M}))$), $z_n\uparrow \mathbf{1}$ ($z'_n\uparrow \mathbf{1}$). Then (i). There exists a unique $x\in LS(\mathcal{M})$, such that $xz_n=x_nz_n$ for all $n\in\mathbb{N}$, in addition, $x_n\stackrel{t(\mathcal{M})}{\longrightarrow} x$; (ii). If $x_nz_nz'_m=y_mz_nz'_m$ for all $n,m\in\mathbb{N}$, then $(x_nz_n-y_nz'_n)\stackrel{t(\mathcal{M})}{\longrightarrow} 0$ for $n\rightarrow\infty$. (i). Consider a neighborhood $V(B,\varepsilon,\delta)$ of zero in topology $t(\mathcal{M})$, where $\varepsilon,\delta>0,\ B\in\Sigma,\ \mu(B)<\infty$ (see the definition of topology $t(\mathcal{M})$ in section 2). Since $z_n^\bot=(\mathbf{1}-z_n)\downarrow 0$, it follows that $\varphi(z_n^\bot)\in W(B,\varepsilon,\delta)$ for $n\geq n(B,\varepsilon,\delta)$. Taking $x\in LS(\mathcal{M}),\ q_n=z_n$, we have $(xz_n^\bot) q_n=0,\ \mathcal{D}(z_n^\bot q_n)=0$, i.e. $xz_n^\bot\in V(B,\varepsilon,\delta)$ for all $x\in LS(\mathcal{M}),\ n\geq n(B,\varepsilon,\delta)$. For $m>n$, we have $$x_mz_m-x_nz_n=x_mz_m-x_mz_n=x_m(z_m-z_n)=x_mz_mz_n^\bot\in V(B,\varepsilon,\delta)$$ for all $n\geq n(B,\varepsilon,\delta)$. It means that $\{x_nz_n\}_{n=1}^\infty$ is a Cauchy sequence in $(LS(\mathcal{M}),t(\mathcal{M}))$. Consequently, there exists $x\in LS(\mathcal{M})$ such that $x_nz_n\stackrel{t(\mathcal{M})}{\longrightarrow} x$. Since $x_nz_n^\bot\in V(B,\varepsilon,\delta)$ for all $n\geq n(B,\varepsilon,\delta)$, it follows that $x_nz_n^\bot\stackrel{t(\mathcal{M})}{\longrightarrow} 0$, and therefore $x_n=x_nz_n+x_nz_n^\bot\stackrel{t(\mathcal{M})}{\longrightarrow} x$. Fixing $k\in\mathbb{N}$, for $n>k$ we have $x_kz_k=x_nz_k\stackrel{t(\mathcal{M})}{\longrightarrow} xz_k$ for $n\rightarrow\infty$, i.e. $xz_k=x_kz_k$ for all $k\in\mathbb{N}$. If $a\in LS(\mathcal{M})$ and $az_n=x_nz_n=xz_n$ for all $n\in\mathbb{N}$, then $0=(a-x)z_n\stackrel{t(\mathcal{M})}{\longrightarrow} (a-x)$, i.e. $a=x$. (ii). If $x_m z_m {z'}_{n}^{\bot} \stackrel{t(\mathcal{M})}{\longrightarrow} 0$ for $n\rightarrow\infty$, $y_nz'_nz_m^\bot\stackrel{t(\mathcal{M})}{\longrightarrow} 0$ for $m\rightarrow\infty$, and $x_nz_n-x_mz_m\stackrel{t(\mathcal{M})}{\longrightarrow} 0$ for $n,m\rightarrow\infty$, then $$\begin{gathered} \begin{split} x_nz_n-y_nz'_n&=x_nz_n-x_mz_m+x_mz_mz'_n+x_mz_m{z'}_n^\bot-y_nz'_n= \\ &=(x_nz_n-x_mz_m)+y_nz_mz'_n+x_mz_m{z'}_n^\bot-y_nz'_n= \\ &= (x_nz_n-x_mz_m)-y_nz'_nz_m^\bot+x_mz_m{z'}_n^\bot\stackrel{t(\mathcal{M})}{\longrightarrow} 0 \end{split}\end{gathered}$$ for $n,m\rightarrow\infty$. Now, we consider a derivation $\delta$ from $S(\mathcal{M})$ into $LS(\mathcal{M})$ and construct an extension $\widetilde{\delta}$ from $LS(\mathcal{M})$ into $LS(\mathcal{M})$. Recall that for an arbitrary operator $x\in LS(\mathcal{M})$ there exists a sequence $\{z_n\}_{n=1}^\infty\subset \mathcal{P}(\mathcal{Z}(\mathcal{M}))$ such that $z_n\uparrow\mathbf{1}$ and $xz_n\in S(\mathcal{M})$ for all $n\in\mathbb{N}$. Since $\delta(xz_n)z_m=\delta(xz_nz_m)$ (see Lemma \[l1\]), the sequence $\{\delta(xz_n)\}_{n=1}^\infty$ is consistent with the sequence $\{z_n\}_{n=1}^\infty$. By Proposition \[p8\](i), there exists a unique $y(x)\in LS(\mathcal{M})$ such that $\delta(xz_n)\stackrel{t(\mathcal{M})}{\longrightarrow} y(x)$ (notation: $y(x)=t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xz_n)$). Set $\widetilde{\delta}(x)=y(x)$. According to Proposition \[p8\](ii), the definition of operator $\widetilde{\delta}(x)$ does not depend on a choice of a sequence $\{z_n\}_{n=1}^\infty\subset\mathcal{P}(\mathcal{Z}(\mathcal{M}))$, for which $z_n\uparrow\mathbf{1}$ and $xz_n\in S(\mathcal{M})$, $n\in\mathbb{N}$. If $x\in S(\mathcal{M})$, then, taking $z_n=\mathbf{1},\ n\in\mathbb{N}$, we obtain $\widetilde{\delta}(x)=\delta(x)$. \[p9\] The mapping $\widetilde{\delta}$ is a unique derivation from $LS(\mathcal{M})$ into $LS(\mathcal{M})$ such that $\widetilde{\delta}(x)=\delta(x)$ for all $x\in S(\mathcal{M})$. Let $x,y\in LS(\mathcal{M})$, and let $z_n,p_n\in\mathcal{P}(\mathcal{Z}(\mathcal{M}))$ be such that$z_n\uparrow\mathbf{1},\ p_n\uparrow\mathbf{1},\ xz_n,yp_n\in S(\mathcal{M}),\ n\in\mathbb{N}$. Observing that $$z_np_n\in\mathcal{P}(\mathcal{Z}(\mathcal{M})),\ (z_np_n)\uparrow\mathbf{1},\ xz_np_n,yz_np_n,(x+y)z_np_n\in S(\mathcal{M}),\ n\in\mathbb{N},$$ we have $$\begin{gathered} \begin{split} \widetilde{\delta}(x+y)&=t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta((x+y)z_np_n)= \\ &=\bigl(t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xz_np_n)\bigl) +\bigl(t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(yz_np_n)\bigl)= \\ &=\widetilde{\delta}(x)+\widetilde{\delta}(y). \end{split}\end{gathered}$$ Similarly, $\widetilde{\delta}(\lambda x)=\lambda\widetilde{\delta}(x)$, $\lambda\in\mathbb{C}$. Further, using convergences $$xz_n\stackrel{t(\mathcal{M})}{\longrightarrow} x,\ yp_n\stackrel{t(\mathcal{M})}{\longrightarrow} y,\ \delta(xz_n)\stackrel{t(\mathcal{M})}{\longrightarrow} \widetilde{\delta}(x),\ \delta(yp_n)\stackrel{t(\mathcal{M})}{\longrightarrow} \widetilde{\delta}(y)$$ and the inclusion $xyz_np_n\in S(\mathcal{M}),\ n\in\mathbb{N}$, we have $$\begin{gathered} \begin{split} \widetilde{\delta}(xy)&= t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xyz_np_n)=t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta((xz_n)(yp_n))=\\ &=t(\mathcal{M})-\lim_{n\rightarrow\infty}(\delta(xz_n)yp_n+xz_n\delta(y_np_n))=\widetilde{\delta}(x)y+x\widetilde{\delta}(y). \end{split}\end{gathered}$$ Consequently, $\widetilde{\delta}: LS(\mathcal{M})\rightarrow LS(\mathcal{M})$ is a derivation, in addition, $\widetilde{\delta}(x)=\delta(x)$ for all $x\in S(\mathcal{M})$. Assume that $\delta_1: LS(\mathcal{M})\rightarrow LS(\mathcal{M})$ is also a derivation for which $\delta_1(x)=\delta(x)$ for all $x\in S(\mathcal{M})$. Let us show that $\widetilde{\delta}=\delta_1$. If $x\in LS(\mathcal{M}),\ z_n\uparrow\mathbf{1},\ xz_n\in S(\mathcal{M}),\ n\in\mathbb{N}$, then, by Lemma \[l1\] and Proposition \[p8\] (i), we obtain $$\begin{gathered} \begin{split} \widetilde{\delta}(x)&=t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xz_n)=t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta_1(xz_n)=\\ &= t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta_1(x)z_n=\delta_1(x). \end{split}\end{gathered}$$ Now, we give the construction of extension of a derivation $\delta: \mathcal{M}\rightarrow LS(\mathcal{M})$ up to a derivation $\widehat{\delta}: S(\mathcal{M})\rightarrow LS(\mathcal{M})$. For each $x\in LS(\mathcal{M})$ set $s(x):=l(x)\vee r(x)$, where $l(x)$ is the left and $r(x)$ is the right support of $x$. If $x=u|x|$ is a polar decomposition of $x\in LS(\mathcal{N})$, then $u\in\mathcal{M}$ [@MCh §2.3] and, due to equalities $l(x)=uu^*,\ r(x)=u^*u$, we have $l(x)\sim r(x)$. We need the following lemma. \[l3\] If $\mathcal{D}$ is a dimension function of a von Neumann algebra $\mathcal{M}$, then for any derivation $\delta$ from $\mathcal{M}$ into $LS(\mathcal{M})$ the following inequality $$\mathcal{D}(s(\delta(x)))\leq 3 \mathcal{D}(s(x))$$ holds for all $x\in\mathcal{M}$. For $x\in\mathcal{M}$ we have $$\begin{gathered} l(\delta(x)s(x))\sim r(\delta(x)s(x))\leq s(x),\\ r(x\delta(s(x)))\sim l(x\delta(s(x)))=l(s(x)x\delta(s(x)))\leq s(x),\end{gathered}$$ i.e. $$l(\delta(x)s(x))\preceq s(x)$$ and $$r(x\delta(s(x)))\preceq s(x),$$ that implies the inequalities (see (D2), (D3)) $$\mathcal{D}(l(\delta(x)s(x)))\leq\mathcal{D}(s(x)),\ \mathcal{D}(r(x\delta(s(x))))\leq\mathcal{D}(s(x)).$$ Since $$\delta(x)=\delta(xs(x))=\delta(x)s(x)+x\delta(s(x)),$$ we have $$s(\delta(x))=s(\delta(x)s(x)+x\delta(s(x)))\leq s(x)\vee l(\delta(x)s(x))\vee r(x\delta(s(x))).$$ Due to (D6), we have $$\mathcal{D}(s(\delta(x)))\leq \mathcal{D}(s(x))+\mathcal{D}(l(\delta(x)s(x)))+\mathcal{D}(r(x\delta(s(x))))\leq 3 \mathcal{D}(s(x)).$$ As in the definition of the topology $t(\mathcal{M})$, denote by $\varphi$ a $*$-isomorphism from $\mathcal{Z}(\mathcal{M})$ onto the $*$-algebra $L^\infty(\Omega,\Sigma,\mu)$, where $\mu$ is a measure satisfying the direct sum property. By Proposition \[plm-spk1\](i), the convergence of the sequence of projections $p_n\stackrel{t(\mathcal{M})}{\longrightarrow} 0$ is equivalent to existence of a sequence $\{z_n\}\subset\mathcal{P}(\mathcal{Z}(\mathcal{M}))$ such that $z_np_n\in \mathcal{P}_{fin}(\mathcal{M})$ for all $n$, $\varphi(z_n^\bot)\stackrel{t(L^\infty(\Omega))}{\longrightarrow} 0$ and $\mathcal{D}(z_np_n)\stackrel{t(L^\infty(\Omega))}{\longrightarrow} 0$. \[l4\] If $\{x_n\}_{n=1}^\infty\subset LS(\mathcal{M}),\ s(x_n)\in\mathcal{P}_{fin}(\mathcal{M}),\ \mathcal{D}(s(x_n))\stackrel{t(L^\infty(\Omega))}{\longrightarrow} 0$, then $x_n\stackrel{t(\mathcal{M})}{\longrightarrow} 0$. Taking $z_n=\mathbf{1}$ for all $n\in\mathbb{N}$, we have $$z_ns(x_n)\in\mathcal{P}_{fin}(\mathcal{M}),\ \varphi(z_n^\bot)=0,\ n\in\mathbb{N},$$ and $$\mathcal{D}(z_ns(x_n))=\mathcal{D}(s(x_n))\stackrel{t(L^\infty(\Omega))}{\longrightarrow} 0.$$ Consequently, $s(x_n)\stackrel{t(\mathcal{M})}{\longrightarrow} 0$ (see Proposition \[plm-spk1\](i)). Since $E^\bot_\lambda(|x_n|)\leq s(x_n)$ for all $\lambda>0,\ n\in\mathbb{N}$, it follows $E_\lambda^\bot(|x_\alpha|) \stackrel{t(\mathcal{M})}{\longrightarrow} 0$, and therefore $x_n\stackrel{t(\mathcal{M})}{\longrightarrow} 0$ (see Remark \[rem\_plm-spk1\]). If $p_n\in \mathcal{P}_{fin}(\mathcal{M})$ and $p_n\downarrow 0$, then $\mathcal{D}(p_n)\in L^0_+(\Omega,\Sigma,\mu)$ (see (D1)) and $\mathcal{D}(p_n)\downarrow 0$ (see (D2) and D(7)), in particular, $\mathcal{D}(p_n) \stackrel{t(L^\infty(\Omega))}{\longrightarrow} 0$. Hence, Lemma \[l4\] implies the following \[cc1\] If $\{p_n\}_{n=1}^\infty\subset \mathcal{P}_{fin}(\mathcal{M}),\ p_n\downarrow 0$, then $p_n\stackrel{t(\mathcal{M})}{\longrightarrow} 0$. \[l6\] Let $x\in S(\mathcal{M}),\ p_n,q_n\in\mathcal{P}(\mathcal{M}),\ p_n\uparrow\mathbf{1},\ q_n\uparrow\mathbf{1},\ xp_n,xq_n\in\mathcal{M},\ p_n^\bot,q_n^\bot\in\mathcal{P}_{fin}(\mathcal{M}),\ n\in\mathbb{N}$. If $\delta: \mathcal{M}\rightarrow LS(\mathcal{M})$ is a derivation, then there exists $\widehat{\delta}(x)\in LS(\mathcal{M})$, such that $$t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xp_n)=\widehat{\delta}(x)=t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xq_n).$$ For $n<m$ we have $$l(x(p_m-p_n))\sim r(x(p_m-p_n))\leq p_m-p_n,$$ and therefore, applying Lemma \[l3\] and properties (D2), (D3), we obtain $$\begin{aligned} \mathcal{D}( & s(\delta(xp_m-xp_n)))= \mathcal{D}(s(\delta(x(p_m-p_n))))\leq 3\mathcal{D}(s(x(p_m-p_n)))\leq\\ & 3\mathcal{D}(l(x(p_m-p_n))\vee (p_m-p_n))\leq 6\mathcal{D}(p_m-p_n)\leq 6\mathcal{D}(p_n^\bot).\end{aligned}$$ Since $\mathcal{D}(p_n^\bot)\in L_+^0(\Omega,\Sigma,\mu)$ (see (D1)) and $\mathcal{D}(p_n^\bot)\downarrow 0$ (see (D7)) it follows that $\mathcal{D}(p_n^\bot)\stackrel{t(L^\infty(\Omega))}{\longrightarrow} 0$ (see (D7)). Hence, $$\mathcal{D}(s(\delta(xp_m)-\delta(xp_n)))\stackrel{t(L^\infty(\Omega))}{\longrightarrow} 0$$ for $n,m\rightarrow\infty$. By Lemma \[l4\], we have that $(\delta(xp_m)-\delta(xp_n))\stackrel{t(\mathcal{M})}{\longrightarrow} 0$ for $n,m\rightarrow\infty$, i.e. $\{\delta(xp_n)\}_{n=1}^\infty$ is a Cauchy sequence in $(LS(\mathcal{M}),t(\mathcal{M}))$. Consequently, there exists $\widehat{\delta}(x)\in LS(\mathcal{M})$, such that $$t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xp_n)=\widehat{\delta}(x).$$ Let us show that $t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xq_n)=\widehat{\delta}(x)$. For each $n\in\mathbb{N}$ we have $$\begin{gathered} \begin{split} (p_n-q_n) & ((p_n-p_n\wedge q_n)\vee (q_n-p_n\wedge q_n))=\\ &=((p_n-p_n\wedge q_n)-(q_n-p_n\wedge q_n))((p_n-p_n\wedge q_n)\vee (q_n-p_n\wedge q_n))\\ &=(p_n-p_n\wedge q_n)-(q_n-p_n\wedge q_n)=p_n-q_n. \end{split}\end{gathered}$$ Hence, $$r(p_n-q_n)\leq ((p_n-p_n\wedge q_n)\vee (q_n-p_n\wedge q_n)).$$ Since $$r(x(p_n-q_n))\leq r(p_n-q_n)$$ and $$l(x(p_n-q_n))\sim r(x(p_n-q_n)),$$ it follows $$\begin{gathered} \begin{split} \mathcal{D}(s(x(p_n-q_n)))&=\mathcal{D}(l(x(p_n-q_n))\vee r(x(p_n-q_n)))\\ &\stackrel{(D6)}{\leq} \mathcal{D}(l(x(p_n-q_n)))+\mathcal{D}(r(x(p_n-q_n))) =2\mathcal{D}(r(x(p_n-q_n)))\cr &\stackrel{(D6)}{\leq} 2\mathcal{D}(p_n-p_n\wedge q_n)+2\mathcal{D}(q_n-p_n\wedge q_n)\leq 4\mathcal{D}(\mathbf{1}-p_n\wedge q_n)\cr &=4\mathcal{D}(p_n^\bot\vee q_n^\bot)\leq 4(\mathcal{D}(p_n^\bot)+\mathcal{D}(q_n^\bot)). \end{split}\end{gathered}$$ Since (see Lemma \[l3\]) $$\mathcal{D}(s(\delta(xp_n)-\delta(xq_n)))=\mathcal{D}(s(\delta(x(p_n-q_n))))\leq 3\mathcal{D}(s(x(p_n-q_n))),$$ we have $$\mathcal{D}(s(\delta(xp_n)-\delta(xq_n)))\leq 12(\mathcal{D}(p_n^\bot)+\mathcal{D}(q_n^\bot))\downarrow 0.$$ By Lemma \[l4\], we obtain $$t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xq_n)=t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xp_n)=\widehat{\delta}(x).$$ Now, equipped with Lemma \[l6\], we may extend any derivation $\delta: \mathcal{M}\rightarrow LS(\mathcal{M})$ up to a derivation $\widehat{\delta}$ from $S(\mathcal{M})$ into $LS(\mathcal{M})$. For each $x\in S(\mathcal{M})$ there exists a sequence $\{p_n\}\in\mathcal{P}(\mathcal{M})$, such that $p_n\uparrow\mathbf{1},\ p_n^\bot\in\mathcal{P}_{fin}(\mathcal{M}),\ xp_n\in\mathcal{M}$ for all $n\in\mathbb{N}$. By Lemma \[l6\], there exists $\widehat{\delta}(x)\in LS(\mathcal{M})$, such that $t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xp_n)=\widehat{\delta}(x)$. In addition, the definition of $\widehat{\delta}(x)$ does not depend on a choice of sequence $\{p_n\}_{n\geq 1}$ satisfying the above mentioned property, in particular, $\widehat{\delta}(x)=\delta(x)$ for all $x\in\mathcal{M}$ (in this case, $p_n=\mathbf{1},\ n\in\mathbb{N}$). \[p10\] The mapping $\widehat{\delta}$ is a unique derivation from $S(\mathcal{M})$ into $LS(\mathcal{M})$, such that $\widehat{\delta}(x)=\delta(x)$ for all $x\in\mathcal{M}$. For $x,y\in S(\mathcal{M})$ select $p_n,q_n\in \mathcal{P}(\mathcal{M}),\ n\in\mathbb{N}$, such that $$p_n\uparrow\mathbf{1},\ q_n\uparrow\mathbf{1},\ p_n^\bot,q_n^\bot\in\mathcal{P}_{fin}(\mathcal{M}),\ xp_n,yq_n\in\mathcal{M}$$ for all $n\in\mathbb{N}$. The sequence of projections $e_n=p_n\wedge q_n$ is increasing, and, in addition, $$\begin{gathered} xe_n=xp_ne_n\in\mathcal{M},\ ye_n=yq_ne_n\in\mathcal{M},\\ e_n^\bot=p_n^\bot\vee q_n^\bot\in\mathcal{P}_{fin}(\mathcal{M}),\mathcal{D}(e_n^\bot)\leq\mathcal{D}(p_n^\bot)+\mathcal{D}(q_n^\bot)\downarrow 0.\end{gathered}$$ The last estimate implies the convergence $e_n^\bot\downarrow 0$ (see (D7)), or $e_n\uparrow\mathbf{1}$. By Lemma \[l6\], we have $$\begin{gathered} \begin{split} \widehat{\delta}(x+y)&=t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta((x+y)e_n)= \\ &=\bigl(t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xe_n)\bigl) +\bigl(t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(ye_n)\bigl)=\widehat{\delta}(x)+\widehat{\delta}(y). \end{split}\end{gathered}$$ Similarly, $\widehat{\delta}(\lambda x)=\lambda\widehat{\delta}(x)$ for all $\lambda\in\mathbb{C}$. Let us show that $\widehat{\delta}(xy)=\widehat{\delta}(x)y+x\widehat{\delta}(y),x,y\in S(\mathcal{M})$. Due to polar decomposition $y=u|y|,\ u^*u=r(y)$, we have $y_n=yE_n(|y|)\in\mathcal{M}$ for all $n\in\mathbb{N}$. Set $$g_n=\mathbf{1}-r(E_n^\bot(|x|)y_n),\ s_n=g_n\wedge E_n(|y|).$$ Since $$g_n^\bot=r(E_n^\bot(|x|)y_n)\sim l(E_n^\bot(|x|)y_n)\leq E_n^\bot(|x|),$$ we obtain $$g_n^\bot\preceq E_n^\bot(|x|).$$ Since $x\in S(\mathcal{M})$, there exists $n_0\in\mathbb{N}$ such that $E_n^\bot(|x|)\in\mathcal{P}_{fin}(\mathcal{M})$ for all $n\geq n_0$, and therefore $g_n^\bot\in\mathcal{P}_{fin}(\mathcal{M})$ for all $n\geq n_0$. The equality $$y_ng_n=E_n(|x|)y_ng_n+E_n^\bot(|x|)y_ng_n=E_n(|x|)y_ng_n$$ implies that $$\begin{gathered} \begin{split} E_{n+1}^\bot(|x|)y_{n+1}s_n&=E_{n+1}^\bot(|x|)E_n^\bot(|x|y_{n+1}E_n(|y|))s_n=\\ &= E_{n+1}^\bot(|x|)(E_n^\bot(|x|)y_nE_n(|y|))s_n=\\& =E_{n+1}^\bot(|x|)(E_n^\bot(|x|)y_ns_n)=E_{n+1}^\bot(|x|)(E_n^\bot(|x|)y_ng_n)s_n=0, \end{split}\end{gathered}$$ in particular, $$s_n\leq \mathbf{1}-r(E_{n+1}^\bot(|x|)y_{n+1})=g_{n+1}$$ for all $n\in\mathbb{N}$. From here and from the inequalities $s_n\leq E_n(|y|)\leq E_{n+1}(|y|)$ it follows that $s_n\leq s_{n+1}$. Since $y\in S(\mathcal{M})$, we have $E_n^\bot(|y|)\in \mathcal{P}_{fin}(\mathcal{M})$ for $n\geq n_1$ for some $n_1\geq n_0$. Hence, $$s_n^\bot=g_n^\bot\vee E_n^\bot(|y|)\in\mathcal{P}_{fin}(\mathcal{M})$$ for $n\geq n_1$ and $$\mathcal{D}(s_n^\bot)\leq\mathcal{D}(g_n^\bot)+\mathcal{D}(E_n^\bot(|y|))\leq (\mathcal{D}(E_n^\bot(|x|))+\mathcal{D}(E_n^\bot(|y|)))\downarrow 0,$$ i.e. $s_n^\bot\downarrow 0$ or $s_n\uparrow\mathbf{1}$. Using Corollary \[cc1\], Lemma \[l6\], the inclusions $xE_n(|x|)\in\mathcal{M},\ yE_n(|y|)\in\mathcal{M}$ and equalities $$\begin{gathered} \begin{split} xys_n&=xyE_n(|y_n|)s_n=xy_ns_n=xy_ng_ns_n= \\ &=xE_n(|x|)y_nq_ns_n= xE_n(|x|)yE_n(|y|)s_n, \end{split}\end{gathered}$$ we obtain $$\begin{gathered} \begin{split} \widehat{\delta}(xy)=&t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xys_n)= t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xE_n(|x|)yE_n(|y|s_n))= \\& =t(\mathcal{M})-\lim_{n\rightarrow\infty}\bigl(\delta(xE_n(|x|))ys_n+xE_n(|x|)\delta(ys_n)\bigl)= \\& =\bigl(t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xE_n(|x|))\bigl)\cdot \bigl(t(\mathcal{M})-\lim_{n\rightarrow\infty}ys_n\bigl)+\cr &+ \bigl(t(\mathcal{M})-\lim_{n\rightarrow\infty}xE_n(|x|)\bigl)\cdot \bigl(t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(ys_n)\bigl)=\widehat{\delta}(x)y+x\widehat{\delta}(y). \end{split}\end{gathered}$$ Consequently, $\widehat{\delta}: S(\mathcal{M})\rightarrow LS(\mathcal{M})$ is a derivation, such that $\widehat{\delta}(x)=\delta(x)$ for all $x\in\mathcal{M}$. Let $\delta_1: S(\mathcal{M})\rightarrow LS(\mathcal{M})$ also be a derivation, for which $\delta_1(x)=\delta(x)$ for all $x\in\mathcal{M}$. If $x\in S(\mathcal{M})$, then $E_n(|x|)\uparrow\mathbf{1},\ xE_n(|x|)\in\mathcal{M},\ n\in\mathbb{N},\ E_n^\bot(|x|)\in\mathcal{P}_{fin}(\mathcal{M})$ for all $n\geq n_3$ for some $n_3\in\mathbb{N}$. Hence, $E_n(|x|)\stackrel{t(\mathcal{M})}{\longrightarrow} \mathbf{1}$ (see Corollary \[cc1\]). Since $(LS(\mathcal{M}),t(\mathcal{M}))$ is a topological algebra, it follows that $$\begin{gathered} \begin{split} \delta_1(x)&=t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta_1(x)E_n(|x|)= \cr &=\bigl(t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta_1(xE_n(|x|))\bigl)- \bigl(t(\mathcal{M})-\lim_{n\rightarrow\infty}x\delta_1(E_n(|x_n|))\bigl)= \cr &=\bigl(t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(xE_n(|x|))\bigl)-\bigl( t(\mathcal{M})-\lim_{n\rightarrow\infty}x\delta(E_n(|x_n|))\bigl)=\cr &= \widehat{\delta}(x)-x(t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(E_n(|x_n|))). \end{split}\end{gathered}$$ Since $\delta(\mathbf{1})=0,\ s(x)=s(-x)$ for $x\in LS(\mathcal{M})$, it follows via Lemma \[l3\], that $$\begin{gathered} \begin{split} \mathcal{D}(s(\delta(E_n(|x|))))&= \mathcal{D}(s(\delta(-E_n(|x|))))=\\ &=\mathcal{D}(s(\delta(\mathbf{1}-E_n(|x|))))\leq 3\mathcal{D}(E_n^\bot(|x|))\downarrow 0. \end{split}\end{gathered}$$ By Lemma \[l4\], we obtain $\delta(E_n(|x|))\stackrel{t(\mathcal{M})}{\longrightarrow} 0$, that implies the equality $\delta_1(x)=\widehat{\delta}(x)$. Propositions \[p9\] and \[p10\] imply the following theorem, which is the main result of this section. \[ext\] Let $\mathcal{A}$ be a subalgebra of $LS(\mathcal{M})$, $\mathcal{M}\subset\mathcal{A}$ and let $\delta: \mathcal{A}\rightarrow LS(\mathcal{M})$ be a derivation. Then there exists a unique derivation $\delta_{\mathcal{A}}: LS(\mathcal{M})\rightarrow LS(\mathcal{M})$ such that $\delta_{\mathcal{A}}(x)=\delta(x)$ for all $x\in\mathcal{A}$. Since $\mathcal{M}\subset\mathcal{A}$, the restriction $\delta_0$ of the derivation $\delta$ on $\mathcal{M}$ is a well-defined derivation from $\mathcal{M}$ into $LS(\mathcal{M})$. Hence, by Propositions \[p9\] and \[p10\], the mapping $\delta_{\mathcal{A}}=\widetilde{\widehat{\delta}}$ is a unique derivation from $LS(\mathcal{M})$ into $LS(\mathcal{M})$ such that $\delta_{\mathcal{A}}(x)=\delta_0(x)$ for all $x\in\mathcal{M}$. Let us show that $\delta_{\mathcal{A}}(a)=\delta(a)$ for every $a\in\mathcal{A}$. If $a\in\mathcal{A}$, then there exists a sequence $\{z_n\}_{n=1}^\infty\subset\mathcal{P}(\mathcal{Z}(\mathcal{M}))$, such that $z_n\uparrow\mathbf{1}$ and $az_n\in S(\mathcal{M}),\ n\in\mathbb{N}$. Since $z_n\stackrel{t(\mathcal{M})}{\longrightarrow}\mathbf{1}$ (see Proposition \[p8\](i)), we have, by Lemma \[l1\], $$\delta_{\mathcal{A}}(a)=t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta_{\mathcal{A}}(a)z_n= t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta_{\mathcal{A}}(az_n),$$ and, similarly, $\delta(a)=t(\mathcal{M})-\lim_{n\rightarrow\infty}\delta(az_n)$. Using the equality $\delta_{\mathcal{A}}(x)=\delta_0(x)=\delta(x)$ for each $x\in\mathcal{M}$, and following the proof of uniqueness of the derivation $\widehat{\delta}$ from Proposition \[p10\], we obtain $\delta_{\mathcal{A}}(az_n)=\delta(az_n)$ for all $n\in\mathbb{N}$, that implies the equality $\delta_{\mathcal{A}}(a)=\delta(a)$. The following corollary immediately follows from Theorems \[main\] and \[ext\]. \[c2\] Let $\mathcal{M}$ be a properly infinite von Neumann algebra, $\mathcal{A}$ is a subalgebra in $LS(\mathcal{M})$ and $\mathcal{M}\subset\mathcal{A}$. Then any derivation $\delta: \mathcal{A}\rightarrow LS(\mathcal{M})$ is continuous with respect to the local measure topology $t(\mathcal{M})$. In particular, Corollary \[c2\] implies that for a properly infinite von Neumann algebra $\mathcal{M}$ any derivation $\delta: S(\mathcal{M})\rightarrow S(\mathcal{M})$ is $t(\mathcal{M})$-continuous. Note, that in case, when $\mathcal{M}$ is of type $I_\infty$, any derivation of $S(\mathcal{M})$ is inner [@AK], and therefore is automatically continuous with respect to the topology $t(\mathcal{M})$. Applications to the algebra $S(\mathcal{M},\tau)$ of $\tau$-measurable operators ================================================================================ Let $\mathcal{M}$ be a semifinite von Neumann algebra acting on Hilbert space $H$, $\tau$ be a faithful normal semifinite trace on $\mathcal{M}$. An operator $x\in S(\mathcal{M})$ with domain $\mathfrak{D}(x)$ is called *$\tau$-measurable* if for any $\varepsilon>0$ there exists a projection $p\in\mathcal{P}(\mathcal{M})$ such that $p(H)\subset \mathfrak{D}(x)$ and $\tau(p^\bot)<\infty$. The set $S(\mathcal{M},\tau)$ of all $\tau$-measurable operators is a $*$-subalgebra of $S(\mathcal{M})$ such that $\mathcal{M}\subset S(\mathcal{M},\tau)$. If the trace $\tau$ is finite, then $S(\mathcal{M},\tau)=S(\mathcal{M})$. The algebra $S(\mathcal{M},\tau)$ is a noncommutative version of the algebra of all measurable complex functions $f$ defined on $(\Omega,\Sigma,\mu)$, for which $\mu(\{|f|>\lambda\})\rightarrow 0$ for $\lambda\rightarrow\infty$. For each $x\in S(\mathcal{M},\tau)$ it is possible to define the generalized singular value function $$\mu_t(x)=\inf\{\lambda>0:\tau(E_\lambda^\bot(|x|)<t\}=\inf\{\|x(\mathbf{1}-e)\|_\mathcal{M}:e\in\mathcal{P(M)},\tau(e)<t\},$$ which allows to define and study a noncommutative version of rearrangement invariant function spaces. For the theory of the latter spaces, we refer to [@DDdP],[@K-S]. Let $t_\tau$ be the measure topology [@Ne] on $S(\mathcal{M},\tau)$, whose base of neighborhoods of zero is given by $U(\varepsilon,\delta)=\{x\in S(\mathcal{M},\tau):\ \exists p\in \mathcal{P}(\mathcal{M}),\ \tau(p^\bot)\leq\delta,\ xp\in\mathcal{M},\ \|xp\|_{\mathcal{M}}\leq\varepsilon \}$, $\varepsilon>0,\ \delta>0$. The pair $(S(\mathcal{M},\tau),t_\tau)$ is a complete metrizable topological $*$-algebra. Here, the topology $t_\tau$ majorizes the topology $t(\mathcal{M})$ on $S(\mathcal{M},\tau)$ and, if $\tau$ is a finite trace, the topologies $t_\tau$ and $t(\mathcal{M})$ coincide [@MCh §§3.4,3.5]. However, if $\tau(\mathbf{1})=\infty$, then on $(S(\mathcal{M},\tau),t_\tau)$ topologies $t_\tau$ and $t(\mathcal{M})$ do not coincide in general [@CM]. For example, when $\mathcal{M}=L^\infty(\Omega,\Sigma,\mu),\tau(f)=\int_\Omega fd\mu,f\in L^\infty_+(\Omega),$ where $\mu$ is a $\sigma$-finite measure, $\mu(\Omega)=\infty$, the topology $t_\tau$ in $S(L^\infty(\Omega),\tau)$ coincide with the topology of convergence in measure $\mu$, and the topology $t(L^\infty(\Omega))$ is the topology of convergence locally in measure $\mu$ (Section 2), in particular, if $A_n\in\Sigma,\mu(A_n)=\infty,n\in\mathbb{N}$ and $\chi_{A_n}\downarrow 0$, then $\chi_{A_n}\xrightarrow{t(L^\infty(\Omega))} 0$, whereas $\chi_{A_n}\stackrel{t_\tau}{\nrightarrow}0$. See the detailed comparison of topologies $t_\tau$ and $t(\mathcal{M})$ in [@CM]. It is proved in [@Ber] that in a properly infinite von Neumann algebra $\mathcal{M}$ any derivation $\delta: S(\mathcal{M},\tau)\rightarrow S(\mathcal{M},\tau)$ is continuous with respect to the topology $t_\tau$. Corollary \[c2\] implies that, in this case, derivation $\delta: S(\mathcal{M},\tau)\rightarrow S(\mathcal{M},\tau)$ is continuous with respect to the topology $t(\mathcal{M})$ too. Now, we give an application of Theorem \[ext\] to derivations defined on absolutely solid $*$-subalgebras of the algebra $LS(\mathcal{M})$. Recall [@BdPS], that a $*$-subalgebra $\mathcal{A}$ of $LS(\mathcal{M})$ is called absolutely solid if conditions $x\in LS(\mathcal{M}),\ y\in\mathcal{A},\ |x|\leq |y|$ imply that $x\in\mathcal{A}$. In [@BdPS Proposition 5.13] it is proved that, if $\delta$ is a derivation on absolutely solid $*$-subalgebra $\mathcal{A}\supset\mathcal{M}$ and $\delta(x)=[w,x]$ for all $x\in\mathcal{A}$ and some $w\in LS(\mathcal{M})$, then there exists $w_1\in\mathcal{A}$, such that $\delta(x)=[w_1,x]$ for all $x\in\mathcal{A}$, i.e. the derivation $\delta$ is inner on the $*$-subalgebra $\mathcal{A}$. This observation and Theorem \[ext\] yield our final result \[c3\] Suppose that all derivations on the algebra $LS(\mathcal{M})$ are inner and let $\mathcal{A}\supset\mathcal{M}$ be an absolutely solid $*$-subalgebra of $LS(\mathcal{M})$. Then all derivations on $\mathcal{A}$ are inner. In particular, any derivation on the algebras $S(\mathcal{M})$ and $S(\mathcal{M},\tau)$ are inner. The result of Corollary \[c3\] extends and generalizes [@AK Theorem 6.8] and [@BdPS Proposition 5.17]. Note, that the conditions of Corollary \[c3\] hold, in particular, for properly infinite von Neumann algebras, which do not have direct summand of type $II_\infty$ [@AAK],[@AK],[@BdPS]. [99]{} S.Albeverio, Sh.Ayupov and K.Kudaibergenov, *Derivations on the algebra of measurable operators affiliated with type $I$ von Neumann algebras*, J. Funct. Anal., [**256**]{} (2009), No. 9, 2917-2943. Sh.A.Ayupov and K.K.Kudaybergenov, *Derivations on algebras of measurable operators*, Infinite Dimensional Analysis, Quantum Probability and Related Topics [**13**]{} (2010), No 2, 305–337. A.F.Ber, B.de Pagter and F.A.Sukochev, *Derivations in algebras of operator-valued functions*, J. Operator Theory, **66** (2011), No. 2, 261–300. A.F.Ber, V.I.Chilin, F.A.Sukochev, *Non-trivial derivations on commutative regular algebras*, Extracta Math. [**21**]{} (2006), 107-147. A.F.Ber, *Continuity of derivations on properly infinite \*-algebras of $\tau$-measurable operators*, Mat. Zametki [**90**]{} (2011) No. 5, 776–780 (Russian). English translation: *Math. Notes*, [**90**]{} (2011) No. 5-6, 758–762. O.Bratelli and D.W.Robinson, *Operator algebras and quantum statistical mechanics 1*, New York: Springer-Verlag, 1979. V.I.Chilin and M.A.Muratov, *Comparision of topologies on $*$-algebras of locally measurable operators*, Positivity (2012). P.G. Dodds, T.K. Dodds and B. de Pagter, *Noncommutative Banach function spaces*, Math. Z., **201** (1989), 583-597. P. Halmos, *Measure Theory*, Springer-Verlag, New York, 1974. R.V.Kadison and J.R.Ringrose, *Fundamentals of the Theory of Operator Algebras II*, Academic Press, Orlando, 1986. N.J. Kalton, F.A. Sukochev, *Symmetric norms and spaces of operators.* J. Reine Angew. Math. **621** (2008), 81–121. A.G.Kusraev, *Automorphisms and derivations in an extended complex $f$-algebra,* Siberian Math. J. [**47**]{} (2006), No. 1, 77–85. M.A.Muratov and V.I.Chilin, *Algebras of measurable and locally measurable operators.* – Kyiv, Pratsi In-ty matematiki NAN Ukraini. [**69**]{} (2007), 390 pp.(Russian). M.A.Muratov and V.I.Chilin, *Central extensions of $*$-algebras of measurable operators.* Reports of the National Academy of Science of Ukraine, [**7**]{} (2009), 24-28.(Russian). E.Nelson, *Notes on non-commutative integration,* J. Funct. Anal. [**15**]{} (1974) 103-116. D.Olesen, *Derivations of $AW^*$-algebras are inner*, Pacific J. Math., [**53**]{} (1974), No.1, 555–561. M. Reed, B. Simon, *Methods of Modern Mathematical Physics. Vol. 1: Functional Analysis,* Academic Press, 1980. K.Saito, *On the algebra of measurable operators for a $AW^*$-algebra, II*, Tohoku Math. J., [**23**]{} (1971), No. 3, 525-534. S.Sakai, *$C^{*}$–algebras and $W^{*}$–algebras*, Springer-Verlag, New York, 1971. S.Sakai, *Operator algebras in dynamical systems. The theory of unbounded derivations in $C^{*}$–algebras*, Cambridge: Cambridge University Press, 1991. S.Sankaran, *The $^{\ast}$-algebra of unbounded operators,* J. London Math. Soc. [**34**]{} (1959), 337–344. I.E.Segal, *A non-commutative extension of abstract integration*, Ann. of Math. [**57**]{} (1953), 401–457. M.Takesaki, *Theory of operator algebras I*, New York, Springer-Verlag, 1979. F.J.Yeadon, *Convergence of measurable operators*, Proc. Camb. Phil. Soc. [**74**]{} (1973), 257–268.
{ "pile_set_name": "ArXiv" }
--- abstract: 'On the Euclidean domains of classical signal processing, linking of signal samples to the underlying coordinate structure is straightforward. While graph adjacency matrices totally define the quantitative associations among the underlying graph vertices, a major problem in graph signal processing is the lack of explicit association of vertices with an underlying quantitative coordinate structure. To make this link, we propose an operation, called the *vertex multiplication*, which is defined for graphs and can operate on graph signals. Vertex multiplication, which generalizes the coordinate multiplication operation in time series signals, can be interpreted as an operator which assigns a coordinate structure to a graph. By using the graph domain extension of differentiation and graph Fourier transform (GFT), vertex multiplication is defined such that it shows Fourier duality, which states that differentiation and coordinate multiplication operations are duals of each other under Fourier transformation (FT). The proposed definition is shown to reduce to coordinate multiplication for graphs corresponding to time series. Numerical examples are also presented.' author: - 'Aykut Koç and Yigit E. Bayiz[^1][^2]' bibliography: - 'archives\_graph.bib' - 'archives\_lct.bib' title: 'Graph Signal Processing: Vertex Multiplication' --- [Shell : Bare Demo of IEEEtran.cls for IEEE Journals]{} Graph Signal Processing, Graph Fourier Transform, duality, coordinate multiplication, vertex multiplication. **EDICS Category: 3-BBND** Introduction ============ Classical digital signal processing (DSP) provides a useful tool for the analysis of signals defined by sampling an Euclidian space such as time-series signals and the ${\mathbb{R}}^2$ plane for images. However, the classical DSP theory is not designed to capture the complicated structures of large networks, such as social and economic networks, networks arising from the world wide web, and sensor networks. The graph signal processing (GSP) provides a new framework that can make the analysis of these large networks possible, [@shuman13emerging; @ortega18gspsurvey; @ribeiro18gsp; @stankovic19gsp1; @stankovic19gsp2; @gavili17shiftoperator; @sandryhaila14bigdata; @sandryhaila14freq; @stankovic19vertexfreq; @chen15sampling; @wang17gspfrt; @wang18gspfrtsampling; @sandryhaila13filtersicassp; @sandryhaila13discretegsp]. The GSP framework aims towards developing efficient methods for analyzing and processing signals with complex underlying structures such as networks or graphs [@ortega18gspsurvey; @sandryhaila13discretegsp]. Graph based extensions to several other areas such as machine learning and analysis of brain signals have also been investigated [@stankovic20gsp3machinelearning; @huang16graphbrain]. Classical signal processing concepts have been extended to the graph domain [@ortega18gspsurvey; @ribeiro18gsp; @stankovic19gsp1]. Specifically, filtering [@stankovic19gsp2; @ribeiro18gsp; @shuman13emerging; @sandryhaila13filtersicassp; @ortega18gspsurvey; @gavili17shiftoperator; @sandryhaila13discretegsp; @hua19graphfilter], frequency analysis [@sandryhaila14freq; @stankovic19gsp2; @grassi17timevertex], sampling [@chen15sampling; @anis14sampling], interpolation [@narang13interpolation], Fourier transform (FT) [@sandryhaila13gfticassp], signal reconstruction [@brugnoli20recons], processing of stationary signals and processes [@perraudin17stationary; @marques17stationary], and multiscale decomposition methods [@zheng16multiscaledec] have been considered for graph signals. Much of the developments in the processing of graph signals rely on extending the definition of FT and frequency analysis to graph signals[@sandryhaila14freq]. This allows the translation of many signal processing algorithms to graph signals, and provided a framework that has proved to be foundational in many novel GSP applications, including filtering, sampling and interpolation theory,[@chen15sampling; @narang13interpolation; @anis14sampling], big data analysis, [@sandryhaila14bigdata], and classification, [@chen14semisuper; @chen13adaptive]. There are two main approaches for extending the FT to the graph domain. The first is derived from the *spectral graph theory* and uses the graph Laplacian. This framework describes the graph Fourier transform (GFT) as a change of basis into the basis of the eigenvectors of the graph Laplacian [@shuman13emerging]. Although this approach is generally considered to be limited to analysis of undirected graphs, there also exist extensions to directed graphs, [@singh16directedlaplacian]. Built upon the *algebraic signal processing*, the second approach is based on the adjacency matrix of the graph and describes the GFT as a change of basis into the eigenvectors of the adjacency matrix [@sandryhaila13discretegsp]. This reflects the intuition that the adjacency matrix is analogous to the discrete shift matrix, and the eigenvalues of the latter form a basis for the discrete Fourier transform (DFT). The second approach, which we have adopted throughout this letter, supports both directed and undirected graphs. The relationships between operators on graphs and their FT counterparts have also been studied, [@sandryhaila13discretegsp; @shuman13emerging; @ortega18gspsurvey; @sandryhaila14freq; @stankovic19gsp2; @grassi17timevertex]. Specifically, important operations like convolution, translation, modulation, dilation, and filtering have been generalized to GSP domain in [@shuman13emerging]. In [@agaskar12uncertaintyongraphs], the uncertainty relation has been generalized to signals defined on graphs. FT relations for some of the familiar operators such as shift, modulation and dilation have also been studied in [@shuman13emerging]. Yet, Fourier duality for the differentiation on graphs has not been addressed. In this letter, by using the Fourier duality, we propose an operation, called the *vertex multiplication (VM)* for graphs. VM mimics the coordinate multiplication operator of time series signals (${\cal U} f(u) = uf(u)$), which is the Fourier dual of differentiation. While discretization of coordinate multiplication is straightforward, generalization to the graph domain is problematic since the vertices of a graph do not correspond to certain quantitative values, apart from just indices of order. Defined in a matrix form, VM can be interpreted as an operator which assigns a coordinate structure to a graph by associating a *“coordinate vector"* (represented by the columns of the matrix) for each vertex, and can also operate on graph signals (represented as vectors) through matrix multiplication. Since several established metrics and quantitative manipulations can be applied to coordinate vectors, the proposed direct assignment of coordinate structure to the vertex domain can also be instrumental in efforts to define “distance" metrics in the vertex domain, [@ron11relaxationdistance; @lafon06diffusiondistance], study the notion of smoothness of signals on graphs, [@zhu12apprxgraphsignals], localization of signals in the vertex-domain and study of transforms on graphs, [@coifman06diffusionwavelet; @hammond11graphwavelets; @narang12waveletgraph; @shuman13emerging]. The proposed VM generalizes a fundamental operation like coordinate multiplication to graph domain and defines the Fourier duality of differentiation for GSP. Moreover, VM, being defined totally consistent with the circulant and dual structure of the DFT, can also be considered a natural way to overcome a major obstacle in embedding the underlying structure of irregular vertex domain to a quantitative coordinate structure assigned to the vertices. This coordinate association is important in the ongoing generalizations from DSP to GSP. Duality Relation ================ The duality between differentiation and coordinate multiplication operators is particularly important in classical signal processing [@papoulis77book; @cohentannoudji77quantum]. This duality between time (space) and frequency (spatial-frequency) domains is also one of the most fundamental properties of the FT. In mathematical terms, let $\mathcal{U}$, $\mathcal{D}$ and $\mathcal{F}$ denote the coordinate multiplication, differentiation and FT operators, respectively. Continuous manifestations of the former two are: $$\label{eq:operatorU} {\cal U} f(u) = uf(u), \\$$ $$\label{eq:operatorD} {\cal D} f(u) = \frac{1}{2\pi j} \frac{df(u)}{du},$$ where $(2\pi j)^{-1}$ is included to make ${\cal U}$ and ${\cal D}$ precise Fourier duals (the effect of either in one domain is its dual in the other domain). Then, the duality is given as: $$\label{U_fdf} \mathcal{U}=\mathcal{FDF}^{-1}.$$ This relation can also manifest itself between shift and modulation operations, both of which are fundamental properties of Fourier analysis,[@shuman13emerging]. The duality relations of the FT in the classical signal processing theory are crucial for understanding much of the underlying theory as well as for being instrumental in applications, [@koc19ieee]. Therefore, an extension of these relations to GSP is inescapable. The duality relation creates a way to define the coordinate multiplication operator without needing coordinates to be explicitly defined once a differential operator on graphs and GFT are provided. Therefore, the extension of Fourier duality to graphs can be used to define a coordinate structure on graphs where one replaces $\mathcal{D}$ with the differential operator on graphs and $\mathcal{F}$ with GFT. GFT and Differentiation on Graphs ================================= A finite *graph* $G = (\mathcal{V},\mathbf{A})$ is a finite set of $N$ ordered points $\mathcal{V} = \{v_0,v_1\dots v_{N-1}\}$ (called vertices) which are connected to each other according to some relation. The connections in $G$ are represented by an *adjacency matrix* $\mathbf{A}$. The element $a_{ij}$ of $\mathbf{A}$ is the weight of the connection between $i$’th vertex and $j$’th vertex where $i,j = 0,1,...,N-1$. In general, this connection is directed. However for undirected graphs for which the connections are not directed, we have a symmetric $\mathbf{A}$ with $ a_{ij} = a_{ji}$. Any complex valued function $x$ defined on the set of vertices $\mathcal{V}$, i.e.: $ x:\mathcal{V} {\rightarrow}{\mathbb{C}}$ is called a *graph signal*. Since $\mathcal{V}$ is finite, it is convenient to represent $x$ as a vector where each index of $x$ is the value the signal takes on the corresponding vertex: $$\mathbf{x} = [x_0,x_1,\dots,x_{N-1}]^{\top},\quad x_i = x(v_i).$$ When viewed as a vector, operators acting on $x$ can be represented by left multiplication with matrices. Of these operators, the one represented by the adjacency matrix itself is of particular importance, since it implicitly contains the connectivity information of $G$. This operator is called a *graph shift* and extends the cyclic shift operator defined on a periodic time series signal in the DSP which has the graph structure shown in Fig. \[fig:exampleA\] ($G_1$ at the top part). In this case, the adjacency matrix is identical to the cyclic shift[@sandryhaila14freq]: $$\label{eq:tsAdjacency} \mathbf{A} = \begin{bmatrix} 0& 0&0&...&0&1 \\ 1& 0&0&...&0&0 \\ 0& 1&0&...&0&0 \\ .& .&.& &.&. \\ 0& 0&0&...&1&0 \\ \end{bmatrix}.$$ Let $G$ be a graph with adjacency matrix $\mathbf{A}$ and $x$ be a graph signal defined on $G$. Then $\mathbf{A}$ can be written in the Jordan canonical form as: $$\mathbf{A} = \mathbf{V\Lambda V^{-1}}.$$ Then, the *Graph Fourier Transform (GFT)* of $x$, denoted by $\Tilde{x}$, is defined as [@sandryhaila13discretegsp]: $$\Tilde{x} = \mathbf{V^{-1}} x.$$ GFT of a signal $x$ is unique up to the ordering of the Jordan blocks in the Jordan canonical form. If the adjacency matrix $\mathbf{A}$ of $G$ is diagonalizable, then the Jordan Canonical form is identical to diagonalizing $\mathbf{A}$. In this case GFT becomes a change of basis into the basis of eigenvectors of $\mathbf{A}$, and this process is unique up to the ordering of the Fourier basis vectors. It is also easy to see that if we are to take the GFT of a time series graph, which has the adjacency matrix of the form given in Eq. \[eq:tsAdjacency\], then the GFT basis becomes the eigenvectors of the cyclic shift matrix, and the GFT reduces to the DFT. We now consider the definition of differentiation operation on GSP domain, [@dees19unitaryshift]. Let $x^c: {\mathbb{R}}{\rightarrow}{\mathbb{C}}$ be any smooth periodic function with period $T$, and $\mathbf{t} = (t_0,t_1, ..., t_{N-1})$ be an ordered set of numbers selected from the interval $[0,T)$. Then an irregular sampling of $x^c$ with respect to sampling points $\mathbf{t}$ is the finite discrete signal $x\in{\mathbb{R}}^N$ where: $$x_n = x^c(t_n) \quad \forall n \in \{0,1,\dots,N-1\}.$$ Let $S_0$ be the circular forward shift operator defined in the space of finite discrete signals of length $N$. Then the matrix representation of $S_0$ is of the same cyclic shift form given in Eq. \[eq:tsAdjacency\]. Then, if we let the addition and subtraction on the vector indices to be defined in modulo $N$ (i.e. if we write $x_{-1} = x_{N-1}$ and $t_{-1} = t_{N-1} - T$), one can use the Taylor series expansion to write: $$\begin{aligned} (S_0 x)_n = x_{n-1} &= x^c(t_{n-1}) = \sum\limits_{k=0}^{\infty} \bigg(\frac{(t_{n-1} - t_{n})^k}{k!}\frac{d^k}{dt^k}\bigg|_{t_n}x^c\bigg) \nonumber \\ &= \bigg(\bigg(\sum\limits_{k=0}^{\infty} \frac{(t_{n-1} - t_{n})^k}{k!}\frac{d^k}{dt^k}\bigg) x^c\bigg)(t_{n}) \nonumber\\ &= \exp{\Big((t_{n-1} - t_{n})\frac{d}{dt}\Big)} x^c(t_{n}) .\end{aligned}$$ Thus, the discrete differential operator $\bm{\nabla}$ should satisfy: $$\mathbf{S_0} = \exp(-\bm{\Delta}_t\bm{\nabla}),$$ where $\bm{\Delta}_t = {\rm diag}(t_0 - t_{-1}, t_1 - t_0,...,t_{N-1}-t_{N-2})$. Then, the matrix manifestation of the differential operator defined on unequal sampling of $x^c$ resulting in the vector $x$ is defined as: $$\label{diff} \bm{\nabla} = -\bm{\Delta}_t^{-1} \log \mathbf{S_0},$$ where the complex logarithm can be defined on any branch cut as long as it is consistent throughout the analysis. In this letter, we assume the argument to be in the interval $[0,2\pi)$. [.23]{} [.23]{} [.23]{} [.23]{} Graph Vertex Multiplication =========================== Consider a graph $G$ with an adjacency matrix $\mathbf{A}$. Then the eigenvalues of $\mathbf{A}$ can be written as the ordered set $\{r_0e^{j\omega_0},\dots, r_{N-1}e^{j\omega_{N-1}}\}$ where $r_i$’s are the magnitudes and $\omega_i \in [0,2\pi)$ are given in an increasing order. Then for any graph signal $x$ with a GFT of $\Tilde{x}$, the elements of $\Tilde{x}$ can be ordered with respect to their corresponding eigenvalues. That is, if we write the eigenvalue decomposition of $\mathbf{A}$ as $\mathbf{A = V\Lambda V^{-1}}$ where the diagonal elements of $\mathbf{\Lambda}$ are ordered in an increasing order: $\mathbf{\Lambda} = {\rm diag}(r_0e^{-j\omega_0}, r_1e^{-j\omega_1},...,r_{N-1}e^{-j\omega_{N-1}})$. Then, $\Tilde{x} = \mathbf{V^{-1}}x$. Notice that we implicitly assume that $\mathbf{A}$ has eigenvalues with distinct arguments (frequencies) when we write the diagonalization of $\mathbf{A}$. In this ordering, each coordinate of $\Tilde{x}$ are in the order of (possibly irregularly) increasing frequency. Since $\Tilde{x}$ has no two coordinates corresponding to the same frequency, we can always find a smooth function $\Tilde{x}^c: {\mathbb{R}}{\rightarrow}{\mathbb{C}}$ such that $\Tilde{x}$ induces an irregular sampling on $\Tilde{x}^c$: $$\Tilde{x}_n = \Tilde{x}^c(\omega_n).$$ Then the discrete differentiation of $\Tilde{x}$ can be defined as: $$(\bm{\nabla}_F\Tilde{x})_n$$ where $\bm{\nabla}_F$ is the FT domain discrete differential operator: $$\label{deltaF_old} \bm{\nabla}_F = -\bm{\Delta}_\omega^{-1} \log \mathbf{S_0},$$ where $\bm{\Delta}_\omega = {\rm diag}(\omega_0 - \omega_{-1}, \omega_1 - \omega_0,..., \omega_{N-1}-\omega_{N-2})$ with $\omega_{-1} = (\omega_{N-1} - 2\pi)$. Before proceeding, to be able to use the precise duality relation given in Eq. \[U\_fdf\], we first alter the definition for the discrete derivative by dividing Eq. \[deltaF\_old\] by $j$: $$\Tilde{\bm{\nabla}}_F = j\bm{\Delta}_\omega^{-1} \log \mathbf{S_0},$$ which is analogous to the definition of differential operation with constant multiplier $(2\pi j)^{-1}$ as given in Eq. \[eq:operatorD\]. (Please note that $2\pi$ term is already encapsulated in frequency $w$.) Finally, by using the duality in Eq. \[U\_fdf\], we can define the precise Fourier dual of $\tilde{\nabla}_F$ as a new operator called *vertex multiplication* denoted by $\mathcal{U_G}$, in abstract operator notation. The matrix manifestation of $\mathcal{U_G}$, using GFT, is then given by: $$\mathbf{U}_G = \mathbf{V^{-1}}\tilde{\bm{\nabla}}_F\mathbf{V^{}}= j\mathbf{V^{-1}}(\bm{\Delta}_\omega^{-1} \log \mathbf{S_0})\mathbf{V}.$$ The proposed VM operator $\mathcal{U_G}$ can be interpreted as a collection of vectors $\mathbf{u_i}$ assigned to each vertex on the graph such that $ \mathbf{U}_G = [\mathbf{u}_0,\mathbf{u}_1, \dots,\mathbf{u}_{N-1}]$. Also, it operates on a graph signal $\mathbf{x}=[x_0, \dots ,x_{N-1}]^\top$ as: $$\label{coordinate vector} \mathbf{y} = \mathbf{U}_G\mathbf{x} = \sum_{i=0}^{N-1}{x_i\mathbf{u_i}}.$$ This definition provides a generalization of the coordinate multiplication operator. As such, the VM operator computes the superposition of the multiplication of the signal value on each vertex with the vector $\mathbf{u_i}$ associated with the same vertex. Hence, the columns of the VM matrix mimic the coordinate values in the coordinate multiplication operator in DSP (Eq. \[eq:operatorU\]). Thus, we shall call these columns $\mathbf{u_i}$ the *coordinate vector* of the i’th vertex. Due to the summation in Eq. \[coordinate vector\], the coordinate vector of a vertex has a global effect on the behavior of the VM, and the output graph signal value $\mathbf{y_i}$ depends on values of the input signal at all vertices through the coordinate vectors. The coordinate multiplication of classical DSP can be interpreted as a special case of VM. The matrix representing the coordinate multiplication, denoted by $\mathbf{U}$, is diagonal. It is composed of one-hot column vectors with the non-zero entries being only at the corresponding indices of each vertex. Then the effect of each coordinate vector in Eq. \[coordinate vector\] can be represented locally and thus, the diagonal elements of $\mathbf{U}$ can be assigned to each vertex as proper coordinates. This can be stated more formally in the following lemma: **Lemma 1:** For time series graphs, Eq. \[coordinate vector\] is equivalent to the coordinate multiplication of the classical DSP. **Proof:** Consider the graph representation of a time-series. The adjacency matrix in this case is equivalent to the forward time shift, i.e. $\mathbf{A} = \mathbf{S_0}$. The eigenvalues of $\mathbf{A}$ are then equally spaced on the unit circle in the complex plane and are of the form $e^{j\omega_k}$ where $\omega_k = 2k\pi/N$. Thus we have $$\begin{split} \bm{\Delta}_\omega &= {\rm diag}( \frac{2\pi(0-(-1))}{N}, \frac{2\pi(1-0)}{N},..., \\ & \frac{2\pi((N-1)-(N-2))}{N}) = \frac{2\pi}{N}\mathbf{I}, \end{split}$$ where $\mathbf{I}$ is the $N\times N$ identity matrix. Then, the Fourier domain derivative becomes $-\frac{N}{2\pi} \log \mathbf{S_0}$. This leads to the following vertex multiplication operator: $$\begin{split} \mathbf{U_G} = j\mathbf{V^{-1}}(\frac{N}{2\pi}\log \mathbf{S_0})\mathbf{V} = j\frac{N}{2\pi}\log(\mathbf{V^{-1}} \mathbf{S_0}\mathbf{V)}, \end{split}$$ where $\mathbf{V^{-1}}$ reduces to the DFT. Using elementary properties of the DFT, we can obtain $$\begin{split} \mathbf{V^{-1}} \mathbf{S_0}\mathbf{V} = {\rm diag}(e^{-j\omega_0}, e^{-j\omega_1},..., e^{-j\omega_{N-1}}) = \mathbf{\Lambda}. \end{split}$$ Then, $\mathbf{U_G}$ is the diagonal matrix with entries $\frac{N\omega_k}{2\pi}$. Finally, $$\mathbf{U_G} = \mathbf{U} = {\rm diag}(0,1,2,...,N-1),\,$$ which is consistent with the discrete coordinate multiplication operator, which can be written as $(\mathbf{U}x)_n = nx_n.$ $\blacksquare$ Numerical Examples ================== The effects of the underlying graphs on the corresponding VM operators and their subsequent effects on a graph signal are numerically demonstrated in Fig. \[fig:example\]. $G_1$ is the time series graph. We have chosen $G_2$ such that it deviates from time series by only having $v_0$ makes an extra connection to $v_2$; and $G_3$ such that each vertex $v_i$ has connections to the $v_{i+1}$ and $v_{i+2}$. It is immediately clear that any deviation from the time series graph yields to the matrix manifestation of the VM operator to have complex components. Fig. \[fig:example\] also presents the vertex-multiplied output graph signals shown by colormaps on the vertices. In fact, experimentation with similar graph structures seems to imply that the largest cycle in the structure has a significant effect on the resulting VM matrix. Further study of this relation can lead to methods for detecting the largest cycles in graph structures. Since VM enables us to define quantitative coordinate vectors, possibilities for manipulations are endless. As an example, in Fig. \[fig:coor\], $L_{1}$-norms of the columns of the VM matrices are also plotted, i.e. given $ \mathbf{U}_G = [\mathbf{u}_0,\mathbf{u}_1, \dots,\mathbf{u}_{N-1}]$, $\mathbf{u}_i$’s are assigned to each vertex and coordinates are calculated by $||\mathbf{u}_i||_1$. Both $L_{1}$-norms and their normalized values to the reference coordinates ($0,1,...,7$) of the time series by using the scaling $7*\left( ||\mathbf{u}_i||_1 - \min_{i}||\mathbf{u}_i||_1 \right)/\left( \max_{i}||\mathbf{u}_i||_1- \min_{i}||\mathbf{u}_i||_1\right )$ are plotted. Intuitive behavior can be observed in Fig. \[fig:coor\]. First, the extra connection between $v_0$ and $v_2$ at the very beginning of the graph almost merges $v_0$ and $v_1$ so that their coordinates become close to each other; then the linearly increasing coordinate structure continues as in the time series. Second, $G_3$ leads to an interesting coordinate pattern due to the regular structure of the graph. Still, a linearly increasing diagonal coordinate structure exists with the abrupt deviation where central vertices $v_3$ and $v_5$ seem to exchange their coordinates. [0.24]{} ![Coordinates are calculated by $L_1$-norm of the coordinate vectors for $G_1$, $G_2$, and $G_3$. (Left: $L_1$-norms. Right: $L_1$-norms normalized to the time-series coordinate interval $[0,7]$.[]{data-label="fig:coor"}](coor-eps-converted-to.pdf "fig:"){width="\linewidth"} \[fig:coorA\] [.24]{} ![Coordinates are calculated by $L_1$-norm of the coordinate vectors for $G_1$, $G_2$, and $G_3$. (Left: $L_1$-norms. Right: $L_1$-norms normalized to the time-series coordinate interval $[0,7]$.[]{data-label="fig:coor"}](coor_normalized-eps-converted-to.pdf "fig:"){width="\linewidth"} \[fig:coorB\] Conclusions =========== We proposed a generalization of the coordinate multiplication operation to the graph domain, called vertex multiplication. By using Fourier duality, the proposed VM is in consistence with the classical signal processing in which differentiation and coordinate multiplication are duals. VM can be interpreted as an operator that assigns a coordinate structure to a graph by assigning each vertex a coordinate vector. These coordinate vectors, which are intrinsically consistent with the FT theory and its dual structure, can be manipulated to further assign single coordinates to the vertices or for other purposes. We showed that the VM reduces to the coordinate multiplication for time series signals. Given an adjacency matrix, such an explicit coordinate association can be helpful in the ongoing generalizations from classical signal processing to GSP. It may also lead to new theoretical and computational endeavors, and deepen our theoretical understanding of the link between vertex and frequency domains with possible insights and applications to the notion of smoothness, distance metrics and localization in the vertex-domain, and transform designs for signals on graphs. [^1]: Yigit E. Bayiz is with the Department of Electrical and Electronics Engineering, Bilkent University, Bilkent, Ankara, Turkey [^2]: Aykut Koç (Senior Author) is with the Department of Electrical and Electronics Engineering, Bilkent University, Bilkent, Ankara, Turkey, and also with the National Magnetic Resonance Research Center (UMRAM), Bilkent University, Bilkent, Ankara, Turkey, e-mail: [email protected]
{ "pile_set_name": "ArXiv" }
--- abstract: 'We partially confirm a conjecture of Donaldson relating the greatest Ricci lower bound $R(X)$ to the existence of conical Kähler-Einstein metrics on a Fano manifold $X$. In particular, if $D\in |-K_X|$ is a smooth simple divisor and the Mabuchi $K$-energy is bounded below, then there exists a unique conical Kähler-Einstein metric satisfying ${\mathrm{Ric}}(g) = \beta g + (1-\beta) [D]$ for any $\beta \in (0,1)$. We also construct unique smooth conical toric Kähler-Einstein metrics with $\beta=R(X)$ and a unique effective $\mathbb{Q}$-divisor $D\in [-K_X]$ for all toric Fano manifolds. Finally we prove a Miyaoka-Yau type inequality for Fano manifolds with $R(X)=1$.' address: - 'Department of Mathematics, Rutgers University, Piscataway, NJ 08854' - 'Department of Mathematics, Rutgers University, Newark, NJ 07102' author: - Jian Song - Xiaowei Wang title: 'The greatest Ricci lower bound, conical Einstein metrics and the Chern number inequality' --- [^1] Introduction ============ The existence of Kähler-Einstein metrics has been a central problem in Kähler geometry since Yau’s celebrated solution [@Y1] to the Calabi conjecture. Constant scalar curvature metrics with conical singularities have been extensively studied in [@Mc; @Tr; @LT] for Riemann surfaces. In general, we can consider a pair $(X, D)$ for an $n$-dimensional compact Kähler manifold and a smooth complex hypersurface $D$ of $X$. A conical Kähler metric $g$ on $X$ with cone angle $2\pi \beta$ along $D$ is locally equivalent to the following model edge metric $$g= \sum_{j=1}^{n-1} dz_j \otimes d\bar z_j + |z_n|^{-2(1-\beta)} d z_n \otimes d\bar z_n,$$ if $D$ is defined by $z_n=0$. Applications of conical Kähler metrics are proposed [@Ts; @T94] to obtain various Chern number inequalities. Recently, Donaldson developed the linear theory to study the existence of canonical conical Kähler metrics in [@D4]. And Brendle in [@Br] solved Yau’s Monge-Ampère equations for conical Kähler metrics with cone angle $2\pi \beta$ for $\beta\in (0,1/2)$ along a smooth divisor $D$. The general case is settled by Jeffres, Mazzeo and Rubinstein [@JMR] for all $\beta\in (0,1)$. As an immediate consequence, there always exists conical Kähler-Einstein metrics with negative or zero constant scalar curvature with cone angle $2\pi \beta$ along a smooth divisor $D$ for $\beta\in (0,1)$. When $X$ is a Fano manifold, Donaldson’proposes to study the conical Kähler-Einstein equation $$\label{KE-D} {\mathrm{Ric}}(\omega) = {\beta}\omega + (1-{\beta}) [D],$$ where $D$ is smooth simple divisor in the anticanonical class $[-K_X]$ and $\beta \in (0, 1)$. One of the motivations is that one can study the existence problem for smooth Kähler-Einstein metrics on $X$ by deforming the cone angle. Such an approach can be regarded as a variant of the standard continuity method. In particular, since Tian and Yau have already established the existence of a complete Ricci-flat Kähler metric on the non-compact manifold $X\setminus D$ in [@TY], one would expect that is solvable for ${\beta}$ sufficient small. This is comfirmed successfully by Berman in [@Be]. Now the question is how large ${\beta}$ can be. The largest ${\beta}$ is closely related to the following holomorphic invariant known as the greatest Ricci lower bound first introduced by Tian in [@T92]. \[Rg\] Let $X$ be a Fano manifold. The greatest Ricci lower bound $R(X)$ is defined by $$ R(X)=\sup\{\beta\mid {\mathrm{Ric}}(\omega)\geq \beta \omega, \text{ for some smooth K\"{a}hler metric }\omega\in c_1(X)\}.$$ It is proved by Szekelyhidi in [@Sze] that $[0, R(X))$ is the maximal interval for the continuity method to solve the Kähler-Einstein equation on a Fano manifold $X$. In particular, it is independent of the choice for the initial Kähler metric when applying the continuity method. The invariant $R(X)$ is explicitly calculated for ${\mathbb{P}}^2$ blown up at one point by Szekelyhidi [@Sze], and for all toric Fano manifolds by Li [@LiC]. It is well-known that Mabuchi $K$-energy being bounded from below implies $R(X)=1$ and it is proved in [@MS] that $R(X)=1$ implies $X$ being $K$-semistable. The following conjecture is proposed by Donaldson in [@D4] to relate $R(X)$ to the existence of conical Kähler-Einstein metrics. \[D-conj\]There does not exist a conical Kähler-Einstein metric solving (\[KE-D\]) if $\beta\in (R(X),1]$, while there exists one if $\beta\in (0, R(X))$. This conjecture can be considered as a different geometric interpretation of the invariant $R(X)$. Another importance of the conjecture lies in the fact that it supplies a new approach to the Yau’s conjecture [@Y3] on the equivalence of existence of Kähler-Einstein metric for Fano manifolds and certain algebro-geometric stability condition, which is refined and extended by Tian [@T97] and Donaldson [@D]. The algebro-geometric aspect of Conjecture \[D-conj\] has been studied by Li [@LiC3], Sun [@Sun], Odaka-Sun [@OS] and Berman [@Be-l]. In particular, the notion of Log $K$-stability is introduced in [@LiC3] and [@Sun] as the algebro-geometric obstruction to solve equation . In particular, $R(X)$ can be applied to test the Log $K$-stability of $X$ when it is toric Fano. In [@Be-l], Berman proves that Log $K$-stability is a necessary condition to the solution of . This naturally leads to the Log version of Yau-Tian-Donaldson conjecture, that is, to establish the equivalence of the solvability of and the Log $K$-stability of $(X,D)$ (cf.[@LiC3] and [@OS]). An interesting observation made by Sun in [@Sun] is that $K$-stability implies Log $K$-stability. Now we describe the main results of the present work. The first one is to partially confirm Conjecture \[D-conj\]. We consider a more general class of conical Kähler-Einstein metrics with smooth simple divisors in any pluricanonical systems, and remove the assumption in [@D4] on $D$ by showing there exists no holomorphic vector field tangential to $D$ (c.f. Theorem \[nohol\]). \[main1\] Let $X$ be a Fano manifold and $R(X)$ be the greatest lower bound of Ricci curvature of $X$. 1. For any $\beta \in [R(X), 1]$ and any smooth simple divisor $D\in |-mK_X|$ for some $m\in \mathbb{Z}^+$, there does not exist a smooth conical Kähler-Einstein metric $\omega$ satisfying $$\label{KE} {\mathrm{Ric}}(\omega) = \beta \omega + \frac{ 1-\beta }{m} [ D ], $$ if $R(X)<1$. 2. For any $\beta \in (0, R(X))$, there exist a smooth simple divisor $D\in |-mK_X|$ for some $m \in \mathbb{Z}^+$ and a smooth conical Kähler-Einstein metric $\omega$ satisfying . The second part of the theorem is not completely satisfactory in the sense that one would like to have an $m$ that is independent of $\beta\in (0, R(X))$. In the case when the Mabuchi $K$-energy is bounded below, or more generally $R(X)=1$, we show that $D$ and $\beta$ do not rely on $\beta$. \[main2\] Let $X$ be a Fano manifold. If the Mabuchi $K$-energy $\mathcal{M}$ is bounded below and if $D\in |-K_X|$ is a smooth simple divisor, then for any $\beta\in (0, 1)$ there exists a smooth conical Kähler-Einstein metric satisfying the conical Kähler-Einstein equation $${\mathrm{Ric}}(\omega) = \beta \omega + (1-\beta) [D].$$ In general, if the paired Mabuchi $K$-energy $\mathcal{M}_{D, R(X)}$ is bounded below for a smooth simple divisor $D \in |-m K_X |$ for some $m\in \mathbb{Z}^+$, then for any $\beta\in (0, R(X))$ there exists a smooth conical Kähler-Einstein metric satisfying equation (\[KE\]). We would like to remark that most results of Theorem \[main1\] and \[main2\] are independently obtained by Li and Sun [@LS]. Theorem \[main2\] might have many applications. In particular, if the Mabuchi $K$-energy is bounded below, there exists a sequence of conical Kähler-Einstein metrics ${\mathrm{Ric}}(g_\epsilon ) = (1-\epsilon) g + \epsilon [D]$ as $ \epsilon \rightarrow 0 $. $(X, g_\epsilon)$ might converge in Gromov-Hausdorff topology to a $\mathbb{Q}$-Fano variety $X_\infty$ coupled with a canonical Kähler-Einstein metric. Theorem \[main2\] also holds if $R(X)=1$ and $D\in |-mK_X|$ for some $m\geq 2$ as in the following proposition. By Bertini’s theorem, there always exists a smooth simple divisor $D\in |-mK_X|$ for $m$ sufficiently large. \[rmain1\]Let $X$ be a Fano manifold and $D\in |-mK_X|$ be a smooth simple divisor for some $m\geq 2$. Then for any $\beta\in \left(0, \frac{(m-1)\cdot R(X)}{m-R(X)}\right)$, there exists a smooth conical Kähler-Einstein metric $\omega$ satisfying for $D$. In particular, when $R(X)=1$, equation with $D$ is solvable for any ${\beta}\in (0,1)$. The invariant $R(X)$ can also be identified as the optimal constant for the nonlinear Moser-Trudinger inequality. Let $X$ be a Fano manifold and $\omega\in c_1(X)$ be a smooth Kähler metric on $X$. One can define the following $F$-functional by $$F_{\omega, \beta} = J_\omega(\varphi) - \frac{1}{V} \int_X \varphi \omega^n - \frac{1}{\beta} \log \frac{1}{V} \int_X e^{-\beta \varphi} \omega^n,$$ where $$J_\omega(\varphi)=\frac{\sqrt{-1}}{V}\sum_{i=0}^{n-1}\frac{i+1}{n+1}\int_X{\partial}\varphi\wedge{\overline{\partial}}\varphi\wedge\omega^i\wedge\omega^{n-1-i}_\varphi,$$ is the Aubin-Yau functional, $\omega_\varphi=\omega+\sqrt{-1}{\partial}{\overline{\partial}}\varphi>0$ and $V= \int_X \omega^n$. As a corollary of Theorem \[main1\], we can establish a connection between $R(X)$ and the Moser-Trudinger inequality. 1. If $\beta \in (0, R(X))$, $F_{\omega, \beta} $ is bounded below and $J$-proper (see Definition \[strong\]) on $PSH(X, \omega) \cap L^\infty(X)$, or equivalently, there exist $\epsilon, C_\epsilon>0$ such that the following Moser-Truding inequality holds$$\int_X e^{-\beta \varphi} \omega^n \leq C_\epsilon e^{ (\beta-\epsilon) J_\omega (\varphi) - \frac{\beta}{V} \int_X \varphi \omega^n}$$ for $\varphi\in PSH(X,\omega)\cap L^\infty(X)$. 2. If $\beta\in (R(X), 1)$, then $$\inf_{PSH(X, \omega)\cap L^\infty(X)} F_{\omega, \beta}(\cdot) = - \infty.$$ It is first proved in [@T97; @TZ] for the properness of $F$-functional on Fano Kähler-Einstein manifolds without holomorphic vector fields. The $J$-properness for $F$ is conjectured in this case in [@T97] and is later proved in [@PSSW]. The presence of the smooth divisor $D$ for $\beta\in (0,1]$ blocks holomorphic vector fields as shown in Theorem \[nohol\]. It is interesting to ask if $F_{\omega, \beta}$ is always bounded below on $PSH(X, \omega) \cap L^\infty(X)$ if $\beta = R(X)$. In the case of toric Fano manifolds, $F_{\omega, \beta}$ is indeed bounded from below if $\beta=R(X)$ as a corollary of the following theorem (Corollary \[torlowbd\]). A more interesting problem will be to understand the limiting behavior of the conical Kähler-Einstein metrics as $\beta\rightarrow R(X)$ as the holomorphic vector field will appear in the limiting space. The following theorem serves an example for the above speculation. \[main3\] Let $X$ be a toric Fano manifold. Then there exist a unique effective toric $\mathbb{Q}$-divisor $D\in |-K_X|$ and a unique smooth toric conical Kähler metric $\omega$ satisfying $${\mathrm{Ric}}(\omega) = R(X) \omega + (1-R(X)) [D]. $$ Furthermore, $R(X)$ is the largest number for $\beta\in (0, 1]$ such that $$\label{toreqnb} {\mathrm{Ric}}(\omega) = \beta \omega + (1-\beta) [D_\beta] $$ admits a smooth conical toric solution $\omega_\beta$ for an effective toric $\mathbb{R}$-divisor $D_\beta$ in $[-K_X]$. We remark that the divisor $D$ can not be smooth, instead it is a union of effective smooth toric $\mathbb{Q}$-divisors with simple normal crossings. Theorem \[main3\] is closely related to the results of [@LiC2] with a different approach for the limiting behavior of the continuity method. The proof of Theorem \[main3\] relies on the toric setting introduced in [@D1; @D2] and the estimates in [@WZ]. For $\beta > R(X)$, there still exists a smooth conical solution for equation (\[toreqnb\]), however, $D_\beta$ won’t be effective and so the Ricci current of the conical metric cannot be positive. In Theorem \[main3\], $R(X)$ will be explicitly calculated as in [@LiC] and $D$ is determined by Lemma \[divisord\] and Lemma \[torickefr\] . For example, when $X$ is ${\mathbb{P}}^2$ blown up at one point, it admits a toric ${\mathbb{P}}^1$ bundle $\pi: X \rightarrow {\mathbb{P}}^1$, then $R(X)=6/7$ and $D = 2D_\infty + (H_1 + H_2)/2$, where $D_\infty$ is the section at the infinity, $H_1$ and $H_2$ are the two ${\mathbb{P}}^1$ fibres invariant under the torus action. This seems to suggest that Donaldson’s conjecture might only hold for smooth simple divisors lying in the pluri-anti-canonical system. In fact, it is shown in [@LS] Theorem \[main3\] can be applied to prove Conjecture \[D-conj\] in the toric case when one is allowed to replace $|-K_X|$ by the linear system of a suitable power of $-K_X$. Finally, we will give some applications of Theorem \[main2\]. In general, the conical Kähler metrics do not have bounded curvature tensors, as they might blow up near the conical divisor, particularly when the cone angle is greater than $\pi$. However, we can show that if the Ricci curvature of a conical Kähler metric is bounded, its curvature tensors are always bounded in $L^2$-norm. \[chpro\] Let $X$ be a Kähler manifold and $D$ be a smooth simple divisor on $X$. Let $g$ be a smooth conical Kähler metric on $X$ with cone angle $2\pi \beta$ along $D$ with $\beta\in (0,1)$. If the Ricci curvature of $g$ is bounded, the $L^2$-norm of the curvature tensors of $g$ is also bounded. When we say the Ricci curvature of a conical Kähler metric $g$ is bounded, it means that the Ricci curvature of $g$ is uniformly bounded on $X\setminus D$ by ignoring the mass along $D$ in the Ricci current. Proposition \[chpro\] enables us to define Chern characters, and in particular the Chern numbers for those conical Kähler metrics and derive corresponding Gauss-Bonnet and signature formulas for Kähler surfaces with conical singularities along a smooth holomorphic curve $\Sigma$. This is related to recent results of Atiyah and Lebrun [@AL] for smooth Riemannian $4$-folds. In fact, the bound of $L^2$-norm of the curvature tensors only depend on the scalar curvature bound and topological invariants such as intersection numbers among $D$, the first, second Chern classes. When the greatest Ricci lower bound $R(X)$ is $1$, we have the Miyaoka-Yau type inequality holds for $X$ as in the Kähler-Einstein case [@Mi; @Y2]. In such a case, $X$ is $K$-semistable [@MS] and the result below reflects the general perspective connecting $K$-semistability and the Chern number inequality. \[main4\] Let $X$ be a Fano manifold. If $R(X)=1$, then the Miyaoka-Yau type inequality holds $$c_2(X) \cdot c_1(X)^{n-2} \geq \frac{n} {2(n+1)} c_1(X)^n. $$ In general, if $D\in |-K_X|$ is a smooth simple divisor and if the paired Mabuchi $K$-energy $\mathcal{M}_{D, \beta}$ is bounded below, then $$ c_2(X) \cdot c_1(X)^{n-2} \geq \frac{n\beta^2} {2(n+1)} c_1(X)^n. $$ Parallel argument can be applied to give a complete proof of the Chern number inequality for smooth minimal models of general type using conical Kähler-Einstein metrics. This approach is first proposed by Tsuji [@Ts], while the analytic estimates seem missing. We remark that the first complete proof for smooth minimal models of general type is due to Zhang [@ZhY] by using the Kähler-Ricci flow. $R(X)$ and conical Kähler-Einstein metrics ========================================== Paired energy functionals ------------------------- In this section, we recall the paired energy functionals originally introduced in [@Be]. \[fdef\] Let $X$ be a Fano manifold and $\omega\in c_1(X)$ be a Kähler current with bounded local potential and $\Omega_\theta $ be a nonnegative real $(n,n)$-current on $X$ whose curvature $$\theta = -{\sqrt{-1}\partial{\overline{\partial}}}\log \Omega_\theta \in c_1(X)$$ is a nonnegative $(1,1)$-current. Let $\Omega_\omega$ be a nonnegative real $(n,n)$-current on $X$ defined by $${\sqrt{-1}\partial{\overline{\partial}}}\log \Omega_\omega = \omega.$$ Suppose that for some $\beta\in (0, 1]$, $$\int_X (\Omega_\omega)^\beta (\Omega_\theta)^{1-\beta} = V = c_1(X)^n.$$ We define the paired $F$-functional by $$\label{F} F_{\omega,\theta, \beta}(\varphi)= J_\omega(\varphi)- \frac{1}{V} \int_X \varphi \omega^n -\frac{1}{\beta}\log\frac{1}{V}\int_X \left( e^{-\varphi } \Omega_\omega \right)^\beta \left( \Omega_\theta\right)^{1-\beta} ,$$ for $\varphi \in PSH(X, \omega) \cap L^\infty(X)$, where $$J_\omega(\varphi)=\frac{\sqrt{-1}}{V}\sum_{i=0}^{n-1}\frac{i+1}{n+1}\int_X{\partial}\varphi\wedge{\overline{\partial}}\varphi\wedge\omega^i\wedge\omega^{n-1-i}_\varphi,$$ is the Aubin-Yau $J$-functional and $\omega_\varphi=\omega+\sqrt{-1}{\partial}{\overline{\partial}}\varphi\geq 0$. The Euler-Lagrangian equation for (\[F\]) is given by $$\label{criteqn1} (\omega+{\sqrt{-1}\partial{\overline{\partial}}}\varphi)^n = (e^{-\varphi} \Omega_\omega)^\beta (\Omega_\theta)^{1-\beta}, $$ and the corresponding curvature equation is $$\label{criteqn2} {\mathrm{Ric}}(\omega_\varphi ) = \beta \omega_\varphi + (1-\beta) \theta. $$ When $\beta =1$, $F_{\omega, \theta, \beta} (\varphi) = F_\omega$ is the original Ding’s functional [@Di]. The paired $F$-functional also satisfies the cocycle condition by slightly modifying the proof for the original $F$-functional. \[cocy\] $F_{\omega, \theta, \beta} $ satisfies the following cocycle condition $$F_{\omega, \theta, \beta} (\varphi) - F_{\omega_\psi, \theta, \beta} (\varphi - \psi) = F_{\omega, \theta, \beta}(\psi) $$ for any $\varphi, \psi \in PSH(X, \omega) \cap L^\infty(X)$. If we write $F(\omega, \omega_\varphi) = F_{\omega, \theta,\beta}(\varphi)$, then the cocycle condition is equivalent to the following $$F(\omega, \omega_\varphi) + F(\omega_\varphi, \omega_\psi) + F(\omega_\psi, \omega) =0.$$ \[strong\] We say a functional $G( \cdot )$ is $J$-proper on $PSH(X, \omega)\cap L^\infty(X)$ if there exist ${\delta},\ C_{\delta}>0$ such that $$G(\varphi)\geq {\delta}J_{\omega}(\varphi)-C_{\delta}$$ for all $ \varphi\in {\mathrm{PSH}}(X,\omega)\cap L^\infty(X)$. Let $X$ be a Fano manifold and $D$ be a smooth simple divisor in $|-mK_X|$. Let $s$ be a defining section of $[D]$. Since $s\in H^0(X, -mK_X)$, $$\Omega_D= |s|^{- \frac{2}{m}}=(s\otimes \overline{s})^{ - \frac{1}{m}}$$ can be considered as a smooth nonnegative real $(n,n)$-form with poles along $D$ of order $m^{-1}$. Obviously, ${\mathrm{Ric}}(\Omega_D) = -{\sqrt{-1}\partial{\overline{\partial}}}\log \Omega_D = m^{-1} [D]$. We then define the following notations for conveniences. \[modf\] Let $D\in |-mK_X|$ be a smooth simple divisor for some $m \in \mathbb{Z}^+$. We define $$\begin{aligned} \mathcal{F}_{\omega, \beta}(\varphi) &=& F_{\omega, m^{-1}[D], \beta} (\varphi)\\ &=& J_{\omega}(\varphi) - \frac{1}{V} \int_X \varphi \omega^n - \frac{1}{\beta} \log \frac{1}{V} \int_X e^{-\beta \varphi} (\Omega_\omega)^{\beta} (\Omega_D)^{1-\beta} \nonumber $$ and $$F_{\omega, \beta} (\varphi) = J_{\omega}(\varphi) - \frac{1}{V} \int_X \varphi \omega^n - \frac{1}{\beta} \log \frac{1}{V} \int_X e^{-\beta \varphi} \omega^n. $$ In order to relate the Moser-Trudinger inequality to $R(X)$, we introduce Let $X$ be a Fano manifold and $\omega\in c_1(X)$ be a smooth Kähler metric. We define the optimal [*Moser-Trudinger constant*]{} by $${\mathfrak{mt}}(X)=\sup\left\{{\beta}\in (0, 1] \mid \inf_{PSH(X, \omega)\cap L^\infty(X)} F_{\omega, \beta}(\cdot)> - \infty \right\}.$$ It is straightforward to verify that the invariant ${\mathfrak{mt}}(X)$ does not depend on the choice of the Kähler metric $\omega\in c_1(X)$. We also define the following paired Mabuchi $K$-energy for conical Kähler metrics first introduced in [@Be]. Here a conical Kähler metric is called smooth if it a polyhomogenous Kähler edge metric defined by Jeffres, Mazzeo and Rubinstein in [@JMR] and we let $C^\infty_{D, \beta}(X)$ denote the set of all smooth conical Kähler metrics with cone angle $2\pi \beta$ along the smooth divisor $D$. Let $X$ be a Fano manifold. Suppose $\omega$ and $\omega_\varphi$ are two smooth conical Kähler metrics in $c_1(X)$ with cone angle $2\pi (1-(1-\beta)/m)$ along a smooth simple divisor $D \in |-mK_X|$. The paired Mabuchi $K$-energy is defined by $$\mathcal{M}_{\omega, \beta} (\varphi)= \frac{1}{V} \int_X \log \frac{\omega_\varphi^n}{\omega_n} - \beta (I_\omega-J_\omega)_\omega (\varphi) + \frac{1}{V} \int_X h_\omega (\omega^n - \omega_\varphi^n), $$ where $h_\omega$ is the Ricci potential of $\omega$ defined by ${\sqrt{-1}\partial{\overline{\partial}}}h_\omega = {\mathrm{Ric}}(\omega) - \omega$, and $$I_\omega (\varphi) = \sqrt{-1} \sum_{i=0}^{n-1} \int_X \partial \varphi \wedge {\overline{\partial}}\varphi \wedge \omega^i \wedge \omega_\varphi^{n-i-1}$$ is the Aubin-Yau $I$-functional. It is proved in [@Be] that if the conical Kähelr-Einstein equation is solvable for the data $(D, \beta)$, both ${\mathcal{F}}_{\omega, \beta}$ and $\mathcal{M}_{\omega, \beta}$ are bounded below $PSH(X, \omega)\cap C^\infty(X)_{D,{\beta}}$. Furthermore, if one is bounded below, the other must also be bounded below, and conversely, if either of the functionals is $J$-proper, the Monge-Ampère equation associated to the conical Kähler-Einstein equation admits a bounded solution [@Be]. Moreover, the solution is a smooth conical Kähler-Einstein metric in the sense of [@JMR]. Pluri-anticanonical system -------------------------- In this section, we will remove the assumption on the smooth simple divisor $D$ in [@D4] and construct conical Kähler-Einstein metrics by deforming the angle along the continuity method. The following lemma is an immediate consequence of the adjunction formula and Bertini’s Theorem. \[noholV\] Let $X$ be a Fano manifold of $\dim X\geq 2$. For any sufficiently large $m\in \mathbb{Z}^+$, there exists a smooth simple divisor $D\in \left| -mK_X \right|$. In particular, $$c_1(D)=(1-m)c_1(X)\mid_D = \frac{1-m}{m} [D]|_D.$$ This shows that the smooth simple divisor $D$ lies in $ |-mK_X|$ is a Calabi-Yau manifold if and only if $m=1$. \[nohol\] Let $X$ be a Fano manifold of $\dim X\geq 2$ and $D$ be a smooth simple divisor in $|-mK_X|$ for some $m\in \mathbb{Z}^+$. Then there does not exist any holomorphic vector field tangential to $D$. For $\dim X=2$ it follows from the classification that if $X$ admits a nontrivial holomorphic vector field, $X$ is isomorphic to ${\mathbb{P}}^2$, ${\mathbb{P}}^1\times {\mathbb{P}}^1$ or ${\mathbb{P}}^2$ blown up at $1,2$ or $3$ points. $D\in |-mK_X|$ implies that $g(D)\geq 1$. For ${\mathbb{P}}^2$ blown up at $0, 1, 2$ or $3$ points, any holomorphic vector field on $X$ is the lifting of a holomorphic vector field on ${\mathbb{P}}^2$ fixing the blown-up points. Hence any smooth invariant divisor of such holomorphic vector fields on $X$ must be ${\mathbb{P}}^1$ with $g(D)=0$. It is also straightforward to check that any invariant divisor on ${\mathbb{P}}^1\times {\mathbb{P}}^1$ must also be $P^1$. So from now on let us assume that $\dim X\geq 3$. First, we claim that if there exists such a holomorphic vector field, it must vanish along $D$. Since $X$ is Fano, we have $\pi_1(X)=0$ and hence $\pi_1(D)=0$ by Lefschetz hyperplane theorem and our assumption that $\dim X\geq 3$. Since $m>0$, either $c_1(D)<0$ or $D$ is a simply connected Calabi-Yau manifold by Lemma \[noholV\]. In both cases $D$ does not admit any nontrivial holomorphic vector field. So our claim is proved. Second, we claim there is no holomorphic vector field vanishing along $D$. It suffices to show that $$\label{h0} H^0(X,TX\otimes K_X^ m)=0 $$ thanks to the following exact sequence $$\begin{CD} 0@>>>TX\otimes K_X^{m}@>>>TX@>>>TX\mid_D@>>>0. \end{CD}$$ Since $TX\otimes K_X\cong \Omega^{n-1}_X$, we have $$H^0(X,TX\otimes K_X^{ m})=H^0(X,\Omega^{n-1}_X\otimes K_X^{\otimes(m-1)})).$$ Now if $m>1$ then the right hand side is $0$ by Kodaira-Akizuki-Nakano vanishing theorem and the fact that $K_X$ is negative. For $m=1$, equation (\[h0\]) follows from $H^0(X,\Omega^{n-1}_X)\cong H^{n-1}(X,\sO_X)=0$ by Kodaira vanishing theorem and $X$ being Fano. Theorem \[nohol\] is speculated by Donaldson [@D4] and is proved in [@Be] for the special case when the holomorphic vector field is Hamiltonian and $m=1$. Combined with the openness result in [@D4], we immediately have the following corollary. \[open\] Let $X$ be a Fano manifold and $D\in |-mK|$ be a smooth simple divisor for some $m\in \mathbb{Z}^+$. If there exists a smooth conical Kähler-Einstein metric satisfying $${\mathrm{Ric}}(g) = \beta g + \frac{1-\beta}{m} [D]$$ for some $\beta\in (0,1)$, then there exists $\epsilon >0$ such that for any $\beta'$ with $|\beta-\beta'|< \epsilon$, there exists a smooth conical Kähler-Einstein metric $g'$ satisfying ${\mathrm{Ric}}(g') = \beta' g + \frac{1-\beta'}{m} [D].$ The $\alpha$-invariant and the Moser-Trudinger inequality --------------------------------------------------------- Let $X$ be a Fano manifold and $D$ be a smooth simple divisor in $|-mK_X|$ for some $m\in \mathbb{Z}^+$. Let $\omega' \in c_1(X) $ be a smooth Kähler form and let $\Omega_{\omega'} $ be a smooth volume form on $X$ such that $${\mathrm{Ric}}(\Omega_{\omega'}) = -{\sqrt{-1}\partial{\overline{\partial}}}\log \Omega_{\omega'}= \omega' .$$ We now apply the continuity method and consider the following family of equations for $\beta\in [0, 1]$. $$\label{cont} (\omega' + {\sqrt{-1}\partial{\overline{\partial}}}\varphi_t)^n = e^{-t\varphi_t} ( \Omega_{\omega'} )^\beta (\Omega_D)^{1-\beta}, ~~~t\in [0, \beta]. $$ We let $$S=\{ t\in [0, \beta]\mid \text{(\ref{cont}) is solvable for some } t \text{ with } \omega_t \text{ a smooth conical K\"ahler metric } \}.$$ By the results in [@JMR], $0\in S$ and $S$ is open. Let $\omega_t = \omega + \varphi_t$ for any $t\in S$. The curvature equation of (\[cont\]) is given by $${\mathrm{Ric}}(\omega_t) = t\omega_t + (\beta-t)\omega' + \frac{1-\beta}{m} [D] \geq t \omega_t.$$ Hence the Green’s function for $\omega_t$ is uniformly bounded below by $t$ for all $t\in S$ [@JMR]. Furthermore, let $\Delta_t$ be the Laplace operator associated to $\omega_t$. Then $$\Delta_t \dot \varphi_t = -\varphi_t - t \dot \varphi_t.$$ Following the argument for smooth case with slight modification to the conical Kähler metrics, one can show the following proposition. It is proved in a more general setting in [@Be]. \[lowbd1\] Let $X$ be a Fano manifold and $D\in |-mK_X|$ be a smooth simple divisor. 1. If there exists $\beta \in (0, 1]$ and a smooth conical Kähler-Einstein metric $\omega_{KE}$ satisfying $${\mathrm{Ric}}(\omega_{KE}) = \beta \omega_{KE} + \frac{1-\beta}{m} [D],$$ then the paired $F$-functional $${\mathcal{F}}_{\omega_{KE}, \beta} (\varphi)= J_{\omega_{KE}} (\varphi) - \frac{1}{V} \int_X \varphi \omega_{KE} ^n - \frac{1}{\beta} \log \frac{1}{V} \int_X \left( e^{-\varphi } \Omega_{\omega_{KE}} \right)^\beta ( \Omega_D)^{1-\beta} .$$ is uniformly bounded below for all $\varphi \in PSH(X, \omega_{KE} ) \cap L^\infty(X)$. 2. If $\omega\in c_1(X)$ is a smooth Kähler metric and ${\mathcal{F}}_{\omega, \beta} (\varphi) $ is $J$-proper on $PSH(X, \omega) \cap L^\infty(X)$ for some $\beta\in (0, 1]$, then there exists a unique smooth conical Kähler metric $\omega_{KE} $ solving $${\mathrm{Ric}}(\omega_{KE} ) = \beta \omega_{KE} + \frac{1-\beta}{m} [D].$$ Same argument in the proof of Proposition \[lowbd1\] can be applied to prove the following lemma if replacing $m^{-1} [D]$ by a smooth Kähler metric $\theta \in c_1(X)$. \[lowerbd2\] Let $X$ be a Fano manifold and $\theta$ be a smooth Kähler metric in $c_1(X)$. 1. If there exists a smooth Kähler metric $\omega_\theta$ on $X$ satisfying $${\mathrm{Ric}}(\omega_\theta) = \beta \omega_\theta + (1-\beta)\theta$$ for some $\beta\in (0, 1]$. Then $$F_{\omega_\theta, \theta, \beta} (\varphi) = J_{\omega_\theta}(\varphi) - \frac{1}{V} \int_X \varphi \omega_\theta^n - \frac{1}{\beta} \log \frac{1}{V} \int_X e^{-\beta\varphi} (\Omega_{\omega_\theta})^\beta (\Omega_\theta)^{1-\beta}$$ is uniformly bounded below on $PSH(X, \omega) \cap L^\infty(X)$. 2. If $\omega\in c_1(X)$ is a smooth Kähler metric and $F_{\omega, \theta, \beta} (\varphi) $ is $J$-proper on $PSH(X, \omega) \cap L^\infty(X)$ for some $\beta\in (0, 1]$, then there exists a unique smooth Kähler metric $\omega_\theta $ solving $${\mathrm{Ric}}(\omega _\theta ) = \beta \omega_\theta + (1-\beta)\theta.$$ The $\alpha$-invariant is introduced by Tian [@T87] to obtain a sufficient condition for the existence of Kähler-Einstein metrics on Fano manifolds. It is shown by Demailly [@De] that the $\alpha$-invariant coincides with the log canonical threshold in birational geometry. It is natural to relate the log canonical threshold for pairs to the paired $\alpha$-invariant. It is first introduced in [@Be] as a generalization for the $\alpha$-invariant. Let $X$ be a Fano manifold and $D\in |-mK_X|$ be a smooth simple divisor. Let $s$ be the defining section of $[D]$ and $h$ be a smooth hermitian metric for $-mK_X$. Let $\omega \in c_1(X)$ be a smooth Kähler metric. Then we define the paired $\alpha$-invariant for $\beta \in (0,1]$ by $$\alpha_{D, \beta} (X)= \sup \left\{ \alpha>0 ~\left|~ \sup_{ \varphi \in PSH(X, \omega)\cap L^\infty(X) } \int_X |s|_h^{-\frac{2(1-\beta)}{m} } e^{-\alpha\beta({\varphi-\sup\varphi})} \omega^n <\infty \right. \right\}. $$ It is straightforward to check that the invariant $\alpha_{D, \beta}$ does not depend on the choice of $\omega \in c_1(X)$. The following theorem is due to Berman [@Be] for an effective bound on $\alpha_{D, \beta}$ to construct conical Kähler-Einstein metrics by combining the results in [@JMR]. \[al\] There exists $\beta_D \in (0, 1]$ such that for all $\beta\in (0,\beta_D]$ we have $$\alpha_{D, \beta} (X)> \frac{n}{n+1} . $$ In particular, there exists a smooth conical Kähler-Einstein metric $\omega \in c_1(X)$ satisfying $${\mathrm{Ric}}(\omega) = \beta \omega + \frac{ 1-\beta }{m} [D]$$ for $\beta \in (0, \beta_D)$. In [@S04], the first author proves that if the $\alpha$-invariant on an $n$-dimensional Fano manifold is greater $n/(n+1)$, then the $F$-functional is $J$-proper. The following theorem is a generalization for the conical case. \[alprop\] Let $X$ be a Fano manifold and $\omega\in c_1(X)$ be a smooth Kähler metric. If $D\in |-mK_X|$ is a smooth simple divisor and if there exists $\beta\in (0, 1]$ such that $$\alpha_{D, \beta} (X)> \frac{n}{n+1},$$ then the functional $${\mathcal{F}}_{\omega, \beta} (\varphi)= J_{\omega} (\varphi) - \frac{1}{V} \int_X \varphi \omega^n - \frac{1}{\beta} \log \frac{1}{V} \int_X \left( e^{-\varphi } \Omega_{\omega} \right)^\beta ( \Omega_D)^{1-\beta}$$ as in Definition \[modf\], is $J$-proper on $PSH(X, \omega)\cap L^\infty(X)$. We break the proof into the following two steps. 1. Since $\alpha_{D, \beta}(X)>n/(n+1)$, by Theorem \[al\] there exists a smooth conical Kähler-Einstein metric $\omega_{KE}$ satisfying $${\mathrm{Ric}}(\omega_{KE}) = \beta \omega_{KE} + \frac{1-\beta}{m} [D].$$ Let $PSH(X, \omega_{KE}, K)$ be the set of all $\varphi \in PSH(X, \omega_{KE})\cap L^\infty(X) $ such that $$\label{assum} {\mathrm{osc}}_X \varphi = \sup_X \varphi - \inf_X \varphi \leq (n+1)J_{\omega_{KE}} (\varphi) +K. $$ We claim that ${\mathcal{F}}_{\omega_{KE}, \beta}$ is $J$-proper for all $\varphi\in PSH(X, \omega_{KE}, K)$. To see that, take ${\alpha}$ satisfying $$\frac{n\beta}{n+1}<{\alpha}< \beta \min ({\alpha}_{D,{\beta}}, 1)$$ and let $\Omega_D=|s|^{2/m}$ then we have $$\begin{aligned} \int_Xe^{-{\beta}\varphi}\Omega_{\beta}&=&\int_Xe^{-{\alpha}(\varphi-\sup\varphi)+({\alpha}-{\beta})\phi-{\alpha}\sup\varphi} \Omega_D\\ &\leq&C e^{({\alpha}-{\beta})\inf\varphi-{\alpha}\sup\varphi} \int_X e^{-{\alpha}(\varphi-\sup \varphi)}\Omega_D\\ &\leq&C e^{({\alpha}-{\beta})\inf\varphi-{\alpha}\sup\varphi}\end{aligned}$$ by the definition of $\alpha_{D, \beta}$. By assumption , we have $$\begin{aligned} \int_Xe^{-{\beta}\varphi}\Omega_D &\leq &Ce^{({\beta}-{\alpha})(n+1)J_{\omega_{KE}}(\varphi)-{\beta}\sup \varphi)} \\ &\leq &Ce^{(n+1)({\beta}-{\alpha})J_{\omega_{KE}}(\varphi)-\frac{{\beta}}{V}\int_X \varphi\omega_{KE}^n} \\ &=&Ce^{{\beta}J_{\omega_{KE}}(\varphi)-((n+1){\alpha}-n{\beta})J_{\omega_{KE}}(\varphi)-\frac{{\beta}}{V}\int_X \varphi\omega_{KE}^n},\end{aligned}$$ and our claim follows by taking logarithm of both sides of the above inequality. 2. Now we will remove the assumption (\[assum\]) for $\varphi\in PSH(X, \omega)\cap L^\infty(X)$. We first consider all $\varphi \in PSH(X, \omega)$ such that $\omega'= \omega_{KE}+{\sqrt{-1}\partial{\overline{\partial}}}\varphi$ is a smooth conical Kähler metric with cone angle $2\pi (1 - \frac{1-\beta}{m})$ along $D$. By applying the same argument for smooth Kähler-Einstein metrics to solve the continuity method backward, one can show that $$(\omega' + {\sqrt{-1}\partial{\overline{\partial}}}\varphi_t ) ^n = \left( e^{- t \varphi_t } \Omega_{\omega'} \right)^{\beta} \Omega_D^{1-\beta}$$ admits a smooth conical solution for $t\in [0, 1]$ because the implicit function theorem can be applied at $t=1$ due to Theorem \[nohol\]. In particular, $\varphi_1 = - \varphi.$ Let $\omega_t = \omega' + {\sqrt{-1}\partial{\overline{\partial}}}\varphi_t$. Note that for $t\geq 1/2$, the Ricci curvature of $\omega_{KE}+ {\sqrt{-1}\partial{\overline{\partial}}}(\varphi_t - \varphi_1)= \omega_t$ is no less than $\beta/2$. Then the Green’s functions for both $\omega$ and $\omega_t$ are uniformly bounded from below by $-G$ for some positive number $G$, and $$\Delta_{\omega_{KE}} (\varphi -\varphi_1) = tr_{\omega_{KE}} (\omega_t - \omega_{KE}) \geq -n , ~~ \Delta_{\omega_t} (\varphi_t - \varphi_1) \leq n.$$ Then by the Green’s formula, for $t\geq \beta/2$, we have $$\frac{1}{V} \int_X ( \varphi_t - \varphi_1) \omega_{KE}^n - nG \leq ( \varphi_t - \varphi_1) \leq \frac{1}{V} \int_X ( \varphi_t - \varphi_1) \omega_t^n + n G.$$ Hence $${\mathrm{osc}}_X (\varphi_t - \varphi_1) \leq I_{\omega_{KE}} (\varphi_t - \varphi_1) + 2nG \leq (n+1)J_{\omega_{KE}}(\varphi_t- \varphi_1) + 2nG.$$ This implies that $\varphi_t - \varphi_1\in PSH(X, \omega_{KE}, 2n G)$ and then the $J$-properness holds for $\varphi_t - \varphi_1$, and there exist $\delta, C_1, C_2>0$ such that $$\begin{aligned} {\mathcal{F}}_{\omega_{KE}, \beta} (\varphi_t - \varphi_1) &\geq& \delta J_{\omega_{KE}} (\varphi_t - \varphi_1) - C_1\\ &\geq& \frac{\delta}{n+1} {\mathrm{osc}}_X (\varphi_t - \varphi_1) -C_2. $$ Consequently, there exist $C_3>0$ such that $$\begin{aligned} n (1-t) J_{\omega_{KE}} (\varphi) &\geq & (1-t) J_{\omega'}(\varphi_1)\\ &\geq & (1-t) ( I _{\omega'} (\varphi_1) - J_{\omega'}(\varphi_1))\\ &\geq & \int_t^1(I_{\omega'}(\varphi_s) - J_{\omega'}(\varphi_s)) ds\\ &=& {\mathcal{F}}_{\omega', \beta}( \varphi_t) - {\mathcal{F}}_{\omega', \beta} (\varphi_1) \\ &=& {\mathcal{F}}_{\omega_{KE}, \beta} (\varphi_t - \varphi_1) \\ &\geq & \left(\frac{{\alpha}}{{\beta}}-\frac{n}{n+1}\right) {\mathrm{osc}}_X (\varphi_t - \varphi_1) - C_3. $$ The third inequality follows by the increasing monotonicity for $(I_{\omega'} - J_{\omega'})(\varphi_t) $ for $t\in [0, 1]$. Then by applying the cocycle condition and the same argument in the smooth case in [@T97; @TZ], we have $$\begin{aligned} {\mathcal{F}}_{\omega_{KE}, \beta} (\varphi) &= & - {\mathcal{F}}_{\omega', \beta} (\varphi_1)\\ &=& \int_0^1 ( I_{\omega'} (\varphi_t) - J_{\omega'} (\varphi_t)) dt \\ &\geq& (1-t) (I_{\omega'} (\varphi_t) - J_{\omega'}(\varphi_t)) \\ &\geq& \frac{1-t}{n} J_{\omega'} (\varphi_t) \\ &\geq& \frac{1-t}{n} J_{\omega'}(\varphi_1) -\frac{2 (1-t)}{n} {\mathrm{osc}}_X (\varphi_t - \varphi_1) - C_4\\ &\geq&\frac{ 1-t }{n^2} J_{\omega_{KE}} (\varphi) - C_5 (1-t)^2J_{\omega_{KE}} (\varphi) - C_6. $$ Since $C_5$ and $C_6$ are independent of the choice for $t\geq 1/2$, by choosing $t$ sufficiently close to $1$, we can find $\epsilon', C_{\epsilon'}>0$, such that $$\label{FF-ppr} {\mathcal{F}}_{\omega_{KE}, \beta}(\varphi) \geq \epsilon' J_{\omega, \beta}(\varphi) - C_{\epsilon'}$$ for all $\varphi \in PSH(X, \omega_{KE})$ such that $\omega_\varphi$ is a smooth conical Kähler metric with cone angle $2\pi (1-\frac{1-\beta}{m})$ along $D$. The set of such $\varphi$ is dense in $PSH(X, \omega_{KE}) \cap L^\infty(X)$ and so the $J$-properness holds for $PSH(X, \omega_{KE})\cap L^\infty(X)$. 3. Finally, the $J$-properness for any $\omega$ follows from Lemma \[cocy\] and the fact the Green’s function for $\omega_{KE}$ is bounded from below. An interpolation formula ------------------------ In this section, we will prove the following interpolation formula for the $F$-functional to obtain $J$-properness. \[proper4.4\] Let $X$ be a Fano manifold and $D$ a smooth simple divisor in $|-mK_X|$ for some $m\in \mathbb{Z}^+$. Let $\omega$ be a smooth Kähler metric in $c_1(X)$. If there exists $\alpha \in (0, 1]$ such that $$\inf_{PSH(X, \omega)\cap L^\infty(X)} {\mathcal{F}}_{\omega, \alpha} (\cdot ) >-\infty$$ then ${\mathcal{F}}_{\omega, \beta} (\varphi) $ is $J$-proper on $PSH(X, \omega) \cap L^\infty(X)$ for all $\beta \in (0, \alpha)$. We want to show that $${\mathcal{F}}_{\omega, \beta} (\varphi) = J_{\omega} (\varphi) - \frac{1}{V} \int_X \varphi \omega^n - \frac{1}{\beta} \log \frac{1}{V} \int_X (e^{-\varphi}\Omega_\omega)^{\beta} (\Omega_D)^{1-\beta}$$ is $J$-proper for all $\beta\in (0, \alpha)$. First for $0<\tau<{\beta}<{\alpha}$, we write ${\beta}=\tau/p+{\alpha}/q$ for some $1/p+1/q=1$. Then Hölder inequality implies the interpolation $$\begin{aligned} {\mathcal{F}}_{\omega,{\beta}}(\varphi)\!\!\!\!\! &=& \!\!\!J_\omega(\varphi)-\frac{1}{{\beta}}\log\frac{1}{V}\int_X(e^{-\varphi}\Omega_\omega)^{\beta}\Omega_D^{1-{\beta}}\\ &=&\!\!\!\left( \frac{\tau}{{\beta}p}+\frac{{\alpha}}{{\beta}q}\right) J_\omega(\varphi)-\frac{1}{{\beta}}\log\frac{1}{V}\int_X\left(e^{-\varphi}\frac{\Omega_\omega}{\Omega_D}\right)^{\tau/p+{\alpha}/q} \cdot \Omega_D\\ &\geq &\!\!\! \frac{\tau}{{\beta}p}\left( J_\omega(\varphi)-\frac{1}{\tau}\log\frac{1}{V} \int_X(e^{-\varphi}\Omega_\omega)^\tau\Omega_D^{1-\tau} \right) +\frac{{\alpha}}{{\beta}q}\left(J_\omega(\varphi)-\frac{1}{{\alpha}}\log \frac{1}{V}\int_X (e^{-\varphi}\Omega_\omega)^\tau\Omega_D^{1-\tau}\right)\\ &=&\frac{\tau}{{\beta}p}\cdot {\mathcal{F}}_{\omega,\tau}(\varphi)+\frac{\alpha}{{\beta}q}\cdot {\mathcal{F}}_{\omega,{\alpha}}(\varphi)\\ &\geq&\frac{\tau}{{\beta}p}\cdot {\mathcal{F}}_{\omega,\tau}(\varphi) - C_1\\ &\geq& \epsilon J_{\omega} (\varphi)-C.\end{aligned}$$ The last inequality follows from Theorem \[alprop\] by choosing $\tau$ sufficiently small so that $\alpha_{D, \tau} > n/(n+1)$. The same argument in the proof of Proposition \[proper4.4\] can be applied to prove the following lemma after replacing $m^{-1}[D]$ by a smooth Kähler metric $\theta\in c_1(X)$. \[prop4.5\] Let $X$ be a Fano manifold and $\theta$ be a smooth Kähler metric in $c_1(X)$. Let $\omega$ be a smooth Kähler metric in $c_1(X)$. If there exists $\alpha \in (0, 1]$ such that $$F_{\omega, \alpha} (\varphi) = J_{\omega} (\varphi) - \frac{1}{V} \int_X \varphi \omega^n - \frac{1}{\alpha} \log \frac{1}{V} \int_X e^{-\alpha\varphi} (\Omega_\omega)^{1-\alpha} (\Omega_\theta)^{1-\alpha}$$ is bounded below on $PSH(X, \omega) \cap L^\infty(X)$, then $F_{\omega, \beta} (\varphi) $ is $J$-proper on $PSH(X, \omega) \cap L^\infty(X)$ for all $\beta\in (0, \alpha)$. We remark that Lemma \[prop4.5\] also serves as an alternative proof for Theorem 1.1 in [@Sze] relating $R(X)$ and the continuity method. Proof of Theorem \[main1\] -------------------------- Let us prove the first part of Theorem \[main1\]. \[alphaup\]Let $X$ be a Fano manifold and $D$ be a smooth simple divisor in $|-mK_X|$ for some $m\in \mathbb{Z}^+$. If there exist $\beta\in (0, 1]$ and a smooth conical Kähler-Einstein metric $\omega$ satisfying $${\mathrm{Ric}}(\omega) = \beta \omega + \frac{1-\beta}{m} [D],$$ then $$\beta\leq R(X).$$ In particular, the inequality holds if and only if $\beta=1$. Let $\omega_{KE}$ be a smooth conical Kähler-Einstein metric on $X$ satisfying ${\mathrm{Ric}}(\omega_{KE}) = \beta \omega_{KE} + \frac{1-\beta}{m} [D]$. By Proposition \[lowbd1\], we know that ${\mathcal{F}}_{\omega_{KE}, \beta}$ is bounded below. By Proposition \[proper4.4\], ${\mathcal{F}}_{\omega, \beta'}$ is $J$-proper for all $\beta' \in (0, \beta)$. Let $\omega, \theta \in c_1(X)$ be two smooth Kähler metrics on $X$. The $J$-properness of ${\mathcal{F}}_{\omega, \beta'}$ immediately implies the $J$-properness for $$F_{\omega, \theta, \beta'} (\varphi) = J_{\omega} (\varphi) - \frac{1}{v} \int_X \varphi \omega^n - \frac{1}{\beta'} \log \frac{1}{V} \int_X e^{-\beta' \varphi} (\Omega_\omega)^{\beta'} (\Omega_\theta)^{1-\beta'}$$ because $\Omega_D$ is strictly bounded below from $0$. Then by Lemma \[lowerbd2\], there exists a minimizer of $F_{\omega, \theta, \beta'}$ which soloves the equaiton $${\mathrm{Ric}}(\omega)= \beta' \omega + (1-\beta') \theta \geq \beta' \omega .$$ This shows that $R(X)\geq \beta'$ and so $R(X) \geq \beta$. If $\beta= R(X) <1$, there must exist $\epsilon>0 $ and a smooth conical Kähler-Einstein metric $g$ such that ${\mathrm{Ric}}(g) = (\beta+\epsilon) g + (1-\beta-\epsilon)m^{-1} [D]$ by Corollary \[open\]. Then it is a contradiction to the definition of $R(X)$ by repeating the previous argument. We now prove the second part of Theorem \[main1\]. \[alphalow\] Let $X$ be a Fano manifold. Then for any $\beta \in (0, R(X))$, there exist a smooth simple divisor $D\in |-mK_X|$ for some $m\in \mathbb{Z}^+$ and a smooth conical Kähler-Einstein metric $g$ satisfying $${\mathrm{Ric}}(\omega) = \beta \omega + \frac{1-\beta}{m} [D].$$ We break the proof into the following steps. 1. Let $\omega$ and $\theta$ be two smooth Kähler metrics in $c_1(X)$. For any $\beta \in (0, R(X))$, by Szeklyhidi’s result [@Sze], the following family of Monge-Ampère equation $$(\omega+{\sqrt{-1}\partial{\overline{\partial}}}\varphi_t)^n = \left( e^{- t \varphi } \Omega_{\omega} \right) ^{\beta} (\Omega_\theta)^{1-\beta}, ~~~ \int_X (\Omega_{\omega})^\beta (\Omega_\theta)^{1-\beta } = V$$ is solvable for all $t\in [0, 1]$. Then by Lemma \[lowerbd2\], $$F_{\omega, \beta} (\varphi) = J_{\omega} (\varphi) - \frac{1}{V} \int_X \varphi \omega^n - \frac{1}{\beta} \log \int_X e^{-\beta \varphi} (\Omega_{\omega})^\beta (\Omega_\theta)^{1-\beta}$$ is bounded below on $PSH(X, \omega) \cap L^\infty(X)$. By Lemma \[prop4.5\], for any $\beta' \in (0, \beta)$, $F_{\omega, \beta'} (\varphi) $ is $J$-proper. It immediately follows that for any $\beta\in (0, R(X))$, there exist $\epsilon, C_\epsilon>0$ such that for all $\varphi \in PSH(X, \omega) \cap L^\infty(X)$, $$\int_X e^{- \beta \varphi} \omega^n \leq C_\epsilon e^{ (\beta -\epsilon) J_{\omega}(\varphi) - \frac{\beta}{V} \int_X \varphi \omega^n }.$$ 2. Let $D$ be a smooth simple divisor in $|-mK_X|$. We will later choose $m$ sufficiently large. Let $s$ be a defining section of $D$ and $h$ be a smooth hermitian metric on $-mK_X$. For any $\beta \in (0, R(X))$, there exists $\delta>0$ such that $\beta + \delta < R(X)$. Then $$\begin{aligned} \int_X |s|_h^{-\frac{2(1-\beta)}{m} }e^{ - \beta \varphi} \omega^n&\leq & \left(\int_X e^{-(\beta+\delta) \varphi} \omega^n \right)^{\frac{\beta}{\beta+\delta}} \left( \int_X |s|_h^{-\frac{2(1-\beta)(\beta+\delta) }{m\delta} } \omega^n \right)^{\frac{\delta}{\beta+\delta}}\\ &\leq& C_\delta \left(\int_X e^{-(\beta+\delta) \varphi} \omega^n \right)^{\frac{\alpha}{\alpha+\delta}} $$ if we choose $m > \frac{\delta}{(1-\beta)(\beta+ \delta)}$. By the conclusion in Step 1, there exist $\epsilon, C_\epsilon>0$ such that $$\int_X |s|^{-\frac{2(1-\beta)}{m} }e^{ - \beta \varphi} \omega^n \leq C_\epsilon e^{ (\beta -\epsilon) J_{\omega}(\varphi) - \frac{\beta}{V} \int_X \varphi \omega^n }.$$ Equivalently, ${\mathcal{F}}_{\omega, \beta}(\varphi)$ is $J$-proper on $PSH(X, \omega)\cap L^\infty(X)$. By Proposition \[lowbd1\], there exists a unique smooth conical Kähler-Einstein metric $\omega_\beta$ solving $${\mathrm{Ric}}(\omega_\beta) = \beta \omega_\beta + \frac{1-\beta}{m} [D].$$ Now we can relate the optimal Moser-Tridinger constant to the invariant $R(X)$. \[flowbd\] Let $X$ be a Fano manifold and $\omega\in c_1(X)$ be a smooth Kähler metric. 1. If $\beta \in (0, R(X))$, $F_{\omega, \beta} $ is $J$-proper on $PSH(X, \omega) \cap L^\infty(X)$. Equivalently, there exist $\epsilon, C_\epsilon>0$ such that the following Moser-Trudinger inequality holds for all $\varphi \in PSH(X, \omega)\cap L^\infty(X)$ $$\int_X e^{-\beta \varphi} \omega^n \leq C_\epsilon e^{ (\beta-\epsilon) J_\omega (\varphi) - \frac{\beta}{V} \int_X \varphi \omega^n}.$$ 2. If $\beta\in (R(X), 1)$, then $$\inf_{PSH(X, \omega)\cap L^\infty(X)} F_{\omega, \beta}(\cdot) = - \infty.$$ For $\beta\in (0, R(X))$ and a fixed smooth Kähler metric $\theta\in c_1(X)$, there exists a smooth Kähler metric $\omega$ satisfying ${\mathrm{Ric}}(\omega) = \beta\omega + (1-\beta) \theta$. The corollary is an immediate consequence of by modifying the interpolation formula in Proposition \[proper4.4\], after replacing $m^{-1} [D]$ by $\theta$. Immediately, we can show that $R(X)$ and ${\mathfrak{mt}}(X)$ take the same value for a Fano manifold $X$. Let $X$ be a Fano manifold. Then $${\mathfrak{mt}}(X) = R(X) = \sup \{ \beta\in (0, 1) ~|~F_{\omega, \beta}~\text{is~} J\text{-proper~on } PSH(X, \omega)\cap L^\infty(X)\}, $$ where $\omega\in c_1(X)$ is a smooth Kähler metric. Proof of Theorem \[main2\] -------------------------- Before proving Theorem \[main2\], we first quote the following proposition establishing the equivalence for the Mabuchi $K$-energy and the $F$-functional when either of them is bounded below proved in [@LiH] by applying the Kähler-Ricci flow and Perelman’s estimates for the scalar curvature. \[knu\] Let $X$ be a Fano manifold and $\omega\in c_1(X)$ be a smooth Kähler metric. Then Ding’s functional $F_\omega$ is bounded below on $PSH(X, \omega)\cap C^\infty(X)$ if and only if the Mabuchi $K$-energy is bounded below. Proposition \[knu\] holds for the paired Mabuchi $K$-energy and the paired $F$-functional as shown in [@Be]. One can also apply the continuity method for the conical Kähler metrics with positive Ricci curvature as in [@Ru]. Let $PSH(X, \omega) \cap C_{D, \beta}^\infty(X)$ be the set of all bounded $\varphi$ such that $\omega + {\sqrt{-1}\partial{\overline{\partial}}}\varphi$ is a smooth conical Kähler metric with cone angle $2\pi\beta$ along $D$. \[knu2\] Let $X$ be a Fano manifold and $D\in |-mK_X|$ be a smooth simple divisor. Let $\omega$ be a smooth conical Kähler metric in $c_1(X)$ with cone angle $2\pi (1-\beta)/m$ along $D$. Then $$\inf_{ PSH(X, \omega)\cap C^\infty_{D, \frac{1-\beta}{m} } (X) }\mathcal{M}_{\omega, \beta} (\cdot)> -\infty$$ is equivalent to $$\inf_{PSH(X, \omega) \cap L^\infty(X)} \mathcal{F}_{\omega, \beta} (\cdot) > t - \infty.$$ We can now prove Theorem \[main2\]. \[main4\_1\]Let $X$ be a Fano manifold and $D$ be a smooth simple divisor in $|-mK_X|$ for some $m\in \mathbb{Z}^+$. If the paired Mabuchi $K$-energy $\mathcal{M}_{\omega, R(X)} $ on $X$ is bounded below, then for any $\beta\in (0, R(X))$ there exists a smooth conical Kähler-Einstein metric satisfying $$\label{keeqn4} {\mathrm{Ric}}(g) = \beta g + \frac{1-\beta}{m}[D].$$ Let $\omega\in c_1(X)$ be a smooth Kähler metric. Then by Proposition \[knu2\], $\mathcal{F}_{\omega, R(X)}(\varphi)$ is bounded below on $PSH(X, \omega)\cap L^\infty(X)$. Applying the interpolation formula in Proposition \[proper4.4\], ${\mathcal{F}}_{\omega, \beta}$ is $J$-proper on $PSH(X, \omega) \cap L^\infty(X)$ for all $\beta\in (0, R(X))$. The theorem follows by Proposition \[lowbd1\]. When the Mabuchi $K$-energy is bounded below on $X$, for any $\beta\in (0, 1)$, there exists a conical Kähler-Einstein metric satisfying equation (\[keeqn4\]) for $m=1$. In this case, $D$ is a smooth Calabi-Yau hypersurface of $X$. If we only assume $R(X)=1$, we have the same conclusion in Theorem \[main4\_1\] for the linear systems $|-mK_X|$ with $m\geq 2$. Let $X$ be a Fano manifold and $D$ be a smooth simple divisor in $|-mK_X|$ for some $m\geq 2$. Then for any $\beta\in \left(0, \frac{(m-1)\cdot R(X)}{m-R(X)}\right)$, there exists a smooth conical Kähler-Einstein metric $\omega$ satisfying $${\mathrm{Ric}}(\omega) = \beta \omega + \frac{1-\beta}{m} [D].$$ In particular, when $R(X)=1$, we have conical Kähler-Einstein metric for any ${\beta}\in (0,1)$. We prove the case for $R(X)=1$ and the general case follows from the same argument. Let $s$ be the defining section of $|-mK_X|$ and $h$ be a smooth hermitian metric of $-mK_X$. Then $|s|_h^{-(2-\epsilon)}$ is integrable for any $\epsilon>0$. Furthermore, $F_{\omega, (1-\beta)m^{-1}[D], \beta}$ is proper for any smooth Kähler forms $\omega, \theta\in c_1(X)$ if $\beta\in (0,1)$, Then the proposition can be proved by similar interpolation argument in the proof of Theorem \[main4\_1\]. Conical toric Käher-Einstein metrics ==================================== Conical toric Kähler metrics ---------------------------- In this section, we will introduce toric conical Kähler metrics on toric Kähler manifolds and corresponding toric Kähler and symplectic potentials as in [@D1; @D2]. We begin with some basic definitions for toric manifolds. A convex polytope $P$ in $\mathbb{R}^n$ is called a Delzant polytope if a neighborhood of any vertex of $P$ is $SL(n,{\mathbb{Z}})$ equivalent to $\{x_j\geq 0, j=1, ..., n\} \subset \mathbb{R}^n $. $P$ is called an integral Delzant polytope if each vertex of $P$ is a lattice point in $\mathbb{Z}^n$. Let $P$ be an integral Delzant polytope in $\mathbb{R}^n$ defined by $$P=\{ x\in \mathbb{R}^n~|~ l_j (x) >0, j=1, ..., N\}, $$ where $$l_j(x) = v_j \cdot x + \lambda_j$$ and $v_i $ a primitive integral vector in $\mathbb{Z}^n$ and $\lambda_j \in \mathbb{Z}$ for all $j=1, ..., N$. Then $P$ defines an $n$-dimensional nonsingular toric variety by the following observation. For each $n$-dimensional integral Delzant polytope $P$, as in [@D1; @D2], we consider the set of pairs $\{p, v_{p, i} \}$, where $p$ is a vertex of $P$ and the neighboring faces are given by $l_{p, i} (x) = v_{p, i} \cdot x - \lambda_{p, i} >0$ for $i=1, ..., n$. For each $p$, we choose a coordinate chart $\mathbb{C}^n$ with $z=(z_1, ..., z_n)$. Then for any two vertices $p$ and $p'$, there exists $\sigma_{p, p'} \in GL(n, \mathbb{Z})$ such that $$\sigma_{p, p'} \cdot v_{p, i} = v_{p', i}.$$ Furthermore, we have $$\sigma_{p, p'} \cdot \sigma_{p', p''} \cdot \sigma_{p'', p} = 1.$$ Therefore $\sigma_{p, p'}$ serves as the transition function for two coordinate charts over $ (\mathbb{C}^*)^n$. More precisely, let $z=(z_1, ..., z_n)$ and $z'=(z_1', ..., z_n') \in \mathbb{C}^n$ be the coordinates for the chart associated to $p$ and $p'$ respectively. Suppose $\sigma_{p, p'}=(\alpha_{ij})$. Then $$z_i' = \prod_{j} z_j^{\alpha_{ij}}.$$ Each integral Delzant polytope uniquely determines a nonsingular toric variety $X_P$ by such a construction with the data $(p, \{ v_{p, i}\}_{i=1}^n)$. The constant $\lambda_{p, i}$ determines an ample line bundle $L$ over $X_P$, and moreover, $$H^0(X_P, L) =span\{ z^{\alpha}\}_{\alpha \in \mathbb{Z}^n \cap \overline{P}}.$$ Let $\varphi_P = \log (\sum_{\alpha \in \mathbb{Z}^n\cap \overline{P}} |z|^{2\alpha} )$. Then $\omega_P = {\sqrt{-1}\partial{\overline{\partial}}}\varphi_P$ is a smooth Kähler metric on $(\mathbb{C}^*)^n$ and it can be smoothly extended to a smooth global toric Kähler metric on $X_P$ in $c_1(L)$. Then the space of toric Kähler metrics in $c_1(L)$ is equivalent to the set of all smooth plurisubharmonic function $\varphi$ on $(\mathbb{C}^*)^n$ such that $ \varphi- \varphi_P$ is bounded and ${\sqrt{-1}\partial{\overline{\partial}}}\varphi$ extends to a smooth Kähler metric on $X_P$. If we consider the toric Kähler potential $\varphi$ which is invariant under the real torus action, we can view $\varphi$ as a function in $\mathbb{R}^n$ by $$\varphi = \varphi(\rho), ~~~\rho=(\rho_1, ..., \rho_n), ~\rho_i = \log |z_i|^2.$$ One can also define a symplectic potential $u$ on $P$ by $$ u(x) = \sum_{j=1}^N l_j (x) \log l_j(x) + f(x) $$ such that $f(x)\in C^\infty(\overline{P})$ and $u(x)$ is strictly convex in $P$. It is due to Guillemin [@Gu] that the toric Kähler potential and the symplectic potential are related by the Legendre transform $$\varphi (\rho) = \mathcal{L} u (\rho) = \sup_{x\in P} (x\cdot \rho - u(x)), ~ u (x) = \mathcal{L} \varphi(x) = \sup_{\rho\in \mathbb{R}^n } (x\cdot \rho - \varphi(\rho))$$ or equivalently $$\varphi(\rho)= x\cdot \rho - u(x), ~~~ u(x) = x\cdot \rho - \varphi(\rho), ~~~x=\nabla_\rho \varphi(\rho), ~\rho= \nabla_x u(x).$$ Now we would like to generalize the Guillemin condition to toric conical Kähler metrics on $X_P$. This can be considered as a generalization of orbifold Kähler metrics by replacing the finite subgroup by possibly infinite non-discrete subgroup of $(S^1)^n$. Suppose that the integral Delzant polytope is defined by $$P=\{ x\in \mathbb{R}^n ~|~ l_j(x)=v_j \cdot x + \lambda_j >0, ~j=1, ..., N \}$$ with $v_j \in \mathbb{Z}^n$ being a primitive lattice point and $\lambda_j \in \mathbb{Z}$. On each coordinate chart determined by the pair $(p, \{v_{p, i}\}_{i=1}^n)$ associated to a vertex of $P$, we let $z=(z_1, ..., z_n)$ be the coordinates on $\mathbb{Z}^n$. $\{ z_i =0\}$ extends to a smooth toric divisor of $X_P$. Let $D$ be a toric divisor of $X_P$ and suppose $D$ restricted to this coordinate chart is given by $$\sum_{i=1}^n a_i [z_i=0].$$ For any function $f(z)$ invariant under the $(S^1)^n$-action, we can lift it to a function $$\tilde f(w) = f(z)$$ by letting $$|w_i| = |z_i| ^{1/ \beta_i}, ~w=(w_1, ..., w_n) \in \mathbb{C}^n,$$ and clearly $\tilde f(w)$ is also $ (S^1)^n$-invariant. $w\in \mathbb{C}^n$ is then a $\beta$-covering of $z\in \mathbb{C}^n$. Then we can consider the $(S^1)^n$-invariant function space for $k\in \mathbb{Z}^{\geq 0}$ and $\alpha \in [0,1]$ $$C^{k, \alpha}_{\beta, p} = \{ f(z)=f(|z_1|, ..., |z_n|) ~|~ \tilde f(w) \in C^{k, \alpha} (\mathbb{C}^n) \} .$$ This in turn defines the weighted function space $$C^{k, \alpha}_{\beta}(X_P), \beta=(\beta_1, ..., \beta_N)\in (\mathbb{R}^+)^N$$ whose restriction on each chart belongs to $C^{k, \alpha}_\beta$ with respect to the weight $\beta$ and $\beta_j$ corresponding to the divisor induced by $l_j(x)=0$. Then we can define the space of weighted toric Kähler metrics on $X_P$ by considering a Kähler current $\omega$ whose restriction on each chart is given by $$\omega= {\sqrt{-1}\partial{\overline{\partial}}}\varphi_p$$ such that $\varphi_p \in C^{\infty}_{\beta, p}$. Such a weighted toric Kähler metric is naturally a smooth conical Kähler metric with cone angle $2\pi \beta_i$ along $[z_i =0]$ and is called a smooth $\beta$-weighted Kähler metric. The local lifting $\tilde \varphi (w)$ is a smooth plurisubharmonic function on the lifting space $w\in \mathbb{C}^n$. We can also define the space of weighted toric Kähler potential $\varphi$ on $(\mathbb{C}^*)^n$ such that $\varphi - \varphi_P$ is bounded and ${\sqrt{-1}\partial{\overline{\partial}}}\varphi$ extends to a smooth weighted Kähler metric on $X_P$. We now define a weighted $C^\infty_\beta$ [*symplectic potential*]{} $$u(x) = \sum_{j=1}^N\beta_j^{-1} l_j(x) \log l_j(x) + f(x)$$ for $f\in C^\infty(\overline P)$ for $j=1, ..., n$ such that $f\in C^\infty(\overline{P})$ and $u$ is strictly convex in $P$. Then the weighted Kähler potentials and the weighted symplectic potential determine each other uniquely. The following is a straightforward generalization of the Guillemin condition for conical toric Kähler metrics. The weighted $C^\infty_\beta$ toric potential $\varphi$ and the weighted $C^\infty_\beta$ symplectic potential are related by the Legendre transform $$\varphi (\rho) = \mathcal{L} u (\rho) = \sup_{x\in P} (x\cdot \rho - u(x)), ~ u (x) = \mathcal{L} \varphi(x) = \sup_{\rho\in \mathbb{R}^n } (x\cdot \rho - \varphi(\rho))$$ or equivalently $$\varphi(\rho)= x\cdot \rho - u(x) \text{ with } x=\nabla_\rho \varphi(\rho)\text{ and }u(x) = x\cdot \rho - \varphi(\rho)\text{ with } ~\rho= \nabla_x u(x).$$ In particular, if $u(x)=\beta_j^{-1} l_j(x) \log l_j(x) + f(x)$, then the cone angle of the corresponding conical toric Kähler metric is $2\pi \beta_j$ along the toric divisor determined by $l_j(x)=0$. Let $\omega = {\sqrt{-1}\partial{\overline{\partial}}}\phi$ be a smooth $\beta$-weighted Kähler metric and let $u=\mathcal{L} \phi$. Then $u(x)= \sum_j \beta_j^{-1} l_j(x) \log l_j(x) + f(x)$ for some $f\in C^\infty(\overline P)$. Let $P=[0, 1]$ then the associated toric manifold is $X=\mathbb{P}^1$ with the polarization $\mathcal{O}(1)$. We consider the symplectic potential by $$u= (\beta_1)^{-1} x \log x + \beta_2^{-1} (1-x) \log (1-x).$$ Then $$\rho= \log |z|^2 = u'(x) = (\beta_1^{-1}- \beta_2^{-1} ) + \log \frac{ x^{\beta_1^{-1}}}{(1-x)^{\beta_2^{-1}}}, ~\text{or } ~ |z|^2 = \frac{ x^{\beta_1^{-1}}}{(1-x)^{\beta_2^{-1}}} e^{\beta_1^{-1} - \beta_2^{-1}}$$ and so $$x \sim |z|^{2\beta_1} \text{ near } 0, ~~~ (1-x) \sim |z|^{-2\beta_2} \text{ near } \infty .$$ In particular, $x$ is a smooth function in $|z|^{2\beta_1}$ near $z=0$ and $(1-x)$ is a smooth function in $ |z|^{-2\beta_2}$ near $z=\infty$. The Kähler potential $\varphi$ is given by $$\varphi (\rho) = x (\beta_1^{-1}- \beta_2^{-1} ) - \beta_2^{-1} \log (1-x) .$$ Hence ${\sqrt{-1}\partial{\overline{\partial}}}\varphi$ extends to a conical metric with cone angle $2\pi \beta_1$ and $2\pi\beta_2$ at $[z=0]$ and $[z=\infty]$ respectively. When $\beta= \beta_1 = \beta_2$, $\varphi= \beta^{-1} \log (1+ |z|^{2\beta}) $ and the $\omega = 2 {\sqrt{-1}\partial{\overline{\partial}}}\varphi$ is a smooth conical Kähler-Einstein metric in $c_1(\mathbb{P}^1)$ satisfying $${\mathrm{Ric}}(\omega) = \beta \omega + (1-\beta) ([z=0] + [z=\infty]).$$ Let $g$ be a smooth $\beta$-weighted toric Kähler metric on a toric manifold $X_P$. Let $D$ be the toric divisor such that $g$ is a smooth toric Kähler metric on $X\setminus D$. Then for any $k\geq 0$, there exists $C_k>0$ such that for any point $p \in X\setminus D$, $$|\nabla_g^k Rm(g)|_{g} (p) \leq C_k . $$ The calculation of $|\nabla_g^k Rm(g)|_{g} (p)$ can be locally carried out on the $\beta$-covering space for each coordinate chart $(p, \{v_i\}_{i=1}^n)$ as in the orbifold case. All the quantities must be bounded because the $g$ is a smooth toric Kähler metric after being lifted on the covering space. We now can solve a Monge-Ampère equation with smooth $\beta$-weighted data. \[tormaso\] Let $\omega$ be a smooth $\beta$-weighted toric Kähler metric on a toric manifold $X_P$. Then for any smooth $\beta$-weighted function $f$ on $X_P$ with $\int_{X_P} e^{-f} \omega^n = \int_{X_P} \omega^n$, there exists a unique $\beta$-weighted smooth function $\varphi$ satisfying $$\label{toricma1} (\omega+{\sqrt{-1}\partial{\overline{\partial}}}\varphi)^n = e^f \omega^n, ~\sup_{X_P} \varphi =0. $$ We consider the following continuity method for $t\in [0, 1]$ $$\label{toriccon1} (\omega+ {\sqrt{-1}\partial{\overline{\partial}}}\varphi_t)^n = e^{t f + c_t} \omega^n, $$ where $c_t$ is determined by $\int_{X_P} e^{tf + c_t} \omega^n = \int_{X_P} \omega^n.$ Let $$S=\{ t \in [0, 1]~|~ \text{ (\ref{toriccon1}) is solvable for } t \text{ with } \varphi_t \in C^{\infty}_\beta (X_P) \}.$$ Obviously, $0\in S$. $S$ is open by applying the implicit function theorem for the linearized operator $$\Delta_{\beta, t} : C^{k+2, \alpha}_\beta(X_P) \rightarrow C^{k, \alpha}_\beta (X_P).$$ All the local calculation can be carried out in the $\beta$-covering space on each coordinate chart $(p, \{v_i\}_{i=1}^n)$ because all the data involved are invariant under $(S^1)^n$-action. It suffices to prove uniform a priori estimates for $\varphi_t$ in $C^{k}_\beta(X_P)$ for $t\in [0, 1]$. [*$C^0$-estimates.*]{} Let $\Omega$ be a smooth volume form on $X_P$. Then $\frac{e^{tf + c_t} \omega^n}{\Omega}$ lies in $L^{1+\epsilon}(X_P, \Omega)$ for some $\epsilon>0$. By Yau’s Moser iteration [@Y1] adapted to the conical case or by Kolodziej’s $L^\infty$-estimate [@Ko], there exists $C>0$ such that for all $t\in [0, 1]$, if $\varphi_t\in C^\infty_\beta(X)$ solves (\[toriccon1\]), $$||\varphi_t||_{L^\infty(X_P)} \leq C.$$ [*Second order estimates.*]{} We consider $$H_t = \log tr_{\omega} (\omega_t) - A\varphi_t.$$ Suppose at $t \in [0, 1]$, $\sup_{X_P} H_t = H_t (q)$. We lift all the calculation on the $(S^1)^n$-invariant $\beta$-covering space in a fixed local coordinate chart $w\in \mathbb{C}^n$. Standard calculations show that near $\tilde q$, there exists $C>0$ such that $$\begin{aligned} \tilde{\Delta}_{t, \beta} \tilde{H}_t &\geq& - Ctr_{\tilde \omega_t}(\tilde \omega) - An + A tr_{\tilde \omega_t}(\tilde \omega)\\ &\geq& \frac{A}{2} tr_{\tilde \omega_t} (\tilde \omega) - C,\\ $$ where $\tilde q$, $\tilde \omega$ and $\tilde \omega_t$ are the lifting of $q$, $\omega$ and $\omega_t$. By the maximum principle, at $\tilde q$, $tr_{\tilde \omega_t} (\tilde \omega) $ is bounded above by a constant independent of $t\in [0, 1]$. Combining the equation (\[toricma1\]), $tr_{\tilde \omega}(\tilde \omega_t)$ is also bounded above by a constant independent of $t$. Hence there exists $C>0$ such that for for all $t\in [0, 1]$, if $\varphi_t\in C^\infty_\beta(X)$ solves (\[toriccon1\]), $$C^{-1} \omega \leq \omega_t \leq C \omega.$$ [*Higher order estimates.* ]{} Calabi’s third order estimates can be applied the way as in [@Y1; @PSS] by the maximum principle after lifting all the local calculations on the $(S^1)^n$-invariant covering space. The Schauder estimates can also be applied by the bootstrap argument. Eventually, for any $k>0$, there exists $C_k$ such that for all $t\in [0, 1]$, if $\varphi_t\in C^\infty_\beta(X)$ solves (\[toriccon1\]), $$||\varphi_t||_{C^k_\beta(X_P)} \leq C_k.$$ Proof of Theorem \[main3\] -------------------------- An $n$-dimensional integral polytope $P$ is called a Fano polytope if it is a Delzant polytope and $\lambda_i=1$ for each defining function $l_i (x) = v_i \cdot x + \lambda_i$. Immediately, we have $0 \in P$. The toric manifold $X_P$ associated to $P$ is a Fano manifold. Each $(n-1)$-face of $P$ corresponds to a toric divisor of $P$. Then the union $D_P= \sum_j D_j$ for all the boundary divisors lies in $c_1(X)=c_1(-K_X)$, where $D_j$ is the toric divisor induced by the face $\{ l_j(x)=0\}$. In particular, $[D]$ is very ample. The Futaki invariant of $X_P$ with respect to $(S^1)^n$-action is shown in [@M] exactly the Barycenter of $P$ defined by $$P_c = \frac{ \int_P x dV}{\int_P dV},$$ where $dV=dx_1 dx_2 ... d x_n$ is the standard Euclidean volume form. The following theorem is due to Wang and Zhu [@WZ] for the existence of Kähler-Einstein metrics on toric Fano manifolds. There exists a smooth toric Kähler-Einstein metric on a toric Fano manifold $X_P$ if and only if the Barycenter of $P$ coincides with $0$. If the Barycenter is not at the origin, it is also proved in [@WZ] that there exists a toric Kähler-Ricci soliton on $X_P$. The following theorem is proved by Li [@LiC] to calculate the the greatest Ricci lower bound $R(X)$. Let $X_P$ be a toric Fano manifold associated to a Fano Delzant polytop $P$. Let $P_c$ be the Barycenter of $P$ and $Q\in \partial P$ such that the origin $O \in \overline{P_cQ}$. Then the greatest Ricci lower bound of $X_P$ is given by $$R(X_P) = \frac{ |\overline{OQ}| }{ |\overline{P_c Q}| }.$$ For any $\tau \in \mathbb{R}^n $, we define the divisor $D(\tau)$ by $$D(\tau) = \sum_{j=1}^N l_j(\tau) D_j. $$ $D(\tau)$ is a Cartier $\mathbb{R}$-divisor in $c_1(X)$ and it is effective if and only if $\tau\in \overline{P}$. The defining section $s_\tau$ of $D(\tau)$ is given by the monomial $$s_\tau = z^\tau.$$ Although $s_\tau$ is only locally defined, $$|s_\tau|^2=|z|^{2\tau}= e^{\tau\cdot \rho}$$ is globally defined and $|s_\tau|^{-2}$ induces a singular hermitian metric on $-K_X$ and can be viewed as a singular volume form with poles along $D_j$ of order $l_j(\tau)$. We immediately have the following lemma. \[divisord\] If $R(X) <1$, then the $\mathbb{R}$-divisor $D(\tau)$ with $\tau = -\frac{\alpha}{1-\alpha} P_c$ is effective if and only if $\alpha \in [0, R(X)]$. We consider the following real Monge-Ampère equation on $\mathbb{R}^n$ for a convex function $\phi$ $$\label{make} \det (\nabla^2 \phi) = e^{ - \alpha \phi - (1-\alpha) \tau \cdot \rho}. $$ Let $u = \mathcal{L} \phi$ be the symplectic potential. Then $$\det(\nabla^2 u)= \left( \det (\nabla^2 \phi)^n\right)^{-1}$$ and dual Monge-Ampère equation for $u$ is given by $$\det(\nabla^2 u) = e^{-\alpha u + ( \alpha x + (1-\alpha) \tau ) \cdot \nabla u}. $$ If we let $\omega= {\sqrt{-1}\partial{\overline{\partial}}}\phi$, the corresponding curvature equation is given by $${\mathrm{Ric}}(\omega) = \alpha \omega + (1-\alpha) [D(\tau)].$$ \[torickefr\]Suppose there exists a smooth conical Kähler-Einstein metric $\omega= {\sqrt{-1}\partial{\overline{\partial}}}\phi$ satisfying $${\mathrm{Ric}}(\omega) = \alpha \omega + (1-\alpha) [D(\tau)]$$ for some $\alpha \in (0, R(X)]$ and $\tau \in \overline{P}$. Then $$\tau =- \frac{ \alpha}{1-\alpha} P_c, ~for~\alpha\neq 1,~\text{ }~ \tau =0, ~for ~\alpha=1. $$ Furthermore, there exists $f\in C^\infty(\overline{P})$ such that $$\mathcal{L}\phi(x)= \sum_{j=1}^N \beta_j^{-1} l_j (x) \log l_j(x) + f(x), ~~ \beta_j = \frac{ l_j (P_c)}{ l_j(0)} \alpha. $$ Consider the corresponding Monge-Ampère equation $(\det \nabla^2 \phi)^n = e^{-\alpha \phi - (1-\alpha) \tau \cdot \rho}$. The right hand side $e^{-\alpha \phi - (1-\alpha) \tau \cdot \rho}$ is integrable on $\mathbb{R}^n$ and in fact $$\int_{\mathbb{R}^n} e^{-\alpha \phi - (1-\alpha) \tau\cdot \rho} d \rho = \int_{\mathbb{R}^n} \det (\nabla^2 \phi) d \rho = \int_X \omega^n = c_1(X)^n.$$ Then the Monge-Ampère mass $\det(\nabla^2\phi) d\rho$ becomes $dx$ by the moment map and $$\begin{aligned} 0&=& \int_{\mathbb{R}^n} \nabla \left( e^{-\alpha \phi - (1-\alpha) \tau \cdot \rho} \right) d \rho\\ &=& - \int_{\mathbb{R}^n} (\alpha \nabla \phi + (1-\alpha) \tau) e^{-\alpha \phi - (1-\alpha) \tau \cdot \rho} d \rho\\ &=& - \int_P (\alpha x + (1-\alpha) \tau) d x\\ &=& - (\alpha P_c + (1-\alpha) \tau) \int_P dx. $$ Therefore $\tau = \frac{\alpha}{1-\alpha} P_c$ for $\alpha\neq 1$. Suppose $u (x) = \mathcal{L}\phi(x) = \sum_{j=1}^N \beta_j^{-1} l_j(x) \log l_j(x) + f(x).$ The Monge-Ampère equations for $\phi$ and $u$ are given by $$\det(\nabla^2 \phi) = e^{ -\alpha ( \phi + P_c \cdot \rho)}, ~~ \det(\nabla^2 u ) = e^{-\alpha ( u - (x-P_c) \cdot \nabla u) }.$$ Without loss of generality, we assume that $l_1(x)=l_2(x)= ...= l_n(x)=0$ with $l_i(x)=v_i\cdot x+1,\ 1\leq i\leq n$ defines a vertex $p$ of $P$. Then there exists a smooth positive function $F(x)$ on any compact subset $U$ of $\overline{P}$ with $U\cap \{ l_j(x)=0 \} =\phi$ for all $j>n$, such that $$\det ( \nabla^2 u) = \frac{F(x)}{l_1(x) l_2(x)... l_n(x)}.$$ On the other hand, $$u(x) - (x -P_c) \nabla u(x) =\sum_{j=1}^N \beta_j^{-1} l_j( P_c) \log l_j(x) - \sum_{j=1}^N \beta_j^{-1} (x- P_c) \cdot v_j - (x-P_c)\cdot\nabla f(x) $$ and so $$e^{-\alpha( u - (x-P_c)\cdot \nabla u)} = \frac{ e^{ - \alpha (x-P_c) \cdot ( \nabla f(x) + \sum_j \beta_j^{-1} v_j)} }{ \prod_{j=1}^N \left( l_j(x)^{\alpha \beta_j^{-1} l_j(P_c)} \right)}.$$ By comparing the power of $l_j(x)$, we have $$\beta_j = l_j (P_c) \alpha = \frac{l_j(P_c)}{l_j(0)} \alpha$$ since $l_j(0)=1$. This completes the proof of the lemma. Lemma \[torickefr\] tells us that we should consider the following Monge-Ampère equation $$\label{toricidma} \det (\nabla^2 \phi)^n = e^{-\alpha (\phi - P_c \cdot \rho)}. $$ The right hand side of equation (\[toricidma\]) is always integrable because $P_c$ lies in $P$ and $\phi - \log (\sum_k e^{p_k \cdot \rho })$ is bounded on $\mathbb{R}^n$ where $p_k$ runs over all vertices of $P$. For each $\alpha\in (0, 1]$, we define $\beta=\beta(\alpha)= (\beta_1, \beta_2, ..., \beta_N)$ by $\beta_j= \frac{l_j(P_c)}{l_j(0)}\alpha$. \[inikm\]For any $\alpha\in (0, 1]$, there exists a $C^\infty_{\beta(\alpha)}$ conical toric Kähler metric $\omega$ such that $${\mathrm{Ric}}(\omega) = \alpha \theta + (1-\alpha) [ D(\tau)],$$ where $\theta \in c_1(X)$ is a fixed $C^\infty_{\beta(\alpha)}$ toric Kähler metric and $\tau= \frac{\alpha}{1-\alpha} P_c$. In particular, ${\mathrm{Ric}}(\omega) >0$, if $\alpha \in (0, R(X))$. It suffices to prove for $\alpha\in (0,1)$. Let $\hat u (x) = \sum_j \beta_j^{-1} l_j(x) \log l_j(x)$ for $\beta_j = \frac{l_j(P_c)}{l_j(0)} \alpha$. Let $\hat\phi = \mathcal{L}\hat u$ and $\hat\omega = {\sqrt{-1}\partial{\overline{\partial}}}\hat \phi$. Then there exists a $C^\infty_{\beta(\alpha)}$ real valued $(1,1)$-form $\eta \in c_1(X)$ and a divisor $D(\tau)$ with $\tau = -\frac{\alpha}{1-\alpha} P_c$ such that $${\sqrt{-1}\partial{\overline{\partial}}}\log \hat\omega^n = \alpha \eta + (1-\alpha) [D(\tau)].$$ This is because along each $D_j $ defined by $l_j(x)=0$, $\hat\omega^n$ has a pole of order $$1-\beta_j=1-l_j(P_c)\alpha$$ and $$(1-\alpha)D(\tau)=(1-\alpha) \left(\sum_j l_j \left(\frac{\alpha}{1-\alpha} P_c\right) D_j \right)= \sum_j (1- l_j(P_c) \alpha) D_j.$$ Since $\eta\in c_1(X)$ is $C^\infty_{\beta(\alpha)}$, there exists $\psi\in C^\infty_{\beta(\alpha)}$ such that $\eta+ {\sqrt{-1}\partial{\overline{\partial}}}\psi$ is a $C^\infty_{\beta(\alpha)}$ toric Kähler metric. We can assume that $\int_{X_P} e^\psi \hat\omega^n = \int_{X_P} \hat\omega^n$ after a constant translation. Then by Proposition \[tormaso\], the following equation $$(\hat\omega + {\sqrt{-1}\partial{\overline{\partial}}}\varphi)^n = e^{\psi} \hat\omega^n, ~~\sup_{X_P} \varphi =0$$ admits a unique $C^\infty_{\beta(\alpha)}$ solution $\varphi$. By letting $\omega = \hat\omega + {\sqrt{-1}\partial{\overline{\partial}}}\varphi$ and $\theta = \eta+ {\sqrt{-1}\partial{\overline{\partial}}}\psi$, we have $${\mathrm{Ric}}(\omega) = \alpha\theta + (1-\alpha) [D_(\tau)].$$ By Lemma \[inikm\], we can choose a $C^\infty_{\beta(\alpha)}$ Kähler potential $\phi_0$ with $\omega_0 = {\sqrt{-1}\partial{\overline{\partial}}}\phi_0$ such that $${\mathrm{Ric}}(\omega_0) = \alpha \theta + (1-\alpha)D(\tau)$$ for a $C^\infty_{\beta(\alpha)}(X_P)$ Kähler metric $\theta \in c_1(X_P)$ and $\tau= \frac{\alpha}{1-\alpha} P_c$ if $\alpha\neq 0$. Let $$w=\frac{1}{\alpha} \left( - \alpha P_c \cdot \rho - \log \det(\nabla^2 \phi_0) \right).$$ Then $${\sqrt{-1}\partial{\overline{\partial}}}w =\theta$$ and $ |w - \hat\phi|$ is uniformly bounded by the argument in Lemma \[inikm\]. This implies that $w$ is a $C^\infty_{\beta(\alpha)}$ Kähler potential and we have $$\det(\nabla^2 \phi_0 ) = e^{-\alpha (w - P_c \cdot \rho) }.$$ We will then define the following continuity method for a family of Monge-Ampère equations for $t\in [0, \alpha]$ $$\label{toriccon2} \det(\nabla^2 \phi_t) = e^{ - t (\phi_t - P_c \cdot \rho) - (\alpha -t) w }. $$ Let $\varphi_t = \phi_t - \phi_0$ and define $C^\infty_{\beta(\alpha)}$ functions $h_{\omega_0}$ and $h_{\theta}$ on $X_P$ by $$- {\sqrt{-1}\partial{\overline{\partial}}}\log \omega_0^n - {\sqrt{-1}\partial{\overline{\partial}}}h_{\omega_0} = \alpha \omega_0 + (1-\alpha)[D(\tau)]$$ and $$- {\sqrt{-1}\partial{\overline{\partial}}}\log \theta^n - {\sqrt{-1}\partial{\overline{\partial}}}h_{\theta} = \alpha \theta + (1-\alpha)[D(\tau)], ~ \int_{X_P} e^{h_\theta}\theta^n = \int_{X_P} \theta^n.$$ Then equation (\[toriccon2\]) is equivalent to $$({\sqrt{-1}\partial{\overline{\partial}}}\omega_0 + {\sqrt{-1}\partial{\overline{\partial}}}\varphi_t)^n = e^{- t\varphi_t } \left( e^{h_{\omega_0}} \omega_0 \right)^{\frac{t}{\alpha}} \left(e^{h_\theta} \theta \right)^{\frac{\alpha-t}{\alpha}}, $$ where $h_{\omega_0}$ and $h_{\theta}$ are $C^\infty_{\beta(\alpha)}$ on $X_P$ with $$- {\sqrt{-1}\partial{\overline{\partial}}}\log \omega_0^n - {\sqrt{-1}\partial{\overline{\partial}}}h_{\omega_0} = \alpha \omega_0 + (1-\alpha)[D(\tau)], ~~ - {\sqrt{-1}\partial{\overline{\partial}}}\log \theta^n - {\sqrt{-1}\partial{\overline{\partial}}}h_{\theta_0} = \alpha \theta + (1-\alpha)[D(\tau)]$$ Let $$S=\{ t \in [0, \alpha]~|~ \text{ (\ref{toriccon2}) is solvable for } t \text{ with } {\sqrt{-1}\partial{\overline{\partial}}}\phi_t \in C^{\infty}_\beta (X_P) \}.$$ Obviously, $\phi_0$ solves (\[toriccon2\]) for $t=0$ and so $S\neq \phi$. Notice that $${\mathrm{Ric}}(\omega_t) = t\omega_t + (\alpha -t) \theta + (1-\alpha) [D(\tau)] \geq t \omega_t$$ for $t\in [0, \alpha]$ and $\tau = \frac{\alpha}{1-\alpha} P_c$ if $\alpha\neq 1$. It implies that the first eigenvalue of the Laplace operator $\Delta_t = {\textnormal{tr}}_{\omega_t} ({\sqrt{-1}\partial{\overline{\partial}}})$ is strictly greater than $t$. By the argument in Proposition \[tormaso\], $S$ is open and it suffices to show that $S$ is closed by proving uniform a priori estimates for $\phi_t - \phi_0$. There exists $C>0$ such that for all $t\in [0, \alpha]$, $$||\phi_t - \phi_0||_{L^\infty(\mathbb{R}^n)} \leq C. $$ We fix some positive $\epsilon_0 \in S$. We let $$\Phi_t =\alpha^{-1} \left( \phi_t - P_c\cdot \rho \right) \text{ and } W= \alpha^{-1} ( w - P_c\cdot \rho).$$ Then the equation (\[toriccon2\]) becomes $$\det \nabla^2 \Phi_t = e^{- (\Phi_t + W)}.$$ Let $W_t= \Phi_t + W$. Immediately, we can see that the moment map with respect to $\Phi_t$ is given by $$F_t: \nabla \Phi_t \rightarrow P-P_c$$ whose image is the translation of $P$ by $-P_c$. In particular, the Barycenter of the new polytope $P- P_c$ coincides with the origin. Suppose $$m_t = W_t(\rho_t) = \inf_{\mathbb{R}^n} W_t (\rho)$$ for a unique $\rho_t\in \mathbb{R}^n$ since $W_t$ is asymopotically equivalent to $ \log (\sum e^{( p_k - P_c) \cdot \rho} )$ where $p_k$ are the vertices of $P$. We can apply the same argument by Wang-Zhu in [@WZ]. First one can show by John’s lemma and the maximum principle (see Lemma 3.1, 3.2 in [@WZ]), that there exists $C>0$ such that for all $t\in [\epsilon_0, \alpha]$, $$m_t = W_t(\rho_t) = \inf_{\mathbb{R}^n} W_t (\rho) \leq C.$$ Then by making use of the fact that the Barycenter of $P-P_c$ lies at the origin $O$, the same argument in [@WZ] (Lemma 3.3) shows that there exists $C>0$ such that for all $t\in [\epsilon_0, \alpha]$, $$|\rho_t| \leq C.$$ This then implies that $$\varphi_t = \alpha^{-1}(\Phi_t - W)$$ is uniformly bounded above for $t\in [\epsilon_0, \alpha]$ by the same argument in Lemma 3.4 in [@WZ]. The uniform lower bound of $\varphi_t$ can be obtained either by the Harnack inequality $$-\inf_{X_P} \varphi_t \leq C(1+\sup_{X_P} \varphi_t)$$ adapted from the smooth case or directly by the argument in Lemma 3.5 in [@WZ]. \[torest2\] For any $k\geq 0$, there exists $C_k>0$ such that for all $t\in [0, \alpha]$, $$||\varphi_t||_{C^k(X_P)} \leq C_k. $$ The Laplace $\Delta_{\beta(\alpha), t} \varphi_t$ is uniformly bounded by Yau’s estimates after lifting the calculations to the $\beta(\alpha)$-covering space as in the proof of Proposition \[tormaso\]. The $C^3$-estimates and the Schauder estimates can be applied in the same way. \[toke\] Let $X_P$ be a toric Fano manifold. 1. For any $\beta \in (0, R(X_P))$, there exist a unique smooth toric conical Kähler-Einstein metric $\omega$ and a unique effective toric $\mathbb{R}$-divisor $D_\beta \in |-K_X|$ satisfying $${\mathrm{Ric}}(\omega) = \beta \omega + (1-\beta)[D_\beta].$$ 2. For $\beta = R(X_P)$, there exists a unique smooth toric conical Kähler-Einstein metric $\omega$ satisfying $$\label{torkeeqn} {\mathrm{Ric}}(\omega ) = R(X_P) \omega + (1-R(X_P)) [D_P]$$ for an effective $\mathbb{Q}$-divisor $D_P$ in $c_1(X)$. In particular, if $D_j$ is the toric divisor associated to the face defined by $l_j(x) =v_j \cdot x + \lambda_j=0$, then $$D_P = \sum_j \frac{1-\beta_j}{1-R(X)} D_j, ~~\beta_j = \frac{ l_j (P_c)} { l_j(0) } R(X_P)$$ and the cone angle of $\omega$ along $D_j$ is $2\pi \beta_i$, if $R(X)<1$. 3. For $\beta \in (R(X), 1]$, there does not exist a smooth toric conical Kähler-Einstein metric $\omega$ satisfying $${\mathrm{Ric}}(\omega) = \beta \omega + (1-\beta) [D],$$ with an effective $\mathbb{R}$-divisor $D_\beta$ in $[-K_X]$. \(1) and (2) are proved by the uniform estimates from Lemma \[torest2\]. If $\beta>R(X_P)$, there still exists a smooth conical Kähler-Einstein metric satisfying $Ric(\omega) = \beta \omega+ (1-\beta)[D_\beta]$ for some toric divisor $D$, however, by Lemma \[divisord\], $D$ is not effective and so (3) is proved. \[torlowbd\] Let $X$ be a Fano toric manifold and $\omega\in c_1(X)$ be a smooth Kähler metric. We define $$F_{\omega, \alpha}(\varphi) = J_\omega(\varphi) -\frac{1}{V}\int_X\varphi \omega^n - \frac{1}{\alpha} \log \frac{1}{V} \int_X e^{-\alpha \varphi} \omega^n$$ for all $\varphi\in C^\infty(X)\cap PSH(X, \omega)$. Suppose $R(X)<1$. Then 1. for $\alpha\in (0,R(X))$, $F_{\omega, \alpha}$ is $J$-proper. 2. for $\alpha = R(X)$, $F_{\omega, \alpha}$ is bounded below. 3. for $\alpha\in (R(X), 1]$, $\inf_{\varphi \in PSH(X, \omega)\cap C^\infty(X)} F_{\omega, \alpha}(\varphi) = - \infty.$ It suffices to prove $(2)$ by Corollary \[flowbd\] . This can be proved by modifying the argument in [@BBGZ; @Be]. By Theorem \[toke\], there exists a unique $(S^1)^n$-invariant $\psi \in L^\infty(X_P) \cap PSH(X_P, \omega)$ satisfying $$(\omega+ {\sqrt{-1}\partial{\overline{\partial}}}\psi) = e^{-\alpha \psi}\mu, ~\text{ }~\mu = \left(\Omega_\omega\right)^\alpha (\Omega_D)^{1-\alpha},$$ where $\Omega_\omega$ is a smooth volume form with ${\sqrt{-1}\partial{\overline{\partial}}}\log\Omega_\omega= -\omega$ and $\Omega_D$ is a positive $(n,n)$-current with $-{\sqrt{-1}\partial{\overline{\partial}}}\log \Omega_D=[D]$ and $D=D\left(\frac{\alpha}{1-\alpha} P_c \right)$. For any $\varphi\in PSH(X, \omega)\cap L^\infty(X)$, let $\varphi_t$ be the weak geodesic $\varphi_t$ joining $\psi$ and $\varphi$ with $\varphi_0=\psi$ and $\varphi_1=\varphi$. Then the modified functional $$f(t)={\mathcal{F}}_{\omega, \alpha}(\varphi_t) = J_\omega(\varphi_t) - \frac{1}{V}\int_{X_P} \varphi_t \omega^n - \frac{1}{\alpha} \log \frac{1}{V}\int_{X_P} e^{-\alpha \varphi}\mu$$ is convex on $[0, 1]$ and $f'(0)\geq 0$ by applying the same argument in Theorem 6.2 in [@BBGZ]. This shows that ${\mathcal{F}}_{\omega, \alpha}$ is bounded below and since $\Omega_D$ is bounded below away from $0$, and therefore $F_{\omega, \alpha}$ is bounded from below as well. Let $X$ be $\mathbb{P}^2$ blown-up at one point. Then $R(X) = 6/7$ shown in [@Sze] and $X$ admits a holomorphic ${\mathbb{P}}^1$ fibre bundle $\pi: X \rightarrow {\mathbb{P}}^1$. Let $D_\infty$ be the infinity section of $\pi$ and $H_1$ and $H_2$ be the two toric $P^1$ fibre of $\pi $. Then the divisor $D_P$ in the equation (\[torkeeqn\]) is given by $$D_P= 2D_\infty + (H_1+H_2)/2.$$ We remark that for $\beta\in (R(X), 1]$, there exists a smooth toric conical Kähler-Einstein metric in $c_1(X_P)$ satisfying ${\mathrm{Ric}}(g) = \beta g + (1-\beta) [D_\beta]$ for some toric divisor $D_\beta$ in $c_1(X_P)$ which can not be effective. However, it is not a conical Kähler metric we are interested in, because the Ricci current of such Kähler-Einstein metric is not bounded below. We also notice that in the case of $\beta = R(X)<1$, the conical Kähler-Einstein metric is not necessarily invariant under the maximal compact subgroup of the automorphism group of $X$ by the following example. Let $X$ be $\mathbb{P}^2$ blown-up at one point. It is a compactification of $\mathbb{C}^2\setminus \{0\}$ by adding one divisor $D_0 =\mathbb{P}^{1}$ at $0$ and another divisor $D_\infty= \mathbb{P}^1$ at $\infty$. The maximum compact group of the automorphism group of $X$ is $U(2)$. The only divisor in $c_1(X)$ fixed by the $U(2)$-action is given by $$D= 3[D_\infty]- [D_0].$$ Therefore, there does not exists a smooth conical Kähler-Einstein metric on $X$ with positive Ricci current. In fact, there does not exist any conical Kähler-Einstein metric for such $D$ by the following observation, although one can look for complete conical Kähler-Einstein metric in this case. Let $z=(z_1, z_2) \in \mathbb{C}^2$ and $\rho= \log |z|^2=\log |z_1|^2 + |z_2|^2 \in \mathbb{R}$. Then a convex function $u=u(\rho)$ satisfies the Calabi symmetry if 1. $u''>0$, 2. there exists $0<a<b$ and $\beta_1, \beta_2>0$ such that $U_0'(0)>0$, $U_\infty'(0)>0$, where $U_0(e^{\beta_1 \rho} )=u(\rho) - a\rho $ and $U_\infty(e^{-\beta_2 \rho}) = u(\rho) - b \rho$. Then ${\sqrt{-1}\partial{\overline{\partial}}}u$ extends to a smooth conical Kähler metric $\omega_u$ on $X$. In particular, if we choose $b=3$ and $a=1$, $\omega_u \in c_1(X)$. The conical Kähler-Einstein equation for $\omega_u$ is given by the following ODE $$u' u'' = e^{2\rho -\alpha u}.$$ Differentiating both sides after taking the log, we have $$\frac{u'''}{u''} + \frac{u''}{u'} = 2 - \alpha u'.$$ Let $x= u'$ and $Y(x) = u''(\rho)$. Then we have $$Y'(x)+ (x^{-1} +\alpha) Y(x) =2, ~Y(1)=Y(3)=0$$ and so $$Y(x) = \frac{2}{\alpha} - \frac{2}{\alpha^2 x} + \frac{c}{x e^{\alpha x}}, ~(3\alpha -1)e^{2\alpha} - \alpha +1=0.$$ Hence by the monotonicity of $(3\alpha -1)e^{2\alpha} - \alpha +1$, $\alpha =0$ but it contradicts the assumption $u'u''(\infty) = 0$. On the other hand, there exist smooth conical Kähler-Einstein metrics in $b[D_\infty]- a[D_0]$ for some $0<a<b$ by solving the above ODE in the same way. The Chern number inequality =========================== Curvature estimates {#curv-est} ------------------- In this section, we will derive some curvature estimates for a smooth conical Kähler metric whose Ricci curvature is bounded. Let $D$ be the smooth simple divisor of $X$. At each point $p$ on $D$, we can use the following holomorphic local coordinates $$z=(y, \xi) = (z_1, ..., z_{n-1}, z_n), ~y=(z_1, ..., z_{n-1}), ~\xi = z_n$$ and $D$ is locally defined by $\xi=0$. We write $\xi = r^{1/\beta} e^{i\theta}$ for $\theta\in [0, 2\pi)$. We use Greek letters $\alpha, \beta, ... $ as indices for $1, ..., n$ and letters $i, j, ...$ for $1, ..., n-1$. Let us recall the following result by Jeffres, Mazzeo and Rubinstein in [@JMR] on a complete asymptotic expansion of smooth conical Kähler metrics. . \[expansion\] ([@JMR Proposition 4.3]) Let $\omega$ be a smooth conical Kähler metric with conical singularity along a smooth simple divisor $D$ of angle $2\pi \beta$. Suppose $\varphi$ is a local potential of $\omega$, i.e., $\omega = {\sqrt{-1}\partial{\overline{\partial}}}\varphi$ in a neighborhood of a conical point $(y, \xi)$, then the asymptotic exapansion of $\varphi$ takes the following form $$\varphi(r, \theta, y)\sim \sum_{j, k, l\geq 0} a_{jkl}(\theta, z) r^{j+\frac{k}{\beta}} (\log r )^l.$$ In particular, if the Ricci curvature of $\omega$ is bounded and $\beta \in (1/2, 1)$, $\varphi$ has the following expansion $$\varphi(r, \theta, y) = a_{00}(y) + (a_{01}(y) \sin \theta + b_{01}(y) \cos\theta) r^{1/\beta} + a_{20} (y) r^2 + O(r^{2+\epsilon}) $$ for some $\epsilon(\beta)>0$. When the Ricci curvature is bounded and $\beta\in (1/2, 1)$, $$\varphi = a(y) + b(y) (\xi + \bar \xi) + \sqrt{-1} c(y) (\xi - \bar \xi) + d(y) |\xi|^{2\beta} + o (|\xi|^{2\beta + \epsilon} )$$ for some $\epsilon>0$. From now on in this section, we will always assume that $g$ is a conical Kähler metric on $X$ with cone angle $2\pi \beta$ for $\beta \in (1/2, 1)$ along the simple smooth divisor $D$, in addition, the Ricci curvature of $g$ is bounded. The following lemma can be obtained by straightforward calculations. \[metor\] Let $g$ be a smooth conical Kähler metric with cone angle $2\pi \beta $ along a smooth simple divisor $D$ for $\beta\in (0, 1)$. Let $o(1)$ be the quantity satisfying $\lim_{|\xi|\rightarrow 0} o(1) =0$. Then $$g_{z_i \bar z_j} \sim \delta_{i j} + o(1)$$ $$g_{\xi \bar \xi} \sim |\xi|^{-2(1-\beta)} + o(|\xi|^{-2(1-\beta)}).$$ $$g_{z_i \bar \xi} \sim O(1).$$ By taking the inverse, we have the following corollary from Lemma \[metor\]. \[asymp-g\] Let $g$ be a smooth conical Kähler metric with cone angle $2\pi \beta $ along a smooth simple divisor $D$ for $\beta\in (0, 1)$. $$g^{z_i \bar z_j} \sim \delta_{ij} + o(1)$$ $$g^{\xi \bar \xi} \sim |\xi|^{2(1-\beta)} + o(|\xi|^{2(1-\beta)}).$$ $$g^{ z_i \bar \xi} \sim |\xi|^{2(1-\beta)}.$$ The following lemmas give the estimates for the curvature tensor of $g$. \[metor2\] Let $g$ be a smooth conical Kähler metric with cone angle $2\pi \beta $ along a smooth simple divisor $D$ for $\beta\in (0, 1)$. If the Ricci curvature of $g$ is bounded, then $$\begin{aligned} \label{cur1} R_{z_i \bar z_j z_k \bar z_l} &\sim& R_{z_i \bar z_j z_k \bar \xi} = O(1)\\ \label{cur2} R_{z_i \bar z_j \xi \bar \xi } &=& O (|\xi|^{-2(1-\beta)}). \\ \label{cur3} R_{\xi \bar z_j \xi \bar z_l } &=& O(|\xi|^{-1}) . \\ \label{cur4} R_{\xi \bar \xi \xi \bar z_l} &=& O(|\xi|^{-1}) .\\ \label{cur5} R_{\xi \bar \xi \xi \bar \xi } &=& O(|\xi|^{-\max (1, 4(1-\beta)) }) . \end{aligned}$$ The estimates (\[cur1\]), (\[cur2\]) and (\[cur3\]) can be shown by straightforward calculation using the curvature formula $$R_{\alpha \bar \beta \gamma \bar \zeta} = - g_{\alpha\bar \beta, \gamma \bar \zeta} + g^{\mu \bar \nu} g_{\alpha \bar \nu, \gamma} g_{\mu \bar \beta, \bar \zeta}.$$ The estimates (\[cur4\]) and (\[cur5\]) follow by combining the boundness of the Ricci curvature and the estimates (\[cur2\]) and (\[cur3\]). Let $g$ be a smooth conical Kähler metric with cone angle $2\pi \beta $ along a smooth simple divisor $D$ for $\beta\in [1/2, 1)$. If the Ricci curvature of $g$ is bounded, then $$R^{z_i}_{\xi, z_k \bar z_l} \sim R^\xi_{\xi, z_k \bar z_l}= O(1).$$ $${R^{z_i}_{z_p, z_k \bar z_l} }\sim R^{\xi}_{z_p, z_k \bar z_l} = O(|\xi|^{2(1-\beta)}).$$ From the curvature estimates, we immediately have the following proposition. Let $g$ be a smooth conical Kähler metric with cone angle $2\pi \beta $ along a smooth simple divisor $D$ for $\beta\in (0, 1)$. If the Ricci curvature of $g$ is bounded, then We have the following pointwise estimates for $|R_m|^2$ $$| Rm(g) |^2_g = O( |\xi|^{-2 + 4(1-\beta)})$$ Consequently, the $L^2$- norm of $Rm(g)$ is bounded, i.e., there exists $C>0$ such that $$\int_X | Rm(g) |^2_ g ~dVol(g) \leq C. $$ Chern forms for conical Kähler metrics -------------------------------------- Let $g$ be a smooth conical Kähler metric with cone angle $2\pi \beta$ along a smooth simple divisor $D$. We let $\theta$ be the connection form on the tangent bundle $TX$ induced by $g$, so locally we may write $$\theta^\alpha_\gamma = g^{\alpha\bar \beta} g_{\gamma \bar \beta, \eta} d z_\eta$$ and $\Omega^{\alpha}_\gamma = {\overline{\partial}}\theta^\alpha_\gamma$ be the curvature form of $\theta$. Then the total Chern class is defined by $$\det (tI + \Omega) = \sum_{i=0}^n t^{2(n-i)} c_i(\Omega)$$ and let $P_i(\Omega_1, ..., \Omega_n)$ being the polarization of $c_i(\Omega)=c_i(X, g)$ for the conical metric $g$. Let $g_0$ be a smooth Kähler metric and $\theta_0$ be the connection induced by $g_0$ as $$(\theta_0)^\alpha_\gamma = (g_0)^{\alpha\bar \beta} (g_0)_{\gamma \bar \beta, \eta} d z_\eta.$$ Then $\Omega_0= \sqrt{-1} {\overline{\partial}}\theta_0$ is the curvature form of $\theta_0$. Let $\theta_t = t\theta + (1-t) \theta_0$ with curvature $$\Omega_t = \bar \partial \theta_t = t \Omega + (1- t) \Omega_0.$$ Then we have $$c_2(X, \omega) - c_2(X, \omega_0) = 2 \sqrt{-1} \int_0^1 {\overline{\partial}}P_2(\theta- \theta_0, \Omega_t) dt.$$ We will construct connections on the divisor $D$ from $\theta$ and $\theta_0$. Let $p$ be a point in the divisor $D$. We can choose holomorphic local coordinates $$z=(z_1, ..., z_n), ~ \xi = z_n$$ such that $D$ is locally defined by $\xi =0$ as in section \[curv-est\]. We define $H$ and $H_0$ locally by $$H_{i\bar j} = g_{i\bar j}, ~ H^{i\bar j} = (H^{-1})_ {i\bar j}, ~~H_{n\bar j} = g_{n \bar j},$$ and $$ (H_0)_{i\bar j} = (g_0)_{i\bar j}, ~ (H_0) ^{i\bar j} = (H_0^{-1})_{i\bar j}, ~~(H_0) _{n\bar j} = (g_0) _{n \bar j}. $$ For each coordinate system $(z,\xi)$ chosen as above, we define $(1, 0)$- forms $\theta_D$ and $\theta_{0, D}$ locally by $$\theta_D = (\theta^i_i)|_D $$ and $$\theta_{0, D} = (\theta_0)^i_i |_D + H^{i\bar j} H_{n\bar j} (\theta_0)^n_i |_D. $$ \[c-th\] The $(1, 0)$-form $$\theta_D = (H^{i\bar j} H_{i \bar j, k}|_D) d z_k = \partial \log \det H|_D $$ defines a global smooth Chern connection of the anti-canonical line bundle of $D$. In particular, its curvature form $\sqrt{-1}{\overline{\partial}}\theta_D$ is a smooth closed real $(1,1)$-form in $c_1(D)$. By Corollary \[asymp-g\], we have $g^{i\bar n}|_D=0$ and hence $$(\theta_D)^i_i = g^{i\bar \beta} g_{i \bar\beta, k } dz_k = (H^{i\bar j} H_{i\bar j, k}|_D )dz_k = \partial (\log \det H|_D) = \partial \log (\omega|_D)^{n-1}.$$ By Proposition \[expansion\], the regular part of $\omega$ restricted to $D$ is a smooth Kähler form from the expansion in Proposition \[expansion\] and for different holomorphic local coordinates $z=(z_1, ..., z_{n-1})$ and $w=(w_1, ..., w_{n-1})$ on $D$, $$\theta_D(z) = \theta_D(w) + \partial \log \left| \det \left( \frac{\partial z_i}{\partial w_j} \right) \right|^2.$$ Therefore $\theta_D $ defines a smooth connection on anti-canonical bundle of $D$. \[c-th0\] The $(1,0)$-form $$\begin{aligned} \theta_{0, D} &=& \left( (\theta_0)^i_i+(H^{i \bar j }H_{n \bar j} )(\theta_0)^n_i \right) |_D\\ &=& \left( (g_0)^{i\bar \beta} (g_0)_{i \bar \beta, k} |_D dz_k + (H^{i \bar j }H_{n \bar j} )(g_0)^{n\bar \beta} (g_0)_{i \bar \beta, k}|_D dz_k\right)\end{aligned}$$ defines a smooth Chern connection of the anticanonical bundle of $D$. In particular, $\sqrt{-1} {\overline{\partial}}\theta_{0,D}$ is a smooth closed real $(1,1)$ form in $c_1(D)$. Since $g_0$ is smooth and the restriction of $H_{i\bar j}$, $H^{i\bar j}$, $H_{n, \bar j}$ to $D$ are all smooth by the asymptotic expansion in Proposition \[expansion\], $\theta_{0, D}$ is locally smooth. It suffices to show that they patch together give rise to a connection. To do that we need to show that the transformation of $\theta_{0,D}$ under different coordinate charts satisfies the cocycle condition for the anticanonical bundle of $D$. Let $(z_1, ..., z_{n-1})$ and $(w_1, ..., w_{n-1})$ be two holomorphic local coordinates for some neighborhood of a point $p$ in $D$. Then they extend to two holomorphic coordinates $(z_1,.., z_{n-1}, z_n=\xi)$, $(w_1, ..., w_{n-1}, w_n=\eta)$ in a neighborhood of $p$ in $X$, where $D$ is locally defined by $\xi=0$ and $\eta=0$. Therefore for $i=1, ..., n-1$, $$\left. \frac{\partial \eta}{\partial z_i} \right |_D= \left.\frac{\partial \xi}{\partial w_i} \right|_D = 0.$$ By letting $$A^\alpha_\gamma= \frac{\partial z_\alpha}{\partial w_\gamma}, ~B^\alpha_\gamma = \frac{\partial w_\alpha}{\partial z_\gamma}, \tilde A^i_k= \frac{\partial z_i}{\partial w_k}, ~\tilde B^i_k = \frac{\partial w_i}{\partial z_k},$$ we obtain along $D$, $$A= \left( \begin{array}{cc} \tilde A & * \\ 0 & \frac{\partial \xi}{\partial \eta} \\ \end{array} \right), ~ B= \left( \begin{array}{cc} \tilde B & * \\ 0 & \frac{\partial \eta}{\partial \xi} \\ \end{array} \right), ~ A=B^{-1}, ~\tilde A = \tilde B^{-1}. $$ Straighforward computations show that $$\theta_{0, D}(z) = \theta_{0, D}(w) + \partial \log |\det \tilde A|^2$$ which completes the proof. For any smooth conical Kähler metric $\omega$ with cone angle $2\pi \beta$ along a smooth divisor $D$, we can define the first and second Chern classes $c_1(X, \omega)$ and $c_2(X, \omega)$. A priori, the intersection numbers among $c_1(X, \omega)$ and $c_2(X, \omega)$ might depend on the choice of $\omega$ even if the Ricci curvature of $\omega$ is bounded. The following proposition relates the $c_1(X, \omega)$ and $c_2(X, \omega)$ to $c_1(X)$ and $c_2(X)$. \[an-char\] Let $D$ be a smooth simple divisor on a Kähler manifold $D$. Suppose $\omega$ is a smooth conical Kähler metric with cone angle $2\pi \beta$ along $D$ with bounded Ricci curvature. Then $$\label{c1} \int_X c_1(X, \omega) \wedge \omega^{n-1} = (c_1(X) -(1-{\beta}) [D]) \cdot [\omega]^{n-1} , $$ $$\label{c2} \int_X c_2(X, \omega) \wedge \omega^{n-2} = \left(c_2(X) + (1-\beta)(-c_1(X) + [D] \right) \cdot [D] ) \cdot [\omega]^{n-2}, $$ and $$\label{c12} \int_X c_1^2(X, \omega) \wedge \omega^{n-2} = (c_1(X) - (1-{\beta})[D])^2 \cdot [\omega]^{n-2}.$$ We break the proof into the following steps. [*Step 1.*]{} Equation (\[c1\]) and () follow easily from that observation that ${\mathrm{Ric}}(\omega)$ is smooth on $X\setminus D$ and it continuously extends to a closed $(1,1)$ form on $X$, which lies in the class of $c_1(X).$ Therefre, $$c_1(X, \omega) = c_1(X) - (1-\beta)[D].$$ [*Step 2.*]{} We first introduce a few notations. Let $\omega_0$ be a smooth Kähler form in the same class of $[\omega]$. Since the curvature tensor can be viewed as the curvature in the tangent bundle, we write $\theta = (\theta^i_j)$, $\theta_0=(\theta^i_j)$ as the Chern connections on the tangent bundle with respect to the Kähler metric $\omega$ and $\omega_0$. Their curvature forms are given by $\Omega$ and $\Omega_0$ with $$\Omega= \sqrt{-1} {\overline{\partial}}\theta, ~~\Omega_0 = \sqrt{-1} {\overline{\partial}}\theta_0.$$ Let $s$ be a defining section of $D$ and $h$ be a smooth hermitian metric on the line bundle associated to $[D]$. We define $$X_\epsilon = \{ p\in X~|~ |s|_h^2(p) > \epsilon^2 \}.$$ Then locally $\xi = fS$ for some holomorphic function and on $\partial X_\epsilon$, we have $$\xi = fs = |fs| e^{ i \sigma} = \epsilon|f| h^{-1/2} e^{\sqrt{-1} \sigma}, ~~\sigma \in [0, 2\pi),$$ and $$d \xi = \sqrt{-1} \xi d\sigma + \epsilon e^{ \sqrt{-1} \sigma} d (|f|h^{-1/2}).$$ Let $\tau = d(|f|h^{-1/2})$. Then $\tau$ is a smooth $1$-form and on $\partial X_{\epsilon}$, $$d\xi = \sqrt{-1} \xi d\sigma + \epsilon \tau|_{\partial X_{\epsilon}}$$ [*Step 3.*]{} At each point $p\in \partial X_\epsilon$, we can apply a linear transformation to $(z_1, ..., z_{n-1})$ such that $g_{i\bar j} = \delta_{ij} $ at $p$ and by rescaling $\xi$ so that $g_{n\bar n} = |\xi|^{-2(1-\beta)}$ near $p$. Let $$H_{i\bar j} = g_{i\bar j}, ~~H^{i\bar j} = (H^{-1})_{i\bar j}, ~~H_{n \bar j} = g_{n \bar j}.$$ then we have $$\label{gxi} |\xi|^{-2(1-\beta)} g^{i\bar n} = |\xi|^{-2(1-\beta)} (g_{n\bar i} + o(1)) (\det(g_{\alpha\beta}) )^{-1} =g_{n\bar i} + o(1) = H^{i \bar j } H_{n\bar j } + o(1)$$ and the connection form $\theta$ has following estimates $$\begin{aligned} \theta^n_n &=& \theta^\xi_\xi= \sum g^{n\bar \beta} g_{n\bar \beta, \alpha} d z_\alpha = -(1-\beta +o(1)) \xi^{-1}d\xi + \sum_i o(1)\cdot dz_i \\ &=& -(1-{\beta}) \sqrt{-1} d\sigma +\sum_i O(1)\cdot d z_i \text{ (since }\xi\sim {\epsilon}\text{ on } {\partial}X_{\epsilon}\text {)}\\ $$ $$\begin{aligned} \theta^i_n &=& \sum g^{i \bar \beta} g_{n \bar \beta, \alpha} d z_\alpha = g^{i \bar n} g_{n \bar n , n} d \xi + o(1)\xi^{-1} d\xi + \sum_i o(1)\cdot d z_i \\ &=&g^{i\bar n} (|\xi|^{-2(1-\beta)})_\xi d\xi + o(1)\cdot\xi^{-1} d\xi + \sum_i o(1)\cdot d z_i \\ &=& -(1-\beta) H^{i\bar j} H_{n \bar j} \xi^{-1} d\xi + o(1)\cdot\xi^{-1} d\xi + \sum_i o(1)\cdot d z_i \\ &=& -\sqrt{-1}(1-\beta) H^{i\bar j} H_{n \bar j} d\sigma + o(1)\cdot d\sigma + \sum_i O(1)\cdot dz_i $$ $$\begin{aligned} \theta^n_i &=& O(1)\cdot d \xi + \sum_i o(1)\cdot dz_i = o(1)\cdot d\sigma +\sum_i o(1)\cdot dz_i \\ \theta^i_k &=& \sum_\alpha O(1)\cdot d z_\alpha = o(1)\cdot d\sigma + \sum_i O(1)\cdot dz_i. $$ On $\partial X_\epsilon$, by our assumption that ${\beta}\in (1/2,1)$ and using we deduce $$\begin{aligned} \Omega^n_n&=& \sqrt{-1} \sum_{\alpha, \beta} R^n_{n, \alpha \bar \beta} dz_\alpha \wedge d\bar z_\beta \\ &=& \sum_\alpha o(1)\cdot d\sigma \wedge dz_\alpha + \sum_\beta o(1)\cdot d\sigma \wedge d\bar z_\beta + \sum_{i, j} O(1)\cdot dz_i \wedge d\bar z_j $$ $$\begin{aligned} \Omega^i_n &=& \left( g^{i\bar \gamma} g_{n\bar \gamma, \alpha} \right)_{\bar \beta} dz_\alpha \wedge d\bar z_\beta\\ &=& \sqrt{-1} (g^{i\bar n} g_{n\bar n, n})_{\bar z_k} dz_k\wedge d\xi + \sqrt{-1}(g^{i\bar j } g_{n\bar j, n})_{\bar z_k} dz_k \wedge d\xi + \sqrt{-1} (g^{i\bar j} g_{n\bar j, k} )_{\bar\xi} d\xi \wedge d z_k \\ &&+ \sqrt{-1} (g^{i\bar j} g_{n\bar j, k} )_{\bar z_l} d\bar z_l\wedge d z_k + \sum_{k,l}o(1)\cdot dz_k \wedge d\bar z_l \\ &=& \sqrt{-1} (1-{\beta})(H^{i\bar j} H_{n\bar j})_{\bar z_k} dz_k \wedge \xi^{-1} d\xi + \sum_k o(1)\cdot dz_k \wedge d\sigma + \sum_{k,l}O(1)\cdot dz_k \wedge d\bar z_l\\ &=& \sqrt{-1}(1-{\beta}) {\overline{\partial}}(H^{i\bar j} H_{n\bar j}) \wedge d\sigma + \sum_k o(1)\cdot dz_k \wedge d\sigma + \sum_{k,l}O(1)\cdot dz_k \wedge d\bar z_l $$ $$\begin{aligned} \Omega^n_i &=& \sqrt{-1} R^n_{i, \alpha \bar\beta} dz_\alpha \wedge d\bar z_\beta \\ &=& \sqrt{-1} R^n_{i, n\bar l} d\xi \wedge d\bar z_l + \sqrt{-1} R^n_{i, k \bar n} dz_k \wedge d\bar \xi \\ && + \sqrt{-1} R^n_{i, k\bar l} dz_k \wedge d\bar z_l + o(1) dz_k \wedge d\bar z_l\\ &=& o(1) dz_k \wedge d\sigma + o(1) dz_k \wedge d\bar z_l\\ $$ $$\begin{aligned} \Omega^i_p &=& \sqrt{-1} R^i_{p, \alpha \bar\beta} dz_\alpha \wedge d\bar z_\beta \\ &=& \sqrt{-1} R^i_{p, n\bar l} d\xi \wedge d\bar z_l + \sqrt{-1} R^i_{p, k \bar n} dz_k \wedge d\bar \xi \\ && + \sqrt{-1} R^i_{p, k\bar l} dz_k \wedge d\bar z_l + o(1) dz_k \wedge d\bar z_l\\ &=& o(1) dz_k \wedge d\sigma + O(1) dz_k \wedge d\bar z_l $$ [*Step 4.*]{} By our assumption, $\omega$ is a smooth Kähler metric with cone angle $2\pi{\beta}$ along $D\in X$ such that it Ricci curvature is bounded on $X\setminus D$. This implies that $$\label{c1-loc} \Omega^{\alpha}_{\alpha}={\mathrm{Ric}}(\omega)=\omega_0-(1-{\beta}) {\sqrt{-1}\partial{\overline{\partial}}}\log|s_D|_h^2$$ where $h$ is a smooth Hermitian metric on the line bundle $\mathcal{O}_X(D)$, $\omega_0$ is a smooth Käher metric and $s_D\in H^0(X,\mathcal{O}_X(D))$ is a defining section for $D$. This implies that $$\int_{X\setminus X_{\epsilon}}\Omega^{\alpha}_{\alpha}\wedge \omega^{n-1} =([\omega_0]-(1-{\beta})[D])\cdot \omega^{n-1}+o({\epsilon})\ ,$$ from which we obtain . Let $\theta_t = t\theta + (1-t) \theta_0$ be the connection on the tangent bundle $TX$. The curvature $$\Omega_t = \sqrt{-1} \bar \partial \theta_t = t \Omega + (1- t) \Omega_0.$$ The transgression formula gives $$c_2(X, \omega) - c_2(X, \omega_0) = 2 \sqrt{-1} \int_0^1 {\overline{\partial}}P_2(\theta- \theta_0, \Omega_t) dt$$ and $$\begin{aligned} && 2 P_2(\theta-\theta_0, \Omega_t)\\ &=& (\theta- \theta_0)^n_n \wedge (\Omega_t)^i_i - (\theta-\theta_0)^i_n \wedge (\Omega_t)^n_i +(\theta- \theta_0)^i_i \wedge (\Omega_t)^n_n -(\theta-\theta_0)^n_i \wedge (\Omega_t)^i_n \\ &&- (\theta-\theta_0)^i_k \wedge (\Omega_t)^k_i\\ &=& -\sqrt{-1}(1-{\beta}) (t {\overline{\partial}}\theta + (1-t) {\overline{\partial}}\theta_0)^i_i \wedge d\sigma \\ && - (1-t) (1-\beta)\sqrt{-1} \left( H^{i\bar j} H_{n\bar j} {\overline{\partial}}(\theta_0)^n_i - (\theta-\theta_0)^n_i \wedge {\overline{\partial}}(H^{i\bar j} H_{n\bar j}) \right) \wedge d\sigma\\ && + o(1) d\sigma \wedge dz_k \wedge d\bar z_l + O(1) dz_i \wedge dz_k \wedge d\bar z_l\\ &=& \sqrt{-1} (1-{\beta})\left(-t{\overline{\partial}}( \theta^i_i) \wedge d\sigma - (1-t) {\overline{\partial}}\left( (\theta_0)^i_i + H^{i\bar j}H_{n\bar j} (\theta_0)^n_i \right) \wedge d\sigma\right) \\ && + o(1) d\sigma \wedge dz_k \wedge d\bar z_l + O(1) dz_i \wedge dz_k \wedge d\bar z_l. $$ Now let $\eta$ be a smooth Kähler form. Then $$\begin{aligned} &&\int_X (c_2(X, \omega) - c_2(X, \omega_0) \wedge\eta^{n-2}\\ &=&2 \sqrt{-1} \int_0^1 \left( \int_X {\overline{\partial}}P_2(\theta - \theta_0, t\Omega + (1-t) \Omega_0) \wedge \eta^{n-2} \right) dt\\ &=&2 \int_0^1 \left( \int_{\partial X_\epsilon} P_2 (\theta- \theta_0, t\Omega + (1-t) \Omega_0) \wedge \eta^{n-2} \right) dt + o(1)\\ &=&(1-{\beta}) \sqrt{-1} \int_0^1 \int_{\partial X_\epsilon} \left(- t {\overline{\partial}}(\theta^i_i) - (1-t) {\overline{\partial}}( (\theta_0)^i_i + H^{i\bar j} H_{n\bar j} (\theta_0)^n_i) \right)\wedge d\sigma \wedge \eta^{n-2} dt + o(1)\\ &=&(1-{\beta}) \sqrt{-1}\int_0^1 \int_D \left( -t {\overline{\partial}}\theta_D - (1-t) {\overline{\partial}}\theta_{0, D} \right) \wedge (\eta|_D) ^{n-2}dt\\ &=&(1-{\beta}) (-c_1(D)/2-c_1(D)/2)\cdot [D] \cdot [\eta]^{n-2} + o(1) \\ &=&(1-{\beta}) (-c_1(X) + [D])\cdot [D]\cdot [\eta]^{n-2} + o(1) . $$ The last three equalities follow from Lemma \[c-th\], Lemma \[c-th0\] and the adjunction formula. And similarly, we have $$\begin{aligned} &&\int_X (c_1^2(X, \omega) - c_1^2(X, \omega_0) \wedge\eta^{n-2}\\ &=& \int_0^1 \left( \int_X {\overline{\partial}}Q(\theta - \theta_0, t\Omega + (1-t) \Omega_0) \wedge \eta^{n-2} \right) dt\\ &=&2(1-{\beta})\int_0^1 \int_{\partial X_\epsilon} \left(- t {\overline{\partial}}(\theta^{\alpha}_{\alpha}) - (1-t) {\overline{\partial}}( (\theta_0)^{\alpha}_{\alpha}) \right)\wedge d\sigma \wedge \eta^{n-2} dt + o(1)\\ \text{( by \eqref{c1-loc} )}&=&(1-{\beta}) \bigg(-c_1(X)+(1-{\beta})[D]-c_1(X)\bigg)\cdot [D] \cdot [\eta]^{n-2} \\ &=&\bigg (-2(1-{\beta})c_1(X)\cdot [D]+(1-{\beta})^2 [D]^2\bigg)\cdot[\eta]^{n-2} $$ which is equivalent to . [Step 5.]{} Suppose that $\omega_0 \in [\omega]$ is a smooth Kähler form. We want to show that $$\label{c2-g} \int_X c_2(X, \omega) \wedge \omega^{n-2} = \int_X c_2(X, \omega) \wedge \omega_0^{n-2}$$ and $$\label{c12-g} \int_X c_1^2(X, \omega) \wedge \omega^{n-2} = \int_X c_1^2(X, \omega) \wedge \omega_0^{n-2} . $$ Since the proofs are parallel to each other, we will only prove the . Let $\varphi $ be defined by $\omega = \omega_0 + {\sqrt{-1}\partial{\overline{\partial}}}\varphi$. $$\int_X c_2(X, \omega) \wedge ( \omega^{n-2} - \omega_0^{n-2}) = \sum_{i=0}^{n-3} \int_X {\sqrt{-1}\partial{\overline{\partial}}}\varphi \wedge c_2(X, \omega)\wedge \omega^i \wedge \omega_0^{n-3-i} .$$ For any $i= 0, 1, ..., n-3$, $$\begin{aligned} && \int_X {\sqrt{-1}\partial{\overline{\partial}}}\varphi \wedge c_2(X, \omega)\wedge \omega^i \wedge \omega_0^{n-3-i} \\ &=&\int_{\partial X_\epsilon} d^c \varphi \wedge c_2(X, \omega)\wedge \omega^i \wedge \omega_0^{n-3-i} + o(1) .\\ $$ Note that $$\begin{aligned} d^c \varphi &=& o(1) d\sigma + O(1) d z_k , \\ \omega &=& o(1) d\sigma \wedge d z_k +o(1) d\sigma \wedge d \bar z_k + O(1) dz_k \wedge d \bar z_l $$ and $$\begin{aligned} && \Omega^i_i \wedge \Omega^j_j \sim \Omega^i_j \wedge \Omega^j_i \sim \Omega^n_n \wedge \Omega^j_j \sim \Omega^n_j \wedge \Omega^j_n \\ &=& o(1) d z_k \wedge d \bar z_ l \wedge dz_p \wedge d \sigma + o(1) d z_k \wedge d \bar z_l \wedge d \bar z_q \wedge d\sigma + O(1) dz_k \wedge d\bar z_l \wedge d z_p \wedge d \bar z_q . $$ Therefore $$\begin{aligned} && \int_X {\sqrt{-1}\partial{\overline{\partial}}}\varphi \wedge c_2(X, \omega)\wedge \omega^i \wedge \omega_0^{n-3-i} \\ &=&\int_{\partial X_\epsilon} d^c \varphi \wedge c_2(X, \omega)\wedge \omega^i \wedge \omega_0^{n-3-i} + o(1)\\ &=& 0 $$ after letting $\epsilon$ tend to $0$. [Step 6.]{} Finally, combining the above estimates, we obtain and the proof of the proposition is completed. One can apply similar argument shows that if $\omega$ is a smooth conical Kähler metric $\omega$ with cone angle $2\pi \beta$ along a smooth simple divisor $D$ and if the Ricci curvature of $\omega$ is bounded, then the $n^{th}$-Chern number $c_n(X, \omega)$ is well-defined and does not depend on the choice of $\omega$ . The Gauss-Bonnet and signature theorems for Kähler surfaces with conical singularities -------------------------------------------------------------------------------------- We first introduce the following definition. \[Eu-Sig\] Let $X$ be a Kähler surface and $\Sigma$ be a smooth holomorphic curve on $X$. If $g$ is a smooth conical Kähler metric with cone angle $2\pi\beta$ along $\Sigma$, we define the corresponding conical Euler number and signature by $$\chi(X,g)=\int_{X\setminus \Sigma} \frac{1}{8\pi^2}\left(\frac{S^2}{24}+|W|^2 -\frac{|\stackrel{\circ}{{\mathrm{Ric}}}|^2}{2} \right)dg$$ and $$\sigma(X,g)=\frac{1}{12\pi^2}\int_{X\setminus \Sigma} (|W^+|^2-|W^-|^2)dg,$$ where $S$ is the scalar curvature of metric $g$, $W$ is the Weyl tensor for $g$ and $\stackrel{\circ}{{\mathrm{Ric}}}$ is the traceless Ricci curvature. In particular, if $\beta=1$, we recover classical characteristic class. The Gauss-Bonnet and the signature theorems are proved in [@AL] for smooth compact Riemannian $4$-folds with specified conical metrics with cone angle $2\pi \beta$ along a smooth embedded Riemann surface. In particular, As an immediate consequence of Proposition \[an-char\] and Definition \[Eu-Sig\] above, we obtain the following formulas related to the recent result by Atiyah and LeBrun [@AL] by removing the assumption $\beta\in(0,1/3)$ in the Kähler case. Let $g$ be a smooth conical Kähler metric with angle $2\pi\beta$ for $\beta\in (0, 1]$ along a holomorphic curve $\Sigma$. If the Ricci curvature of $g$ is bounded, we have $$\begin{aligned} \chi(X,g) &=& \chi(X)-(1-{\beta})\chi(D)\\ \sigma(X,g)&=&\sigma(X)-\frac{1}{3}(1-{\beta}^2)[D]^2. $$ To prove the statement, we apply the identity (cf. [@AL] ) $$\frac{1}{8\pi^2}\left(\frac{S^2}{24}+|W|^2 -\frac{|\stackrel{\circ}{{\mathrm{Ric}}}|^2}{2} \right)dg =c_2(X,g)$$ and $$\frac{1}{12\pi^2}(|W^+|^2-|W^-|^2)dg=\frac{1}{3}(c_1^2(X,g)-2c_2 (X,g))$$ and the statement follows from Proposition \[an-char\] The Chern number inequality on Fano manifolds --------------------------------------------- In this section, we will prove Theorem \[main4\]. \[chern3\] Let $X$ be an $n$-dimensional Fano manifold. If $R(X)=1$, then the Miyaoka-Yau type inequality holds $$c_2(X) \cdot c_1(X)^{n-2} \geq \frac{n } {2(n+1)} c_1(X)^n\ .$$ We fix a smooth simple divisor $D\in |-mK_X|$ for some $m\in \mathbb{Z}^+$. Such a divisor always exists by Bertini’s theorem. Then for any $\beta\in (0,1)$, there exists a smooth conical Käher-Einstein metric $\omega$ satisfying ${\mathrm{Ric}}(\omega) = \beta g\omega+ (1-\beta)m^{-1} [D]$. By Chern-Weil theory, if we let $$\stackrel{\circ} R_{i\bar j k\bar l} = R_{i\bar j k\bar l} - \frac{tr(R)}{n(n+1)} (g_{i\bar j } g_{k\bar l} + g_{i\bar l} g_{k\bar j} ),$$ $$\stackrel{\circ}{Ric}= Ric - \frac{ tr(Ric)}{n} g$$ be the traceless curvature and Ricci curvature tensor, we have $$\left(\frac{2(n+1)}{n} c_2(X, \omega) - c_1^2(X, \omega) \right) \cdot [\omega]^{n-2} = \frac{1}{n(n-1)} \int_X \left( \frac{n+1}{n} |\stackrel{\circ} R |^2 - \frac{ n^2-2}{n^2} |\stackrel{\circ}{Ric}|^2 \right) \omega^n. $$ We then have $$\left(\frac{2(n+1)}{n} c_2(X, \omega) - c_1^2(X, \omega) \right) \cdot [\omega]^{n-2} \geq 0.$$ By Proposition \[an-char\], we have $$c_1(X, \omega) =\beta c_1(X) , ~~~ c_2(X, \omega)= c_2(X)+ (1-\beta) (- c_1 (X) + [D]) \cdot [D].$$ This implies that $$\left( \frac{2(n+1)}{n} c_2(X) - \beta^2 c^2_1(X) \right) \cdot (\beta c_1(X) ) ^{n-2} \geq 0.$$ The theorem then follows by letting $\beta \rightarrow 1$. We also have the following lemma when $R(X)<1$ with the same argument in the proof of Proposition \[chern3\]. \[chern4\] Let $D\in |-K_X| $ be a smooth simple divisor. If there exists a conical Kähler-Einstein metric $g_\beta$ for some $\beta \in (0, R(X))$ satisfying $${\mathrm{Ric}}(g) =\beta g + (1-\beta) [D],$$ then $$c_2(X) \cdot c_1(X)^{n-2} \geq \frac{n \beta^2} {2(n+1)} c_1(X)^n\ .$$ Theorem \[main4\] is proved by combining Proposition \[chern3\], Lemma \[chern4\] and Theorem \[main2\]. The Chern number inequality on minimal manifolds of general type ---------------------------------------------------------------- The following proposition is first claimed in [@Ts], although the analytic part in the proof does not seem to be complete. The first complete proof seems to be given by Zhang [@ZhY] using the Ricci flow. In this section, we apply Proposition \[an-char\] to complete Tsuji’s original approach. Let $X$ be a smooth minimal model of general type. Then $$\left( \frac{2(n+1)}{n} c_2(X) - c_1^2(X) \right) \cdot (-c_1(X))^{n-2} \geq 0. $$ In particular, if the equality holds, the canonical Kähler-Einstein metric is a complex hyperbolic metric on the smooth part of the canonical model of $X$. Fix a smooth simple ample divisor $D$ on $X$. Since $[K_X]+\epsilon [D]$ is a Kähler class for any $\epsilon>0$, there exists a smooth conical Kähler-Einstein metric in $[K_X] + \epsilon [D]$ with conical singularity along $D$ satisfying $${\mathrm{Ric}}( \omega ) = -\omega + \epsilon [D].$$ By the same argument in the proof of Proposition \[chern3\], $$( \frac{2(n+1)}{n} c_2(X, \omega ) - c_1^2(X, \omega ) ) \cdot (-c_1(X, \omega))^{n-2} \geq \int_X |\stackrel{\circ} R(\omega)|^2 \omega^n\geq 0.$$ By standard argument from [@EGZ; @ZhZ], $\omega$ converges as $\epsilon \rightarrow 0$ to the unique Kähler-Einstein metric $g_{can}$ on the canonical model $X_{can}$ of $X$ in $C^\infty$ local topology away from the exceptional locus the pluricanonical system. In particular, $\omega_{can}$ is smooth on the smooth part of $X_{can}$. By letting $\epsilon$ tends to $0$, we have $$\label{chern6} \left( \frac{2(n+1)}{n} c_2(X) - c_1^2(X) \right) \cdot (-c_1(X))^{n-2} \geq \int_X |\stackrel{\circ}R(\omega_{can})|^2 \omega_{can}^n\geq 0, $$ where $X_{can}^\circ$ is the smooth part of $X_{can}$. When the equality holds, $\stackrel{\circ}R (g_{can})$ vanishes on $X_{can}^\circ$ and so $g_{can}$ must a complex hyperbolic metric on $X^\circ_{can}$. Then by the estimate (\[chern6\]), we immediately obtain an $L^2$-bound for the curvature tensor of the canonical Kähler-Einstein metric on the regular part of the canonical model associated to a smooth minimal model of general type. Discussions =========== In this section, we make some speculations on the limiting behavior of the conical Kähler-Einstein metrics as $\beta$ tends to $R(X)$. Let $X$ be a Fano manifold with $R(X)=1$. Then by Proposition \[rmain1\], there exists a smooth simple divisor $D\in |-mK_X|$ for some $m\in \mathbb{Z}^+$ such that for any $\epsilon>0$, there exists a smooth conical Kähler-Einstein metric $g_\epsilon$ with ${\mathrm{Ric}}(g_\epsilon) = (1-\epsilon)g_\epsilon + \epsilon m^{-1} [D]$ for all $\epsilon \in (0, 1)$. We conjecture that $(X, g_\epsilon)$ converges to a $\mathbb{Q}$-Fano variety $(X_\infty, g_\infty)$ coupled with a canonical Kähler-Einstein metric $g_\infty$ in Gromov-Hausdorff topology. The above conjecture is related to the recent result in [@TW], where the Kähler-Ricci flow is combined with the continuity method to produce a limiting Einstein metric space when $R(X)=1$. We also make a more general conjecture when $R(X)\neq 1$. Let $X$ be a Fano manifold and $D$ be a smooth simple divisor in $|-mK_X|$ for some $m\in \mathbb{Z}^+$. We consider the conical Ricci flow defined by $${\frac{\partial g}{\partial t}} = - {\mathrm{Ric}}(g) + \beta g + m^{-1}(1-\beta) [D] $$ starting with a smooth conical Kähler metric $g_0\in c_1(X)$ with cone angle $2\pi (1-(1-\beta)m^{-1})$ along $D$. Then for some $m\in \mathbb{Z}^+$ and a generic choice of $D$, we have 1. If $\beta \in (0, R(X) )$, the flow converges to a smooth conical Kähler Einstein metric on $X$ with conical singularity along $D$. 2. If $\beta = R(X)$, the flow converges to a singular Kähler-Einstein metric on a paired $\mathbb{Q}$-Fano variety $(X_\infty, D_\infty)$ with conical singularities along an effective $\mathbb{Q}$-divisor $D_\infty \in [-K_{X_\infty}]$. 3. If $\beta \in (R(X), 1]$, the flow converges to a singular Kähler-Ricci soliton on a paired $\mathbb{Q}$-Fano variety $(X_\infty , D_\infty)$ with conical singularities along an effective $\mathbb{Q}$-divisor $D_\infty \in [ - K_{X_\infty}]$. [**Acknowledgements**]{} We would like to thank D.H. Phong, Jacob Sturm, Valentino Tosatti, Ved Datar and Bin Guo for many stimulating discussions. After we finished the first draft of the paper, we were kindly informed by Tosatti that most of the results in Theorem \[main1\] and \[main2\] are independently obtained by Chi Li and Song Sun [@LS]. We would also like to thank Chi Li and Song Sun for sending us their preprint and for their interesting comments. [99]{} Atiyah, M. and Lebrun, C. [*Curvature, Cones, and Characteristic Numbers*]{}, arXiv:1203.6389 Aubin, T. [*Équations du type Monge-Ampère sur les variétés kählériennes compactes*]{}, Bull. Sci. Math. (2) [**102**]{} (1978), no. 1, 63–95 Bando, S. [*The K-energy map, almost Einstein Kähler metrics and an inequality of the Miyaoka-Yau type*]{}, Tohuku Math. Journ. [**39**]{} (1987), 231–235 Bando, S. and Mabuchi, T. [*Uniqueness of Einstein Kähler metrics modulo connected group actions*]{}, Adv. Stud. in Pure Math. [**10**]{} (1987), 11–40 Berman, B. [*A thermodynamical formalism for Monge-Ampere equations, Moser-Trudinger inequalities and Kähler-Einstein metrics*]{}, arXiv:1011.3976 Berman, B. [*K-polystability of Q-Fano varieties admitting Kähler-Einstein metrics*]{}, arXiv:1205.6214 Berman, R., Boucksom, S, Guedj, V. and Zeriahi,, A. [*A variational approach to complex Monge-Ampere equations*]{}, arXiv:0907.4490 Brendle, S. [ *Ricci flat Kähler metrics with edge singularities*]{}, arXiv:1103.5454 Demailly, J-P. [*Appendix to I. Cheltsov and C. ShramovÕs article ÒLog canonical thresholds of smooth Fano threefoldsÓ : On TianÕs invariant and log canonical thresholds*]{}, Uspekhi Mat. Nauk, 63:5(383) (2008), 73–180 Ding, W., [*Remarks on the existence problem of positive Kähler-Einstein metrics*]{}, Math. Ann. [**282**]{} (1988), no. 3, 463–471 Donaldson, S.K. [*Scalar curvature and stability of toric varieties*]{}, J. Differential Geom. 62 (2002), 289–349 Donaldson, S.K. [*Extremal metrics on toric surfaces: a continuity method*]{}, J. Differential Geom. 79 (2008), no. 3, 389–432 Donaldson, S.K. [*Kähler geometry on toric manifolds, and some other manifolds with large symmetry*]{}, Handbook of geometric analysis. No. 1, 29Ð75, Adv. Lect. Math. (ALM), 7, Int. Press, Somerville, MA, 2008 Donaldson, S.K. [ *Stability, birational transformations and the Kähler-Einstein problem*]{}, arXiv:1007.4220 Donaldson, S.K. [*Kähler metrics with cone singularities along a divisor*]{}, arXiv:1102.1196 Eyssidieux, P., Guedj, V. and Zeriahi, A. [*Singular Kähler-Einstein metrics*]{}, J. Amer. Math. Soc. 22 (2009), no. 3, 607–639 Guillemin, V. [*Kähler structures on toric varieties*]{}, J. Differential Geom. 40 (1994), no. 2, 285–309 Jeffres,T., Mazzeo,R. and Rubinstein, Y. [*Kähler-Einstein metrics with edge singularities with appendix by Rubinstein Y. and Li, C.*]{}, arXiv:1105.5216 Kobayashi, R. [*Kähler-Einstein metric on an open algebraic manifold*]{}, Osaka J. Math. 21 (1984), no. 2, 399–418 Kolodziej,S. [*The complex Monge-Ampere equation*]{}, Acta Math. 180 (1998),69–117. Li, C. [*Greatest lower bounds on Ricci curvature for toric Fano manifolds*]{} Adv. Math. 226 (2011), no. 6, 4921–4932. Li, C. [*On the limit behavior of metrics in continuity method to Kähler-Einstein problem in toric Fano case*]{}, arXiv:1012.5229 Li, C. [*Remarks on logarithmic K-stability*]{}, arXiv:1104.0428 Li, C. and Sun, S. [*Conic Kähler-Einstein metric revisited*]{}, preprint Li, H. [*On the lower bound of the K-energy and F-functional*]{}, Osaka J. Math. 45 (2008), no. 1, 253–264 Luo, F. and Tian, G. [*Liouville equation and spherical convex polytopes*]{}, Proc. Amer. Math. Soc. 116 (1992), 1119–1129 Mabuchi, T. [*K-energy maps integrating Futaki invariants*]{}, Tôhoku Math. Journ. [**38**]{} (1986), 575–593 McOwen, R.C. [*Point singularities and conformal metrics on Riemann surfaces*]{}, Proc. Amer. Math. Soc. 103 (1988), 222–224 Miyaoka, Y. [*On the Chern numbers of surfaces of general type*]{}, Invent. Math., 42 (1977), 225–237. Munteanu, O. and Szekelyhidi, G. [*On convergence of the Kähler-Ricci flow* ]{}, Comm. Anal. Geom. 19 (2011), n. 5, 887–904 Odaka, Y. and Sun,S. [*Testing log K-stability by blowing up formalism*]{}, arXiv:1112.1353 Phong, D.H., Sesum, N. and Sturm, J. [*Multiplier ideal sheaves and the KŠhler-Ricci flow*]{}, Comm. Anal. Geom. 15 (2007), no. 3, 613–632 Phong, D.H., Sturm, J., Song, J. and Weinkove, B. [*The Moser-Trudinger inequality on Kähler-Einstein manifolds*]{}, Amer. J. Math. 130 (2008), no. 4, 1067–1085 Rubinstein, Y. [*On energy functionals, Kähler-Einstein metrics, and the Moser-Trudinger-Onofri neighborhood*]{}, J. Funct. Anal. 255 (2008), no. 9, 2641–2660 Siu, Y.-T. [*Lectures on Hermitian-Einstein metrics for stable bundles and Kähler-Einstein metrics*]{}. DMV Seminar, 8. Birkhäuser Verlag, Basel, 1987 Song, J. [*The $\alpha$-invariant on certain surfaces with symmetry groups*]{}, Trans. Amer. Math. Soc. [**357**]{} (2005), no. 1, 45–57 Song, J. and Tian, G. [*Canonical measures and Kähler-Ricci flow*]{}, J. Amer. Math. Soc. 25 (2012), no. 2, 303–353 Song, J. and Weinkove, B. [*Energy functionals and canonical Kähler metrics*]{}, Duke Math. J. 137 (2007), no. 1, 159–184 Sun, S. [*Note on K-stability of pairs*]{}, arXiv:1108.4603 Szekelyhidi, G. [*Greatest lower bounds on the Ricci curvature of Fano manifolds*]{}, Compositio Math. 147 (2011), 319–331 Tian, G. [*On Kähler-Einstein metrics on certain Kähler manifolds with $c_1(M)>0$*]{}, Invent. Math. [**89**]{} (1987), 225–246 Tian, G. [*On stability of the tangent bundles of Fano varieties*]{}, Internat. J. Math. 3 (1992), no. 3, 401–413 Tian, G. [*Kähler-Einstein metrics on algebraic manifolds. Transcendental methods in algebraic geometry*]{}, 143Ð185, Lecture Notes in Math., 1646, Springer, Berlin, 1996 Tian, G. [*Kähler-Einstein metrics with positive scalar curvature*]{}, Invent. Math. [**137**]{} (1997), 1–37 Tian, G. and Wang, B. [ *On the structure of almost Einstein manifolds*]{}, arXiv:1202.2912 Tian, G. and Yau, S.T., [*Existence of Kähler-Einstein metrics on complete Kähler manifolds and their applications to algebraic geometry*]{}, Mathematical aspects of string theory (San Diego, Calif., 1986), 574Ð628, Adv. Ser. Math. Phys., 1, World Sci. Publishing, Singapore, 1987 Tian, G. and Zhu, X. [*A nonlinear inequality of Moser-Trudinger type*]{}, Calc. Var. [**10**]{} (2000), 349–354 Troyanov, M. [*Prescribing curvature on compact surfaces with conic singularities*]{}, Trans. Amer. Math. Soc. 324 (1991), 793–821 Tsuji, [*Stability of tangent bundles of minimal algebraic varieties*]{}, Topology 27 (1988), no. 4, 429–442 Wang, X. and Zhu, X,, [*Kähler-Ricci solitons on toric manifolds with positive first Chern class*]{}, Adv. Math. 188 (2004), no. 1, 87–103 Yau, S.-T. [*On the Ricci curvature of a compact Kähler manifold and the complex Monge-Ampère equation, I*]{}, Comm. Pure Appl. Math. [**31**]{} (1978), no. 3, 339–411 Yau, S.T. [*Calabi’s conjecture and some new results in algebraic geometry*]{}, Proc. Natl. Acad. Sci. USA, 74 (1977), 1798–1799 Yau, S.-T. [*Open problems in geometry*]{}, Proc. Symposia Pure Math. [**54**]{} (1993), 1–28 (problem 65) Zhang, Y. [*Miyaoka-Yau inequality for minimal projective manifolds of general type*]{}, Proc. Amer. Math. Soc. 137 (2009), no. 8, 2749–2754 Zhang, Z. [*On degenerate Monge-Ampère equations over closed Kähler manifolds*]{}, Int. Math. Res. Not. 2006, Art. ID 63640, 18 pp [^1]: The first named author is supported by National Science Foundation grants DMS-0847524 and a Sloan Foundation Fellowship.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Using Monte Carlo integration methods, we describe the behavior of the exact four-spinon dynamic structure function $S_{4}$ in the antiferromagnetic spin 1/2 Heisenberg quantum spin chain as a function of the neutron energy $\omega$ and momentum transfer $k$. We also determine the four-spinon continuum, the extent of the region in the $\left( k,\omega \right)$ plane outside which $S_{4}$ is identically zero. In each case, the behavior of $S_{4}$ is shown to be consistent with the four-spinon continuum and compared to the one of the exact two-spinon dynamic structure function $S_{2}$. Overall shape similarity is noted.' author: - 'A. Abada' - 'B. Si-Lakhal' title: 'MONTE CARLO STUDY OF FOUR-SPINON DYNAMIC STRUCTURE FUNCTION IN ANTIFERROMAGNETIC HEISENBERG MODEL' --- Introduction ============ The spin $s=\frac{1}{2}$ Heisenberg quantum spin chain describes the magnetic properties of quasi-one-dimensional antiferromagnetic compounds like KCuF$_{3}$ [@hirakawa-kurogi]. The spin dynamics is experimentally investigated using inelastic neutron scattering [@Exp]. From a theoretical standpoint, the quantity of interest is the dynamic structure function (DSF) $S$ of two local spin operators. This is because the magnetic scattering cross section per magnetic site is directly proportional to it [@squires-lovesey]. The Heisenberg model has been studied quite intensively [@PSD; @Baxter], and because of the presence of the $\mathrm{U}_{q}(\widehat{\mathrm{sl}}_{2})$ symmetry [@JimMi], a number of exact results are available. Static properties need only the Yang-Baxter relation [@Baxter], whereas dynamic correlation functions require the additional notion of vertex operators and exploit bosonization methods [@CorFun]. One thus obtains compact expressions for form factors [@JimMi]. Regarding the dynamic structure function, the focus has so far been on $S_{2}$, the two-spinon contribution to the total $S$. First there has been the Anderson (semi-classical) spin-wave theory [@And], an approach based on an expansion in powers of $1/s$ and hence, exact only in the classical limit $s=\infty $. It can describe with some satisfaction compounds with higher spins [@TMMC], but fails in the quantum limit $s=\frac{1}{2}$. For this latter system, the Müller ansatz has been proposed, built mainly from finite-chain calculations and symmetry considerations [@Muller]. It is an approximate expression for $S_{2}$ that accounts satisfactorily for many aspects of the phenomenology [@Exp]. More recently, an exact expression for $S_{2}$ has been obtained, making extensive use of the $\mathrm{U}_{q}(\widehat{\mathrm{sl}}_{2})$ symmetry [@BCK], and a comparison with the Müller ansatz shows that it gives a better account of the data [@KMB; @BKM]. Beyond the two-spinon DSF $S_{2}$ are of course all the other $S_{2p>2}$ contributions. The first one to look at is the four-spinon dynamic structure function $S_{4}$, and the purpose of the present article is to describe its behavior. An exact expression for $S_{4}$ has been derived in [@ABS], see also [@Boug]. We intend to use that expression to describe the behavior of $S_{4}$ as a function of the neutron energy transfer $\omega$ and neutron momentum transfer $k$. We must mention that a preliminary investigation into the behavior of $S_{4}$ has already been initiated in [@ABSS]. There we have used quadratures to perform the integrations involved in the expression of $S_{4}$. But because of slow convergence of the algorithms we wrote, we could describe the behavior of $S_{4}$ only as a function of $k$, and only for relatively small values of $\omega$. In the present work, the integrations are performed using Monte Carlo methods, and this makes it possible not only to study $S_{4}$ as a function of $k$ for a wider range of values of $\omega$, but obtain a description of $S_{4}$ as a function of $\omega$ for a wide range of values of $k$ as well. It is however important to note that, for the same values of $\omega$, the behavior of $S_{4}$ as a function of $k$ we obtain here by Monte Carlo methods is similar and consistent with the one we obtained in [@ABSS] using quadratures. Also, we systematically carry a comparison of each result we obtain regarding $S_{4}$ to a corresponding one regarding $S_{2}$. The reason is that there is good familiarity in the literature with the two-spinon DSF and so, such a comparison allows faster acquaintance with the four-spinon contribution. To have the discussion as clear as possible, we scale both $S_{4}$ and $S_{2}$ to one. All results concerning $S_{2}$ are already known [@BCK; @KMB]; only those concerning $S_{4}$ are new. This article is organized as follows. After these introductory remarks, we describe in the next section the Heisenberg model and give the definition of the dynamic structure function. We write its decomposition in $2p$-spinon contributions and give the expressions of $S_{2}$ and $S_{4}$. In section three, we first determine the four-spinon continuum, the region in the $\left( k,\omega \right)$ plane outside which $S_{4}$ is identically zero. We will see that it extends beyond the spin-wave des Cloizeaux and Pearson boundaries. To the best of our knowledge, this result is a first direct and explicit exact theoretical confirmation that the total $S$ for the infinite chain tails outside the spin-wave continuum. Next in this section is a description of the behavior of $S_{4}$ as a function of $\omega$ for fixed values of $k$ followed by a comparison with the corresponding behavior of $S_{2}$. It is seen that there is consistency with the four-spinon continuum and that the overall shape (not the details) of $S_{4}$ is similar to the one of $S_{2}$, each in its own respective continuum. Last is carried a description of the behavior of $S_{4}$ as a function of $k$ for fixed values of $\omega$ and a comparison with the corresponding one of $S_{2}$. Here too similarity between the two overall shapes is found as well as consistency with the four-spinon continuum. Section four includes concluding remarks and indicates few directions in which this work can be carried forward. This paper is a continuation of the work [@ABS]. We use the same notation except for a slight modification in (\[FS4\]) where here we introduce the function $h$ instead of the function $f$ with the relation $f\equiv \exp \left( -h\right)$. Four-spinon dynamic structure function ====================================== The antiferromagnetic spin-$\frac{1}{2}$ $XXX$ Heisenberg chain is defined as the isotropic limit of the $XXZ$ anisotropic Heisenberg Hamiltonian: $$H=-\frac{1}{2}\hspace{-3pt}\sum_{n=-\infty }^{\infty }\left( \sigma _{n}^{x}\sigma _{n+1}^{x}\hspace{-3pt}+\sigma _{n}^{y}\sigma _{n+1}^{y} \hspace{-3pt}+\Delta \sigma _{n}^{z}\sigma _{n+1}^{z}\right) \,. \label{hamiltonian}$$ $\Delta =(q+q^{-1})/2$ is the anisotropy parameter and the isotropic antiferromagnetic limit is obtained as $\Delta \rightarrow -1^{-}$, or equivalently $q\rightarrow -1^{-}$. Here $\sigma _{n}^{x,y,z}$ are the usual Pauli matrices acting at the site $n$ of the chain. The exact diagonalization of this Hamiltonian is performed directly in the thermodynamic limit. This is necessary if we want to exploit the $\mathrm{U}_{q}(\widehat{\mathrm{sl}}_{2})$ quantum group symmetry present in the model [@JimMi]. One consequence is the appearance of two vacuum states $|0\rangle _{i}$, $i=0,1$ due to two different boundary conditions on the infinite chain. The Hilbert space $\mathcal{F}$ consists of $n$-spinon energy eigenstates $|\xi _{1},...,\xi _{n}\rangle _{\epsilon _{1},...,\epsilon _{n};\,i}$ such that: $$H|\xi _{1},...,\xi _{n}\rangle _{\epsilon _{1},...,\epsilon _{n};\,i}=\sum_{j=1}^{n}e(\xi _{j})|\xi _{1},...,\xi _{n}\rangle _{\epsilon _{1},...,\epsilon _{n};\,i\;}, \label{energy-eigenstates}$$ where $e(\xi _j)$ is the energy of spinon $j$ and $\xi _{j}$ is a spectral parameter living on the unit circle. In the above relation, $\epsilon _{j}=\pm 1$. The translation operator $T$ which shifts the spin chain by one site acts on the energy eigenstates in the following manner: $$T|\xi _{1},...,\xi _{n}\rangle _{\epsilon _{1},...,\epsilon _{n};\,i}=\prod_{i=1}^{n}\tau (\xi _{i})|\xi _{1},...,\xi _{n}\rangle _{\epsilon _{1},...,\epsilon _{n};1-i\;}, \label{action-of-T}$$ where $\tau (\xi _{j})=e^{-ip(\xi _{j})}$ and $p(\xi _{j})$ is the lattice momentum of spinon $j$. The exact expressions of the spinon energy and lattice momentum in terms of the spectral parameter are known in the literature [@Fadeev-Takhtajan; @JimMi; @ABS]. We are interested in their $XXX$ limit and it is given below in eq (\[dispersion-relation\]). The completeness relation in $\mathcal{F}$ reads: $$\mathbf{I}=\sum_{i=0,1}\sum_{n\geq 0}\sum_{\{\epsilon _{j}=\pm 1\}_{j=1,n}} \frac{1}{n!}\oint \prod_{j=1}^{n}\frac{d\xi _{j}}{2\pi i\xi _{j}}\;|\xi _{1},...,\xi _{n}\rangle _{\epsilon _{1},...,\epsilon _{n};\,i\;i;\,\epsilon _{1},...,\epsilon _{n}}\langle \xi _{1},...,\xi _{n}|\,. \label{completeness-relation}$$ The two-point dynamic structure function is defined as the Fourier transform of the zero-temperature vacuum-to-vacuum two-point function. The transverse DSF is therefore defined by: $$S^{i,+-}(\omega ,k)=\int_{-\infty }^{\infty }dt\sum_{m\in \mathbb{Z}} e^{i(\omega t+km)}\,_{i}\langle 0|\sigma _{m}^{+}(t)\,\sigma _{0}^{-}(0)|0\rangle _{i}\;, \label{definition-of-S}$$ where $\omega$ and $k$ are the neutron energy and momentum transfer respectively, and $\sigma ^{\pm }$ denotes $(\sigma ^{x}\pm i\sigma ^{y})/2$. The DSF satisfies the following relations: $$S^{i,+-}(\omega ,k)=S^{i,+-}(\omega ,-k)=S^{i,+-}(\omega ,k+2\pi )\,, \label{symmetries-of-DSF}$$ expressing reflection symmetry and periodicity. Inserting the completeness relation (\[completeness-relation\]) and using the Heisenberg relation: $$\sigma _{m}^{x,y,z}(t)=\exp \left( iHt\right) \,T^{-m}\sigma _{0}^{x,y,z}(0)\,T^{m}\,\exp \left( -iHt\right) \,, \label{heisenberg-relation}$$ we can write the transverse DSF as the sum of $n$-spinon contributions: $$S^{i,+-}(\omega ,k)=\sum_{n\,\,\mathrm{even}}S_{n}^{i,+-}(\omega ,k)\;, \label{sum-over-n-spinons}$$ where the $n$-spinon DSF $S_{n}$ is given by: $$\begin{aligned} S_{n}^{i,+-}(\omega ,k) &=&\frac{2\pi }{n!}\sum_{m\in \mathbb{Z}} \,\sum_{\epsilon _{1},...,\epsilon _{n}}\oint \prod_{j=1}^{n}\frac{d\xi _{j}} {2\pi i\xi _{j}}\,e^{im(k+\sum_{j=1}^{n}p_{j})}\,\delta \left( \omega -\sum_{j=1}^{n}e_{j}\right) \, \nonumber \\ &&\times \;X_{\epsilon _{n},...,\epsilon _{1}}^{i+m}(\xi _{n},...,\xi _{1})\;X_{\epsilon _{1},...,\epsilon _{n}}^{1-i}(-q\xi _{1},...,-q\xi _{n})\;, \label{expression-of-Sn}\end{aligned}$$ a relation in which $X^{i}$ denotes the form factor: $$X_{\epsilon _{1},...,\epsilon _{n}}^{i}(\xi _{1},...,\xi _{n})\equiv \,_{i}\langle 0|\sigma _{0}^{+}\left( 0\right) |\xi _{1},...,\xi _{n}\rangle _{\epsilon _{1},...,\epsilon _{n};\,i}\;. \label{form-factor}$$ In relation (\[expression-of-Sn\]), $i+m$ is to be read modulo 2. Note that each $S_{n}$ satisfies the symmetry relations (\[symmetries-of-DSF\]). An exact expression for the form factor $X^{i}$ is known [@smirnov; @JimMi]. To arrive at the result, one has to exploit extensively the infinite-dimensional representation of $\mathrm{U}_{q}(\widehat{\mathrm{sl}}_{2})$ and bosonize the relevant vertex operators in order to be able to manipulate in a systematic way traces of these operators which ultimately yield the correlation functions. Using this form factor, it is possible to give an exact expression for the $n$-spinon DSF in the anisotropic case [@ABS] and determine its isotropic limit [@Boug], obtained via the replacement [@JimMi; @ABS]: $$\xi =ie^{-2i\varepsilon \rho }\,;\qquad q=-e^{-\varepsilon }\,,\qquad \varepsilon \rightarrow 0^{+}\,, \label{Isotropic-limit}$$ where $\rho$ becomes the spectral parameter suited for this limit. The expressions of the energy $e$ and momentum $p$ in terms of $\rho$ then read: $$e(\rho )=\frac{\pi }{\cosh (2\pi \rho )}=-\pi \sin \,p\;;\quad \cot \,p=\sinh (2\pi \rho )\;;\quad -\pi \leq p\leq 0\;. \label{dispersion-relation}$$ It turns out that the transverse two-spinon DSF $S_{2}$ does not involve a contour integration, see (\[expression-of-Sn\]). Its exact expression has been derived in [@BCK]. It reads: $$S_{2}^{+-}(\omega ,k-\pi )=\frac{1}{4}\frac{e^{-I(\rho )}}{\sqrt{\omega _{2u}^{2}-\omega ^{2}}}\,\Theta (\omega -\omega _{2l})\,\Theta (\omega _{2u}-\omega )\;, \label{S2exact}$$ where $\Theta$ is the Heaviside step function and the function $I(\rho )$ is given by: $$I(\rho )=\int_{0}^{+\infty }\frac{dt}{t}\frac{\cosh (2t)\,\cos (4\rho t)-1} {\,\sinh (2t)\cosh (t)}\,e^{t}\,. \label{I-de-rho}$$ $\omega _{2u(l)}$ is the upper (lower) bound of the two-spinon excitation energies called the des Cloizeaux and Pearson (dCP) [@Muller; @BCK] upper (lower) bound or limit. They read: $$\omega _{2u}=2\pi \sin \frac{k}{2};\qquad \quad \omega _{2l}=\pi \,|\sin \,k|\,. \label{dCP}$$ The quantity $\rho $ is related to $\omega$ and $k$ by the relation: $$\cosh \,\pi \rho =\sqrt{\frac{\omega _{2u}^{2}-\omega _{2l}^{2}}{\omega ^{2}-\omega _{2l}^{2}}}\,, \label{relation-rho-omega-k}$$ which is obtained using eq (\[dispersion-relation\]) and the energy-momentum conservation laws. The properties of $S_{2}$ have been discussed in [@KMB; @BKM] where a comparison with the Müller ansatz [@Muller] is carried. The four-spinon DSF $S_{4}$ involves only one contour integration and its expression is given in [@ABS]. For $0\leq k\leq \pi $ it reads: $$S_{4}^{+-}(\omega ,k-\pi )=C_{4}\int_{-\pi }^{0}dp_{3}\int_{-\pi }^{0}dp_{4}\,F(\rho _{1},...,\rho _{4})\,. \label{S4}$$ For other values of $k$, it extends by symmetry using (\[symmetries-of-DSF\]). $C_{4}$ is a numerical constant irrelevant for the present work since we will scale $S_4$ to unity, and the integrand $F$ is given by: $$F(\rho _{1},...,\rho _{4})=\sum_{(p_{1},\,p_{2})}\frac{\exp [-h(\rho _{1},...,\rho _{4})]\,\sum_{\ell =1}^{4}|g_{\ell }(\rho _{1},...,\rho _{4})|^{2}}{\sqrt{W_{u}^{2}-W^{2}}}\,. \label{FS4}$$ The different quantities involved in this expression are as follows. $$\begin{aligned} W &=&\omega +\pi \,(\sin \,p_{3}+\sin \,p_{4}\,)\,; \nonumber \\ W_{u} &=&2\pi \left\vert \sin \,\left( K/2\right) \right\vert \,,\qquad K=k+p_{3}+p_{4}\,; \nonumber \\ \cot \,p_{j} &=&\sinh (2\pi \rho _{j})\,,\qquad -\pi \leq p_{j}\leq 0\,. \label{arguments-of-F}\end{aligned}$$ The function $h$ is given by: $$h(\rho _{1},...,\rho _{4})=\sum_{1\leq j<j^{^{\prime }}\leq 4}I\,(\rho _{jj^{\prime }})\,, \label{f}$$ where $\rho _{jj^{\prime }}=\rho _{j}-\rho _{j^{\prime }}$ and the function $g_{\ell }$ reads: $$\begin{aligned} g_{\ell } &=&(-1)^{\ell +1}(2\pi )^{4}\sum_{j=1}^{4}\cosh \,\,(2\pi \rho _{j}) \nonumber \\ &&\times \sum_{m=\Theta (j-\ell )}^{\infty }\frac{\prod_{i\neq \ell }(m- \frac{1}{2}\Theta (\ell -i)+i\rho _{ji})}{\prod_{i\neq j}\pi ^{-1}\sinh (\pi \rho _{ji})}\prod_{i=1}^{4}\frac{\Gamma (m-\frac{1}{2}+i\rho _{ji})}{\Gamma (m+1+i\rho _{ji})}\,, \label{gl}\end{aligned}$$ where $\Theta$ is the Heaviside step function. In (\[FS4\]), the sum $\sum_{(p_{1},p_{2})}$ is over the two pairs $(p_{1},p_{2})$ and $(p_{2},p_{1})$ solutions of the energy-momentum conservation laws: $$W=-\pi (\sin \,p_{1}+\sin \,p_{2})\,;\qquad \quad K=-p_{1}-p_{2}\,. \label{conservation-laws}$$ They read: $$(p_{1},p_{2})=\left( -K/2+\arccos \left( W/\left[ 2\pi \sin \left( K/2\right) \right] \right) \,,\;-K/2-\arccos \left( W/\left[ 2\pi \sin \left( K/2\right) \right] \right) \right) . \label{solution-conservation-laws}$$ Note that the solution in (\[solution-conservation-laws\]) is allowed as long as $W_{l}\leq W\leq W_{u}$ where $W_{u}$ is given in (\[arguments-of-F\]) and: $$W_{l}=\pi |\sin \,K|\,. \label{Wl}$$ The (analytic) behavior of the function $F$ in (\[FS4\]) is discussed in [@ABS]. It is shown that the series $g_{\ell }$ is convergent. It is also shown that $g_{\ell }$ stays finite when two $\rho _{i}$’s or more get equal. Since the function $\exp \left( -h\right) $ goes to zero in these regions [@KMB], the integrand $F$ of $S_{4}$ is regular there. Furthermore, it is shown that $F$ is exponentially convergent when one of the $\rho _{i}$’s goes to infinity, which means the two integrals over $p_{3} $ and $p_{4}$ in (\[S4\]) do not yield infinities. All these analytic results help secure safe numerical manipulations. Behavior of exact four-spinon DSF ================================= From now on, we restrict ourselves to the interval $0\leq k\leq \pi $. All forthcoming results can be carried to the other intervals of $k$ using the symmetry relations (\[symmetries-of-DSF\]). Also, we scale both $S_{4}$ and $S_{2}$ to appropriate units in order to display conveniently their respective behaviors. Four-spinon continuum --------------------- The first feature we discuss is the ‘four-spinon continuum’, by analogy with the two-spinon (or the spin-wave) continuum. It is the extent of the region in the $(k,\omega )$-plane outside which $S_{4}$ is identically zero. Remember that from (\[S2exact\]), $S_{2}$ is confined to the region $\omega _{2l}(k)\leq \omega \leq \omega _{2u}(k)$, where $\omega _{2l,u}(k)$ are the dCP boundaries given in (\[dCP\]). From the condition $W_{l}\leq W\leq W_{u}$ mentioned after (\[solution-conservation-laws\]), we deduce that in order for $S_{4}$ to be nonzero identically, we must have $\omega _{4l}(k)\leq \omega \leq \omega _{4u}(k)$, where: $$\begin{aligned} \omega _{4l}(k) &=&3\pi \sin (k/3)\quad \mathrm{for}\quad 0\leq k\leq \pi /2\,; \nonumber \\ \omega _{4l}(k) &=&3\pi \sin (k/3+2\pi /3)\quad \mathrm{for}\quad \pi /2\leq k\leq \pi \,; \nonumber \\ \omega _{4u}(k) &=&4\pi \cos (k/4)\quad \mathrm{for}\quad 0\leq k\leq \pi \,. \label{dCP4}\end{aligned}$$ We see that $\omega _{4l}(k)$ and $\omega _{4u}(k)$ are a sort of four-spinon dCP boundaries for $S_{4}$. The two and four-spinon continua are drawn in figure 1 below. We immediately notice that the four-spinon continuum is not restricted to the region between the two-spinon dCP branches, which means, a fortiori, that the full $S$ is also not confined to the spin-wave continuum. This is a direct and explicit theoretical confirmation of the ‘tail’ of the dynamic structure function observed outside the spin-wave continuum in finite-chain numerical calculations [@Muller] and the phenomenology [@Exp]. Due to the six and higher spinon contributions, it is actually legitimate to expect the full DSF to tail even further outside the four-spinon continuum, with arguably much smaller values. For example, in the interval $0\leq k/\pi <0.51741$, there is a narrow region between $\omega _{2u}$ and $\omega _{4l}$ inside which both $S_{2}$ and $S_{4}$ are identically zero whereas the total $S$ may have (very small) nonzero values, something that could eventually be checked in finite-chain calculations. However, it is not possible to estimate exactly the continua corresponding to the $S_{n>4}$ without manageable explicit formulae for these. But what is already clear from our study is the fact that indeed, the spin-wave continuum is not restrictive to the total dynamic structure function. There are further general features we can also read from figure 1 without having recourse to detailed calculations. For example, we see that inside the interval $0\leq k/\pi <\allowbreak 0.51741$, the four-spinon continuum lies entirely above the two-spinon continuum. This means that for this interval, $S_{2}$ may be accepted as a good approximation for the total $S$ between the spin-wave boundaries and $S_{4}$ between $\omega _{4l}$ and $\omega _{4u}$. For $0.51741\leq k/\pi \leq 1$ however and as $k$ increases, there is increasing overlap between the two continua so that inside the spin-wave continuum, we may expect $S$ to divert in perhaps a non-negligible way from $S_{2}$. Behavior as a function of the energy transfer --------------------------------------------- Next we describe the behavior of $S_{4}$ as a function of $\omega $ for fixed values of $k$. Figure 2 displays the shapes of $S_{4}$ for respectively $k/\pi =1/4,1/2,3/4$ and $1$. As an illustration, take for example the case $k/\pi =1/4$. We see on figure 2 that $S_{4}$ is zero until $\omega /\pi \simeq 0.68$, which corresponds to the beginning of the four-spinon continuum $\omega _{4l}(\pi /4)/\pi =3\sin (\pi /12)= 0.77646$. We obtain a slightly smaller value when reading directly from figure 2 because of fitting. Note that it is important to check each time consistency with the four-spinon continuum. This is because the lower and upper branches $\omega _{4l,u}(k)$ are not imposed explicitly in the integration algorithm of $S_{4}$ unlike the case of $S_{2}$ where the corresponding expression (\[S2exact\]) incorporates explicitly the two-spinon dCP boundaries $\omega _{2l,u}(k)$. For $S_{4}$, only the conditions $W_{l}\leq W\leq W_{u}$ noted just before (\[Wl\]) are imposed. Starting from $\omega /4\pi \simeq 0.17$, $S_{4}$ jumps sharply from zero to a maximum at $\omega /4\pi \simeq 0.21$ (read from figure 2). Then it decreases with two apparent local minima until it becomes negligible at roughly $\omega /4\pi \simeq 0.6$. The upper branch of the four-spinon continuum at $k=\pi /4$ is $\omega _{4u}(\pi /4)/4\pi =\cos (\pi /16)=\allowbreak 0.9808$. Hence, $S_{4}$ is practically negligible (but not identically zero) for values of $\omega /4\pi $ between about $0.6$ and $0.9808$. The description of $S_{4}$ as a function of $\omega$ for the other values of $k$ can be carried along the same lines and each time, consistency with the four-spinon continuum is checked. The shapes are all similar to one another: steep increase from zero to a maximum value, followed by a ‘wiggled’ slower decrease to zero. Furthermore, it is interesting to notice that for all values of $k$, the overall shape of $S_{4}$, *not* the detail, is roughly similar to the one of $S_{2}$, represented as a function of $\omega$ for the same values of $k$ in figure 3. For instance, for $k/\pi =1/4$, $S_{2}$ starts very sharply at about $\omega /2\pi \simeq 0.35$, which corresponds to the start of the two-spinon continuum at $\omega _{2l}(\pi /4)/2\pi = 0.35355$. It reaches a maximum before decreasing more slowly towards zero. The only difference worth mentioning is the decrease of $S_{4}$ after its absolute maximum which presents local minima and maxima whereas the decrease of $S_{2}$ is always ‘smooth’. This may originate from the ‘richer’ structure of the expression of $S_{4}$ with respect to that of $S_{2}$. Behavior as a function of the momentum transfer ----------------------------------------------- Last we describe the behavior of $S_{4}$ as a function of $k$ for fixed values of $\omega$. In figure 4 are plotted the graphs of $S_{4}$ in terms of $k$ for $\omega /\pi =1/4$, $1/2$, $3/4$, $1$ respectively. Let us describe for example the case $\omega /\pi =1/4$. We see on figure 4 that $S_{4}$ is zero until we reach the value $k/\pi \simeq 0.88$. On the other hand, the four-spinon continuum lies outside the interval $0.07967\leq k/\pi \leq 0.92033$. In the region $0\leq k/\pi \leq 0.0796\,7$, figure 4 shows no discernible finite values for $S_{4}$, only a very thin ‘trace’ that would be more visible with a better resolution. This means that $S_{4}$ is negligible for those small values of $k$. For larger values of $\omega$ though, $S_{4}$ picks up clear finite values in the interval $0\leq k\leq 3\arcsin (\omega /3\pi )$; see the other graphs on figure 4. Those values get larger as $\omega$ increases. Returning to the case $\omega /\pi =1/4$, we see that $S_{4}$ starts from zero at $k/\pi \simeq 0.88$ (slightly smaller than the exact value $0.92033$ because of fitting), rises sharply to a maximum and then decreases. The behavior of $S_{4}$ for the other values of $\omega$ is also consistent with the four-spinon continuum. Take for example the case $\omega /\pi =1/2$. The four-spinon continuum indicates that $S_{4}$ is identically zero for $0.1599\leq k/\pi \leq 0.84008$. We see indeed small values for $S_{4}$ from $k/\pi =0$ up to a little before 0.2, and $S_{4}$ rising again from zero a little after $k/\pi =0.8$ to a local maximum, then to an absolute maximum before decreasing. Here too $S_{2}$ has a similar overall behavior within its own (two-spinon) continuum. Figure 5 displays the behavior of $S_{2}$ with respect to $k$ for the same fixed values of $\omega $. But as $\omega$ increases, we notice the richer structure of $S_{4}$ with respect to the corresponding one for $S_{2}$. This is due to the more involved expression of $S_{4}$. In fact, for larger values of $\omega$, the integrand $F$ in (\[S4\]) is nonvanishing in larger and larger areas in the $\left( p_{3},p_{4}\right)$ plane. For illustration, compare the behavior of $F$ shown in figure 6 for $\left( k,\omega \right) =\left( 0.5\pi ,2\pi \right)$ with the one shown in figure 7 for $\left( k,\omega \right) =\left( 0.5\pi ,3\pi \right)$. In this regard, we recall that the quadrature-based algorithms written in [@ABSS] were based on this observation and took advantage of the fact that for fairly small values of $\omega$, the integrand $F$ was negligible in large areas of the $\left( p_{3},p_{4}\right)$ plane. This is the reason why those algorithms could not be carried to larger values of $\omega$, or be able to describe efficiently a behavior as a function of the neutron energy $\omega$ itself. As already mentioned, we have used here Monte Carlo techniques and we can do better, but consistency with [@ABSS] is realized, i.e., we have the same behavior of $S_{4}$ as a function of $k$ for the same small values of $\omega$. Conclusion ========== In this work, we have described the behavior of the exact four-spinon dynamic structure function $S_{4}$ in the antiferromagnetic isotropic Heisenberg quantum spin chain at zero temperature as a function of the neutron energy $\omega$ and momentum transfer $k$. We have also determined the four spinon continuum, the region outside which $S_{4}$ is identically zero. The discussion was carried in the form of a comparison with the corresponding behavior of the exact two-spinon dynamic structure function $S_{2}$, already known in the literature. Figures 8 and 9 summarize these two behaviors where $S_{4}$ (8) and $S_{2}$ (9) are drawn in the $\left( k,\omega \right) $ plane. Recall that we have scaled them down to one. There are four directions in which one may wish to carry forward with this work. The first is the anisotropic case. The model is exactly soluble and we do have generic expressions for $S_{n}$ in the form of contour integrals in the spectral parameters complex planes [@Boug]. The difficulty here is that the integrands involve much more complicated functions which are already present in $S_{2}$, and one should expect intricate complexities in this more general case. The second direction in which one may want to push forward is the situation where there is an external magnetic field. There are finite-chain calculations in this regard, [@Muller] and more recently the works [@KM-KBM]. But one has to remember that the model in not exactly solvable in this case. So one may want to try small perturbations around the zero-field limit solution. The third direction is the finite-temperature case. Here too there are finite-chain results and it is interesting to see those effects on $S_{2}$ and $S_{4}$. The fourth direction is to look into the situation of a spin-one chain. The model is still exactly solvable and exploiting the quantum group symmetry, compact expressions for the form factors are available [@idzumi; @bougourzi-weston; @bougourzi-any-spin]. [99]{} K. Hirakawa and Y. Kurogi, Prog. Theor. Phys. **S46** (1970) 147. D.A. Tennant, R.A. Cowley, S.E. Nagler and A.M. Tsvelik, Phys. Rev. **B**52 (1995) 13368; D.A. Tennant, S.E. Nagler, D. Weltz, G. Shirane and K. Yamada, Phys. Rev. **B**52 (1995) 13381; D.A. Tennant, T.G. Perring, R.A. Cowley and S.E. Nagler, Phys. Rev. Lett. 70 (1993) 4003; S.E. Nagler, D.A. Tennant, R.A. Cowley, T.G. Perring and S.K. Satija, Phys. Rev. **B**44 (1991) 12361. G.L. Squires, ‘*Introduction to the theory of thermal neutron scattering’*, Cambridge University Press, 1996; S.W. Lovesey, ‘*Theory of neutron scattering from condensed matter’*, Clarendon, Oxford, 1987. Some (not *all*) of the literature related to the Heisenberg model comprises W. Heisenberg, Z. Phys. 49 (1928) 619; H. Bethe, Z. Phys. 71 (1931) 205; L. Hulthén, Arkiv. Mat. Astron. Fysik A 1126 (1938) 1; E.H. Lieb and D.C. Mattis, J. Math. Phys. 3 (1962) 749; J. des Cloizeaux and J.J. Pearson, Phys. Rev. 128 (1962) 2131; R.B. Griffiths, Phys. Rev. 133 (1964) A 768; C.N. Yang and C.P. Yang, Phys. Rev. 150 (1966) 321; 150 (1966) 327; 151 (1966) 258; Th. Niemeijer, Physica 36 (1967) 377; E. Barouch, B.M. McCoy and D.B. Abraham, Phys. Rev. **A**4 (1971) 2331; M. Gaudin, Phys. Rev. Lett. 26 (1971) 1301; M. Takahashi, Prog. Theor. Phys. 46 (1971) 401; L.A. Takhtajan and L.D. Faddeev, Russ. Math. Surveys 34 (1979) 11; B.M. McCoy, J.H.H. Perk and R.E. Shrock, Nucl. Phys. **B**220 (1983) 35; O. Babelon, H.J. de Vega and C.M. Viallet, Nucl. Phys. **B**220 (1983) 13; G. Müller and R.E. Shrock, Phys. Rev. **B**29 (1984) 288; J.M.R. Roldan, B.M. McCoy and J.H.H. Perk, Physica 136A (1986) 255; V.E. Korepin, A.G. Izergin and N.M. Bogoliubov, ‘*The Quantum Inverse Scattering Method and Correlation Functions*’, Cambridge University Press, 1993; F.H.L. Essler, H. Frahm, A.G. Izergin and V.E. Korepin, Comm. Math. Phys. 174 (1994) 191; V.E. Korepin, A.G. Izergin, F.H.L. Essler and D. Uglov, Phys. Lett. **A**190 (1994) 182. R.J. Baxter, ‘*Exactly Solved Models in Statistical Mechanics*’, Academic Press, 1982. M. Jimbo and T. Miwa, ‘*Algebraic Analysis of Solvable Lattice Models*’, American Mathematical Society, 1994. O. Davies, O. Foda, M. Jimbo, T. Miwa and A. Nakayashiki, Comm. Math. Phys. 151 (1993) 89; I.B. Frenkel and N.H. Jing, Proc. Natl. Acad. Sci. 85 (1988) 9373; A. Abada, A.H. Bougourzi and M.A. El Gradechi, Mod. Phys. Lett. A8 (1993) 715; A.H. Bougourzi, Nucl. Phys. **B**404 (1993) 457; A.H. Bougourzi, ‘*Bosonization of quantum affine groups and its application to the higher spin Heisenberg model’*, `q-alg/9706015`. P.W. Anderson, Phys. Rev. 86 (1952) 694. M.T. Hutchings, G. Shirane, R.J. Birgeneau and S.L. Holt, Phys. Rev. **B**5, (1972) 1999. G. Müller, H. Thomas, H. Beck and J.C. Bonner, Phys. Rev. **B**24 (1981) 1429. A.H. Bougourzi, M. Couture and M. Kacir, Phys. Rev. **B**54 (1996) 12669. M. Karbach, G. Müller and A.H. Bougourzi, ‘*Two-spinon dynamic structure factor of the one-dimensional $S=1/2$ Heisenberg antiferromagnet*’, `cond-mat/9606068`. A.H. Bougourzi, M. Karbach and G. Müller, ‘*Exact two-spinon dynamic structure factor of the one-dimensional* $s=1/2$ *Heisenberg-Ising antiferromagnet*’, `cond-mat/9712101`. A. Abada, A.H. Bougourzi and B. Si-lakhal, Nucl. Phys. **B**497 \[FS\] (1997) 733. A.H. Bougourzi, Mod. Phys. Lett **B**10 (1996) 1237. A. Abada, A.H. Bougourzi, B. Si-Lakhal and S. Seba, ‘*Four-Spinon Dynamical Correlation Function in Isotropic Heisenberg Model’*, `cond-mat/9802271`, unpublished. L.D. Faddeev and L.A. Takhtajan, J. Soviet. Math. 24 (1984) 241. F.A. Smirnov, ‘*Form Factors in Completely Integrable Models of Quantum Field Theory*’ World Scientific, Singapore, 1992. M. Karbach, D. Biegel and G. Müller, ‘*Quasiparticles governing the zero-temperature dynamics of the 1D spin-1/2 Heisenberg antiferromagnet in a magnetic field*’, `cond-mat/0205142`; M. Karbach and G. Müller, ‘*Line shape predictions via Bethe ansatz for the one dimensional spin-1/2 Heisenberg antiferromagnet in a magnetic field*’, `cond-mat/0005174`. M. Idzumi, Int. J. Mod. Phys. **A**9 (1994) 4449; ‘*Correlation functions of the spin 1 analog of the XXZ model*’, `hep-th/9307129`. A.H. Bougourzi and R.A. Weston, Nucl. Phys. **B**417 (1994) 439. A.H. Bougourzi, ‘*Bosonization of quantum affine groups and its application to higher spin Heisenberg model*’, `q-alg/9706015`.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We consider quantum entanglement of three accelerating qubits, each of which is locally coupled with a real scalar field, without causal influence among the qubits or among the fields. The initial states are assumed to be the GHZ and W states, which are the two representative three-partite entangled states. For each initial state, we study how various kinds of entanglement depend on the accelerations of the three qubits. All kinds of entanglement eventually suddenly die if at least two of three qubits have large enough accelerations. This result implies the eventual sudden death of all kinds of entanglement among three particles coupled with scalar fields when they are sufficiently close to the horizon of a black hole.' author: - Yue Dai - Zhejun Shen - Yu Shi title: Quantum entanglement in three accelerating qubits coupled to scalar fields --- published as Phys. Rev. D [**94**]{}, 025012 (2016). Introduction ============ How quantum states are affected by gravity or acceleration is a subject of longstanding interest [@mukhanov]. In the presence of a black hole, the physical vacuum is that in the Kruskal-Szekers coordinates, which are nonsingular and cover the whole Schwarzschild spacetime, while a remote observer can observe particles. Known as Hawking radiation, it underlies the paradox of information loss [@hawking]. Analogously, an accelerating object coupled with a field detects a thermal bath of particles of this field, even though this field is in the Minkowski vacuum [@davies; @fulling; @dewitt; @unruh]: this is known as the Unruh effect. For a composite quantum system, the characteristic quantum feature is quantum entanglement. An interesting question is how quantum entanglement among objects coupled with fields is affected by the Unruh effect. Various investigations were made on two entangled detectors, one or both of which accelerate [@matsas1; @doukas1; @doukas2; @huyu; @hu0; @dai]. Naturally, one may wonder about the situation of three entangled field-coupled qubits. This is interesting and nontrivial because there are various different types of entanglement among three qubits $A$, $B$, and $C$. There is bipartite entanglement between two qubits, and there is bipartite entanglement between one qubit and the remaining two qubits as one party. Most interestingly, there is tripartite entanglement among all three qubits, which cannot be reduced to any combination of all kinds of bipartite entanglement [@bennett]. This is profound and important in understanding many-body correlations. In quantum information theory, recent years witnessed much development in quantifying entanglement in terms of some measures. A convenient one is the so-called negativity, which is the sum of the negative eigenvalues of the partial transpose of the density matrix, ranging from $0$ to $1/2$ [@vidal]. It can be used to quantify various kinds of bipartite entanglement in a three-qubit state. Twice the negativity that quantifies the bipartite entanglement between a qubit and the remaining two qubits is called a one-tangle, ranging from $0$ to $1$. Twice the negativity quantifies the bipartite entanglement between two qubits is called a two-tangle, ranging from $0$ to $1$. Using all the one-tangles and two-tangles, one can define a measure of tripartite entanglement called a three-tangle, ranging from $0$ to $1$ [@ou]. In the present paper, by using these negativity-based entanglement measures, we make detailed investigations on how various types of entanglement in three field-coupled qubits vary with their accelerations, which have implications on particles near the horizon of a black hole. It is well known that there are two inequivalent types of tripartite entanglement [@duer], typified, respectively, by the GHZ state $$|\mathrm{GHZ}\rangle = \frac{1}{\sqrt{2}}(|000\rangle + |111\rangle).$$ and the W state $$|\mathrm{W}\rangle=\frac{1}{\sqrt{3}}(|001\rangle + |010\rangle+|100\rangle).$$ So both the GHZ and W states are considered in this paper. Method ====== We consider three qubits $A$, $B$, and $C$ far away from each other. For simplicity, it is assumed that each qubit $q$ ($q=A,B,C$) is coupled only locally with the field $\Phi_q$ around it, as described by the Unruh-Wald model [@unruhwald]. The Hamiltonian of each qubit $q$ itself is $H_q=\Omega_q Q_q^\dagger Q_q$, where the creation operator $Q_q^\dagger$ and annihilation operator $Q_q$ are defined by $Q_q | 0 \rangle_q = Q_q^\dagger | 1 \rangle_q = 0$, $Q_q^\dagger | 0 \rangle_q = | 1 \rangle_q$, and $Q_q | 1 \rangle_q = | 0\rangle_q$, and $\Omega_q$ is the energy difference between the two eigenstates. Its coupling with the field is described by $H_{I_q}(t_q) = \epsilon_q ( t_q ) \int_{\Sigma_q} {\Phi_q ( \mathbf{x} ) [ {\psi_q ( \mathbf{ x} )Q_q + {\psi_q^*} ( \mathbf{x} ){Q_q^\dag }} ]\sqrt { - g} {d^3}x}$, where $\mathbf{x}$ and $t_q$ are spacetime coordinates in the comoving frame of the qubit, the integral is over the spacelike Cauchy surface $\Sigma_q$ at the given time $t_q$, $\epsilon_q ( t_q )$ is the coupling constant with a finite duration, and $\psi_q (\mathbf{ x} )$ is a smooth nonvanishing function within a small region around the qubit. The total Hamiltonian of the system is thus $$\sum_{q=A,B,C} (H_q+H_{\Phi_q}+ H_{I_q}),$$ where $H_{\Phi_q}$ is the Klein-Gordon Hamiltonian for $\Phi_q$. It is assumed that the qubits are far away from each other such that during the interaction times, there is no physical coupling or influence between the fields around different qubits or between a qubit and the field around another qubit. We make this assumption to avoid the issue of a global time slice and those complications caused by the different accelerations of the qubits. Consequently, one can describe the quantum state of the qubits by considering each qubit in its comoving reference frame. The interesting case where all qubits are coupled with the same field will be explored in a future work. Therefore, in our consideration, after a time duration longer than the interacting times $T_q \gg 1/\Omega_q$, the state of the whole system in the interaction picture is transformed by $$U_{A}\otimes U_{B}\otimes U_{C},$$ where $U_{q}$ is the unitary transformation acting on qubit $q$ and the field $\Phi_q$ in its neighboring region. It can be obtained that [@dai] $$U_{q} \approx 1+iQ_{q} a^\dagger (\Gamma_q^*) - i Q_{q}^\dagger a(\Gamma_q^*),$$ where $a(\Gamma_q^*)$ and $a^\dagger(\Gamma_q^*)$ are the annihilation and creation operators of $\Gamma_q^*$, with $$\Gamma_q (x) \equiv -2i \int [G_{Rq}(x;x')-G_{Aq}(x;x')]\epsilon_q(t')e^{i\Omega_q t}\psi_q^*(\mathbf{x}')\sqrt{-g'} d^4x',$$ where $G_{Rq}$ and $G_{Aq}$ are the retarded and advanced Green functions of the field $\Phi_q$ [@unruhwald]. For each field $\Phi_q$, it has been argued that approximately the qubit $q$ is only coupled with the field mode $\Gamma_q^*$, with frequency $\Omega_q$, with the other modes decoupled [@unruhwald; @dai]. Consider the Fock state $|n\rangle_{\Gamma^*_q}$ containing $n$ particles in the mode $\Gamma^*_{q}$, as observed in the Rindler wedge confining qubit $q$. We have [@dai] $$\begin{aligned} U_{q} | 0 \rangle_q | n \rangle_{\Gamma^*_q}&= & | 0 \rangle_q | n \rangle_{\Gamma^*_q} - i\sqrt n \mu_q | 1 \rangle_q | {n - 1} \rangle_{\Gamma^*_q}, \label{s1} \\ U_{q} | 1 \rangle_q | n \rangle_{\Gamma^*_q} & = & | 1 \rangle_q | n \rangle_{\Gamma^*_q} + i\sqrt {n + 1} \mu_q^* | 0 \rangle_q | {n + 1} \rangle_{\Gamma^*_q}, \label{s2}\end{aligned}$$ where $\mu_q \equiv \langle \Gamma_q^*,\Gamma^*_q\rangle$. For an arbitrary mode $\chi$, $\langle \Gamma_q^*,\chi\rangle = \int \epsilon_q(t) e^{i\Omega_q t}\psi_q^*(\mathbf{x}) \chi(t,\mathbf{x})\sqrt{-g} d^4x$ [@unruhwald]. Suppose the initial state of the three qubits to be $|\Psi_i\rangle$. Without causal connection either between the qubits or between the fields, each qubit independently detects a thermal bath of the Unruh particles determined by its own acceleration. With each qubit in its own Rindler wedge, the initial state of the whole system, as observed by the observers comoving with the qubits respectively, is described by the density matrix $$\rho_i = \rho_{\Gamma^*_A}\otimes \rho_{\Gamma^*_B} \otimes \rho_{\Gamma^*_C}\otimes |\Psi_i\rangle\langle \Psi_i|,$$ where $\rho_{\Gamma^*_q}$ is the density matrix of the mode $\chi_{\Gamma^*_q}$ of the field around qubit $q$, and the decoupled modes are neglected. For a uniformly moving qubit $q$, $$\rho_{\Gamma^*_q}=|0\rangle_{\Gamma^*_q}\langle 0|,$$ because the uniformly moving qubit sees a Minkowski vacuum. For an accelerating qubit, $$\rho_{\Gamma^*_q}=\eta_q\sum_{n_q} e^{ - 2\pi n_q \Omega_q/a_q}|n_q\rangle_{\Gamma^*_q}\langle n_q|,$$ where $a_q$ is the acceleration of qubit $q$, and $n_q$ denotes the particle number of mode $\Omega_q$, $\eta_q \equiv \sqrt {1 - {e^{ - 2\pi \Omega_q/a_q}}}$. The final state of the system with respect to the comoving observers is $$\rho_f = U_{C}^\dagger U_{B}^\dagger U_{A}^\dagger \rho_i U_{A} U_{B}U_{C},$$ from which we obtain the reduced density matrix of the three qubits $$\rho_{ABC} = Tr_{{\Gamma^*_A},{\Gamma^*_B},{\Gamma^*_C}} (\rho_f).$$ Then we study the dependence of various types of entanglement on the accelerations. The one-tangle between qubit $\alpha$ and the remaining qubits $\beta$ and $\gamma$ is $${\cal N}_{\alpha(\beta\gamma)} \equiv \|\rho_{ABC}^{T_\alpha}\|-1,$$ where $T_\alpha$ represents a partial transpose with respect to $\alpha$, and $\|\rho\|\equiv \textrm{Tr}\sqrt{\rho\rho^\dagger}$ represents the trace norm of $\rho$. The two-tangle between qubits $\alpha $ and $\beta$ is $${\cal N}_{\alpha\beta} \equiv \|\rho_{\alpha\beta}^{T_A}\|-1,$$ where $\rho_{\alpha\beta}= \textrm{Tr}_\gamma \rho_{ABC}$ is the reduced density matrix of $\alpha$ and $\beta$. The three-tangle is $$\pi \equiv \frac{1}{3}\sum_{\alpha=A,B,C}\pi_\alpha,$$ where $$\pi_\alpha \equiv {\cal N}_{\alpha(\beta\gamma )}^2-{\cal N}_{\alpha\beta}^2- {\cal N}_{\alpha\gamma}^2. \label{pia}$$ Note that a monogamy relation $${\cal N}_{\alpha(\beta\gamma )}^2 \geq {\cal N}_{\alpha\beta}^2+ {\cal N}_{\alpha\gamma}^2 \label{mono}$$ is always valid, and is the basis for the definition (\[pia\]). In the following, for the GHZ and W states, we study various cases of the accelerations of the three qubits. Note the permutation symmetry of each of these two states. GHZ state ========= First we consider the initial state to be the GHZ state, $$\label{ints} |\Psi_i\rangle = |\mathrm{GHZ}\rangle.$$ In the GHZ state, tracing over one qubit always yields a disentangled two-qubit state. On the other hand, the coupling between each qubit and the field around it does not increase the interqubit entanglement. Therefore, each two-tangle always remains zero, $${\cal N}_{AB} = {\cal N}_{BC} = {\cal N}_{AC} = 0.$$ $C$ accelerating ---------------- Let us assume qubit $C$ accelerates while $A$ and $B$ move uniformly. In this case, the density matrix of the three qubits is obtained as $$\begin{split} \rho_{ABC} =&\eta_C^2\sum\limits_{{n_C}} {\frac{{{e^{ - 2\pi {n_C}{\Omega _C}/{a_C}}}}}{{{Z_{{n_C}}}}} } \left[ {\left( {1 + \left( {{n_C} + 1} \right)|\mu _A|^2|\mu _B|^2|\mu _C|^2} \right)\left| {000} \right\rangle \left\langle {000} \right|} \right.\\ & + \left| {000} \right\rangle \left\langle {111} \right| + \left| {111} \right\rangle \left\langle {000} \right| + \left| {111} \right\rangle \left\langle {111} \right| + \left( {{n_C} + 1} \right)|\mu _B|^2|\mu _C|^2\left| {100} \right\rangle \left\langle {100} \right|\\ & + \left( {{n_C} + 1} \right)|\mu _A|^2|\mu _C|^2\left| {010} \right\rangle \left\langle {010} \right| + \left( {{n_C}|\mu _C|^2 + |\mu _A|^2|\mu _B|^2} \right)\left| {001} \right\rangle \left\langle {001} \right|\\ & + \left( {{n_C} + 1} \right)|\mu _C|^2\left| {110} \right\rangle \left\langle {110} \right| + |\mu _B|^2\left| {101} \right\rangle \left\langle {101} \right|\left. { + |\mu _A|^2\left| {011} \right\rangle \left\langle {011} \right|} \right], \end{split}$$ where $$\begin{split} {Z_{{n_C}}} =& 2+ |\mu _A|^2 + |\mu _B|^2 + \left( {2{n_C} + 1} \right)|\mu _C|^2 + |\mu _A|^2|\mu _B|^2\\& +\left( {{n_C} + 1} \right)|\mu _A|^2|\mu _C|^2 + \left( {{n_C} + 1} \right)|\mu _B|^2|\mu _C|^2 + \left( {{n_C} + 1} \right)|\mu _A|^2|\mu _B|^2|\mu _C|^2. \end{split}$$ In the GHZ state, any qubit is maximally entangled with the other two qubits as a single party. Hence the one-tangles are $${\cal N}_{A(BC)}={\cal N}_{B(AC)} = {\cal N}_{C(AB)}=1.$$ When $a_C\neq 0$, the entanglement decreases. ${\cal N}_{A(BC)}$ decreases with the increase of the acceleration-frequency ratio (AFR) $a_C/\Omega_C$ until its sudden death, as shown in Fig. \[g1\]. This is the phenomenon of entanglement sudden death [@yu]. The result here on one-tangles extends the previous result of bipartite entanglement [@dai] from pure states to mixed states. However, ${\cal N}_{A(BC)}={\cal N}_{B(AC)}$ approaches zero asymptotically, presumably because for these two one-tangles, $C$ is only one of the two qubits constituting a party, with the other qubit moving uniformly. In the present case, since the two-tangles remain zero, the three-tangle turns out to be the average of the sum of the three one-tangles, as depicted in Fig. \[g2\]. Note that due to the sudden death of ${\cal N}_{A(BC)}$, there is a sudden change in the three-tangle, after which the three-tangle is just $2{\cal N}_{A(BC)}/3=2{\cal N}_{B(AC)}/3$. $B$ and $C$ accelerating ------------------------ In the case where $B$ and $C$ accelerate while $A$ moves uniformly, we obtain $$\begin{split} {\rho _{ABC}} =& \eta_B^2\eta_C^2\sum\limits_{{n_B},{n_C}} {\frac{{{e^{ - 2\pi \left( {n_B{\Omega _B}/{a_B} + {n_C}{\Omega _C}/{a_C}} \right)}}}}{{{Z_{{n_B},{n_C}}}}}} \left[ {\left| {000} \right\rangle \left\langle {111} \right| + \left| {111} \right\rangle \left\langle {000} \right|} \right.\\ &+ \left| {111} \right\rangle \left\langle {111} \right| + \left( {1 + \left( {{n_B} + 1} \right)\left( {{n_C} + 1} \right){{\left| {{\mu _A}} \right|}^2}{{\left| {{\mu _B}} \right|}^2}{{\left| {{\mu _C}} \right|}^2}} \right)\left| {000} \right\rangle \left\langle {000} \right|\\ &+ \left( {{n_B} + 1} \right)\left( {{n_C} + 1} \right){\left| {{\mu _B}} \right|^2}{\left| {{\mu _C}} \right|^2}\left| {100} \right\rangle \left\langle {100} \right| + \left( {{n_B} + 1} \right){\left| {{\mu _B}} \right|^2}\left| {101} \right\rangle \left\langle {101} \right|\\ &+ \left( {{n_C} + 1} \right){\left| {{\mu _C}} \right|^2}\left| {110} \right\rangle \left\langle {110} \right| + \left( {{n_B}{n_C}{{\left| {{\mu _B}} \right|}^2}{{\left| {{\mu _C}} \right|}^2} + {{\left| {{\mu _A}} \right|}^2}} \right)\left| {011} \right\rangle \left\langle {011} \right|\\ &+ \left( {{n_B}{{\left| {{\mu _B}} \right|}^2} + \left( {{n_C} + 1} \right){{\left| \mu _A \right|}^2}{{\left| {{\mu _C}} \right|}^2}} \right)\left| {010} \right\rangle \left\langle {010} \right|\\ &+ \left. {\left( {{n_C}{{\left| {{\mu _C}} \right|}^2} + \left( {{n_B} + 1} \right){{\left| \mu _A \right|}^2}{{\left| {{\mu _B}} \right|}^2}} \right)\left| {001} \right\rangle \left\langle {001} \right|} \right], \end{split}$$ where $$\begin{split} {Z_{{n_B},{n_C}}} =& 2 + |\mu _A|^2 + (2{n_B}+1)|\mu _B|^2 + (\frac{3}{2}{n_C}+1)|\mu _C|^2 \\ &+ \left( {{n_B} + 1} \right)|\mu _A|^2|\mu _B|^2 + \left[ n_B n_C + \left( {{n_B} + 1} \right)\left( {{n_C} + 1} \right)\right]|\mu _B|^2|\mu _C|^2 \\ &+\left( {{n_C} + 1} \right)|\mu _A|^2|\mu _C|^2 + \left( {{n_B} + 1} \right)\left( {{n_C} + 1} \right)|\mu _A|^2|\mu _B|^2|\mu _C|^2. \end{split}$$ The dependence of ${\cal N}_{A(BC)}$ on $a_B/\Omega_B$ and $a_C/\Omega_C$ is symmetric, as shown in Fig. \[g3\]. When one of the accelerations is zero, ${\cal N}_{A(BC)}$ decreases towards zero asymptotically, as discussed in the preceding subsection. When both are nonzero, ${\cal N}_{A(BC)}$ decreases quickly towards zero, reaching sudden death, at finite values of the two AFRs. \ As shown in Fig. \[g4\], ${\cal N}_{B(AC)}$ strongly depends on $a_B/\Omega_B$ and suddenly dies at a finite value of $a_B/\Omega_B$, while it depends on $a_C/\Omega_C$ weakly, especially when $a_B/\Omega_B$ is so large that ${\cal N}_{B(AC)}$ is close to zero. This is because $B$ is one party by itself, while $C$ is only one of the two qubits constituting the other party. ${\cal N}_{C(AB)}$ can be obtained from ${\cal N}_{B(CA)}$ by exchanging $B$ and $C$, as shown in Fig. \[g5\]. We now look at some two-dimensional (2D) cross sections of the three-dimensional (3D) plots in Fig. \[g3to5\]. Figure \[g6\] is for $a_B/\Omega_B=a_C/\Omega_C$, and hence ${\cal N}_{B(AC)}={\cal N}_{C(AB)}$. Among the three one-tangles, ${\cal N}_{A(BC)}$ is the smallest, presumably because $B$ and $C$ constituting the party ($BC$) both accelerate. The three one-tangles die at the same value of $a_B/\Omega_B=a_C/\Omega_C$. Figure \[g7\] is for $a_C/\Omega_C= 1.5a_B/\Omega_B$, while Fig. \[g8\] is for $a_C/\Omega_C=2a_B/\Omega_B$. The values of $a_B/\Omega_B$ for the sudden death of ${\cal N}_{B(AC)}$ in these two cases are close to that in Fig. \[g6\], as ${\cal N}_{B(AC)}$ is mainly determined by $a_B/\Omega_B$ when it is close to zero. On the other hand, ${\cal N}_{C(AB)}$ dies first, as $a_C/\Omega_C$ is larger than $a_B/\Omega_B$ in these cross sections, while $a_A/\Omega_A=0$. When $a_B/\Omega_B$ is less than the value at which ${\cal N}_{C(AB)}$ suddenly dies, ${\cal N}_{B(AC)} < {\cal N}_{A(BC)}$ as $a_B/\Omega_B > a_A/\Omega_a=0$. When ${\cal N}_{C(AB)}$ suddenly dies, ${\cal N}_{B(AC)} = {\cal N}_{A(BC)}$. When $a_B/\Omega_B$ is larger than the value for the death of ${\cal N}_{C(AB)}$, $ {\cal N}_{A(BC)} < {\cal N}_{B(AC)}$, until ${\cal N}_{B(AC)}$ dies at a larger value of $a_B/\Omega_B$. The above-mentioned feature in the case of $a_B/\Omega_B=a_C/\Omega_C$ that the three one-tangles die at the same value of $a_B/\Omega_B$ is a special case, because when it is constrained that ${\cal N}_{C(AB)}={\cal N}_{B(AC)}$, the sudden death of ${\cal N}_{C(AB)}$ implies that of ${\cal N}_{B(AC)}$, as the two are equal. On the other hand, when ${\cal N}_{C(AB)}$ suddenly dies, there must be ${\cal N}_{A(BC)}= {\cal N}_{B(AC)}$, and hence the three have to suddenly die altogether; in other words, the only option for ${\cal N}_{A(BC)}$ to be between these two is that it dies also. We have also examined several cases of given values of $a_B/\Omega_B$, as shown in Figs. \[g9\], \[g10\], and \[g11\], with $a_B/\Omega_B=300$, $750$, and $1200$ respectively. For $a_C/\Omega_C < a_B/\Omega_B$, ${\cal N}_{B(AC)}<{\cal N}_{C(AB)} < {\cal N}_{A(BC)}$. For $a_C/\Omega_C > a_B/\Omega_B$ up to the death of ${\cal N}_{C(AB)}$, ${\cal N}_{C(AB)} < {\cal N}_{B(AC)}<{\cal N}_{A(BC)}$. After the death of ${\cal N}_{C(AB)}$, ${\cal N}_{A(BC)} < {\cal N}_{B(AC)} $. The three-tangle is shown in the 3D plot in Fig. \[g12\]. The condition of the three-tangle sudden death is that each of the two nonzero AFRs should be large enough. The reason is that the three-tangle is now the average of the squares of one-tangles. Hence it suddenly dies when all one-tangles suddenly die. Some 2D cross sections of Fig. \[g12\] are shown in Fig. \[g13\] and Fig. \[g14\]. Note in the cases that $a_B/\Omega_B=300$ and $a_B/\Omega_B=750$ while $a_A=0$, as shown in Fig. \[g14\], the three-tangle only approaches zero asymptotically with the increase of $a_C/\Omega_C$, since the values of $a_B/\Omega_B$ are not large enough. ![\[g12\] Dependence of the three-tangle on the AFRs of qubits $B$ and $C$. Qubit $A$ moves uniformly. The qubits are in the GHZ state.](g12.eps){width="60.00000%"} $A$, $B$, and $C$ all accelerating ---------------------------------- In the case that all qubits accelerate, we have $$\begin{aligned} {\rho _{ABC}} &=& \frac{1}{2}\eta_A^2\eta_B^2\eta_C^2\sum\limits_{{n_A},{n_B},{n_C}} {\frac{{{e^{ - 2\pi \left( {{n_A}{\Omega_A}/{a_A} + {n_B}{\Omega_B}/{a_B} + {n_C}{\Omega_C}/{a_C}} \right)}}}}{{{Z_{{n_A},{n_B},{n_C}}}}} } \nonumber\\ &&\times\left[ {\left( {1 + \left( {{n_A} + 1} \right)\left( {{n_B} + 1} \right)\left( {{n_C} + 1} \right)|\mu _A|^2|\mu_B|^2|\mu_C|^2} \right)\left| {000} \right\rangle \left\langle {000} \right|} \right.\nonumber\\ &&+ \left( {1 + {n_A}{n_B}{n_C}|\mu_A|^2|\mu_B|^2|\mu_C|^2} \right)\left| {111} \right\rangle \left\langle {111} \right| + \left| {111} \right\rangle \left\langle {000} \right|\nonumber\\ &&+ \left| {000} \right\rangle \left\langle {111} \right| + \left( {{n_A}|\mu_A|^2 + \left( {{n_B} + 1} \right)\left( {{n_C} + 1} \right)|\mu _B|^2|\mu_C|^2} \right)\left| {100} \right\rangle \left\langle {100} \right|\nonumber\\ &&+ \left( {{n_B}|\mu_B|^2 + \left( {{n_A} + 1} \right)\left( {{n_C} + 1} \right)|\mu_A|^2|\mu_C|^2} \right)\left| {010} \right\rangle \left\langle {010} \right|\nonumber\\ &&+ \left( {{n_C}|\mu_C|^2 + \left( {{n_A} + 1} \right)\left( {{n_B} + 1} \right)|\mu_A|^2|\mu_B|^2} \right)\left| {001} \right\rangle \left\langle {001} \right|\nonumber\\ &&+ \left( \left( {{n_A} + 1} \right)|\mu_A|^2 + {{n_B}{n_C}|\mu_B|^2|\mu_C|^2 } \right)\left| {110} \right\rangle \left\langle {110} \right|\nonumber\\ &&+ \left( \left( {{n_B} + 1} \right)|\mu_B|^2 + {{n_A}{n_C}|\mu_A|^2|\mu_C|^2 } \right)\left| {101} \right\rangle \left\langle {101} \right|\nonumber\\ &&+\left. { \left( \left( {{n_C} + 1} \right)|\mu_C|^2 + {{n_A}{n_B}|\mu_A|^2|\mu_B|^2 } \right)\left| {011} \right\rangle \left\langle {011} \right|} \right],\nonumber\\* \,\end{aligned}$$ where $$\begin{split} {Z_{{n_A},{n_B},{n_C}}} =& 2 + \left[{n_A}{n_B}{n_C} + (n_A+1)(n_B+1)(n_C+1)\right]|\mu _A|^2|\mu _B|^2|\mu_C|^2\\ &+ (2{n_A}+1)|\mu_A|^2 + \left[ {n_B}{n_C} + (n_B+1)(n_C+1)\right]|\mu _B|^2|\mu_C|^2\\ &+ (2{n_B}+1)|\mu_B|^2 + \left[ {n_A}{n_C}+(n_A+1)(n_C+1)\right]|\mu _A|^2|\mu_C|^2\\ &+ (2{n_C}+1)|\mu_C|^2 + \left[ {n_A}{n_B}+(n_A +1)(n_B+1)\right]|\mu _A|^2|\mu_B|^2, \\ \end{split}$$ ${\eta_q} = \sqrt {1 - {e^{ - 2\pi {\Omega_q}/{a_q}}}}$, $(q = A, B, C)$. Figure \[g17to18\] is for $a_A/\Omega_A=100$ and $a_C/\Omega_C=2 a_B/\Omega_B$, where we show the three one-tangles. The plots have three intersections. Intersection (1) is at $a_A/\Omega_A = a_C/\Omega_C$ and thus ${\cal N}_{A(BC)} = {\cal N}_{C(AB)}$. Intersection (2) is at $a_A/\Omega_A = a_B/\Omega_B$ and thus ${\cal N}_{A(BC)} = {\cal N}_{B(AC)}$. Intersection (3) is another point where ${\cal N}_{A(BC)} = {\cal N}_{B(AC)}$. In the cases that qubit $A$ moves uniformly, intersection (3) is at the value of $a_B/\Omega_B$ where ${\cal N}_{C(AB)}$ suddenly dies, as shown in Figs. \[g7\], \[g8\], \[g9\], and \[g10\]. Now the nonzero $a_A$ delays this intersection. The three-tangle suddenly dies after all three one-tangles become zero, as shown in Fig. \[g18\]. In Fig. \[15to16\] we show the one-tangles and the three-tangle for $a_B/\Omega_B=1.2 a_A/\Omega_A$, $a_C/\Omega_C=1.5 a_A/\Omega_A$. The three one-tangles suddenly die successively, and afterwards the three-tangle becomes zero. After ${\cal N}_{C(AB)}$ suddenly dies, ${\cal N}_{A(BC)}$ and ${\cal N}_{B(AC)}$ do not intersect. This is because $a_A/\Omega_A$ is so large that the would-be intersection is shifted to some value of $a_A/\Omega_A$ larger than the value of sudden death of each of them. W state ======= Now we consider the initial state to be the W state, $$|\Psi_i\rangle =|\mathrm{W}\rangle.$$ In the W state, the one-tangles are ${\cal N}_{A(BC)} = {\cal N}_{B(AC)} = {\cal N}_{C(AB)} = 2\sqrt{2}/3$. The two-tangles are nonzero now, because for the W state, tracing out one qubit does not yield a separable state. One obtains ${\cal N}_{AB} = {\cal N}_{BA} = {\cal N}_{AC} = {\cal N}_{CA} = {\cal N}_{BC} = {\cal N}_{CB} =(\sqrt{5}-1)/3 \approx 0.412$. The three-tangle is less than 1, because the two-tangles are nonzero. One obtains $\pi =4(\sqrt{5}-1)/9 \approx 0.549$. $C$ accelerating ---------------- First we consider the case that only qubit $C$ accelerates, while $A$ and $B$ move uniformly. We obtain $$\begin{split} {\rho _{ABC}} =& \eta_C^2\sum\limits_{{n_C}} {\frac{{{e^{ - 2\pi {n_C}{\Omega _C}/{a_C}}}}}{{{Z_{{n_C}}}}}} \left[ {|001\rangle \langle 010| + |001\rangle \langle 100|} \right. + |100\rangle \langle 001| + |100\rangle \langle 010|\\ &+ |100\rangle \langle 100| + |010\rangle \langle 010| + \left( {|\mu _A|^2 + |\mu _B|^2 + \left( {{n_C} + 1} \right)|\mu _C|^2} \right)|000\rangle \langle 000|\\ &+ |010\rangle \langle 001| + {n_C}|\mu _C|^2\left( {|101\rangle \langle 101| + |011\rangle \langle 101| + |101\rangle \langle 011| + |011\rangle \langle 011|} \right)\\ &+\left.{|010\rangle \langle 100| + \left( {1 + {n_C}|\mu _A|^2|\mu _C|^2 + {n_C}|\mu _B|^2|\mu _C|^2} \right)|001\rangle \langle 001|} \right], \end{split}$$ where $${Z_{{n_C}}} = 3+|\mu_A|^2 + |\mu_B|^2 +(3{n_C}+1)|\mu_C|^2 + {n_C}|\mu_A|^2|\mu_C|^2 + {n_C}|\mu_B|^2|\mu_C|^2.$$ The entanglement among the qubits decreases when $C$ accelerates. As shown in Fig. \[w1\], the one-tangle ${\cal N}_{C(AB)}$ suddenly dies at a certain value of $a_C/\Omega_C$. This is similar to the GHZ state. However, differing from the GHZ state, with $a_A/\Omega_A = a_B/\Omega_B = 0$, ${\cal N}_{A(BC)} = {\cal N}_{B(CA)}$ has a minimum at a certain value of $a_C/\Omega_C$, but then increases towards a nonzero asymptotical value. As shown in Fig. \[w2\], with the increase of $a_C/\Omega_C$, ${\cal N}_{AC}= {\cal N}_{BC}$ decreases and suddenly dies at a certain value, while ${\cal N}_{AB}$ remains constant since it has nothing to do with $C$. As shown in Fig. \[w3\], with the increase of $a_C/\Omega_C$, the three-tangle first decreases towards a minimum, and then increases towards a nonzero asymptotic value. This can be inferred from the features of the one-tangles and two-tangles, according to the definition of the three-tangle. In the limit of $a_C/\Omega_C \to \infty$, $$\rho_{ABC}\to\frac{1}{3}\left[ |000\rangle\langle000|+|011\rangle\langle011|+|101\rangle\langle101|+|101\rangle\langle011|+|011\rangle\langle101|\right],$$ so the asymptotic values of the one-tangles are $$\begin{aligned} \lim\limits_{a_C\to\infty}\mathcal{N}_{A(BC)} &=& \lim\limits_{a_C\to\infty}\mathcal{N}_{B(AC)} = \frac{2}{3},\\ \lim\limits_{a_C\to\infty}\mathcal{N}_{C(AB)} &=& 0.\end{aligned}$$ The two-tangles are the following: $\mathcal{N}_{AB}=(\sqrt{5}-1)/3$, which is a constant independent of $a_C/\Omega_C$; $\mathcal{N}_{AC}=\mathcal{N}_{BC}=0$ after its sudden death. Consequently, the three-tangle asymptotically approaches $4(\sqrt{5}-1)/27 \approx 0.183$. $B$ and $C$ accelerating ------------------------- In the case that $B$ and $C$ accelerate while $A$ moves uniformly, we have $$\begin{aligned} {\rho _{ABC}} &=& \eta_B^2\eta_C^2\sum\limits_{{n_B},{n_C}} {\frac{{{e^{ - 2\pi \left( {{n_B}{\Omega _B}/{a_B} + {n_C}{\Omega _C}/{a_C}} \right)}}}}{{{Z_{{n_B},{n_C}}}}}} \left[ {|001\rangle \langle 010| + |001\rangle \langle 100|} \right. + |100\rangle \langle 001|\nonumber\\ &&+ |100\rangle \langle 010| + |010\rangle \langle 001| + |010\rangle \langle 100| + |100\rangle \langle 100| + n_B n_C|\mu_B|^2|\mu_C|^2|111\rangle \langle 111|\nonumber\\ &&+ {n_B}|\mu_B|^2|011\rangle \langle 110| + {n_C}|\mu_C|^2|101\rangle \langle 101| + {n_B}|\mu_B|^2|110\rangle \langle 110|\nonumber\\ &&+ {n_C}|\mu_C|^2|011\rangle \langle 101| + \left( {|\mu_A|^2 + \left( {n_B + 1} \right)|\mu_B|^2 + \left( {n_C + 1} \right)|\mu_C|^2} \right)|000\rangle \langle 000|\nonumber\\ &&+ \left( {1 + {n_C}|\mu_A|^2|\mu_C|^2 + \left( {n_B + 1} \right)n_C|\mu_B|^2|\mu_C|^2} \right)|001\rangle \langle 001|\nonumber\\ &&+ n_C|\mu_C|^2|101\rangle \langle 011| + \left( {1 + n_B|\mu _A|^2|\mu _B|^2 + n_B\left( {n_C + 1} \right)|\mu _B|^2|\mu_C|^2} \right)|010\rangle \langle 010|\nonumber\\ &&+ n_B|\mu_B|^2|110\rangle \langle 011| + \left. {\left( {{n_B}{n_C}|\mu_A|^2|\mu _B|^2|\mu _C|^2 + n_B|\mu_B|^2 + n_C|\mu _C|^2} \right)|011\rangle \langle 011|} \right],\nonumber\\* \,\end{aligned}$$ where $$\begin{split} {Z_{{n_B},{n_C}}} =& 1 + \frac{1}{3}|\mu _A|^2 + \left(n_B + \frac{1}{3}\right)|{\mu _B}|^2 + \left(n_C + \frac{1}{3}\right)|{\mu _C}|^2\\ &+ \frac{1}{3}{n_B}|\mu_A|^2|\mu_B|^2 + \frac{1}{3}{n_C}|{\mu_A}|^2|\mu_C|^2+ \left({n_B}{n_C}+\frac{1}{3}n_B+\frac{1}{3}n_C\right)|\mu_B|^2|{\mu_C}|^2\\ &+ \frac{1}{3}\left( n_A + 1 \right){n_B}{n_C}|\mu _A|^2|\mu _B|^2|\mu _C|^2. \end{split}$$ The one-tangle ${\cal N}_{A(BC)}$ is symmetric with respect to $a_B/\Omega_B$ and $a_C/\Omega_C$, as shown in Fig. \[w4\]. ${\cal N}_{B(AC)}$ is shown in Fig. \[w5\], and ${\cal N}_{C(AB)}$ can be obtained from ${\cal N}_{B(AC)}$ by exchanging $B$ and $C$, as shown in Fig. \[w6\]. These symmetries are common with the GHZ state. The two-tangle ${\cal N}_{AB}$ is shown in Fig. \[w7\], and ${\cal N}_{AC}$ can be obtained from ${\cal N}_{AB}$ by replacing $B$ with $C$, as shown in Fig. \[w8\]. ${\cal N}_{BC}$ is shown in Fig. \[w9\]. All exhibit sudden death. To see the dependence on $a_B/\Omega_B$ more clearly, we examined the two-dimensional cross sections of the 3D plots with $ a_C/\Omega_C=a_B/\Omega_B$ (Fig. \[aB=aC\]), $a_C/\Omega_C= 1.5a_B/\Omega_B$ (Fig. \[1.5aB=aC\]) and $a_C/\Omega_C= 2a_B/\Omega_B$ (Fig. \[2aB=aC\]). In these cases, in which $a_c/\Omega_C \propto a_B/\Omega_B$, the behavior of the one-tangles is similar to that in the GHZ state: each one-tangle suddenly dies at certain values of $a_B/\Omega_B$. For $0=a_A/\Omega_A <a_B/\Omega_B <a_C/\Omega_C$, ${\cal N}_{A(BC)} > {\cal N}_{B(AC)}>{\cal N}_{C(AB)}$ until the sudden death of the smallest one ${\cal N}_{C(AB)}$, and afterwards ${\cal N}_{B(AC)} > {\cal N}_{C(AB)}$. See Fig. \[w10\], Fig. \[w11\], and Fig. \[w12\]. A feature different from GHZ is that in addition to the sudden death, there is also a sudden change. We also look at one-tangles for some given values of $a_B/\Omega_B$, as shown in Fig. \[w13\] for $a_B/\Omega_B=300$ and in Fig. \[w14\] for $a_B/\Omega_B=750$. In both figures, a feature in common with the GHZ state is that when the smallest one-tangle (namely, ${\cal N}_{C(AB)}$) suddenly dies, the other two one-tangles exchange the order of magnitude. Compared with the GHZ state, a new feature is that ${\cal N}_{A(BC)}$ and ${\cal N}_{B(AC)}$ are not monotonic with respect to $a_C/\Omega_C$. As $a_B/\Omega_B$ is larger in Fig. \[w14\] than in Fig. \[w13\], the one-tangles decrease faster. In Fig. \[w13\], only ${\cal N}_{C(AB)}$ has sudden death, while the minima of the other two one-tangles are nonzero. In Fig. \[w14\], both ${\cal N}_{C(AB)}$ and ${\cal N}_{A(BC)}$ have sudden death, but ${\cal N}_{A(BC)}$ revives and increases at larger values of $a_C/\Omega_C$, because $B$ and $C$ constitute one party, and $B$ is fixed to be not large enough. In each of the two figures, ${\cal N}_{B(AC)}$ has a nonzero minimum, as $a_B/\Omega_B$ is now fixed, while $C$ is only one of the two qubits constituting the other party. Note that $C$ is the qubit on which the acceleration-frequency can always be large enough. Now we look at the two-tangles. The 2D cross sections at $a_C/\Omega_C=a_B/\Omega_B$, $a_C/\Omega_C= 1.5a_B/\Omega_B$, and $a_C/\Omega_C= 2a_B/\Omega_B$ are shown in Figs. \[w16\], \[w28\], and \[w18\], respectively. ${\cal N}_{AB}$ decreases with $a_B/\Omega_B$, ${\cal N}_{AC}$ decreases with $a_C/\Omega_C$, while ${\cal N}_{BC}$ decreases with both. For a given value of $a_B/\Omega_B$, ${\cal N}_{BC}$ is the smallest two-tangle in each of these cross sections. For the same value of $a_B/\Omega_B$, ${\cal N}_{AC}$ in Fig. \[w16\] is larger than in Fig. \[w28\], which is larger than in Fig. \[w18\], because $a_C/\Omega_C$ in Fig. \[w16\] is smaller than in Fig. \[w28\], which is smaller than in Fig. \[w18\]. Each two-tangle suddenly dies when the AFR of one or both of the parties are large enough. This is consistent with the behavior of entanglement of a two-qubit system [@dai]. In Fig. \[w19\] and Fig. \[w20\] we show the cases of $a_B/\Omega_B=300$ and $a_B/\Omega_B=750$, respectively. In Fig. \[w19\], ${\cal N}_{AB}$, measuring the entanglement between $A$ and $B$, is a constant independent of $a_C/\Omega_C$. In Fig. \[w20\], because $a_B/\Omega_B$ is large enough, ${\cal N}_{AB}$ and ${\cal N}_{BC}$ both remain zero, independent of $a_C/\Omega_C$. Finally, let us look at the three-tangle, whose 3D plot is shown in Fig. \[w21\]. To clearly see how it is different from the three-tangle of the GHZ state (Fig. \[g12\]), we examine some 2D cross sections. In Fig. \[w27\], in which it is set that $a_C/\Omega_C \propto a_B/\Omega_B$, the three-tangle dies when $a_B/\Omega_B$ is large enough, and the larger $a_C/\Omega_C$, the quicker the decrease of the three-tangle with $a_B/\Omega_B$. This feature is similar to that of the GHZ state \[Fig. \[g13\]\]. But on the other hand, when two of the three qubits’ accelerations are not large enough—for example, when it is fixed that $a_A/\Omega_A=0$ while $a_B/\Omega_B=300$ or $750$ \[Fig. \[w31\]\]—with the increase of $a_C/\Omega_C$, the three-tangle decreases to a nonzero minimum, and then increases towards an asymptotic value. Also refer to Fig. \[w3\] for the case of $a_A/\Omega_A=a_B/\Omega_B=0$. Therefore it is implied that the AFRs of at least two qubits should be large enough in order that the three-tangle has sudden death. Recall that in the GHZ state, when the AFRs of two qubits are not large enough, the three-tangle approaches zero asymptotically while the AFR of the remaining qubit tends to infinity \[Figs. \[g2\] and  \[g14\]\]. This is a difference between the two states. ![\[w21\] Dependence of the three-tangle on the AFR of qubits $B$ and $C$, while qubit $A$ moves uniformly. The qubits are in the W state.](w21.eps){width="60.00000%"} $A$, $B$, and $C$ all accelerating ---------------------------------- Now we consider the case that all three qubits accelerate, obtaining $$\begin{aligned} {\rho _{ABC}} &=& \eta_A^2\eta_B^2\eta_C^2\sum\limits_{{n_A},{n_B},{n_C}} {\frac{{{e^{ - 2\pi \left( {{n_A}{\Omega _A}/{a_A} + {n_B}{\Omega _B}/{a_B} + {n_C}{\Omega _C}/{a_C}} \right)}}}}{{{Z_{{n_A},{n_B},{n_C}}}}}} \left[ {|001\rangle \langle 010|} \right.\nonumber\\ &&+ |001\rangle \langle 100| + |100\rangle \langle 001| + |100\rangle \langle 010| + |010\rangle \langle 001| + |010\rangle \langle 100|\nonumber\\ &&+ n_B|\mu_B|^2|011\rangle \langle 110| + n_A|\mu_A|^2|101\rangle \langle 110| + n_C|\mu_C|^2|011\rangle \langle 101|\nonumber\\ &&+ n_A|\mu_A|^2|110\rangle \langle 101| + n_C|\mu_C|^2|101\rangle \langle 011| + n_B|\mu _B|^2|110\rangle \langle 011|\nonumber\\ &&+ \left( {\left( n_A + 1 \right)|\mu_A|^2 + \left( n_B + 1 \right)|\mu_B|^2 + \left( n_C + 1 \right)|\mu_C|^2} \right)|000\rangle \langle 000|\nonumber\\ &&+ \left( {1 + \left( n_A + 1 \right)n_C|\mu_A|^2|\mu_C|^2 + \left( n_B + 1\right)n_C|\mu_B|^2|\mu_C|^2} \right)|001\rangle \langle 001|\nonumber\\ &&+ \left( {1 + \left( n_A + 1 \right)n_B|\mu_A|^2|\mu _B|^2 + n_B\left( n_C + 1\right)|\mu_B|^2|\mu_C|^2} \right)|010\rangle \langle 010|\nonumber\\ &&+ \left( {\left( n_A + 1 \right)n_Bn_C|\mu_A|^2|\mu_B|^2|\mu_C|^2 + n_B|\mu_B|^2 + n_C|\mu_C|^2} \right)|011\rangle \langle 011|\nonumber\\ &&+ \left( {1 + n_A\left( n_B + 1 \right)|\mu_A|^2|\mu _B|^2 + n_A\left( n_C + 1 \right)|\mu _A|^2|\mu_C|^2} \right)|100\rangle \langle 100|\nonumber\\ &&+ \left( {n_A\left( n_B + 1 \right)n_C|\mu _A|^2|\mu _B|^2|\mu_C|^2 + n_A|\mu_A|^2 + n_C|\mu_C|^2} \right)|101\rangle \langle 101|\nonumber\\ &&+ \left( {n_A|\mu_A|^2 + n_B|\mu _B|^2 + n_An_B\left( n_C + 1 \right)|\mu_A|^2|\mu_B|^2|\mu_C|^2} \right)|110\rangle \langle 110|\nonumber\\ &&+ \left.{\left( {n_An_B|\mu_A|^2|\mu_B|^2 + n_An_C|\mu_A|^2|\mu_C|^2 + n_Bn_C|\mu_B|^2|\mu_C|^2} \right)|111\rangle \langle 111|} \right],\nonumber\\* \,\end{aligned}$$ where $$\begin{split} {Z_{{n_A},{n_B},{n_C}}} =& 3 + (3n_An_Bn_C + n_An_B + n_An_C + n_Bn_C)|\mu_A|^2|\mu_B|^2|\mu_C|^2 \\ &+ (3n_A+1)|\mu_A|^2 + (3n_B n_C+n_B+n_C)|\mu_B|^2|\mu_C|^2 \\ &+ (3n_B+1)|\mu_B|^2 + (3n_A n_C+n_A+n_C)|\mu_A|^2|\mu_C|^2\\ &+ (3n_C+1)|\mu_C|^2 + (3n_A n_B+n_A+n_B)|\mu_A|^2|\mu_B|^2. \end{split}$$ As an example, we specify $a_B/\Omega_B = 1.2a_A/\Omega_A$, $a_C/\Omega_C = 1.5a_A/\Omega_A$. One-tangles, two-tangles, and three-tangle are shown in Figs. \[w40\], \[w41\], and Fig. \[w42\], respectively. The behavior of the one-tangles and three-tangle of the W state is similar to that of the GHZ state as shown in Fig. \[15to16\], but under the same conditions, the values of these tangles are smaller in the W state. Like in the GHZ case, ${\cal N}_{C(AB)}$ suddenly dies first, because of the larger value of $a_C/\Omega_C$. Afterwards, ${\cal N}_{A(BC)}$ and ${\cal N}_{B(AC)}$ do not intersect because $a_A/\Omega_A$ is too large. The three-tangle suddenly dies when all one-tangles become zero. Note that when a one-tangle dies, the relevant two two-tangles have already died because of the relation (\[mono\]). Also note that a sudden change occurs in the three-tangle when the one-tangles and two-tangles suddenly die. The cases that one or more AFRs are not large enough can be inferred from the cases when only one or two qubits accelerate discussed above. Summary ======= In this paper, we have studied how the Unruh effect influences various kinds of entanglement of three qubits, each of which is coupled with an ambient scalar field. The initial states have been considered to be two typical three-qubit entangled states—the GHZ and the W states—which represent two types of tripartite entanglement. We have studied the cases of one qubit accelerating, two qubits accelerating, and three qubits accelerating. The details are summarized in Table \[table\]. A two-tangle measures bipartite entanglement between two qubits, with the other qubit traced out. For the GHZ state as the initial state, the two-tangles always remain zero. For the W state as the initial state, the two-tangle between two qubits remains a nonzero constant if these two qubits move uniformly. If at least one of the two qubits accelerates, their two-tangle suddenly dies when the AFR of one or both of these two qubits are large enough. When two qubits move uniformly while the other accelerates, the one-tangle between the accelerating qubit and the party consisting of the two uniformly moving qubits suddenly dies if the AFR of the accelerating qubit is large enough. However, each of the other two one-tangles—between one uniformly moving qubit and the party consisting of the other uniformly moving qubit and the accelerating qubit—approaches an asymptotic value as the AFR tends to infinity. For the GHZ state, the asymptotic value is zero. For the W state, the asymptotic value is nonzero and is larger than the local minima existing at finite values of the AFR. [|c|l|l|l|]{} & &\ &1-tangles & ------------------------------------------------------ ${\cal N}_{C(AB)}$ SD at a finite $a_C/\Omega_C$. ${\cal N}_{A(BC)}={\cal N}_{B(AC)} \rightarrow 0$ as $a_C/\Omega_C\rightarrow \infty$. ------------------------------------------------------ & ----------------------------------------------------------------------- ${\cal N}_{C(AB)}$ SD at a finite $a_C/\Omega_C$. ${\cal N}_{A(BC)}={\cal N}_{B(AC)}$ is non- monotonic, with nonzero local minima, $\rightarrow \frac{2}{3}$ as $a_C/\Omega_C\rightarrow\infty$. ----------------------------------------------------------------------- \ &2-tangles & $0$. & ---------------------------------------------- ${\cal N}_{AB}$ remains constant. ${\cal N}_{AC}={\cal N}_{BC}$ SD at a finite $a_C/\Omega_C$. ---------------------------------------------- \ &3-tangle & ------------------------------------------------------------ $\rightarrow 0$ as $a_C/\Omega_C\rightarrow \infty$. There is a sudden change where ${\cal N}_{C(AB)}$ SD. ------------------------------------------------------------ & -------------------------------------------------- Nonmonotonic, with a nonzero local minimum at a finite $a_C/\Omega_C$, $\rightarrow (\sqrt{5}-1)/27$ as $a_C/\Omega_C\rightarrow \infty$. -------------------------------------------------- \ &1-tangles & --------------------------------------------------- ${\cal N}_{A(BC)}$ SD at finite values of both $a_B/\Omega_B$ and $a_C/\Omega_C$. ${\cal N}_{\beta (A\gamma)}$ SD at a finite value of $a_\beta/\Omega_\beta$, while weakly varies with $a_\gamma/\Omega_\gamma$. --------------------------------------------------- & ---------------------------------------------------------------------- ${\cal N}_{A(BC)}$ eventually SD if both $a_B/\Omega_B$ and $a_C/\Omega_C$ are large enough. It is nonmonotonic, with nonzero local minima, and can SD and revive with the increase of one AFR while the other is not large enough. ${\cal N}_{\beta (A\gamma)}$ SD when $a_\beta/\Omega_\beta$ is large enough. It has nonzero local minima and has sudden changes if $a_\beta/\Omega_\beta$ is not large enough no matter how large is $a_\gamma/\Omega_\gamma$. ---------------------------------------------------------------------- \ &2-tangles &$0$. & -------------------------------- SD when one or both of the two relevant qubits have large enough AFRs. -------------------------------- \ &3-tangle & --------------------------------------------- SD if $a_B/\Omega_B$ and $a_C/\Omega_C$ are both large enough. --------------------------------------------- & ---------------------------------------------- SD if $a_B/\Omega_B$ and $a_C/\Omega_C$ are both large enough. It has nonmonotonicity and local nonzero minima, $\rightarrow$ a nonzero asymptotic value when only one AFR $\rightarrow \infty$ while the other nonzero AFR is not large enough. ---------------------------------------------- \ &1-tangles & --------------------------------------------------------------- ${\cal N}_{\alpha(\beta\gamma)}$ SD if the AFR of $\alpha$ or AFRs of both $\beta$ and $\gamma$ are large enough. --------------------------------------------------------------- & --------------------------------------------------------------- ${\cal N}_{\alpha(\beta\gamma)}$ SD if the AFR of $\alpha$ or AFRs of both $\beta$ and $\gamma$ are large enough. There exists nonmonotonicity. --------------------------------------------------------------- \ &2-tangles &$0$. & ---------------------------- SD when one or both of the two relevant qubits have large enough AFRs. ---------------------------- \ &3-tangle & ----------------------------- SD after all 1-tangles SD, that is, if at least two of three qubits have large enough AFRs. ----------------------------- & ------------------------------------- SD after all 1-tangles SD, that is, if at least two of three qubits have large enough AFRs. ------------------------------------- \ When two qubits accelerate while the other moves uniformly, the one-tangle between the uniformly moving qubit and the party consisting of the two accelerating qubits suddenly dies if the AFRs of both of the two accelerating qubits are large enough. Each of the other two one-tangles—between one accelerating qubit and the party consisting of the other accelerating qubit and the uniformly moving qubit—suddenly dies when the AFR of the qubit which is by itself one party is large enough. It weakly depends on the AFR of the other accelerating qubit. These features are common in the GHZ and the W states. When all qubits accelerate, for both the GHZ and the W states, the one-tangle ${\cal N}_{\alpha(\beta\gamma)}$ suddenly dies if $a_\alpha/\Omega_\alpha$ is large enough, or both $a_\beta/\Omega_\beta$ and $a_\gamma/\Omega_\gamma$ are large enough. Therefore, generally speaking, all the one-tangles eventually suddenly die if at least two of the three qubits have large enough AFRs. When all the one-tangles suddenly die, all the two-tangles must have also died, as dictated by the monogamy relation (\[mono\]), and consequently the three-tangles also suddenly die. The main difference between the W state and the GHZ state is that for the W state, there exists nonmonotonicity with respect to the AFR of each qubit alone. It is well known that near the horizon $r=2m$ of a black hole, the Schwarzschild metric can be approximated as the Rindler metric with the acceleration $a = \frac{m}{{{r^2}}}{ ( {1 - \frac{{2m}}{r}} )^{ - 1/2}}$, while the uniform movement corresponds to free falling into the black hole [@rindler]. Therefore, the above result can be translated to be near the horizon of a black hole, with $r\approx 2m[1-\frac{1}{(4m a)^2}]^{-1}$ corresponding to the acceleration $a$. The calculations in this paper imply that near the horizon of a black hole, for three qubits coupled with scalar fields, all the one-tangles and then all the two-tangles and the three-tangle eventually die if at least two of the three qubits are close enough to the horizon $2m$. That is, all kinds of entanglement of the field-coupled qubits are eventually killed by the black hole horizon. Finally, we conjecture that for $N$ particles, each of which is coupled with a scalar field, all kinds of entanglement suddenly die if $N-1$ particles are close enough to the horizon of a black hole. This work was supported by the National Science Foundation of China (Grant No. 11374060). [99]{} V. F. Mukhanov and S. Winitzki, [*Introduction to Quantum Effects in Gravity*]{}, (Cambridge University Press, Cambridge, England, 2007). S. W. Hawking, Commun. Math. Phys. [**43**]{}, 199 (1975). S. A. Fulling, Phys. Rev. D [**7**]{}, 2850 (1973). P. C. W. Davies, J. Phys. A [**8**]{}, 609 (1975). B. S. DeWitt, Phys. Rep. [**19**]{}, 295 (1975). W. G. the Unruh, Phys. Rev. D [**14**]{}, 870 (1976). S.-Y. Lin, C.-H. Chou, and B. L. Hu, Phys. Rev. D. [**78**]{}, 125025 (2008). A. G. S. Landulfo and G. E. A. Matsas, Phys. Rev. A [**80**]{}, 032315 (2009). J. Doukas and L. C. L. Hollenberg, Phys. Rev. A [**79**]{}, 052109 (2009). J. Doukas and B. Carson, Phys. Rev. A 81, 062320 (2010). J. Hu and H. Yu, Phys. Rev. A [**91**]{}, 012327 (2015). Y. Dai, Z. Shen, and Y. Shi, J. High Energy Phys. 09 (2015) 071. C. H. Bennett, S. Popescu, D. Rohrlich, J. A. Smolin, and A. V. Thapliyal, Phys. Rev. A [**63**]{}, 012307 (2000). G. Vedal and R. F. Werner, Phys. Rev. A [**65**]{}, 032314 (2002). Y. C. Ou and H. Fan, Phys. Rev. A [**75**]{}, 062308 (2007). W. Dür, G. Vidal and J. I. Cirac, Phys. Rev. A [**62**]{}, 062314 (2000). W. G. the Unruh and R. M. Wald, Phys. Rev. D [**29**]{}, 1047 (1984). T. Yu and J. H. Eberly, Science [**23**]{}, 598 (2009). W. Rindler, Am. J. Phys. [**34**]{}, 1174 (1966).
{ "pile_set_name": "ArXiv" }
--- abstract: 'The resolvent of supersymmetric Dirac Hamiltonian is studied in detail. Due to supersymmetry the squared Dirac Hamiltonian becomes block-diagonal whose elements are in essence non-relativistic Schrödinger-type Hamiltonians. This enables us to find a Feynman-type path-integral representation of the resulting Green’s functions. In addition, we are also able to express the spectral properties of the supersymmetric Dirac Hamiltonian in terms of those of the non-relativistic Schrödinger Hamiltonians. The methods are explicitly applied to the free Dirac Hamiltonian, the so-called Dirac oscillator and a generalization of it. The general approach is applicable to systems with good and broken supersymmetry.' author: - 'G. Junker' - 'A. Inomata' title: 'Path Integral and Spectral Representations for Supersymmetric Dirac-Hamiltonians' --- Introduction ============ The idea of supersymmetry (SUSY) in non-relativistic quantum mechanics was first introduced by Nicolai [@Nic76] in 1976 for analyzing spin systems with a $(0+1)$-dimensional version of supersymmetric quantum field theories. SUSY quantum mechanics became popular when Witten’s model was introduced in 1981 [@Wit81]. Since then SUSY has become an important tool in studying properties of models in quantum mechanics and statistical physics [@Junker1996]. Jackiw [@Jackiw1984] observed that the square of a 2-dimensional Dirac particle subjected to an external magnetic field is directly related to the non-relativistic Pauli-Hamiltonian of the same system. This observation triggered a more extended study of SUSY for the Dirac Hamiltonians [@Hughesetal86; @Cooper1988]. Indeed, under certain conditions the Dirac Hamiltonian may be treated as supercharge of some non-relativistic system in SUSY quantum mechanics[@Thaller1988; @Beckers1990; @Thaller1991]. Such a SUSY structure in turn proofs to be very useful in studying the spectral properties of the Dirac Hamiltonian. For an overview of relativistic quantum systems exhibiting SUSY, see ref. \[\]. In recent years, SUSY has also become a key concept in characterizing condensed matter systems known as topological superconductors[@Groverelal2014] and also plays a crucial role in explaining quantum phenomena in carbon-based nano-structures like nano-tubes and graphene. In the latter the two-dimensional Dirac Hamiltonian is used in characterizing its electronic transport properties.[@Sarma2011] Here SUSY plays a crucial role, for example, in understanding phenomena like the unconventional quantum Hall effect.[@Ezawa2006] In the present paper, we study the resolvent of the Dirac Hamiltonian with the SUSY structure. We show that the Green’s functions resulting from the resolvent can be determined by means of path integrals for the effective non-relativistic Lagrangians. The spectral properties of the Dirac Hamiltonian can be derived in a simple way from those of the associated Schrödinger Hamiltonians. We also present explicit applications to a few examples of the Dirac system. In Section II, a stage is set for studying the SUSY properties of the Dirac Hamiltonian. In particular the spectral properties of a supersymmetric Dirac Hamiltonian are shown to be given explicitly via those of a corresponding non-relativistic $N=2$ SUSY Hamiltonian. Section III deals with the path integral representation of the Green’s functions based on the effective Schrödinger Hamiltonian. The path integral used is formally equivalent to Feynman’s, in which the classical action is replaced by the effective action corresponding to the effective Schrödinger Hamiltonian. In Sections IV to VI are devoted to solving examples which include the free Dirac electron, the Dirac oscillator and a generalized Dirac oscillator. Concluding remarks are given in Section VII. Supersymmetric Dirac Hamiltonians ================================= Supersymmetric (SUSY) quantum mechanics, as defined by Witten [@Wit81], is characterized by Hamiltonian $H_{SUSY}$ and a set of self-adjoint operators $Q_i$ ($i=1,2,\ldots, N$) acting on some Hilbert space ${\cal H}$ and satisfying (see, e.g., ref. \[\]), $$\label{SUSYAlgebra} \{ Q_i, Q_j \} = H_{SUSY} \delta_{ij}, \quad i,j=1,2,\ldots,N \, ,$$ where $\{A,B\}:=AB+BA$ denotes the anti-commutator. To study the SUSY-structure of the Dirac system, we restrict ourselves to the case of $N=2$ which involves only the three operators, $Q_1,Q_2$ and $H_{SUSY}$ acting on ${\cal H}$. Here, instead of dealing with $Q_1$ and $Q_2$, we define complex supercharges $$\label{complexQ} Q = \frac{1}{\sqrt{2}}\left(Q_1 + {\textrm{i}}Q_2\right)\, , \quad Q^\dag = \frac{1}{\sqrt{2}}\left(Q_1 - {\textrm{i}}Q_2\right)\, .$$ These operators and the SUSY Hamiltonian close the superalgebra [@Nic76], $$\label{SUSYAlgebra2} \{ Q, Q^\dag \} = H_{SUSY} \, , \quad Q^2 = 0 =\left(Q^\dag\right)^2 \, .$$ In addition, we introduce the so-called Witten operator $W$ which is a unitary (non-trivial) involution on ${\cal H}$, commutes with the SUSY Hamiltonian and anticommutes with the complex supercharges, that is, $$\label{SUSYAlgebra3} \begin{array}{c} W^\dag = W \, , \quad W^2=1 \, , \quad [W,H_{SUSY}]=0 \, , \\[2mm] \{W,Q\} = 0 =\{W,Q^\dag\} . \end{array}$$ See ref. \[\]. The projection operators defined by $P^\pm=(1\pm W)/2$ decompose ${\cal H}$ into the eigenspaces ${\cal H}^\pm$ of the Witten operator with eigenvalues $\pm 1$. Namely, $$\begin{aligned} {\cal H} &=& {\cal H}^+\oplus {\cal H}^-\, ,\\ {\cal H}^\pm &=& P^\pm {\cal H}= \{|\Psi^\pm\rangle\in{\cal H}:W|\Psi^\pm\rangle=\pm|\Psi^\pm\rangle\} .\end{aligned}$$ In the representation in which $W$ is diagonal, i.e., $$\label{W} W = \left( \begin{array}{cc} 1 & 0 \\ 0 & -1 \end{array} \right),$$ the complex supercharges take the form, $$\label{Q} Q = \left( \begin{array}{cc} 0 & D \\ 0 & 0 \end{array} \right),\quad Q^\dag = \left( \begin{array}{cc} 0 & 0 \\ D^\dag & 0 \end{array} \right),$$ where $D$ is not self-adjoint in general. The SUSY Hamiltonian is usually given by $$\label{Hsusy} H_{SUSY}=\{Q,Q^\dag\}= \left( \begin{array}{cc} DD^\dag & 0 \\ 0& D^\dag D \end{array} \right).$$ However, more generally, within the framework of the SUSY algebra (\[SUSYAlgebra3\]), we may choose the SUSY Hamiltonian in the form, $$\label{Hsusy2} H_{SUSY}=a\{Q,Q^\dag\} + \varepsilon_0\textbf{1}= \left( \begin{array}{cc} aDD^\dag +\varepsilon_0& 0 \\ 0& aD^\dag D+\varepsilon_0 \end{array} \right),$$ where $a$ and $\varepsilon_0$ are arbitrary real constants with $a>0$. In studying the Dirac problem such a modification is not imperative but useful. We treat the two diagonal elements $H^+_{SUSY}=aDD^\dag+\varepsilon_0$ and $H^-_{SUSY}=aD^\dag D+\varepsilon_0$ as the SUSY partner Hamiltonians. As in the usual SUSY case the partner Hamiltonians are isospectral. Let $\varepsilon^\pm$ and $|\Psi_\varepsilon^\pm\rangle$ denote the eigenvalues and eigenstates of $H^\pm_{SUSY}$, that is, $$\label{SpecPropHsusy} H^\pm_{SUSY}|\Psi_\varepsilon^\pm\rangle = \varepsilon^\pm |\Psi_\varepsilon^\pm\rangle\,.$$ It is obvious that $\varepsilon^\pm\geq \varepsilon_0$ and $\varepsilon^+=\varepsilon^-=\varepsilon$ for $\varepsilon>\varepsilon_0$. The eigenstates $|\Psi_\varepsilon^\pm\rangle$ for the same eigenvalue $\varepsilon>\varepsilon_0$ are related by the SUSY transformations $$\label{SUSYtrafogen} \begin{array}{rl} D^\dag|\Psi_\varepsilon^+\rangle = &\sqrt{(\varepsilon-\varepsilon_0)/a}|\Psi_\varepsilon^-\rangle\, ,\\[2mm] D|\Psi_\varepsilon^-\rangle = & \sqrt{(\varepsilon-\varepsilon_0)/a}|\Psi_\varepsilon^+\rangle \,. \end{array}$$ For $\varepsilon=\varepsilon_0$ there might exist eigenstates in ${\cal H}^-$ or in ${\cal H}^+$ or in both. In that case SUSY is said to be unbroken. If $\varepsilon_0$ does not belong to the spectrum of $H^-_{SUSY}$ and $H^+_{SUSY}$ SUSY is said to be broken. Now we consider the Dirac Hamiltonian $$\label{HD} H_D=c{\bm{\alpha}}\cdot\left(\mathbf{p}-\frac{e}{c}\mathbf{A}\right)+\beta m_0c^2.$$ Here ${\bm{\alpha}}$ and $\beta$ are the $4\times 4$ Dirac matrices, $c$ is the speed of light, $m_0$ the rest mass and $e$ the charge of an electron. Notice that the scalar potential is missing in (\[HD\]). In the present paper, we study only the case where the scalar potential is absent. We employ the standard representation of the Dirac matrices $$\label{alpha&beta} {\bm{\alpha}}=\left( \begin{array}{cc} 0 & {\bm{\sigma}}\\ {\bm{\sigma}}& 0 \end{array} \right)\, ,\quad \beta=\left( \begin{array}{cc} 1 & 0 \\ 0 & -1 \end{array} \right)\, ,$$ where ${\bm{\sigma}}$ is the unit vector based on the Pauli matrices, and let $$\label{D} D=c{\bm{\sigma}}\cdot\left(\textbf{p}-\frac{e}{c}\textbf{A}\right)\, ,\quad M_0=m_0c^2\, .$$ Then we can express the Dirac Hamiltonian (\[HD\]) in the form $$\label{HD2} H_D=\left(Q+Q^\dag\right)+M_0W=\sqrt{2}Q_1+M_0W$$ or $$\label{HD3} H_D=\left( \begin{array}{cc} M_0 & D \\ D^\dag & -M_0 \end{array} \right).$$ The operator $D$ specified by (\[D\]) is not self-adjoint unless the vector potential $\textbf{A}$ is real. Evidently, in Witten’s sense, $H=(H_D-M_0W)^2$ is the SUSY Hamiltonian of $N=1$ with supercharge $Q_1=(H_D-M_0W)/\sqrt{2}$. In the approach of Thaller [@Thaller1988] and Beckers and Debergh [@Beckers1990], the Dirac Hamiltonian itself is supersymmetric in the sense that it has the odd part $H_D^{odd}=Q+Q^\dag$ and the even part $H_D^{even}=M_0W$ with respect to $W$, obeying $\{H_D^{odd},W\}=0$ and $[H_D^{even},W]=0$. In the preceeding sections, we shall, however, treat the Dirac Hamiltonian (\[HD\]) or (\[HD2\]) in the framework of SUSY quantum mechanics of $N=2$. To this end, we pay attention to the squared Dirac operator which turns out to be (block) diagonal $$\label{HD^2} \begin{array}{rl} \displaystyle H^2_D &= \{Q,Q^\dag\} + M_0^2 \\[2mm] \displaystyle &= \left( \begin{array}{cc} D D^\dag +M_0^2 & 0 \\ 0 & D^\dag D +M_0^2 \end{array} \right). \end{array}$$ If $H_D^2$ is taken as the SUSY Hamiltonian, then its diagonal elements are SUSY partner. Since $DD^\dag$ and $D^\dag D$ behave as $\sim c^2\textbf{p}^2$, we introduce a mass parameter $m>0$ to make the squared Dirac operator look like a Schrödinger Hamiltonian; namely $$\label{HS} H_{SUSY}=\frac{1}{2mc^2}H^2_D= \left( \begin{array}{cc} H^+_{SUSY} & 0 \\ 0 & H^-_{SUSY} \end{array} \right),$$ where $$\label{HS+-} \begin{array}{c} \displaystyle H_{SUSY}^+=\frac{1}{2mc^2}\left(DD^\dag+M^2_0\right)\,,\\[4mm] \displaystyle H_{SUSY}^-=\frac{1}{2mc^2}\left(D^\dag D+M^2_0\right)\,, \end{array}$$ acting on the subspaces ${\cal H}^\pm$, respectively. In other words, we choose $a=1/2mc^2$ and $\varepsilon_0=M^2_0/2mc^2$ in (\[Hsusy2\]). The diagonal elements of the squared Dirac operator modified with the mass parameter are indeed the Schrödinger Hamiltonians for a non-relativistic system of mass $m$ in form. The mass parameter $m$ introduced above may be identified with the electron rest mass $m_0$, but will be left unspecified for a while to take note of its formal and arbitrary nature. Now we can utilize these Schrödinger operators as SUSY partner Hamiltonians by regarding $H_{SUSY}=H_D^2/2mc^2$ as the SUSY Hamiltonian. Furthermore, use of the Schrödinger Hamiltonians enables us to pursue the Feynman-type path integral representation for the Dirac systems as will be discusses in the proceeding sections. It had been noted by Thaller [@Thaller1988], see also ref. \[,\], that the Dirac operator can also be diagonalized by a unitary Foldy-Wouthuysen transformation $U_{\rm FW}$. Let $$\label{UFW} U_{\rm FW}= \frac{|H_D|+WH_D}{\sqrt{2H^2_D+2M_0|H_D|}}$$ be a unitary operator, where $$\label{|HD|} |H_D|=\sqrt{H_D^2}=\sqrt{2mc^2H_{SUSY}}$$ is an even operator, $[W,|H_D|]=0$. It is easy to show that $$\label{Hdiagonal} \begin{array}{rl} H_{\rm FW}&=U_{\rm FW} H_DU_{\rm FW}^\dagger= W|H_D|\\[2mm] &\displaystyle = \left( \begin{array}{cc} \sqrt{D D^\dagger+M_0^2} & 0 \\ 0 & -\sqrt{D^\dagger D +M_0^2} \end{array} \right) , \end{array}$$ which implies that the spectrum of $H_D$ can be obtained from the spectrum of $H_{\rm FW}$. To be a bid more explicit let us denote the upper and lower diagonal elements of $H_{\rm FW}$ by $H^\pm_{\rm FW}$. We may express them via the SUSY Hamiltonians $H^\pm_{SUSY}$ as follows $$\label{HFWpm} H^\pm_{\rm FW}=\pm\sqrt{2mc^2H^\pm_{SUSY}}.$$ Hence the spectrum of $H_D$ which is identical to that of $H_{\rm FW}$ can in turn be obtained from that of $H_{ SUSY}$. Although the transformed $H_{\rm FW}$ is diagonal, it does not have the SUSY structure. The diagonal elements in (\[Hdiagonal\]) do not form the SUSY partners. However, the eigenstates $|\Psi_\varepsilon^\pm\rangle$ of $H^\pm_{SUSY}$ are also eigenstates of $H^\pm_{\rm FW}$. Hence for $\varepsilon > \varepsilon_0$ we have $$\label{PsiHFWpm} H^\pm_{\rm FW}|\Psi_\varepsilon^\pm\rangle = \pm \sqrt{2mc^2\varepsilon}|\Psi_\varepsilon^\pm\rangle\, .$$ In addition in the case of unbroken SUSY $\varepsilon_0=M_0^2/2mc^2$ belongs to the spectrum of either $H^+_{\rm SUSY}$ or $H^-_{\rm SUSY}$ or to both and we have $$\label{HFWPsi0} \begin{array}{l} \displaystyle |\Psi_{\varepsilon_0}^+\rangle\in{\cal H}^+ \Rightarrow H^+_{\rm FW}|\Psi_{\varepsilon_0}^+\rangle =M_0|\Psi_{\varepsilon_0}^+\rangle\, ,\\ \displaystyle |\Psi_{\varepsilon_0}^-\rangle\in{\cal H}^- \Rightarrow H^-_{\rm FW}|\Psi_{\varepsilon_0}^-\rangle =-M_0|\Psi_{\varepsilon_0}^-\rangle\, . \end{array}$$ Hence the spectrum of $H_{\rm FW}$ and therefore also of $H_D$ is symmetric about the origin with possible exceptions at $\pm M_0$, cf. ref. \[\]. The corresponding positive and negative energy eigenstates $|\Psi_{E}^{\rm pos/neg}\rangle$ of $H_D$ can explicitly be calculated from those of $H_{\rm FW}$ by utilizing the unitary transformation (\[UFW\]). That is, $$\label{PsiD} \begin{array}{l} |\Psi_{E}^{\rm pos}\rangle = U_{\rm FW}^\dag|\Psi_{\varepsilon}^{+}\rangle\, ,\quad H_D|\Psi_{E}^{\rm pos}\rangle= \sqrt{2mc^2\varepsilon}|\Psi_{E}^{\rm pos}\rangle\, ,\\ |\Psi_{E}^{\rm neg}\rangle = U_{\rm FW}^\dag|\Psi_{\varepsilon}^{-}\rangle\, ,\quad H_D|\Psi_{E}^{\rm neg}\rangle= -\sqrt{2mc^2\varepsilon}|\Psi_{E}^{\rm neg}\rangle\, , \end{array}$$ and the Dirac eigenvalues $E$ are given by those of the associated SUSY Hamiltonian, $\varepsilon$, via the simple relation $$\label{EDirac} E=\pm\sqrt{2mc^2\varepsilon}\,.$$ With the help of the SUSY transformation (\[SUSYtrafogen\]) the positive and negative energy eigenstates explicitly read $$\label{PsiDexplicit+} |\Psi_{\varepsilon}^{\rm pos}\rangle =\frac{1}{\sqrt{2+2\sqrt{\varepsilon_0/\varepsilon}}} \left( \begin{array}{c} \left(1+\sqrt{\varepsilon_0/\varepsilon}\right)|\Psi_{\varepsilon}^{+}\rangle\\ \sqrt{1-\varepsilon_0/\varepsilon}\,|\Psi_{\varepsilon}^{-}\rangle \end{array} \right)\, ,$$ $$\label{PsiDexplicit-} |\Psi_{\varepsilon}^{\rm neg}\rangle =\frac{1}{\sqrt{2+2\sqrt{\varepsilon_0/\varepsilon}}} \left( \begin{array}{c} -\sqrt{1-\varepsilon_0/\varepsilon}\,|\Psi_{\varepsilon}^{+}\rangle\\ \left(1+\sqrt{\varepsilon_0/\varepsilon}\right)|\Psi_{\varepsilon}^{-}\rangle \end{array} \right)\, .$$ In the case of unbroken SUSY when $M_0\in{\rm spec}(H_D)$ the corresponding eigenstates are given by $|\Psi_{\varepsilon_0}^{\rm pos}\rangle=|\Psi_{\varepsilon_0}^{+}\rangle$ and in the case of $-M_0\in{\rm spec}(H_D)$ by $|\Psi_{\varepsilon_0}^{\rm neg}\rangle=|\Psi_{\varepsilon_0}^{-}\rangle$. In closing this section let us summarize that the spectral properties of a supersymmetric Dirac Hamiltonian (\[HD3\]) are explicitly given via those of the non-relativistic SUSY Hamiltonian (\[HS\]). Path integral representation of the resolvent ============================================= In non-relativistic quantum mechanics, the time-evolution operator can be represented in terms of a Feynman path integral. In fact, the Schrödinger Hamiltonian being quadratic in the momentum makes it possible to define the Feynman path integral properly[@Feyn1948; @Feyn1965]. In the Dirac theory, the Hamiltonian is linear in the momentum. Hence the path integral representation cannot be constructed in a manner analogous to that for the non-relativistic case. Especially, for the SUSY Dirac problem, it is desirable to find an effective SUSY Hamiltonian quadratic in the momentum, corresponding to the Dirac operator. To this end we shall employ the following procedure. Let us start with the resolvent of the Dirac operator (\[HD2\]) $$\label{G} G(z)=\frac{1}{H_D-z},\quad z\in\mathbb{C}\backslash \textrm{spec}(H_D)\,,$$ which is an analytical function of $z$ in the complement of the spectrum of $H_D$. Let us write it in the form $$\label{Gg} G(z)=(H_D+z)g(z^2)\,.$$ Here $$\label{g} g(\zeta)=\frac{1}{H_D^2-\zeta}$$ is the resolvent of the squared Dirac operator $H_D^2$, defined in the $\zeta$-plane. The iterated resolvent $g(\zeta)$ is easier to handle than the original resolvent $G(z)$ is. While the Dirac operator $H_D$ given in (\[HD3\]) is not diagonal, the squared operator $H_D^2$ is diagonal as shown in (\[HD\^2\]). Hence the iterated resolvent $g(\zeta)$ can as well be given in the diagonal form, $$\label{g2} g(\zeta)=\left( \begin{array}{cc} g^+(\zeta) & 0 \\ 0 & g^-(\zeta) \end{array} \right) ,$$ with $$\label{gpm} g^+(\zeta) = \frac{1}{DD^\dag+M_0^2-\zeta}\, , \quad g^-(\zeta) = \frac{1}{D^\dag D+M_0^2-\zeta}\, .$$ The resolvent $G(z)$ of the Dirac operator $H_D$ may be obtained with the help of the diagonal elements of the iterated resolvent, in the form, $$\label{Goperator} \begin{array}{rcl} G(z)&=&\left(H_D + z\right)g(z^2) \\[2mm] &=& \left( \begin{array}{cc} (z+M_0)g^+(z^2) & Dg^-(z^2) \\ D^\dag g^+(z^2) & (z-M_0)g^-(z^2) \end{array} \right). \end{array}$$ The iterated resolvent $g(\zeta)$ as well as $G(z)$ acts on Hilbert space ${\cal H}$. Hence $|\Psi'\rangle=g(\zeta)|\Psi\rangle\in{\cal H}$ for $|\Psi\rangle\in{\cal H}$, which yields on the coordinate base an integral equation, $$\langle\textbf{r}''|\Psi'\rangle=\int_{\mathbb{R}^3}{\textrm{d}}\textbf{r}'\,\langle\textbf{r}''|g(\zeta)|\textbf{r}'\rangle\langle\textbf{r}'|\Psi\rangle \, ,$$ whose kernel $\langle\textbf{r}''|g(\zeta)|\textbf{r}'\rangle$, denote by $g(\textbf{r}'',\textbf{r}';\zeta)$, will be referred to as the iterated resolvent kernel. Since $H_D^2$ in (\[g\]) and hence $H_{SUSY}$ defined by (\[HS\]) are positive semi-definite, the spectrum of the corresponding Schrödinger operator, $\textrm{spec}(H_{SUSY})$, is on the non-negative real axis of the $\zeta$-plane. Considering a contour $C$ encircling counter-clockwise all points corresponding to $\textrm{spec}(H_D^2)$ and using Cauchy’s integral formula, we can find $$\label{contour} {\textrm{e}}^{-{\textrm{i}}tuH^2_D}=-\frac{1}{2\pi {\textrm{i}}}\oint_C{\textrm{d}}\zeta\,{\textrm{e}}^{-{\textrm{i}}tu\zeta}g(\zeta)\, ,$$ where $t$ and $u$ are any real constants. At this point, recall the Schrödinger operator $H_{SUSY}$ is defined by (\[HS\]), we replace $H^2_D$ on the left hand side by $2mc^2H_{SUSY}$. Moreover, we let $u=(2mc^2\hbar)^{-1}$. Then the quantity on the left hand side of (\[contour\]) can be understood as the unitary evolution operator of the system with the Schrödinger Hamiltonian $H_{SUSY}$ if $t$ is identified with the time parameter. In the coordinate representation, (\[contour\]) may be written as $$\label{contour2} K(\textbf{r}'',\textbf{r}';t)=-\frac{1}{2\pi {\textrm{i}}}\oint_C{\textrm{d}}\zeta\,{\textrm{e}}^{-{\textrm{i}}t\zeta /2mc^2\hbar}g(\textbf{r}'',\textbf{r}';\zeta)\, ,$$ which relates the resolvent kernel to Feynman’s kernel (or the propagator), $$\label{propagator} K(\textbf{r}'',\textbf{r}';t)=\langle \textbf{r}''|{\textrm{e}}^{-{\textrm{i}}tH_{SUSY}/\hbar}|\textbf{r}'\rangle\, .$$ We also note that the iterated resolvent can be expressed as $$\label{gint} g(\zeta)={\textrm{i}}u\int_0^\infty{\textrm{d}}t \exp\left\{-{\textrm{i}}t u (H_D^2-\zeta)\right\}\, .$$ The integral on the right hand side converges for $\textrm{Im}(u\zeta)>0$ and $\zeta\notin\textrm{spec}(H_D^2)$. Namely it converges on the upper half of the $\zeta$ plane if $u>0$, or on the lower half plane if $u<0$. With our choice $u=(2mc^2\hbar)^{-1}$, we consider the integral defined only for $\textrm{Im}(\zeta)>0$. In the coordinate representation, (\[gint\]) takes the form $$\label{gcoordiante} g(\boldsymbol{r}'',\boldsymbol{r}';\zeta)=\frac{{\textrm{i}}}{2mc^2\hbar}\int_0^\infty {\textrm{d}}t\ P_\zeta(\boldsymbol{r}'',\boldsymbol{r}';t)\,,$$ where $$\label{Pzeta} P_\zeta(\boldsymbol{r}'',\boldsymbol{r}';t)=\langle\boldsymbol{r}''|\exp\left\{-({\textrm{i}}t/\hbar)\left(H_{SUSY}-\zeta/2mc^2\right)\right\}|\boldsymbol{r}'\rangle\,,$$ which we shall refer to as the promotor[@Inomata1986]. The promotor for the Hamiltonian $H_{SUSY}$ is the same in form as the propagator for the effective Hamiltonian $H_{eff}(\zeta)=H_{SUSY}-\zeta/2mc^2$. As the propagator is given in terms of Feynman’s path integral, so is the promotor. While the Schrödinger Hamiltonian itself is self-adjoint, the effective Hamiltonian is not. However, the Hamiltonian has to be self-adjoint for the time evolution operator, but path integration of the promotor does not require self-adjointness of the effective Hamiltonian. The corresponding diagonal elements of the iterated resolvent kernel are given by $$\label{gpm2} \begin{array}{rcl} g^\pm(\boldsymbol{r}'',\boldsymbol{r}';\zeta) &=& \langle\boldsymbol{r}''|g^\pm(\zeta)|\boldsymbol{r}'\rangle\\[2mm] &=&\displaystyle\frac{{\textrm{i}}}{2m c^2\hbar}\int_0^\infty {\textrm{d}}t\ P^\pm_\zeta(\boldsymbol{r}'',\boldsymbol{r}';t) \end{array}$$ where $$\label{Ppm} P^\pm_\zeta(\boldsymbol{r}'',\boldsymbol{r}';t)=\langle\boldsymbol{r}''|\exp\{-({\textrm{i}}t/\hbar)H_{eff}^\pm(\zeta)\}|\boldsymbol{r}'\rangle$$ and $$\label{Heff} \begin{array}{rcl} H_{eff}^+(\zeta) &=&\displaystyle \frac{1}{2mc^2}\left(D D^\dag +M_0^2-\zeta\right)\, ,\\[2mm] H_{eff}^- (\zeta) &=& \displaystyle\frac{1}{2mc^2}\left(D^\dag D +M_0^2-\zeta\right)\, . \end{array}$$ One of the merits of utilizing the iterated resolvent is that the kernel of its components can be represented by means of Feynman’s path integral for effective non-relativistic systems. Let ${\cal L}_\zeta^\pm(\dot{\boldsymbol{r}},\boldsymbol{r},t)$ represent the classical Lagrangian associated with the effective Hamiltonian $H_{eff}^\pm(\zeta)$. Then the promotors can be expressed as Feynman’s path integral $$\label{PI} P^\pm_\zeta(\boldsymbol{r}'',\boldsymbol{r}';t)=\int_{\boldsymbol{r}(0)=\boldsymbol{r}'}^{\boldsymbol{r}(t)=\boldsymbol{r}''}{\cal D}\boldsymbol{r} \exp\left\{\frac{{\textrm{i}}}{\hbar}\int_0^t{\textrm{d}}s\ {\cal L}_\zeta^\pm(\dot{\boldsymbol{r}},\boldsymbol{r},s) \right\}.$$ If the system with the effective non-relativistic Lagrangian is path-integrable, the diagonal elements of the iterated resolvent kernel can be determined via (\[gpm2\]). The resolvent kernel $G(\boldsymbol{r}'',\boldsymbol{r}';z)=\langle\boldsymbol{r}''|G(z)|\boldsymbol{r}'\rangle$, or Green’s function, of the Dirac operator is given as the coordinate representation of (\[Goperator\]); namely $$\label{Gexplicit} \begin{array}{l} G(\boldsymbol{r}'',\boldsymbol{r}';z)=\\ ~=\displaystyle \left( \begin{array}{cc} (z+M_0)g^+(\boldsymbol{r}'',\boldsymbol{r}';z^2) & D(\boldsymbol{r}'')g^-(\boldsymbol{r}'',\boldsymbol{r}';z^2) \\ D^\dagger(\boldsymbol{r}'')g^+(\boldsymbol{r}'',\boldsymbol{r}';z^2) & (z-M_0)g^-(\boldsymbol{r}'',\boldsymbol{r}';z^2) \end{array} \right). \end{array}$$ where $D(\boldsymbol{r}'')$ is the operator in the $\boldsymbol{r}''$ representation, satisfying $\langle\boldsymbol{r}''|D|\boldsymbol{r}'\rangle=D(\boldsymbol{r}'')\delta(\boldsymbol{r}''-\boldsymbol{r}')$. If the two non-zero elements of the iterated resolvent kernel are found in closed form, then all the elements of Green’s function must be obtained in closed form. In obtaining the above results, the SUSY-structure is a sufficient but not a necessary condition. The recent approach is applicable even to the case where the Dirac operator is non-supersymmetric, i.e., not of the form (\[HD3\]), insofar as its squared Dirac operator is (block) diagonal. A notable example is the Dirac-Coulomb operator for the Dirac electron in the hydrogen atom problem, which is not in the form (\[HD3\]) but whose square is diagonalizable. The path integral for the Dirac-Coulomb problem has been explicitly and exactly calculated without SUSY[@Kayed1984]. The free Dirac Hamiltonian ========================== The simplest example for the general approach discussed above is that of the free Dirac Hamiltonian for which $D^\dag=D=c\boldsymbol{\sigma}\cdot\boldsymbol{p}$ and $M_0=m_0c^2$. Let $m=m_0$ as this choice is physically most natural. For simplicity, we use $m$ for the electronic mass rather than $m_0$ from now on. First we treat this system by using the path integral representation. The effective Hamiltonian is $$\label{Hfree} H_{eff}^\pm(\zeta)=\frac{\boldsymbol{p}^2}{2m}+\frac{mc^2}{2}-\frac{\zeta}{2mc^2}=\frac{\boldsymbol{p}^2}{2m}-\frac{\mu^2(\zeta)}{2m}$$ with $$\label{kz} \mu^2(\zeta)=\zeta/c^2-m^2c^2,$$ The effective Hamiltonian (\[Hfree\]) is nothing more then that of the Schrödinger Hamiltonian for a free non-relativistic particle of mass $m$ with an additive constant acting on ${\cal H}^\pm=L^2(\mathbb{R}^3)\otimes\mathbb{C}^2$. The effective Lagrangian corresponding to (\[Hfree\]) reads $$\label{Lfree} {\cal L}_\zeta^\pm (\dot{\boldsymbol{r}},\boldsymbol{r},t)= \frac{m}{2}\ \dot{\boldsymbol{r}}^2+\frac{\mu^2(\zeta)}{2m}\, .$$ As is well known, the path integral for (\[Lfree\]) can explicitly be calculated [@Feyn1948; @Feyn1965; @Schulman1981], whose result leads the promotor (\[Pzeta\]) to the form $$\label{Pfree} \begin{array}{rl} P^\pm_\zeta(\boldsymbol{r}'',\boldsymbol{r}';t)=&\displaystyle\left(\frac{m}{2\pi{\textrm{i}}\hbar t}\right)^{3/2}\times\\[4mm] &\displaystyle\exp\left\{ \frac{{\textrm{i}}}{\hbar}\left(\frac{m}{2t}(\boldsymbol{r}''-\boldsymbol{r}')^2+\frac{t}{2m}\ \mu^2(\zeta)\right) \right\}. \end{array}$$ With the help of the integral formula No. 3.478.4 in ref. \[\], ${\rm Re}\ a>0$, ${\rm Re}\ b>0$, $$\int_0^\infty{\textrm{d}}t\ t^{-3/2}\exp\left\{-a/t-bt\right\}=\sqrt{\frac{\pi}{a}}\exp\left\{-2\sqrt{ab}\right\},\quad$$ and assuming a small positive imaginary part of the mass, i.e., ${\rm Im}(m)>0$, we can reduce the resolvent kernel for the squared Dirac Hamiltonian, $\boldsymbol{x}=\boldsymbol{r}''-\boldsymbol{r}'$, $$\label{gfree} g^\pm(\boldsymbol{r}'',\boldsymbol{r}';\zeta)=\frac{1}{4\pi|\boldsymbol{x}|(\hbar c)^2}\exp\{({\textrm{i}}/\hbar)\mu(\zeta)|\boldsymbol{x}|\},$$ from which follows via (\[Gg\]) the resolvent kernel for the free Dirac Hamiltonian $$\label{Gfree} \begin{array}{rl} G(\boldsymbol{r}'',\boldsymbol{r}';z)=&\displaystyle \frac{{\textrm{e}}^{({\textrm{i}}/\hbar)\mu(z^2)|\boldsymbol{x}|}}{4\pi|\boldsymbol{x}|(\hbar c)^2}\times\\[4mm] &\displaystyle \left( {\textrm{i}}\hbar c\frac{{\bm{\alpha}}\cdot\boldsymbol{x}}{|\boldsymbol{x}|^2}+c\mu(z^2)\frac{{\bm{\alpha}}\cdot\boldsymbol{x}}{|\boldsymbol{x}|}+\beta mc^2 +z \right). \end{array}$$ This is the well-known result (see e.g. eq. (1.263) in \[\]). The spectral properties of the free Dirac Hamiltonian can directly be read off from those of the free Schrödinger Hamiltonians $$\label{HS+-free} H_{SUSY}^\pm=\frac{\boldsymbol{p}^2}{2m}+\frac{mc^2}{2}$$ using the plane wave solutions. With $\boldsymbol{k}\in\mathbb{R}^3$ denoting the wave vector the spectrum reads $$\label{phihrfee} \varepsilon =\frac{\hbar^2\boldsymbol{k}^2}{2m}+\frac{mc^2}{2}\, .$$ The free Dirac spectrum $E=\pm\sqrt{m^2c^4+c^2\hbar^2\boldsymbol{k}^2}$ is immediately recovered via (\[EDirac\]). The corresponding positive and negative eigenstates are easily be obtained from the non-relativistic plane-waves solutions via the general relations (\[PsiDexplicit+\]) and (\[PsiDexplicit-\]). The Dirac Oscillator ==================== As a second less trivial example let us consider the so-called Dirac Oscillator [@MosSzc1989; @MorZen1989; @Benitez1990; @Quesne1991] characterized by $D =c{\bm{\sigma}}\cdot(\boldsymbol{p}+{\textrm{i}}m\omega\boldsymbol{r})$ and $M_0=mc^2$. With these definitions the corresponding Schrödinger Hamiltonians (\[HS+-\]) read $$\label{HSDO} H_{SUSY}^\pm=\frac{\boldsymbol{p}^2}{2m}+\frac{m}{2}\omega^2\boldsymbol{r}^2 \mp\left(\frac{2\omega}{\hbar}\boldsymbol{S}\cdot\boldsymbol{L}+\frac{3}{2}\hbar\omega\right) +\frac{mc^2}{2}\, ,$$ where $$\boldsymbol{L}=\boldsymbol{r}\times \boldsymbol{p}\, ,\quad \boldsymbol{S}=(\hbar/2)\boldsymbol{\sigma}\, ,$$ denote the orbital and spin angular momentum operator, respectively. Introducing the Spin-Orbit operator $$\label{K} K=2\boldsymbol{S}\cdot\boldsymbol{L}/\hbar^2+1=\boldsymbol{J}^2/\hbar^2-\boldsymbol{L}^2/\hbar^2+1/4\, ,\quad \boldsymbol{J}=\boldsymbol{L}+\boldsymbol{S}\, ,$$ which obviously commutes with $H^\pm_{SUSY}$, $\boldsymbol{J}^2$, $J_3$ and $\boldsymbol{L}^2$, the above pair of Hamiltonians reads $$\label{HSDOK} H_{SUSY}^\pm=\frac{\boldsymbol{p}^2}{2m}+\frac{m}{2}\omega^2\boldsymbol{r}^2 \mp\left(K+\frac{1}{2}\right)\hbar\omega +\frac{mc^2}{2}\, .$$ The ansatz $$\label{PSIDO} \Psi^\pm_{njm_j\sigma}(\boldsymbol{r})=R^{(\sigma)}_{nl}(r)\varphi^{(\sigma)}_{jm_j}(\theta,\phi)\,,$$ where $\varphi^{(\sigma)}_{jm_j}$ denotes the spin spherical harmonics [@BjorkenDrell1964] (see also Appendix A), will reduce the eigenvalue problem of (\[HSDOK\]) to that of the standard non-relativistic harmonic oscillator with an additional spin-orbit coupling in three dimensions. The radial wave functions $R^{(\sigma)}_{nl}$ are identified with the well-known radial wave function of the harmonic oscillator [@Cardoso2003]. The eigenvalues of (\[HSDOK\]) are explicitly given by $$\label{EHO} \begin{array}{rcl} {\cal E}^+_{n,j,\sigma}&=&\hbar\omega\left[2n+j+1-\sigma(j+1)\right]+mc^2/2\, ,\\ {\cal E}^-_{n,j,\sigma}&=&\hbar\omega\left[2(n+1)+j+\sigma j\right]+mc^2/2\, , \end{array}$$ where $n\in\mathbb{N}_0$ denotes the radial quantum number, the total angular momentum quantum number $j=\frac{1}{2},\frac{3}{2},\ldots$ , takes half odd integers and $\sigma = \pm 1$. Hence, we have $\varepsilon_n={\cal E}^+_{n,j,\sigma} = {\cal E}^-_{n-1,j+1,-\sigma}$. As explicated in Appendix A let us note that $l$, $j$ and $\sigma$ are not independent quantum numbers as they are related by $j=l+\sigma/2$. That is for fixed $j$ we have the relation $R^{(\sigma)}_{nl}=R^{(-\sigma)}_{n,l-\sigma}$. For a detailed discussion of the spectral properties we refer to the work by Quesne [@Quesne1991], here we are only interested in the SUSY ground states. Obviously we have $\varepsilon_0={\cal E}_{0,j,1}^+=mc^2/2$. That is SUSY is unbroken and the ground state energy, which belongs to the spectrum of $H^+_{SUSY}$ only, is infinitely degenerate as it does not dependent on the total angular quantum number $j$. Indeed, observing that $$\label{DDO} D^\dag=c\boldsymbol{\sigma}\cdot\left(\boldsymbol{p}-{\textrm{i}}m \omega\boldsymbol{r}\right) = -{\textrm{i}}\hbar c\boldsymbol{\sigma}\cdot\boldsymbol{e}_r \left(\partial_r+\frac{m\omega}{\hbar}r-\frac{K-1}{r}\right)$$ where $\boldsymbol{r}=r\boldsymbol{e}_r$, one realizes that the radial part of the eigenfunctions $\Psi^+_{0,j,m_j,1}$ with $n=0$ and $\sigma=1$, $$R^{(+1)}_{0l}(r)\propto r^{l}\exp\left\{-\frac{m\omega}{2\hbar^2}r^2\right\}\, ,\quad l=j-1/2\, ,$$ are all annihilated by the operator (\[DDO\]). Finally let us note that the SUSY transformations (\[SUSYtrafogen\]) explicitly read $$\label{DOSUSY} \begin{array}{rcl} D^\dag\Psi^+_{n,j,m_j,1} & = & {\textrm{i}}\sqrt{2mc^2\hbar\omega2n}\Psi^-_{n-1,j,m_j,-1} \\[1mm] D^\dag\Psi^+_{n,j,m_j,-1} & = & -{\textrm{i}}\sqrt{2mc^2\hbar\omega2(n+j+1)}\Psi^-_{n,j,m_j,1} \\[1mm] D\Psi^-_{n,j,m_j,-1} & = & -{\textrm{i}}\sqrt{2mc^2\hbar\omega2(n+1)}\Psi^+_{n+1,j,m_j,1} \\[1mm] D\Psi^-_{n,j,m_j,1} & = & {\textrm{i}}\sqrt{2mc^2\hbar\omega2(n+1+j)}\Psi^+_{n,j,m_j,-1} \end{array}$$ and can be derived by explicit application of (\[DDO\]) using the property (\[A3\]) of appendix A and established recurrence relations of the radial functions of the harmonic oscillator [@Cardoso2003]. Path integral representation for the Dirac Oscillator ----------------------------------------------------- The effective Hamiltonians associated with the Dirac oscillator explicitly read $$\label{HeffDO} {\cal H}_{eff}^\pm(\zeta) =\frac{\boldsymbol{p}^2}{2m}+\frac{m}{2}\omega^2\boldsymbol{r}^2 \mp\left(K+\frac{1}{2}\right)\hbar\omega -\frac{\mu^2(\zeta)}{2m}\,$$ Due to the spherical symmetry of above Hamiltonian the promotor associated with it can be expressed in a partial wave expansion of the form $$\label{PDO} \begin{array}{rl} P^\pm_\zeta(\boldsymbol{r}'',\boldsymbol{r}';t)=& \displaystyle \sum_{j\sigma}P^\pm_{\zeta,l}(r'',r';t)\times\\ &\displaystyle \sum_{m_j=-j}^j\varphi^{(\sigma)}_{jm_j}(\theta'',\phi'')\bar{\varphi}_{jm_j}^{(\sigma)}(\theta',\phi') \end{array}$$ where the radial promotor can be expressed in terms of a radial path integral $$\label{PIDO} P^\pm_{\zeta,l}(r'',r';t)=\int_{r(0)=r'}^{r(t)=r''}{\cal D}r \exp\left\{\frac{{\textrm{i}}}{\hbar}\int_0^t{\textrm{d}}s\ {\cal L}_{\rm eff}^\pm(\dot{r},r,t) \right\}$$ with effective radial Lagrangian, $\kappa=\sigma (j+1/2)$, $$\label{LDO} \begin{array}{rl} {\cal L}_{\rm eff}^\pm(\dot{r},r,t) =&\displaystyle \frac{m}{2}\dot{r}^2-\frac{m}{2}\omega^2r^2-\frac{l(l+1)\hbar^2}{2mr^2}+\\[2mm] &\displaystyle\frac{\mu^2(\zeta)}{2m}\pm\hbar\omega(\kappa+1/2) \end{array}$$ which in essence is that of the radial harmonic oscillator in three dimension and its path integration has explicitly been calculated by Peak and Inomata [@Peak1969; @InomataJunker94] resulting in $$\label{PlDO} P^\pm_{\zeta,\ell}(r'',r';t)=\displaystyle \frac{m\omega(r''r')^{1/2}}{{\textrm{i}}\hbar\sin \omega t} \displaystyle\exp\left\{ \frac{{\textrm{i}}m\omega}{2\hbar}(r''^2+r'^2)\cot\omega t +\frac{{\textrm{i}}t\mu^2_\pm(\zeta)}{2\hbar m} \right\}\,{\rm I}_{\ell+1/2}\left(\case{m\omega r'r''}{{\textrm{i}}\hbar\sin\omega t}\right)$$ with $\mu_\pm^2(\zeta)=\mu^2(\zeta)\pm2m\hbar\omega(\kappa+1/2)$ and ${\rm I}_{\ell +1/2}$ denotes the modified Bessel function. The $t$-integration can be performed using the integral formula (derived from 6.669.4 in \[\]) $$\label{t-integral} \displaystyle \int_0^\infty {\textrm{d}}q\frac{{\textrm{e}}^{2 \nu q}}{\sinh q}\exp\left\{-\case{1}{2}(a+b)t\coth q\right\}{\rm I}_{2\rho}\left(\case{t\sqrt{ab}}{\sinh q}\right) \displaystyle = \frac{\Gamma(1/2+\mu-\nu)}{t\sqrt{ab}\Gamma(1+2\mu)}\,{\rm W}_{\nu,\rho}(at){\rm M}_{\nu,\rho}(bt)$$ with integrability conditions $a>b$, ${\rm Re}(1/2+\rho-\nu)>0$ and ${\rm Re}\ t>0$. The explicit integration yields $$\label{glDO} \displaystyle \int_0^\infty {\textrm{d}}t\ P^\pm_{\zeta,\ell}(r'',r',t)= \frac{(r''r')^{-1/2}}{{\textrm{i}}\omega}\frac{\Gamma(\rho-\nu^\pm+1/2)}{\Gamma(1+2\rho)} {\rm W}_{\nu^\pm ,\rho}(r_>^2 m\omega/\hbar){\rm M}_{\nu^\pm ,\rho}(r_<^2 m\omega/\hbar)$$ where $\rho=\ell/2+1/4$, $\nu^\pm(\zeta)=\mu_\pm^2(\zeta)/4m\omega\hbar$, $r_>=\max\{r'',r'\}$, $r_<=\min\{r'',r'\}$ and ${\rm W}_{\rho,\nu}$ and ${\rm M}_{\rho,\nu}$ are the Whittaker functions. We finally arrive at the iterated Green’s function for the Dirac oscillator $$\label{gDO} g^\pm(\boldsymbol{r}'',\boldsymbol{r}';\zeta)=\displaystyle \sum_{j\sigma}g^\pm_{\ell}(r'',r';\zeta) \displaystyle\sum_{m_j=-j}^j\varphi_{jm_j}^{(\sigma)}(\theta'',\phi'')\bar{\varphi}_{jm_j}^{(\sigma)}(\theta',\phi')$$ with $$\label{glDO2} g^\pm_{\ell}(r'',r';\zeta)=\displaystyle\frac{\Gamma(\rho-\nu^\pm(\zeta)+1/2)}{2mc^2\hbar\omega(r''r')^{1/2}\Gamma(1+2\rho)} {\rm W}_{\nu^\pm,\ \rho}(r_>^2 m\omega/\hbar){\rm M}_{\nu^\pm,\ \rho}(r_<^2 m\omega/\hbar)\,,$$ and $\bar{\varphi}$ denoting the complex conjugated and transposed spinor of $\varphi$. From the poles of the Green’s function in the complex $z$-plane we may deduce the spectrum. These poles occur in the Gamma function of the numerator of the iterated radial Green’s function (\[glDO2\]), that is, for $$\label{DOpoles} \nu^\pm_n(\zeta)=n+\rho+1/2,\quad n\in \mathbb{N}_0\,.$$ Hence the poles occur at $$\label{DOSpec} \textstyle \zeta^\pm=m^2c^4+2mc^2\hbar\omega\left[2n+j+\case{3}{2}\pm\case{1}{2}\mp\sigma\left(j+\case{1}{2}\right)-\case{\sigma}{2}\right]$$ and the spectrum of the Dirac oscillator explicitly reads $$\begin{array}{l} E_{n,j,\sigma}^+ = mc^2\sqrt{1+\frac{2\hbar\omega}{mc^2}\left[2n+(j+1)-\sigma(j+1)\right]}\, ,\\[2mm] E_{n,j,\sigma}^- = -mc^2\sqrt{1+\frac{2\hbar\omega}{mc^2}\left[2(n+1)+j+\sigma j\right]}\,. \end{array}$$ Obviously we have $E^-_{n-1,j+1,-\sigma}=-E^+_{n,j,\sigma}$, which implies that the spectrum is symmetric about the origin with the exception of the ground state energy $E^+_{0,j,1}=mc^2$ which belongs only to the positive part of the spectrum. Note that the spectrum can also be obtained directly from (\[EHO\]) using the established relation (\[EDirac\]). A generalisation of the Dirac Oscillator ======================================== Let us now consider a generalisation of the Dirac oscillator characterised by $$\label{gDOU} D=c\boldsymbol{\sigma}\cdot\left(\boldsymbol{p}+{\textrm{i}}\hbar\boldsymbol{\nabla}U\right)\, , \quad M_0=mc^2\,.$$ Here the “superpotential” $U$ is initially assumed to be an arbitrary function on $\mathbb{R}^3$. The corresponding Schrödinger Hamiltonians now read $$\label{HSU} H_{SUSY}^\pm=\frac{\boldsymbol{p}^2}{2m}+\frac{\hbar^2}{2m}\left( \left(\boldsymbol{\nabla}U\right)^2\mp\Delta U\right) \mp\frac{2}{m}\boldsymbol{S}\cdot\left(\boldsymbol{\nabla}U\times\boldsymbol{p}\right)\,.$$ Obviously SUSY is unbroken if one of the functions $$\label{Psi0U} \Psi^\pm_{0,\lambda}(\boldsymbol{r})\propto\exp\{\mp U(\boldsymbol{r})\}\chi_\lambda$$ belongs to the Hilbert space ${\cal H}^\pm$, respectively. In the above $\chi_\lambda$ is an arbitrary constant 2-spinor. Note, that both $\Psi^+_{0,\lambda}$ and $\Psi^-_{0,\lambda}$ cannot be simultaneous ground states as, for example, a rapidly increasing superpotential $U(\boldsymbol{r})\to \infty$ for $\boldsymbol{r}\to \infty$ will lead to an normalizable $\Psi^+_{0,\lambda}$ but then $\Psi^-_{0,\lambda}$ is not square integrable. In order to be a bid more explicit let us consider that the superpotential is a spherical symmetric function. In that case $\boldsymbol{\nabla}U=U'(r)\boldsymbol{e}_r$ and thus $U'$ describes a so-called tensor potential [@Furnstahl1998]. The quadratic superpotential $U(r) = (m\omega/2\hbar)r^2$ gives rise to a linear tensor potential and characterises the previously discussed Dirac Oscillator. For a recent discussion on generalisation of the Dirac oscillator see, for example, [@Akcay2009; @Zarrin2010]. For a general spherical symmetric $U$ the corresponding Schrödinger Hamiltonians take the form $$\label{gU} H^\pm_{SUSY} =\frac{\boldsymbol{p}^2}{2m}+\frac{\hbar^2}{2m} \left(U'\,^2(r)\mp U''(r)\right)\mp\frac{\hbar^2U'(r)}{mr}K.$$ The adjoint of above Dirac operator (\[gDOU\]) reduces to $$\label{gDDO} D^\dag = -{\textrm{i}}\hbar c\boldsymbol{\sigma}\cdot\boldsymbol{e}_r \left(\partial_r-\frac{K-1}{r}+U'(r)\right)$$ and the condition for unbroken SUSY, that is $D^\dag\Psi^+_{0,\lambda}=0$ with $\Psi^+_{0,\lambda}=R^{(+1)}_{0l}\varphi^{(+1)}_{jm_j}$ leads to a radial wave function of the form $$R^{(+1)}_{0l}\propto r^{l}\exp\left\{-U(r)\right\} ,\quad l=j-1/2\, ,$$ which becomes normalizable if the potential $U$ is well-behaved at the origin and increases rapidly enough for large $r$. Note that for $\sigma = -1$ there exists no normalizable SUSY ground state. Let us also mention that for a rapidly decreasing superpotential, $U(r)\to -\infty$ as $r\to \infty$, SUSY is also unbroken with ground state $\Psi^-_{0,\lambda}=R^{(-1)}_{0l}\varphi^{(-1)}_{jm_j}$ belonging to the sector $\sigma = -1$ and $$R^{(-1)}_{0l}\propto r^{l+1}\exp\left\{U(r)\right\} ,\quad l=j+1/2\, .$$ In either case the SUSY ground state is infinitely degenerate as in the special case of the Dirac oscillator. The case of a linear superpotential ----------------------------------- Let us consider as a simple non-trivial case that of a linear superpotential $$\label{linU} U(r) = \gamma r\, , \quad \gamma >0\, .$$ which leads to the following pair of Schrödinger Hamiltonians $$\label{gDOH} { H}^\pm _{SUSY} =\frac{\boldsymbol{p}^2}{2m}+\frac{\hbar^2}{2m} \left(\gamma^2 \mp\frac{2\gamma}{r}K\right)\, .$$ As the eigenvalues $\kappa=\sigma(j+1/2)$ of the spin-orbit operator $K$ are non-zero integers, the above Hamiltonian ${ H}_{SUSY}^+$ characterizes a quantum particle under the influence of a Coulomb-like spin-orbit interaction which is attractive for the positive eigenvalues of $K$ where spin and orbital angular momentum are aligned, $\sigma = +1$, and repulsive for the case $\sigma = -1$. The situation for ${H}_{SUSY}^-$ is just the opposite. Let’s first consider the subspace $\sigma = +1$ and make the following ansatz for the eigenvalue problem $$\label{linplus} {H}_{SUSY}^\pm R_{nl}^{(+)}\varphi^{(+)}_{jm_j}={\cal E}^\pm_{n,j,+1} R_{nl}^{(+)}\varphi^{(+)}_{jm_j}\,.$$ As in this subspace the spin-orbit operator takes the positive eigenvalues $\ell + 1$ this eigenvalue problem is reduced to that of the standard non-relativistic Coulomb problem $$\label{Coul+} \begin{array}{r} \displaystyle \left( \frac{\boldsymbol{p}^2}{2m}\mp\frac{\hbar^2\gamma(l +1)}{mr}+\frac{\hbar^2\gamma^2}{2m} \right) R_{nl}^{(+)}\varphi^{(+)}_{jm_j}=~~~~~\\ {\cal E}^\pm_{n,j,+1} R_{nl}^{(+)}\varphi^{(+)}_{jm_j} \end{array}$$ with effective charge $e^2=\hbar^2\gamma(l + 1)/m$. Hence the eigenvalue problem is immediately solved when the $R^{(\sigma)}_{nl}$ are taken to be the radial eigenfunction for the Coulomb problem. With the well-known eigenvalues of the non-relativistic Coulomb problem $-(me^4/2\hbar^2)(n+l+1)^{-2}$ the discrete eigenvalues are obtained $$\label{Edisc+} {\cal E}^+_{n,j,+1}=\frac{\hbar^2\gamma^2}{2m}\left(1-\left(\frac{\ell + 1 }{n+\ell+1}\right)^2\right)\, ,\quad n\in\mathbb{N}_0.$$ The corresponding continuous spectrum of ${ H}_{SUSY}^+$ coincides with the full spectrum of ${H}_{SUSY}^-$ and may be written as $$\label{Econt} {\cal E}^\pm_{\boldsymbol{k},+1}=\frac{1}{2m}\left(\hbar^2\boldsymbol{k}^2+\hbar^2\gamma^2\right)\, , \boldsymbol{k}\in\mathbb{R}^3\, .$$ In a similar manner one finds the discrete spectrum for the $\sigma = -1$ subspace where $l=j+1/2\neq 0$ $$\label{Edisc-} {\cal E}^-_{n,j,-1}=\frac{\hbar^2\gamma^2}{2m}\left(1-\left(\frac{\ell}{n+1+\ell}\right)^2\right)\, .$$ Obviously we have recovered again the relation $\varepsilon={\cal E}^+_{n,j,+1} = {\cal E}^-_{n+1,j,-1}$ as expected. This finally leads to the discrete eigenvalues of the Dirac operator as follows $$\label{Elin+} \textstyle E_{n,j,+1}^+=mc^2\left[1+\frac{\hbar^2\gamma^2}{m^2c^2}\left(1-\left(\frac{j+1/2}{n+j+1/2}\right)^2\right)\right]^{1/2}\, ,$$ $$\label{Elin-} \textstyle E_{n,j,-1}^-=-mc^2\left[1+\frac{\hbar^2\gamma^2}{m^2c^2}\left(1-\left(\frac{j+1/2}{n+j+3/2}\right)^2\right)\right]^{1/2} \, .$$ The continuous spectrum is given by $$\label{Elinc} E^\pm_{\boldsymbol{k}}=\pm mc^2\left[1+\frac{\hbar^2\gamma^2}{m^2c^2} + \frac{\hbar^2\boldsymbol{k}^2}{m^2c^2}\right]^{1/2}\,.$$ In addition the positive and negative Dirac eigenstates for the linear superpotential directly follow from those of the non-relativistic Coulomb problem via the general relations established in Section II. Finally, let us note that a logarithmic superpotential of the form $U(r)=\gamma\ln(r/r_0)$ also leads to an exactly solvable SUSY Hamiltonian. Path integral representation for the case of a linear superpotential -------------------------------------------------------------------- Similar to the previous discussion on the Dirac oscillator the promotor can be expressed in terms of partial waves (cf. eq. (\[PDO\]) and (\[PI\])) with effective radial Lagrangian $$\label{Llin} {\cal L}_{\rm eff}^\pm(\dot{r},r) =\frac{m}{2}\dot{r}^2-\frac{l(l+1)\hbar^2}{2mr^2}-\frac{\hbar^2}{2m}\left(\gamma^2\pm\frac{2\gamma\kappa}{r}\right)+\frac{\mu^2(\zeta)}{2m}$$ For the $1/r$-potential the path integral cannot directly be evaluated. However, with the help of the so-called local space time transformation [@Inomata1986; @Junker1990; @Inomata1992] it may be reduced to that of a harmonic potential. To be more explicit, let’s introduce a new radial variable $s=\sqrt{r}$ and a new time $\tau$ with ${\textrm{d}}t = 4s^2{\textrm{d}}\tau$ the integral expression (\[gpm2\]) for the radial Green’s function (\[gDO\]) can be put into the form $$\label{gLin} \begin{array}{rl} g^\pm_\ell(r',r',\zeta) &\displaystyle =\frac{{\textrm{i}}}{2mc^2\hbar}\int_0^\infty{\textrm{d}}t P_{\zeta\ell}^\pm(r'',r',t)\\[4mm] &\displaystyle =\frac{{\textrm{i}}\sqrt{r''r'}}{mc^2\hbar}\int_0^\infty{\textrm{d}}\tau \tilde{P}_{\zeta\ell}^\pm(s'',s',\tau) \end{array}$$ where $$\label{PIlin2} \tilde{P}_{\zeta\ell}^\pm(s'',s',\tau)=\int_{s'=s(0)}^{s''=s(\tau)}{\cal D}s \exp\left\{\frac{{\textrm{i}}}{\hbar}\int_0^\tau{\textrm{d}}\tau\tilde{\cal L}^\pm_{eff}\right\}$$ now represents a path integral for a harmonic oscillator in the new space-time co-ordinates as the new effective Lagrangian reads $$\label{Llin2} \tilde{\cal L}^\pm_{eff} = \frac{m}{2}\left(\frac{{\textrm{d}}s}{{\textrm{d}}\tau}\right)^2 -\frac{\Lambda(\Lambda+1)\hbar^2}{2ms^2} -\frac{m}{2}\Omega^2s^2 \pm\frac{4\hbar^2\gamma\kappa}{m}$$ where $\Lambda = 2\ell+1/2$ and $\Omega=(2/m)\sqrt{\hbar^2\gamma^2-\mu^2(\zeta)}$. Hence we immediately obtain the result $$\label{Pl} \begin{array}{rl} \tilde{P}^\pm_{\zeta,\ell}(s'',s',\tau)= &\displaystyle \frac{m\Omega(s''s')^{1/2}}{{\textrm{i}}\hbar\sin \Omega \tau}{\rm I}_{\Lambda+1/2}\left(\case{m\Omega s's''}{{\textrm{i}}\hbar\sin\Omega \tau}\right)\times\\ &\displaystyle \exp\left\{ \frac{{\textrm{i}}m\Omega}{2\hbar}(s''^2+s'^2)\cot\Omega t \pm \frac{4{\textrm{i}}\hbar \gamma \kappa}{m}\tau \right\} \end{array}$$ which obviously is identical in form with that of the Dirac oscillator discussed in section V subsection A. Hence we can directly perform the $\tau$-integration using formula (\[t-integral\]) resulting in $$\label{glDO3} \begin{array}{rl} g^\pm_{\ell}(r'',r',\zeta)=&\displaystyle \frac{(r''r')^{1/4}}{mc^2\hbar\Omega}\frac{\Gamma(\rho-\nu^\pm+1/2)}{\Gamma(1+2\rho)}\times\\[4mm] &{\rm W}_{\nu^\pm,\rho}(r_> m\Omega/\hbar){\rm M}_{\nu^\pm,\rho}(r_< m\Omega/\hbar). \end{array}$$ where $\rho=\Lambda/2+1/4=\ell+1/2$ and $\nu^\pm=\pm 2\hbar\gamma\kappa/\Omega m$. Again the poles of $g^\pm_\ell$ will lead to the eigenvalues of the associated Dirac Hamiltonian and occur at the poles of the Gamma function in the numerator, that is, at $\nu^\pm = n+\ell+1$ with $n\in\mathbb{N}_0$. For the positive energy sector they can only occur for positive $\kappa=\ell+1$ that is $\sigma = 1$ and with $\nu^+=2\hbar\gamma(\ell+1)/\Omega m$ the final result reads $$E^+_{n,\ell , 1}=\sqrt{m^2c^4+\hbar^2c^2\gamma^2\left(1+\left(\frac{\ell+1}{n+\ell+1}\right)^2\right)}.$$ For the negative energy sector the occurrence of poles requires that $\kappa = -\ell < 0$, $\sigma =-1$ and leads to $$E^-_{n,\ell ,-1}=-\sqrt{m^2c^4+\hbar^2c^2\gamma^2\left(1+\left(\frac{\ell}{n+\ell+1}\right)^2\right)}.$$ Finally let us note that the Green’s function also leads to the continuous spectrum of the system. In fact, the branch cut occurring in the definition of $\Omega$ when taken along the negative half line can be parameterized by $\hbar^2\gamma^2-\mu^2(\zeta)=-\lambda^2$ with $\lambda\in\mathbb{R}_+$ and directly leads to $$E^\pm_{\boldsymbol{k}}=\pm\sqrt{m^2c^4+\hbar^2\gamma^2c^2+\hbar^2\boldsymbol{k}^2c^2}\, ,$$ where we have identified $\lambda = \hbar|\boldsymbol{k}|$. These eigenvalues are indeed identical to those given in (\[Elin+\]), (\[Elin-\]) and (\[Elinc\]) by noting the relation $j=l+\sigma/2$. It is worth remarking that a logarithmic superpotential of the form $U(r)=\gamma\ln(r/r_0)$ as mentioned above the path integration can be done in closed form, too. Conclusions =========== In this paper we have studied the resolvent kernel, or Green’s function, of supersymmetric Dirac Hamiltonians. It has been shown that the iterated resolvent $g$ of the squared Dirac Hamiltonian allows for an explicit path-integral representation of Feynman’s type. This is due to the fact that squared Dirac Hamiltonian becomes block diagonal and each block is represented by an effective Hamiltonian which is of the form of a non-relativistic Schrödinger Hamiltonian. The final resolvent kernel $G$ is then simply obtained by differentiation, cf. eq. (\[Gexplicit\]). We have explicitly treated the free particle case leading to a path integral of the non-relativistic free particle; the Dirac oscillator whose path integral representation was that of a non-relativistic harmonic oscillator with an additional spin-orbit coupling; and a generalisation of the Dirac oscillator where we have studied the special case of a linear superpotential leading us to a non-relativistic path integral of the Coulomb type, which can explicitly be calculated. The path integral representation may also be useful in obtaining semi-classical approximations of a supersymmetric Dirac Hamiltonian, for example, in case were the path integral is not exactly solvable. We have also shown in this paper that once the spectral properties, i.e. the eigenvalues and eigenfunctions, of the associated non-relativistic SUSY partner Hamiltonians $H^\pm_{SUSY}$ are explicitly known, the spectral properties of the Dirac Hamiltonian immediately follow. The eigenvalues $E$ of the Dirac Hamiltonian are explicitly given in terms of the eigenvalues $\varepsilon$ of the non-relativistic $H^\pm_{SUSY}$ via the simple relation (\[EDirac\]). The corresponding eigenstates are given by those of $H^\pm_{SUSY}$ via the relation (\[PsiDexplicit+\]) and (\[PsiDexplicit-\]). In the main text we have only considered systems with unbroken SUSY. However, the general approach is valid for good and broken SUSY. In Appendix B we briefly realise the free Dirac Hamiltonian in the so-called supersymmetric representation of the Dirac matrices leading to a broken SUSY as discussed in the general approach. The resulting partner Hamiltonians (\[HSbroken\]) are identical to those of the unbroken realisation (\[HS+-free\]) discussed in the main text. Our examples were all based on a Dirac Hamiltonian in three dimensions but the discussion of sections II and III are also applicable to Dirac Hamiltonians in one and two dimensions in an obvious way. Appendix C briefly discusses the one-dimensional Dirac oscillator and its generalization, the relativistic Witten model. Further examples of supersymmetric Dirac Hamiltonians are briefly discussed in the textbook by Thaller [@TH1992]. In particular the Dirac particle in an external magnetic field, cf. eq. (\[HD\]), exhibits SUSY and its corresponding partner Hamiltonians are given by the usual Pauli-Hamiltonian. Hence, whenever that Pauli problem is solvable, as for example for a constant magnetic field, the corresponding Dirac problem is also solved. Finally let us mention that our approach also applies to radial Dirac Hamiltonians in flat space as well as in central back grounds like de Sitter and anti-de Sitter space [@Cotaescu99]. It might also be applied to cases were no exact SUSY structure and Foldy-Wouthuysen transformation exists using proper non-relativistic approximations as recently discussed by Jentschura and Noble [@JN14]. Spin Spherical Harmonics ======================== In this appendix we briefly review the basic properties of the spin spherical harmonics following Bjorken and Drell [@BjorkenDrell1964] (see also the textbook by Thaller [@TH1992]) $$\label{RSH} \varphi^{(\sigma )}_{jm_j}(\theta,\phi) = \left( \begin{array}{c} \sqrt{\frac{l+1/2+\sigma m_j}{2l+1}}\,Y_l^{m_j-1/2}(\theta,\phi) \\ \sigma \sqrt{\frac{l+1/2-\sigma m_j}{2l+1}}\,Y_l^{m_j+1/2}(\theta,\phi) \end{array} \right)\, ,$$ where $Y_l^{m_l}$ denotes the usual spherical harmonics defined on the unit sphere $S^2$. In the above definition the total and orbital angular momentum quantum numbers $j$ and $l$ are not independent and obey the relation $j=l+\sigma/2$, where $\sigma = \pm 1$. Note that $l\in\mathbb{N}_0$ and $l=0$ is only allowed for the case $\sigma = +1$. The $\varphi^{(\sigma)}_{jm_j}$ are simultaneous eigenfunctions of $\boldsymbol{J}^2$, $J_3$, $\boldsymbol{L}^2$, and the spin orbit operator $K$. $$\label{varphi} \begin{array}{rcll} \boldsymbol{J}^2 \varphi^{(\sigma)}_{jm_j} & = & j(j+1)\hbar^2 \varphi^{(\sigma)}_{jm_j}\, ,\quad & j\in\{\frac{1}{2},\frac{3}{2},\frac{5}{2}\ldots \}\, ,\\[1mm] J_3 \varphi^{(\sigma)}_{jm_j} & = &m_j\hbar \varphi^{(\sigma)}_{jm_j} \, ,\quad & -j \leq m_j \leq j\, ,\\[1mm] \boldsymbol{L}^2 \varphi^{(\sigma)}_{jm_j} &=& l(l+1)\hbar^2 \varphi^{(\sigma)}_{jm_j}\, ,\quad & j=l+\sigma/2 \, , \\[1mm] K \varphi^{(\sigma)}_{jm_j} &= & \kappa\varphi^{(\sigma)}_{jm_j}\,,&\kappa=\sigma(j+1/2)\, . \end{array}$$ They form a complete orthogonal set on the Hilbert space $L^2(S^2)\otimes\mathbb{C}^2$ and thus are most suitable for solving eigenvalue problems of spherical symmetric Hamiltonians like $H_{SUSY}^\pm$ as discussed in the main text. Finally let us mention that they obey the relation [@BjorkenDrell1964] $$\label{A3} \boldsymbol{\sigma}\cdot\boldsymbol{e}_r \varphi^{(\sigma)}_{jm_j} = \varphi^{(-\sigma)}_{jm_j}\,.$$ Also note that in the literature the eigenvalues of the spin-orbit operator are sometimes denoted by $-\kappa$ and the operator is multiplied by $\hbar^2$. For simplicity we prefer a dimensionless $K$ with eigenvalues denoted by $\kappa$. Finally let us mention the obvious relation $$\label{A4} \boldsymbol{\sigma}\cdot\boldsymbol{p} = -{\textrm{i}}\hbar( \boldsymbol{\sigma}\cdot\boldsymbol{e}_r)\left(\partial_r-\frac{K-1}{r}\right) \,,$$ which was used several times in the main text. Free Dirac Hamiltonian with broken SUSY ======================================= Despite the fact that in the main text we only considered systems with unbroken SUSY the general approach is clearly valid for unbroken and broken SUSY. As a simple but instructive example with broken SUSY let us consider again the free Dirac Hamiltonian $$\label{Hfreebroken} H_D=c{\bm{\alpha}}\cdot\boldsymbol{p}-\beta mc^2\,.$$ Using the so-called supersymmetric representation of the Dirac matrices [@TH1992] $$\label{alpha&betaSUSY} {\bm{\alpha}}_s=\left( \begin{array}{cc} 0 & {\bm{\sigma}}\\ {\bm{\sigma}}& 0 \end{array} \right)\, ,\quad \beta_s=\left( \begin{array}{cc} 0 & -{\textrm{i}}\\ {\textrm{i}}& 0 \end{array} \right)\, ,$$ the free Dirac Hamiltonian takes the explicit form $$\label{HfreeBroken} H_D=\left( \begin{array}{cc} 0 & c{\bm{\sigma}}\cdot\boldsymbol{p}-{\textrm{i}}mc^2 \\ c{\bm{\sigma}}\cdot\boldsymbol{p}+{\textrm{i}}mc^2 & 0 \end{array} \right)\, .$$ Hence it is a supersymmetric Dirac operator with $D=c{\bm{\sigma}}\cdot\boldsymbol{p}-{\textrm{i}}mc^2$ and $M_0=0$. In contrast to the discussion in section III here, the supercharge is not self-adjoint but leads to the same SUSY partner Hamiltonians $$\label{HSbroken} H^\pm_{SUSY}=\frac{1}{2mc^2}\left(c^2\boldsymbol{p}^2+m^2c^4\right)\, .$$ However, here $\varepsilon_0=M^2_0/2mc^2=0$ and hence cannot belong to the spectrum of above SUSY Hamiltonians, that is, SUSY is broken for this choice of supercharge. Nevertheless, $H_{SUSY}^\pm$ share the same plan-wave eigenstates as in the unbroken SUSY case with same eigenvalues given by (\[phihrfee\]). One-Dimensional Dirac Operators =============================== The free Dirac operator in one dimension is represented by $2\times 2$ Pauli matrices and explicitly reads $$H_D = c\sigma_1 p + \sigma_3 mc^2 = \left(\begin{array}{cc} mc^2 & cp \\ cp & -mc^2 \end{array} \right)$$ acting on ${\cal H}=L^2(\mathbb{R})\otimes\mathbb{C}^2$. As in the three-dimensional case we may define a generalized Dirac oscillator via the non-minimal substitution $p\rightarrow p-{\textrm{i}}\hbar U'(x)\sigma_3$, where $U:\mathbb{R}\mapsto\mathbb{R}$ denotes the superpotential, giving rise to the standard non-relativistic SUSY Hamiltonians of the Witten model shifted by $\varepsilon_0=mc^2/2$, $$\label{HSgen1D} H^\pm_{SUSY}=\frac{p^2}{2m}+\frac{\hbar^2}{2m}\left(U'\,^2(x)\mp U''(x)\right)+\frac{mc^2}{2}\, ,$$ Hence, whenever the eigenvalue problem of such a non-relativistic SUSY Hamiltonian can be solved one immediately has the solution for the Dirac problem. Indeed, let $\epsilon_n=\varepsilon_n-mc^2/2$ denote the eigenvalues of the unshifted Witten model $H^\pm_{SUSY}-\varepsilon_0$ we immediately have the Dirac spectrum $$\label{DiracWittenSpec} E_n^\pm=\pm mc^2\sqrt{1+\frac{2\epsilon_n}{mc^2}}\,.$$ The corresponding eigenstates are found similarly via relation (\[PsiD\]). In the non-relativistic limit we recover the spectrum of the Witten model, $$\label{NRLimit} \lim_{c\to\infty}\left(E^+_n - mc^2\right)=\epsilon_n\,$$ confirming that the one-dimensional Dirac operator $$\label{DiracWitten} H_D=\left( \begin{array}{cc} mc^2& c\left(p+{\textrm{i}}\hbar U'(x)\right) \\ c\left(p-{\textrm{i}}\hbar U'(x)\right) & -mc^2 \end{array} \right)$$ is indeed the relativistic version of Witten’s non-relativistic SUSY model characterised by the SUSY partner Hamiltonians (\[HSgen1D\]). The special case $U(x)=\frac{m\omega}{2\hbar}x^2$ leads to the one-dimensional Dirac oscillator Hamiltonian $$\label{DO1D} H_D = \left(\begin{array}{cc} mc^2 & c(p+{\textrm{i}}m\omega x)\\ c(p-{\textrm{i}}m\omega x) & -mc^2 \end{array} \right)\,.$$ which obviously is a supersymmetric Dirac Operator with $D=c(p+{\textrm{i}}m\omega x)$, $M_0=mc^2$ and the SUSY Hamiltonians read $$\label{HSDO1D} H^\pm_{SUSY}=\frac{p^2}{2m}+\frac{m}{2}\omega^2 x^2\mp \frac{\hbar\omega}{2}+\frac{mc^2}{2}\, ,$$ coinciding, except of the trivial additive constant $mc^2/2$, with the supersymmetric harmonic oscillator in Witten’s non-relativistic model [@Junker1996]. Here SUSY is unbroken as the ground state energy $\varepsilon_0=M^2_0/2mc^2=mc^2/2$ and belongs to the spectrum of $H^+_{SUSY}$. The positive eigenvalues are $\varepsilon_n =\hbar\omega n+\varepsilon_0$, $n=1,2,3,\ldots$, and the corresponding eigenstates are given by $\Psi^+_n(x)=\langle x|n\rangle$ and $\Psi^-_n(x)=\langle x|n-1\rangle$ where $|n\rangle$ represents the standard $n$-th eigenstate of the one-dimensional harmonic oscillator. Finally, the complete spectrum of the one-dimensional Dirac oscillator reads $$\label{DO1DSpec} E_0^+ = mc^2\, ,\quad E^{\pm}_n =\pm mc^2\sqrt{1+2n\frac{\hbar\omega}{mc^2}}\, ,\quad n\in \mathbb{N}\,.$$ Let us note that the positive part of this spectrum coincides with that of the three-dimensional Dirac operator when $\sigma = 1$ and the negative part coincides with the negative part of the three-dimensional case if $\sigma = -1$ is chosen. For a detailed study of the dynamical behavior for this model we refer to ref. \[\], which compares the dynamics generated by above $H_D$ with that generated by the corresponding diagonalized Hamiltonian $H_{\rm FW}$. [10]{} H. Nicolai, [J. Phys. A]{}[**9**]{}, 1497 (1976). E. Witten, [Nucl. Phys. B]{} [**188**]{}, 513 (1981). G. Junker, [*Supersymmetric Methods in Quantum and Statistical Physics*]{}, (Springer, Berlin, 1996). R. Jackiw, [ Phys. Rev. D]{} [**29**]{}, 2375 (1984). R.J. Hughes, V.A. Kostelecký and M.M. Nieto, Phys. Rev. D [**34**]{}, 1100 (1986). F. Cooper, A. Khare, R. Musto and A. Wipf, [Ann. Phys. ]{} [**187**]{}, 1 (1988). B. Thaller, J. Math. Phys. [**29**]{}, 249 (1988). J. Beckers and N. Debergh, [Phys. Rev. D]{} [**42**]{}, 1255 (1990). B. Thaller, in A. Boutet de Monvel, P. Dita, G. Nenciu and R. Purice (eds), [*Recent Developments in Quantum Mechanics*]{}, Mathematical Physics Studies (A Supplementary Series to Letters in Mathematical Physics), vol 12, (Springer, Dordrecht, 1991) 351. B. Thaller, [*The Dirac Equation*]{}, (Springer, Berlin, 1992). T. Grover, D.N. Sheng and A. Vishwanath, Science [**344**]{}, 280 (2014). S.D. Sarma, S. Adam, E.H. Hwang and E. Rossi, Rev. Mod. Phys. [**83**]{} 407 (2011). M. Ezawa, Phys. Lett. A [**372**]{}, 924 (2008). R.P. Feynman, [Rev. Mod. Phys. ]{}[**20**]{}, 367 (1948). R.P. Feynman and A.R. Hibbs, [*Quantum Mechanics and Path Integrals*]{}, (McGraw-Hill, New York, 1965). A. Inomata, in M.C. Gutzwiller, A. Inomata, J.R. Klauder and L. Streit eds., [*Path Integrals from meV to MeV*]{}, (World Scientific, Singapore, 1986). M.A. Kayed and A. Inomata, [Phys. Rev. Lett.]{} [**53**]{}, 107 (1984). L.S. Schulman, [*Techniques and Applications of Path Integration*]{}, (John Wiley & Sons, New York, 1981). I.S. Gradsteyn and I.M Ryzhik, [*Table of Integrals, Series, and Products*]{}, Edited by A. Jeffrey and D. Zwillinger, (Academic Press, New York,2007) 7th ed. M. Moshinsky and A. Szczepaniak, [J. Phys. A]{}[**22**]{}, L817 (1989). M. Morena and A. Zentella, [J. Phys. A]{}[**22**]{}, L821 (1989). J. Benítez, R.P. Martínez y Romero, H.N. Núñez-Yépez and A.L. Salas-Brito, [Phys. Rev. Lett. ]{}[**64**]{}, 1643 (1990). C. Quesne [ Int. J. Mod. Phys. A]{} [**6**]{}, 1567 (1991). J.D. Bjorken and S.D. Drell, [*Relativisitc Quantum Mechanics*]{}, (McGraw-Hill, New York, 1964). J.L. Cardoso and R. Álvarez-Nodarse, [J. Phys. A]{}[**36**]{}, 2055 (2003). D. Peak and A. Inomata, [J. Math. Phys. ]{}[**10**]{}, 1422 (1969). A. Inomata and G. Junker, in R. Wilson and E.A. Tanner eds., [ *Noncompact Lie groups and their physical applications*]{}, (Kluwer, Dordrecht, 1994) 199. R.F. Furnstahl, J.J. Rusnal and B.D. Serot, Nucl. Phys. A [**632**]{}, 607 (1998). H. Akcay, Phys. Lett. A [**373**]{}, 616 (2009). S. Zarrinkamar, A.A. Rajabi and H. Hassanabadi, Ann. Phys. [**325**]{}, 2522 (2010). G. Junker, [J. Phys. A]{}[**23**]{}, L881 (1990). A. Inomata, H. Kuratsuji and C.C. Gerry, [*Path Integrals and Coherent States of SU(2) and SU(1,1)*]{} (World Scientific, Singapore, 1992). I.I. Cotăescu, Phys. Rev. [**D60**]{}, (1998). U.D. Jentschura and J.H. Noble, [J. Phys. A]{}[**47**]{}, 045402 (1990). F.M. Toyama, Y. Nogami and F.A.B. Coutinho, [J. Phys. A]{}[**30**]{}, 2585 (1997).
{ "pile_set_name": "ArXiv" }
--- abstract: | Using a recently developed piecewise flat method, numerical evolutions of the Ricci flow are computed for a number of manifolds, using a number of different mesh types, and shown to converge to the expected smooth behaviour as the mesh resolution is increased. The manifolds were chosen to have varying degrees of homogeneity, and include Nil and Gowdy manifolds, a three-torus initially embedded in Euclidean four-space, and a perturbation of a flat three-torus. The piecewise flat Ricci flow of the first two are shown to converge to analytic and numerical partial differential equation solutions respectively, with the remaining two flowing asymptotically to flat metrics. [[**Keywords:**]{} Piecewise linear, geometric flow, Ricci curvature, triangulations ]{} [[**Mathematics Subject Classification:**]{} 53C44, 57Q15, 57Q55, 57R12, 65D18]{} author: - Rory Conboye bibliography: - 'SimpGeom.bib' title: ' Piecewise flat Ricci flow of compact without boundary three-manifolds ' --- Introduction ============ The Ricci flow is a uniformizing flow on manifolds, acting like a heat flow. It was developed by Richard Hamilton in the 1980’s [@HamRF] as an approach to proving the Thruston geometrization conjecture, and therefore the Pioncaré conjecture, an endeavour that was eventually successful with the work of Grisha Pereleman in the early 2000’s [@Pereleman1; @Pereleman2]. Since then, Ricci flow has remained an important tool for investigating the interplay between geometry and topology, and has also found a number of applications in image analysis, from facial recognition [@GuZengFacial] to cancer detection [@GuCancer], and in space-time physics [@WoolgarRFphys; @WiseRF-BH; @WiseRF-CFT]. Applications such as these require a robust numerical approach, and preliminary work has even shown the potential for numerical simulations to inform analytical Ricci flow research [@GarfIsen]. Two dimensional numerical Ricci flow has seen a lot of progress recently, from Chow and Luo’s combinatorial Ricci flow [@CombRF] to the discrete surface Ricci flow of Gu et al. [@GuDiscSurfRF; @GuSurveyDRF]. These approaches use piecewise linear approximations of smooth surfaces with the Ricci flow given by a conformal deformation, utilizing the equivalence of the Ricci and scalar curvatures in two dimensions. In three dimensions, Garfinkle & Isenberg [@GarfIsen; @GarfIsen08] have used finite difference methods to find the critical behaviour for a one-parameter family of spherically symmetric neck pinch geometries on $S^3$. Crucially, this was done before the problem was completely understood analytically. A three-dimensional piecewise flat approach was also introduced a number of years ago [@SRF], based on intuitions gained from Regge calculus [@Regge], and applied to the same neck pinch geometries [@MNeckPinch1]. However, its effectiveness seems to be restricted to highly symmetric geometries, with its computational mesh adapted to these symmetries. The computations in this paper are based on a new piecewise flat approach developed in [@PLCurv]. Significantly, this approach has no symmetry restrictions, a particularly broad freedom in the choice of mesh, and gives results that are independent of such mesh choices to certain levels of precision. The Ricci flow evolution consists of a set of independent equations for each edge in the graph, giving it the potential to be highly parallelizable. The equations depend on a piecewise flat Ricci curvature, given in a computationally efficient combinatorial form, which has already been shown to converge to the smooth Ricci curvature with increasing mesh resolutions in [@PLCurv] for a number of different manifolds and mesh types. There are also substantial advantages in the use of piecewise flat manifolds, including the retaining of a manifold structure after discretization, avoiding issues with coordinate singularities due to the coordinate-independent structure, and having the topology completely determined by the piecewise flat graph. Evolutions of this piecewise flat Ricci flow are computed for a number of different situations, including: 1. four different manifolds with different levels of homogeneity, 2. three different types of triangulation mesh, each with different orientations of edges, 3. three different mesh resolutions for each triangulation type. The results of these computations show that there is no dependence on particular manifold symmetries, or on specially adapted triangulations. The computations also show that equivalent results are given for different triangulation types, at least up to a certain level of precision, and that the behaviour for higher mesh resolutions converges to the expected smooth behaviour. To ensure that the effects of the different triangulation types and mesh resolutions can be compared directly, the equations are evolved using a simple Euler method with a consistent number and size of time steps for each manifold. The four manifolds were chosen to have a variety of different characteristics, but all are compact, with topologies close to a three-torus. The first manifold is a Nil geometry, one of Thurston’s eight homogeneous geometries, with analytic solutions for the normalized and non-normalized Ricci flow found by Isenberg & Jackson [@IsenJackRFHomog] and Knopf & McLeod [@KnMcModel] respectively. The second manifold is a Gowdy three-geometry, one of the first non-positive-definite curvature geometries shown to have a convergent Ricci flow by Carfora, Isenberg & Jackson [@CarfIsenGowdy]. This has a two-dimensional isometry group, reducing the non-normalized Ricci flow equations to a set of two coupled partial differential equations (PDEs) for which a numerical solution can easily be found. A three-torus initially embedded in Euclidean four-space was chosen for the third manifold, having a one-dimensional isometry group, with a perturbation of a flat three-torus with no continuous isometries chosen for the last. The latter two manifolds are expected to Ricci flow asymptotically to a flat three-torus. The paper begins with an introduction to the new piecewise flat Ricci flow method, summarising the main results and details from [@PLCurv]. This is followed by a description of the three different types of triangulation. Each of the four manifolds is then dealt with separately in sections \[sec:Nil\] to \[sec:Pert\], with subsections to introduce the smooth manifolds, provide specific details about the triangulations, and both display and discuss the results of the Ricci flow computations. Ricci flow and triangulations ============================= Piecewise flat Ricci flow {#sec:PFRF} ------------------------- Simplicial piecewise flat manifolds in three dimensions are formed by joining Euclidean tetrahedra together, identifying the triangular faces of neighbouring segments. The topology of such a manifold is completely determined by the resulting graph, and the geometry by the set of edge-lengths. While each pair of neighbouring tetrahedra form a consistent Euclidean space, the dihedral angles of the tetrahedra meeting at a given edge may not necessarily sum to $2 \pi$ radians, with the difference known as the deficit angle $$\epsilon_\ell := 2 \pi - \sum_t \theta_t,$$ for the dihedral angles $\theta_t$ at the edge $\ell$. A smooth manifold can then be approximated by first constructing a tetrahedral lattice, using geodesic segments as edges, and then defining a piecewise flat manifold using the same graph, with the edge-lengths defined by the lengths of the corresponding geodesic segments. A good approximation will have uniformly small deficit angles, which can be achieved by using a lattice with high resolution in areas of high curvature. ![The deficit angle at an edge, and a cross section of the region around an edge showing the vertex and edge volumes.[]{data-label="fig:tet"}](fig1.pdf) On a smooth manifold $M$, the Ricci flow changes the metric $g$ according to the Ricci curvature $Rc$ at each point, with an optional second term giving a normalized Ricci flow, $$\frac{d g}{d t} = - 2 Rc \, , \qquad {\text or} \qquad \frac{d g}{d t} = - 2 Rc + \frac{2}{3} \widetilde{R}_M \, g \, ,$$ where the volume can be kept constant with the use of the average scalar curvature $\widetilde{R}$ of the manifold. Both equations tend to reduce the value of the metric components associated with high positive Ricci curvature and increase those associated with high negative Ricci curvature, with the rate depending on the strength of the curvature. It was shown in [@PLCurv] that the effect of the Ricci flow on the lengths of geodesic segments can be given by the integral of the Ricci curvature tangent to the segment at each point. This naturally leads to a piecewise flat approximation of the smooth Ricci flow as a set of independent equations for the fractional change in the edge-lengths, $$\label{eq:PFRF} \frac{1}{|\ell|} \frac{d |\ell|}{d t} = - Rc_\ell \, , \qquad \frac{1}{|\ell|} \frac{d |\ell|}{d t} = - Rc_\ell + \frac{1}{3} \widetilde{R}_S \, ,$$ for a piecewise flat approximation $Rc_\ell$ and an average scalar curvature $\widetilde{R}_S$ over the piecewise flat manifold $S$. These equations will continue to give a good approximation for the smooth Ricci flow as long as the lattice gives uniformly small deficit angles. All that remains is to give a piecewise flat approximation for the average Ricci curvature along each geodesic segment, and the average scalar curvature over the piecewise flat manifold. The first is given in terms of the sectional curvature $K_\ell^\perp$ orthogonal to the edge $\ell$, and the scalar curvature $R_v$ at the vertices $v_1$ and $v_2$ bounding $\ell$. These are in turn defined over volumes $V_v$ and $V_\ell$ associated with the vertex $v$ and edge $\ell$ respectively, which will be defined shortly. From [@PLCurv], \[eq:RandK\] $$\begin{aligned} Rc_\ell &= \frac{1}{4} (R_{v_1} + R_{v_2}) - K_\ell^\perp \, , \label{eq:Rc_l} \\ K_\ell^\perp &= \frac{1}{V_\ell} \left( |\ell| \, \epsilon_\ell + \sum_j \frac{1}{2} |\ell_j| \cos^2 (\theta_j) \epsilon_j \right) , \label{eq:K_l} \\ R_v \ &= \frac{1}{V_v} \sum_{i} |\ell_i| \, \epsilon_i \, , \qquad \widetilde{R}_S = \frac{2}{V_S} \sum_{i} |\ell_i| \, \epsilon_i \, . \label{eq:R_v}\end{aligned}$$ The indices $i$, $j$ correspond to the edges intersecting the volumes $V_v$ and $V_\ell$ respectively, with $\epsilon_i$, $\epsilon_j$ representing deficit angles and $\theta_j$ giving the angle between $\ell_j$ and the edge $\ell$. The expressions for $K_\ell^\perp$ and $R_v$ were found by constructing volume-integrals of each curvature, with the deficit angles shown to represent local surface-integrals of the sectional curvature, and then divided by the volume to give an average. The expression for $\widetilde{R}_S$ can also be seen to come directly from the Regge action [@Regge], divided by the volume $V_S$ of the piecewise flat manifold. The volumes $V_v$ and $V_\ell$ are specifically chosen to ensure that the expressions above are stable to small perturbations of the volume boundaries, or of the lattice itself, and contain a large enough sampling of edges and deficit angles. As such, the vertex volumes $V_v$ are defined to form a dual tessellation of the piecewise flat manifold, with barycentric volumes used for the computations in this paper. The edge-volumes $V_\ell$ are defined as the union of the vertex volumes on either end of the edge, bounded by surfaces which are orthogonal to $\ell$ at each vertex. A cross-section of these volumes can be seen in figure \[fig:tet\]. Computations in [@PLCurv] have successfully shown these expressions to converge to their corresponding smooth curvature values as the mesh resolution is increased, for a variety of different manifolds, and provide good approximations for reasonably sized meshes. The computations in this paper should provide even more support for the effectiveness of these piecewise flat curvature expressions. Similar expressions have also been developed for the extrinsic curvature on piecewise flat manifolds using the same approach [@PLExCurv], with computations also showing good approximation of, and convergence to, their corresponding smooth curvatures. Triangulation types {#sec:Tri} ------------------- For the purposes of this paper, a triangulation-type is defined as a family of simplicial graphs on a manifold, where the number of tetrahedra can be scaled in some regular way by choosing different members of the family. This scaling provides control over mesh refinements and is used to show the convergence of piecewise flat constructions to their smooth counterparts. In this section, three such triangulation types will be defined on a three-torus ($T^3$) topology, with a fundamental domain of size $[0,1]^3$ for a set of coordinates $x$, $y$, $z$. These can be directly applied to most of the manifolds, with minor adjustments required in some cases. The three different triangulation types provide a variety of edge orientations, ensuring that computation results are not based on favourable alignments alone, and provide differing levels of suitability for different dual tessellations. The building blocks for each triangulation type are defined below, and shown in figure \[fig:blocks\]. 1. The *cubic* block is cube-shaped and composed of six tetrahedra. There are seven edges for a given vertex, three along the $x$, $y$ and $z$ coordinates, three face-diagonals and a body-diagonal, with the tetrahedra specified so that the orientation of the face-diagonals on opposite sides agree. This is the most simple construction but it only borders on being a Delaunay triangulation for a flat metric, with the circumcenters of all six tetrahedra coinciding at the centre of the cube. 2. A *skew* block has the same structure as the cubic block, but with the vertices in the $x$ and $z$ directions skewed so that $v_x = (1, -1/3,0)$ and $v_z = (-1/3,-2/9,1)$. This block forms a different tiling of $T^3$, and gives a strongly Delaunay triangulation for a flat metric. 3. A *diamond* block is constructed by forming a set of four tetrahedra in a diamond shape around each coordinate edge, with the edges in the outer ring parallel with the other coordinate directions. This block contains two distinct vertices, instead of the single vertex for the other two blocks, and 14 edges, six that are parallel with the three coordinates and eight forming a set of four body-diagonal-type edges. ![The three different block types, with a slight separation of the three diamond shapes forming the diamond block, and the six tetrahedra of the cubic block on the far left.[]{data-label="fig:blocks"}](fig2.pdf) In each case, a single block can be used as a simplicial graph of $T^3$ by identifying faces on opposite sides. A regular grid of blocks can then be used to refine the mesh, with faces on opposite sides of the grid identified. Once the graph has been defined the vertices can be labelled, with each edge, triangle and tetrahedron then specified by the list of vertices in its closure. Since ambiguities in labelling can arise for grids that have less than three vertices in a single direction (for example all vertices in a single cubic block will have the same label), in these cases the grid is duplicated to form a slightly larger covering space before labelling, with any geometric information also duplicated but not the computations. Unfortunately the piecewise flat Ricci flow equations are not stable when applied directly to the cubic and skew triangulation-types, but this instability can be avoided by treating the interior of each block as flat. In practice this is done be re-defining the body-diagonal edge-lengths so that the deficit angles at these edges are essentially zero. The piecewise flat Ricci flow is still performed on a simplicial triangulation, where dihedral angles and volumes can be computed in a standard way, but with the lengths of the body-diagonals adjusted for each time-step to give an essentially flat interior for each block. Using this approach, there have been no instabilities in any of the piecewise flat Ricci flow computations performed in this article. Details of the instability and proofs for the stabilizing method will be provided in a related paper. The true advantage of piecewise flat approximations is in the complete encoding of the topology in the simplicial graph. Once the boundary identifications have been made and the labels assigned, the boundaries vanish. There are no boundary conditions required for functions on piecewise flat manifolds, since the graph already has the appropriate topology. There is also no need for any coordinate systems, once geodesic lengths have been computed they just become properties of the graph. As well as avoiding coordinate singularities or translations between multiple charts, the geometry is also clear from the lengths of the edges and does not need to be separated from the characteristics of a coordinate system. This even led Tullio Regge to title his 1961 paper [@Regge] as “General relativity without coordinates”. Nil manifolds {#sec:Nil} ============= Smooth manifold --------------- Thurston’s geometrization conjecture states that any closed 3-manifold can be decomposed into a set of irreducible parts, each of which admits a locally homogeneous geometry. The universal covers of these geometries form a set of eight model geometries, one of which is the Nil geometry, the geometry of the continuous Heisenberg group. The Heisenberg group is a nilpotent Lie group, and can be represented by the group of $3 \times 3$ matrices of the form $$\label{eq:Hgp} \left( \begin{array}{ccc} 1 & x & z \\ 0 & 1 & y \\ 0 & 0 & 1 \end{array} \right)$$ with real number entries $x$, $y$ and $z$, which can then be seen as a set of coordinates on $\mathbb{R}^3$. Analytic solutions for the normalized Ricci flow of the Nil geometry were first given by Isenberg and Jackson [@IsenJackRFHomog], with the quasi-convergence of the non-normalized Ricci flow later studied by Knopf and McLeod [@KnMcModel]. These approaches make use of work by Milnor [@Milnor76] and by Ryan and Shepley [@RyanShepley75], where an orthogonal frame of 1-forms $\{ \theta^a \}$ can be found in which the metric is diagonal $$\label{eq:nilDiag} g = A (\theta^1)^2 + B (\theta^2)^2 + C (\theta^3)^2 ,$$ and the structure constants $C^a_{b d}$ (such that $d \theta^a = C^a_{b d} \theta^b \wedge \theta^d$) are all zero except for $C^1_{2 3} = - C^1_{3 2} = \lambda$, for some constant $\lambda$. The Ricci curvature is also diagonal in this frame, which reduces both the normalized and non-normalized Ricci flows to a set of ordinary differential equations for the metric components $A$, $B$ and $C$, with the frame and structure constants invariant to the flows. The constant $\lambda$ can be seen as a scaling factor on the $z$ coordinate in the matrix representation of the Heisenberg group (\[eq:Hgp\]), $$\label{eq:Hgp2} \left( \begin{array}{ccc} 1 & x & z/\lambda \\ 0 & 1 & y \\ 0 & 0 & 1 \end{array} \right).$$ In [@IsenJackRFHomog] this constant is taken as the volume element ($\lambda = \sqrt{A B C}$), coming from a weighted Levi-Civita antisymmetric tensor in equation (4), which is invariant under the volume-preserving normalized Ricci flow. Solutions for $A$, $B$ and $C$ are then given in equations (29). A value of $-2$ is used for $\lambda$ in [@KnMcModel], with solutions for the non-normalized Ricci flow given in equations (15). The latter solutions can also be found in the Ricci flow textbooks [@ChowKnopfRF; @ChowLuNiRF], in equations (1.8) and (4.62) respectively. It can easily be shown that the scalar curvature $R = -\frac{\lambda^2 A}{2 B C}$, with solutions for the metric functions under the normalized Ricci flow then given by the equations: $$\label{eq:NilSolnIJ} A = A_0 \left(1 - \frac{16}{3} R_0 t\right)^{-1/2}, \quad B = B_0 \left(1 - \frac{16}{3} R_0 t\right)^{1/4}, \quad C = C_0 \left(1 - \frac{16}{3} R_0 t\right)^{1/4}, $$ and under the non-normalized Ricci flow by: $$\label{eq:NilSolnKM} A = A_0 (1 - 6 R_0 t)^{-1/3}, \quad B = B_0 (1 - 6 R_0 t)^{1/3}, \quad C = C_0 (1 - 6 R_0 t)^{1/3}, $$ for initial values $A_0$, $B_0$ and $C_0$, and the initial scalar curvature $R_0 = -\frac{\lambda^2 A_0}{2 B_0 C_0}$. The normalized Ricci flow solutions agree with those of [@IsenJackRFHomog] for $\lambda = \sqrt{A B C}$, and the non-normalized solutions agree with [@KnMcModel] for $\lambda = -2$. For computational purposes, the initial metric (\[eq:nilDiag\]) is chosen so that $A_0 = B_0 = C_0 = 1$. In the coordinates given by the matrix representation of the Heisenberg group in equation (\[eq:Hgp2\]), the metric takes the form $$\label{eq:gNil} d s^2 = d x^2 + d y^2 + (d z + \lambda \, x \, d y)^2.$$ The orthogonal frame is obtained from these coordinates by the identification $\{\theta^1, \theta^2, \theta^3\} = \{d z, dy + \lambda \, x \, d z, d x\}$. In order to obtain a compact computational domain, the continuous Heisenberg group can be quotiented by the finite Heisenberg group, with integer entries in the upper-right triangle. This gives a fundamental domain with the $x$ and $y$ coordinates having a range of $[0, 1]$ and the $z$ coordinate a range of $[0, \lambda]$. The $yz$ planes have a two-torus topology, while the identification of the $yz$ faces at $x=0$ and $x=1$ requires a sort of *twist*, as shown in figure \[fig:NilTri\]. Triangulating manifold ---------------------- For a single cubic block triangulation, the decomposition into six tetrahedra is done in a slightly different way to section \[sec:Tri\], see figure \[fig:NilTri\], so that the face-diagonals match the face relations for the fundamental domain. The geometry is isometric in the $y z$-planes, with a standard two-torus topology, so triangulations only require a single block in the $y$ and $z$ directions. Although the Nil geometry is completely homogeneous, the metric representation in (\[eq:gNil\]) depends on the $x$-coordinate, so increased resolutions are given by joining copies of the standard cubic block along the $x$-direction, with the complete collection of blocks still spanning $[0, 1]$. This has the effect of sub-dividing the *twist* into multiple parts, and reduces the size of the deficit angles. Vertices, edges and triangles on opposite faces are then identified, paying attention to the $yz$-face relations. Adapting the skew and diamond triangulation types to the Nil face relations is not so straight forward, so only the cubic triangulation type will be used here. However, since the $y$-edges are not orthogonal to the $z$-edges beyond $x = 0$, each block in the grid can be considered as a slightly different triangulation type. ![Relations between the $yz$-faces at $x=0$ and $x=1$ for $\lambda = 1$, and single and three-block triangulations with these face relations. The orthogonal frames for $\{\theta^i\}$ are also indicated for different values of $x$ along the three-block triangulation.[]{data-label="fig:NilTri"}](fig3.pdf) In order to compare with the analytic results of Isenberg & Jackson [@IsenJackRFHomog] and Knopf & McLeod [@KnMcModel], the normalized Ricci flow is computed for triangulations with $\lambda = 1$ and the non-normalized Ricci flow for triangulations with $\lambda = -2$. Grids of one, two and three blocks are then used for both situations. For the $\lambda = 1$ case, these triangulations have a range of $[0,1]$, $[0,1/2]$ and $[0,1/3]$ respectively in the $y$ and $z$ directions, in order to keep the tetrahedra close to regular. The original domain can then be recovered with four and nine copies for the two and three block triangulations. For the $\lambda=-2$ case, the twist is in the opposite direction, so the triangulations are a reflection in the $x y$-plane of those shown in figure \[fig:NilTri\]. A range of $[0,1/2]$ is used in the $x$-direction here, to keep from stretching the cubes too much, with the $y$ and $z$ directions given ranges of $[0,1/2]$, $[0,1/4]$ and $[0,1/6]$ for the three different grid sizes. Again, the original domain can be recovered with grids of copies with dimensions $2 \times 2 \times 4$, $2 \times 4 \times 8$ and $2 \times 6 \times 12$ in the $x$, $y$ and $z$ directions. Results of evolutions --------------------- Evolutions of the normalized and non-normalized piecewise flat Ricci flow (\[eq:PFRF\]) are computed using the Euler method, with 120 steps of size 0.005 for each triangulation. Since Nil manifolds are homogeneous, the values of the metric functions can be given by the square of the geodesic lengths for a unit coordinate length in the $x$ and $z$ directions for $C$ and $A$ respectively, and in the $y$-direction at $x=0$ for $B$. The resulting values are graphed in figure \[fig:Nilt\] as piecewise linear curves, along with the analytic solutions from equations (\[eq:NilSolnIJ\]) and (\[eq:NilSolnKM\]). The piecewise flat curves can be seen to give good approximations for the analytic solutions, even for the single block triangulations with only six tetrahedra, and there is a clear convergence to the analytic solutions as the number of blocks is increased. ![Graphs of the metric functions in time for both the normalized and non-normalized Ricci flows, all showing convergence to the analytic solutions as the resolution is increased.[]{data-label="fig:Nilt"}](fig4.pdf) The piecewise flat values for $A$, $B$ and $C$ have also been fitted to a general function matching the form of the analytic equations (\[eq:NilSolnIJ\]) and (\[eq:NilSolnKM\]), $$\label{eq:NilParam} f(t) = (1 + a \, t)^b .$$ The resulting best-fit parameters for $a$ and $b$ are shown in table \[tab:NilParam\], for each of the functions and both evolutions, with $R$-squared values of greater than $0.999999$ in all cases. These again show a clear convergence to the analytic solutions from [@IsenJackRFHomog] and [@KnMcModel], as the resolution is increased, with very close approximations for the three-block triangulations in particular. -- ------------ --------- ---------- --------- --------- --------- --------- Normalized a b a b a b 1-Block $2.037$ $-0.547$ $2.057$ $0.265$ $2.029$ $0.270$ 2-Block $2.460$ $-0.515$ $2.578$ $0.247$ $2.449$ $0.260$ 3-Block $2.585$ $-0.506$ $2.674$ $0.246$ $2.596$ $0.252$ Analytic $2.667$ $-0.5$ $2.667$ $0.25$ $2.667$ $0.25$ Non-norm. a b a b a b 1-Block $ 9.50$ $-0.352$ $ 9.88$ $0.348$ $ 9.68$ $0.354$ 2-Block $11.43$ $-0.338$ $11.67$ $0.332$ $11.48$ $0.339$ 3-Block $11.94$ $-0.335$ $12.14$ $0.329$ $11.92$ $0.334$ Analytic $12$ $-0.333$ $12$ $0.333$ $12$ $0.333$ -- ------------ --------- ---------- --------- --------- --------- --------- : The best fit parameters for expressions of the form (\[eq:NilParam\]), showing a clear convergence to the analytic solutions, with $R$-squared values greater than $0.999999$ in all cases. []{data-label="tab:NilParam"} Gowdy manifold {#sec:Gowdy} ============== Smooth manifold --------------- The Gowdy manifolds are a family of 3-geometries that can be foliated into isometric 2-surfaces and are periodic in the direction orthogonal to these surfaces. These manifolds first appeared as the spatial part of a cosmological space-time model introduced by Robert Gowdy [@Gowdy], with dynamics capable of containing gravitational waves. They were then used by Carfora, Isenberg and Jackson [@CarfIsenGowdy] to show that the Ricci flow can be convergent for manifolds with non-positive-definite curvature. With coordinates $x$ and $y$ spanning the isometric 2-surfaces, and $\theta$ orthogonal to them, the metric can be written in the general form $$d s^2 = e^{f + W} d x^2 + e^{f- W} d y^2 + e^{2 a} d \theta^2 \, ,$$ with a constant function $f$, and functions $W$ and $a$ depending on the value of $\theta$. The non-normalized Ricci flow of this metric can be represented by a coupled set of partial differential equations (PDEs) for these functions: $$\partial_t f = 0 , \quad \partial_t a = \frac{1}{2} e^{-2 a} \left(\partial_\theta W\right)^2 , \quad \partial_t W = e^{-2 a} \left( \partial_\theta^2 W - \partial_\theta a \cdot \partial_\theta W \right) .$$ These equations were easily solved numerically for initial-time functions $$f_0 = 0 , \qquad a_0 = 0, \qquad W_0 = 0.1 \sin \theta ,$$ requiring nothing more than the built-in PDE solver in *Mathematica*. The normalized Ricci flow equations are not so straight forward to solve, particularly in ensuring that $f$ is constant for each value of $t$. As a result, only the non-normalized Ricci flow is investigated here, though the volume is also close to being invariant for the non-normalized flow. It should be noted that there is no extra difficulty in computing the piecewise flat normalized Ricci flow. Triangulating manifold ---------------------- The Gowdy manifold is the simply connected covering space of a compact $T^3$ manifold with fundamental domain $[0,s_x] \times [0,s_y] \times [0, 2\pi]$, for any values $s_x$ and $s_y$. This can be used as a compact domain for the piecewise flat computations. Since the manifold is isometric in each $x y$-plane, triangulations only require a single block in the $x$ and $y$ directions. Different resolutions for cubic and skew triangulation types are provided by grids of $6$, $12$ and $24$ blocks in the $\theta$ direction, with $3$, $6$ and $12$ diamond blocks having the same number of vertices and edges. To keep the tetrahedra close to regular, both $s_x$ and $s_y$ are chosen to be $2$, $1$, $1 \over 2$ and $1 \over 4$ for the $3$, $6$, $12$ and $24$ block grids respectively, with $4$, $16$ and $64$ copies of the latter three grids giving the same domain for all triangulations. The piecewise flat curvatures for the cubic and skew triangulations have already been shown in [@PLCurv] for the initial manifold. ![ A grid of six cubic blocks along the $\theta$-direction for the lowest resolution triangulation, with the vertices, edges and triangles on opposite sides identified to give a $T^3$ topology. []{data-label="fig:GowdyTri"}](fig5.pdf) Results of evolutions --------------------- The piecewise flat Ricci flow (\[eq:PFRF\]) was applied to each of the triangulations, using the Euler method with $35$ time-steps of size $0.02$. Piecewise-linear graphs for the flow of the piecewise flat scalar curvature at $\theta=\pi/3$ are shown on the left-hand side of figure \[fig:Gowdyt\], together with the scalar curvature from the numerical PDE solution. The right-hand side of the figure shows the Ricci curvatures along the $y$-direction at $\theta=\pi/3$. The $\theta$ value of $\pi/3$ was chosen since the initial curvatures are all non-zero here, and there is a vertex at this value of $\theta$ for all of the triangulations. The $y$-edge was selected for the Ricci curvature as this is the only edge that is part of all three triangulation types. The graphs show convergence to the PDE solutions for both curvatures and all three triangulation types as the resolution is increased. For each resolution, the curves are also almost exactly the same across all three triangulation types, aside from the scalar curvature for the diamond type triangulations. ![Piecewise-linear graphs of the curvature values in time, showing a clear convergence of the piecewise flat to the PDE solution as the resolution is increased.[]{data-label="fig:Gowdyt"}](fig6.pdf) Since it was shown that the curvature decays exponentially to zero by Carfora et al. [@CarfIsenGowdy], exponential functions were fitted to the time evolution of the scalar and Ricci curvatures graphed in figure \[fig:Gowdyt\], with the decay rates for the best-fit functions shown in table \[tab:GowdyParam\]. Each column can be seen to converge to the PDE value, with good approximations for even the lowest resolutions, and very similar values for each resolution across all three triangulation types, particularly for the Ricci curvature. The $R$-squared values of the fitted functions are greater than $0.99999$ for all but the scalar curvatures of the diamond triangulations, which are no lower than $0.9997$. The larger errors and lower $R$-squared values may be due to the overlapping of edges along the $\theta$-direction for the diamond blocks, giving a larger $\theta$-interval for the averaging of the scalar curvature and requiring more blocks to resolve it better. -- -------------- --------- --------- --------- --------- --------- --------- $R$ $Rc$ $R$ $Rc$ $R$ $Rc$ 6-vertices $1.673$ $0.921$ $1.714$ $0.921$ $3.57$ $0.915$ 12-vertices $1.856$ $0.988$ $1.880$ $0.989$ $2.67$ $0.986$ 24-vertices $1.941$ $1.006$ $1.950$ $1.006$ $2.156$ $1.005$ PDE solution $2.003$ $1.003$ $2.003$ $1.003$ $2.003$ $1.003$ -- -------------- --------- --------- --------- --------- --------- --------- : Decay rates for the best-fit exponential functions of the scalar curvature and the Ricci curvature along the $y$-direction, both at $\theta = \pi/3$. The $R$-squared values are at least $0.99999$ for all but the diamond scalar curvatures, and each column shows a convergence to the corresponding PDE value as the resolution is increased.[]{data-label="tab:GowdyParam"} Visual representations of the Ricci curvature as a function of $\theta$ are shown in figure \[fig:GowdyRc\] for times $t = 0$ and $t = 0.7$, with different edges used for each triangulation type. The $\theta$-coordinate was selected for the cubic triangulations, since $Rc(\hat \theta, \hat \theta) \equiv - R$, corresponding with the scalar curvature in figure \[fig:Gowdyt\] and table \[tab:GowdyParam\] above. The $y$-edge was chosen for the diamond triangulation type, corresponding to the Ricci curvature in figure \[fig:Gowdyt\] and table \[tab:GowdyParam\], and the body-diagonal for the skew triangulations. All six graphs show good approximations for all resolutions, and a convergence to the PDE curves as the resolutions are increased. The shape of the first two sets of graphs remain mostly unchanged, with the scale decreasing by about a quarter and a half respectively, agreeing with the decay rates in table \[tab:GowdyParam\]. Despite the lengths of the body-diagonals being re-defined to give zero deficit angles, required for the stability of the flow as mentioned at the end of section \[sec:Tri\], the piecewise flat curves are no further from the PDE curve than for the other edges. This reinforces the robustness of the curvature constructions in (\[eq:RandK\]). These curves also show the piecewise flat Ricci flow producing more than just a scale change, with the change in shape between the two times resulting from a change in the orientation of the edges, which is perfectly matched by the piecewise flat curvatures. ![Graphs of the Ricci curvature as a function of $\theta$, for $t=0$ and $t = 0.7$, showing convergence to the PDE solutions as the resolution is increased in all cases.[]{data-label="fig:GowdyRc"}](fig7.pdf) To show that the edges graphed in figure \[fig:GowdyRc\] are not special cases, the errors for the Ricci curvatures were computed at all edges for both $t = 0$ and $t = 0.7$, with respect to the corresponding PDE solution values. There is evidence that the errors scale with the overall curvature of the manifold, so the mean absolute values of the errors are given as percentages of an average of the Ricci curvature for the corresponding value of $t$. This average is found by first taking the square-root of the tensor square of the Ricci curvature and then averaging over $\theta$, $$\label{eq:GRcAve} Rc_{\textrm ave} := \frac{1}{2 \pi} \int_0^{2 \pi} \sqrt{Rc^2} \, d \theta \, .$$ In table \[tab:GowdyPercErr\], these percentage errors can be seen to decrease in all cases as the resolution is increased. The errors also have similar values across all triangulation types, showing a certain level of independence to this choice for both the curvature values and the Ricci flow. -- ------------- ----------- ----------- ----------- ----------- ----------- ----------- $t = 0$ $t = 0.7$ $t = 0$ $t = 0.7$ $t = 0$ $t = 0.7$ 6-Vertices $12.4 \%$ $7.4 \%$ $ 9.0 \%$ $5.4 \%$ $ 9.0 \%$ $3.5 \%$ 12-Vertices $3.3 \%$ $2.4 \%$ $2.5 \%$ $2.1 \%$ $2.7 \%$ $1.2 \%$ 24-Vertices $0.8 \%$ $1.2 \%$ $1.1 \%$ $1.4 \%$ $0.8 \%$ $0.7 \%$ -- ------------- ----------- ----------- ----------- ----------- ----------- ----------- : The mean absolute error of the Ricci curvature over all edges in each triangulation, as a percentage of the mean absolute value of the curvature at each time (\[eq:GRcAve\]), showing a decrease everywhere as the resolution is increased.[]{data-label="tab:GowdyPercErr"} There is a notable reduction in the errors from $t=0$ to $t=0.7$ for the lower resolution triangulations, the top row of table \[tab:GowdyPercErr\]. This is likely due to the under-approximation of the initial curvature magnitudes, as seen in figure \[fig:GowdyRc\], giving a slower decay rate and reducing these errors in time. The effect indicates a general stabilizing behaviour in the piecewise flat Ricci flow, with errors *reducing* over time, at least for manifolds where the curvature magnitude is decreasing everywhere. Three-torus initially embedded in $\mathbb{E}^4$ {#sec:Torus} ================================================ Smooth manifold --------------- A regular three-torus can be embedded in Euclidean four-space $\mathbb{E}^4$ by beginning with a circle of radius $r_\theta$ and rotating it around another circle of radius $r_\phi > r_\theta$ in a plane orthogonal to it. The resulting two-torus can then be rotated around a third circle with radius $r_\psi > r_\phi$, in a plane orthogonal to both of the previous circles. For a poloidal-type coordinate system, with angular coordinates $\theta$, $\phi$ and $\psi$ for each circle, the metric induced by the embedding is $$\label{eq:T3metric} d s^2 = r_\theta^2 \ d \theta^2 + (r_\phi + \cos \theta)^2 \ d \phi^2 + (r_\psi +(r_\phi + \cos \theta) \cos \phi)^2 \ d \psi^2 \ .$$ This manifold is isometric along the $\psi$-coordinate, and the integral curves of the $\theta$-coordinate vector fields are the same everywhere, with values of $2 \pi r_\theta$. The minimum and maximum $\phi$ integral curves have lengths $2 \pi (r_\phi - r_\theta)$ and $2 \pi (r_\phi + r_\theta)$ with minimum and maximum lengths of $2 \pi (r_\psi - r_\phi)$ and $2 \pi (r_\psi + r_\phi)$ for the $\psi$ integral curves. These can be seen as generalizations of the inner and outer perimeters of a two-torus in $\mathbb{E}^3$, as shown in figure \[fig:T3Tri\]. The smooth Ricci flow of the metric above is expected to tend asymptotically to a flat three-torus, with the minimum and maximum orbits both approaching the same value asymptotically, for each coordinate. Triangulating manifold ---------------------- For the piecewise flat computations, the radii were chosen to have the values $$\label{eq:T3radii} r_\theta = 1 \, , \qquad r_\phi = 2 \, , \qquad r_\psi = 4 \, .$$ Since the manifold is isometric in the $\psi$-direction, only a single layer of blocks in a $\theta \phi$-plane is required. For the cubic block, grids of size $4 \times 6$, $6 \times 6$ and $6 \times 8$ are used, with grids of size $3 \times 4$, $4 \times 4$ and $4 \times 5$ for the diamond block. These give the same number of vertices and edges for the smallest and largest grids for both triangulation types. The entire range of the $\theta$ and $\phi$ coordinates is covered by all triangulations, but to keep the blocks regular in shape only $1/(2 n_x)$ of the $\psi$-coordinate, where $n_x$ is the number of blocks along the $\theta$-direction. This means that $8$, $12$ and $12$ copies of the cubic triangulations will be needed to cover the entire manifold, and $6$, $8$ and $8$ copies of the diamond triangulations. Unfortunately, the skew triangulation would require at least three layers of blocks in the $\psi$-direction in order to have vertices and edges on opposite sides match, so only the cubic and diamond triangulation types are used for this manifold. ![A two-torus embedded in $\mathbb{E}^3$, showing the inner and outer perimeters tangent to the $\phi$-coordinate vector fields, and the lowest resolution grids for both the cubic and diamond triangulation types, with opposite sides identified to form a $T^3$ topology.[]{data-label="fig:T3Tri"}](fig8.pdf) Results of evolution -------------------- The triangulations are evolved according to the non-normalized piecewise flat Ricci flow equations (\[eq:PFRF\]), using the Euler method with $80$ steps of size $0.05$. In order to help visualize the flow, the lengths of the edges along both the minimum and maximum $\phi$ and $\psi$ integral curves are summed for each time step. Although the $\theta$ integral curves all have equivalent lengths to begin with, the Ricci curvatures are not invariant along the $\phi$-coordinate, so different length curves emerge with the flow. The total lengths of the minimum and maximum integral curves in time, for all three resolutions, are graphed together in figure \[fig:Torust\] for each coordinate and each triangulation type. The graphs show the integral curve lengths converging asymptotically toward the same value for each coordinate, indicating that the manifold is flowing toward a three-torus with consistent dimensions. The values at $t=4$ are also given in table \[tab:TorusEnds\], providing bounds for the final three-torus dimensions for each triangulation. Both the curves and table values can also be seen to converge to the same shape and values for both triangulation types as the resolution is increased, sometimes from different directions. This suggests that the higher resolution values are closer to the smooth values, with both triangulation types agreeing to certain levels of precision. ![The flow of the maximum and minimum integral curve lengths for each coordinate, tending asymptotically toward the same value in all cases. The curves for both triangulation types also tend to converge to the same shape and values as the resolutions are increased.[]{data-label="fig:Torust"}](fig9.pdf) -- ---------- --------- --------- --------- --------- --------- --------- Cubic Min Max Min Max Min Max 24-Block $6.639$ $6.912$ $11.67$ $12.70$ $17.36$ $28.16$ 36-Block $6.729$ $6.934$ $11.81$ $12.48$ $17.89$ $28.26$ 48-Block $6.734$ $6.925$ $12.00$ $12.66$ $18.14$ $28.23$ Diamond Min Max Min Max Min Max 12-Block $6.668$ $7.132$ $11.65$ $12.55$ $17.51$ $27.44$ 16-Block $6.672$ $7.072$ $11.68$ $12.34$ $18.22$ $27.58$ 24-Block $6.682$ $7.024$ $12.02$ $12.69$ $18.28$ $27.77$ -- ---------- --------- --------- --------- --------- --------- --------- : The minimum and maximum perimeter lengths at $t=4$, bounding the dimensions of the limiting flat three-torus, with both triangulation types tending towards the same values. []{data-label="tab:TorusEnds"} To show that the manifold is actually flowing to a flat three-torus, the average of the absolute values of both the Ricci curvature and deficit angles, weighted by the edge-lengths, are also computed for each time step. The equations for these averages are $$\label{eq:RcDaAve} \widetilde{Rc_{\ell}} := \frac{\sum_{\ell} |\ell| |Rc_\ell|}{\sum_{\ell} |\ell|} \ , \qquad \widetilde{\epsilon_\ell} := \frac{\sum_{\ell} |\ell| |\epsilon_\ell|}{\sum_{\ell} |\ell|} \ .$$ The results are graphed in figure \[fig:TorusAveRc\], showing that both flow asymptotically to zero. The deficit angles can also be seen to reduce as the resolution is increased, for both triangulation types and at all times. ![ The average magnitudes of both the Ricci curvature and deficit angles, all tending asymptotically to zero. The Ricci curvature also tends toward the same curve for both triangulation types, and the deficit angles are smaller for increased resolutions at all times. []{data-label="fig:TorusAveRc"}](fig10.pdf) Perturbation of a flat three-torus {#sec:Pert} ================================== Smooth manifold --------------- A simple perturbation of a flat three-torus has been used to give a manifold without any continuous isometries. For a three-torus topology, with coordinates $x$, $y$ and $z$ ranging from $0$ to $1$, the metric was chosen to be $$\label{eq:PertMetric} d s^2 = (1 + 0.2 \sin (\pi x) \sin (\pi y) \sin (\pi z))\left(d x^2 + d y^2 + d z^2\right) \ .$$ As with both the Gowdy manifold and the three-torus, this manifold is expected to Ricci flow asymptotically back to a flat three-torus. Triangulating manifold ---------------------- Without any continuous isometries, there are no reductions that can be made for the triangulation grids. The cubic blocks are arranged in grids of size $2^3$, $3^3$ and $4^3$, with the diamond type triangulations starting at a single block with grids of size $2^3$ and $3^3$ as well. The vertices and edges of the skew blocks must align with other vertices and edges when opposite faces are identified. To make this easier for smaller grid sizes the skew block is adapted a little with the vertices in the $x$ and $z$ directions skewed so that $v_x = (1, -1/2, 0)$ and $v_z = (-1/2, -1/4, 1)$. This still requires even numbers of blocks in each direction, so only two grids of size $2 \times 2 \times 2$ and $4 \times 4 \times 4$ are used. ![ The highest resolution grids for the cubic, diamond and skew triangulation types, with the identification between the two $x y$-faces indicated for the skew triangulation. []{data-label="fig:PertTri"}](fig11.pdf) Results of evolution -------------------- The triangulations are evolved using the Euler method, with $50$ steps of size $0.002$, for the non-normalized piecewise flat Ricci flow. The smooth manifold is expected to flow asymptotically to a flat three-torus, so the lengths of the integral curves for each coordinate should all converge to the same value. Though the manifold does not have any continuous isometries, it is discretely symmetric for any permutation of the coordinates, so the three sets of coordinate integral curves should be equivalent. Since all three triangulation types have edges aligned with the $y$-coordinate, the minimum and maximum lengths of the $y$ integral curves have been computed for each time step, and are graphed on the left-hand side of figure \[fig:Pertt\]. ![ On the left, the minimum and maximum integral curve lengths along the $y$-coordinate show convergence to the same value for each triangulation type. The graphs on the right show the curvature flowing asymptotically to zero, and tending to the same shape for all three triangulation types as the resolution is increased. []{data-label="fig:Pertt"}](fig12.pdf) For the initial smooth manifold, the minimum integral curve lengths occur at $(x, z) = (0, 0)$ and the maximum lengths at $(x, z) = (0.5, 0.5)$. All of the triangulations have edges along the minimum-length integral curves, and the graphs show convergence to the same time-curve as the resolutions are increased. The middle resolution cubic triangulation time-curve has been omitted from the graph, since there are no $y$-edges at $(x, z) = (0.5,0.5)$, and the skew triangulations use a set of edges that are slightly offset, giving shallower curves for the middle graph on the left-hand side of figure \[fig:Pertt\]. The curves for the maximum and minimum lengths tend asymptotically to the same value for all of the triangulations, indicating that the manifold tends to a three-torus with consistent dimensions. The values of these lengths at $t = 0.1$ are also given in table \[tab:PertEnds\], giving bounds for the dimensions of the limiting manifold, with both the minimum and maximum values converging towards similar values for all three triangulation types as the resolution is increased. -- ------------- ---------- ---------- ---------- ---------- ---------- ---------- Resolutions Min Max Min Max Min Max Lowest $1.0074$ $1.0177$ $1.0096$ $1.0148$ $1.0190$ $1.0290$ Middle $1.0106$ $-$ $-$ $-$ $1.0117$ $1.0139$ Highest $1.0114$ $1.0133$ $1.0115$ $1.0129$ $1.0118$ $1.0131$ -- ------------- ---------- ---------- ---------- ---------- ---------- ---------- : Values for both the minimum and maximum lengths in the $y$-direction at $t=0.1$, giving bounds on the dimensions of the limiting flat three-torus for each approximation. []{data-label="tab:PertEnds"} The right-hand side of figure \[fig:Pertt\] shows the average magnitude of the Ricci curvature over all the edges of each triangulation, according to equation (\[eq:RcDaAve\]). These also show the Ricci flow is tending towards a flat manifold, with all three triangulation types tending towards the same time-curves as the resolution is increased. Conclusion {#sec:Con} ========== The computations in sections \[sec:Nil\] to \[sec:Pert\] have successfully demonstrated the effectiveness of the piecewise flat Ricci flow introduced in [@PLCurv] for a variety of manifolds, ranging from completely homogeneous to no continuous symmetries. These computations have shown: - a clear convergence to the smooth Ricci flow as the mesh resolution is increased, - good approximations for the smooth flow for reasonable resolutions, - equivalent results for different types of mesh, to appropriate levels of precision, and - no dependence on any symmetry properties of the manifolds. Specifically, the piecewise flat Ricci flow of the Nil manifolds converge to the analytic solutions first appearing in [@IsenJackRFHomog] and [@KnMcModel], for the normalized and non-normalized Ricci flow respectively, with errors on the order of $1\%$ for the fitted parameters in table \[tab:NilParam\] using only 18 tetrahedra. The approximations for the Gowdy manifold show convergence to a set of PDE solutions, with Ricci curvature errors of about $1\%$ at $t = 0.7$ for 144 tetrahedra. For the three-torus initially embedded in $\mathbb{E}^4$ the average magnitude of the Ricci curvature goes to zero and the minimum and maximum lengths along each coordinate asymptotically approach each other. Finally, the perturbed flat three-torus asymptotically flows back to a flat three-torus, with all three triangulation types approaching the same bounding values for the dimensions of the limiting three-torus as the resolution is increased. Other properties have also become apparent from some of the individual computations. Even very low resolution triangulations still give appropriate behaviour for the flow, as can be seen from the single cubic block triangulations of the Nil manifolds, and the single diamond block in figure \[fig:Pertt\], which is likely due to the topology being fixed by the piecewise flat graphs. In some cases, different triangulations can be used to give upper and lower bounds on the smooth solution, with the top two graphs in figure \[fig:Torust\] showing time-curves for two different triangulation types approaching the same curve but from different directions. The robustness of the piecewise flat curvature constructions is particularly evident from the lower two graphs in figure \[fig:GowdyRc\], especially since the deficit angles are zero at these edges due to adaptations to the triangulations for stability purposes. The lower resolution triangulations in table \[tab:GowdyPercErr\] also show a robustness for the piecewise flat Ricci flow of the Gowdy model, with over-approximations of the curvature decreasing quicker and under-approximations slower, also seen in figure \[fig:GowdyRc\].
{ "pile_set_name": "ArXiv" }
**Counting numerical semigroups** E. Kunz and R. Waldi [**Abstract.**]{} We are interested in formulas for the number of elements in certain classes of numerical semigroups [**Key words.**]{} Numerical semigroup (symmetric, pseudo-symmetric, of maximal embedding dimension), Apéry set, genus, polyhedral cone, lattice point, quasi-polynomial, Ehrhart’s theorem, generating function, lattice path. [**2010 Mathematics Subject Classification.**]{} 20M14, 05A15 [**1. Introduction**]{} For an integer $p\ge 3$ let $\mathfrak H_p$ be the set of all numerical semigroups containing $p$. Using the notion of Apéry set one can construct a bijective map from $\mathfrak H_p$ onto the set of all lattice points of a certain polyhedral cone $C_p\subset \mathbb R^{p-1}$ of dimension $p-1$ (\[Ku\],\[RGGB\]). The lattice points in the interior $C_p^0$ of $C_p$ are in one-to-one correspondence with the $H\in\mathfrak H_p$ of maximal embedding dimension $p$. Apéry’s characterization of symmetric semigroups \[A\] allows to show that the symmetric $H\in\mathfrak H_p$ are mapped onto the lattice points of certain $\lfloor\frac{p}{2}\rfloor$-dimensional closed faces of $C_p$. Here we describe also the distribution of the lattice points corresponding to the pseudo-symmetric semigroups (Proposition 3.2). Further semigroups correspond to the intersection of $C_p$ or its faces with hyperplanes, hence with the lattice points in polyhedrons, and their number can be expressed by quasi-polynomials (Theorem 4.2). This is the case for the $H\in\mathfrak H_p$ with fixed genus $g$. Also the semigroups $H\in\mathfrak H_p$ containing another number $q$ which is prime to $p$ can be described using hyperplane sections of $C_p$ (Example 4.1b)). We denote this set of numerical semigroups by $\mathfrak H_{pq}$. We want to study the degree, the leading term and a quasi-period of the involved Ehrhart quasi-polynomials. For the semigroups $H\in \mathfrak H_{pq}$ the leading term is constant and gives therefore an asymptotic estimate for $q\to \infty$ of the number of the $H\in \mathfrak H_{pq}$ (Proposition 5.2). Similarly for the $H\in \mathfrak H_p$ of maximal embedding dimension $p$ and the symmetric $H\in \mathfrak H_{pq}$ (Propositions 4.5 and 4.6). The following figure illustrates the situation in the simplest case $p=3$. (-1.1,-1.1)(7.1,10.6) (-1.1,-1.1)(7.1,10.6) (0,0)(0,10.3) (0,0)(7,0) (0,0)(1,0) (1,1)(2,1)(3,1) (1,2)(2,2)(3,2)(4,2)(5,2) (2,3)(3,3)(4,3)(5,3) (2,4)(3,4)(4,4)(5,4) (3,5)(4,5)(5,5) (3,6)(4,6)(5,6) (4,7)(5,7) (4,8)(5,8) (5,9) (5,10) (-0.33,-0.67)(5.25,10.5) (-0.33,-0.67)(6,2.5) (.2,.1)(6,3) (.5,0)(5.25,9.5) (2.4,4.6)(5,2) \[180\](-.4,-.6)[$v$]{} \[0\](5.2,5)[$C_3$]{} \[180\](2,1.1)[$\delta_1$]{} \[180\](1,2)[$\delta_2$]{} Figure 1: Geographical distribution of the $H\in \mathfrak H_3$. Using another visualization of the $H\in \mathfrak H_{pq}$ by lattice paths in the plane (\[KKW\],\[KW\]) recursion formulas for the number of semigroups $H\in \mathfrak H_{pq}$ which interest us can be derived, see Section 5. Finally in Section 6 explicit formulas are given for $p\le 5$. For related counting problems, see \[Ka\], \[BGP\]. [**2. The polyhedral cone $C_p$ and its faces**]{} We recall some facts about $C_p$ which are relevant for us. Details can be found in \[Ku\] or \[RGGB\]. For $H\in \mathfrak H_p$ let $\{h_1,\dots,h_{p-1}\}=\text{Ap}(H,p)$ be the Apéry set of $H$ with respect to $p$, that is, $h_i$ is the smallest element of $H$ in $i+p\mathbb N$ for $i=1,\dots,p-1$. Write $h_i=i+\mu_ip$. The semigroup $H$ is uniquely determined by $(\mu_1,\dots,\mu_{p-1})$ since $H=<p,h_1,\dots,h_{p-1}>$. The points $(\mu_1,\dots,\mu_{p-1})\in \mathbb N^{p-1}$ are the solutions in $\mathbb N^{p-1}$ of the system of linear inequalities $$\begin{cases} X_i+X_j\ge X_{i+j}\qquad\qquad (i+j<p)\cr X_i+X_j\ge X_{i+j-p}-1\quad\ (i+j>p)\end{cases}\leqno(1)$$ Let $C_p$ be the solution set of (1) in $\mathbb R^{p-1}$. This is a polyhedral cone with vertex $v:=(-\frac{1}{p},-\frac{2}{p}\dots,-\frac{p-1}{p})$ and $$\mu:\mathfrak H_p\to C_p\cap\mathbb N^{p-1}\quad (H\mapsto (\mu_1,\dots,\mu_{p-1}))$$ is a bijection of $\mathfrak H_p$ onto the set of lattice points of $C_p$. For $H\in \mathfrak H_p$ we call $\mu(H)$ the lattice point associated to $H$, and for $P=\mu(H)$ we say that $H$ is the semigroup belonging to the point $P$. By a [*face*]{} of $C_p$ we always understand an open face. Its dimension is the dimension of the smallest affine space containing it. We consider the interior $C_p^0$ of $C_p$ and the vertex $v$ also as faces. We have $C_p^0\cap\mathbb N^{p-1}=(1,\dots,1)+C_p\cap \mathbb N^{p-1}$ (\[Ku\], 1.4d). $C_p$ and $C_p^0$ have dimension $p-1$. The [*facets*]{} (1-codimensional faces) correspond bijectively to the hyperplanes $$\begin{cases} E_{ij}: X_i+X_j-X_{i+j}=0\qquad\qquad (i+j<p)\cr E_{ij}: X_i+X_j-X_{i+j-p}+1=0\quad\ (i+j>p)\end{cases}$$ Hence for even $p$ there are $\frac{(p-1)^2-1}{2}$ facets, for odd $p$ their number is $\frac{(p-1)^2}{2}$. The translation of $C_p$ by the vector $-v$ leads to a polyhedral cone $C_p^*:=-v+C_p$ in $\mathbb R_+^{p-1}$ with vertex at the origin. It is the solution set of the system $X_i+X_j\ge X_{i+j}\ (i+j\ne p\ \text{with indices reduced modulo}\ p)$. Therefore $C_p^*\subset C_p$. On each face of $C^*_p$ there are lattice points. This is true in particular for its edges (1-dimensional faces). On each such edge there is a [*primitive lattice point*]{} $\delta$, that is a point with relatively prime integral coordinates, and all other lattice points on the edge are integral multiples of $\delta$. The set of all these $\delta$ is called the [*canonical system of representatives*]{} for the edges of $C_p^*$. Let $m(H)$ denote the multiplicity and edim$(H)$ the embedding dimension of a numerical semigroup $H$. The lattice points in the interior of $C_p$ correspond to the $H\in \mathfrak H_p$ with edim($H)=m(H)=p$ (\[Ku\], 2.4b), called in modern language the [*semigroups of maximal embedding dimension*]{} $p$. Let $S$ be a $d$-dimensional face of $C_p$, $S^*$ the face of $C_p^*$ parallel to $S$ and $\bar S$ resp. $\bar S^*$ the topological closures of these faces. The generating function (Hilbert series) for the lattice points on $S$ is the formal power series $$H_S(X_1,\dots,X_{p-1})=\sum_{\mu\in S\cap\mathbb N^{p-1}}X^{\mu}.$$ The generating functions $H_{\bar S}$ for $\bar S$, $H_p$ for $C_p$ and $H_p^0$ for the interior $C_p^0$ of $C_p$ are defined accordingly. [**2.1. Remark.**]{} (\[Ku\], 2.1) $H^0_p(X_1,\dots,X_{p-1})=X_1\cdots X_{p-1}H_p(X_1,\dots,X_{p-1})$. Let $\{\delta_1,\dots,\delta_t\}$ be the canonical system of representatives for the edges of $C_p^*$ and $\delta_1,\dots,\delta_r$ the vectors belonging to $\bar{S^*}$. Using slack variables $T_{ij}$ the system (1) of Section 1 becomes a linear system of equations $$X_i+X_j-X_{i+j}-T_{ij}=0\ (i+j<p)$$ $$X_i+X_j-X_{i+j-p}-T_{ij}=-1\ (i+j>p).$$ To it the results of Stanley \[S1\] Chap.I can be applied. One obtains [**2.2. Proposition.**]{} $H_{\bar S}$ can be written in the form $$H_{\bar S}=\frac{F_{\bar S}}{\prod_{i=1}^r(1-X^{\delta_i})},\ F_{\bar S}\in \mathbb Z[X_1,\dots,X_{p-1}]$$ Similarly $H_p$ has the form $$H_p=\frac{F_p}{\prod_{i=1}^t(1-X^{\delta_i})},\ F_p\in \mathbb Z[X_1,\dots,X_{p-1}].$$ Further deg$_{X_i}H_{\bar S}\le -1$ and deg$_{X_i}H_p\le -1\ (i=1,\dots,p-1)$. See \[Ku\], 3.2 and 3.3 for details. The proof of \[Ku\], 3.2 contains the false statement that $R_{\overline{S^*}}:=\mathbb C[\{X^{\mu}\}_{\mu\in \overline{S^*}\cap\mathbb N^{p-1}}]$ as an algebra over $\mathbb C$ is generated by $X^{\delta_1},\dots,X^{\delta_r}$. However the algebra is a finitely generated module over $A:=\mathbb C[X^{\delta_1},\dots,X^{\delta_r}]$. Then also the module $M_{\bar S}:=\bigoplus_{\mu\in \bar S\cap \mathbb N^{p-1}}\mathbb C\cdot X^{\mu}$ is a finitely generated $A$-module, which is what is actually used in the proof. [**3. Symmetric and pseudo-symmetric semigroups**]{} Apéry’s characterization of symmetric semigroups in terms of their Apéry sets (\[A\]) allows to describe the lattice points in $C_p$ belonging to such semigroups. Let $$t:=\begin{cases}p-1,\ p\ \text{odd}\cr \frac{p}{2},\qquad p\ \text{even}\end{cases}.$$ [**3.1. Proposition**]{} (\[Ku\], 2.9) There are $t$ faces $S_1,\dots,S_t$ of dimension $\lfloor\frac{p}{2}\rfloor$ such that any symmetric $H\in\mathfrak H_p$ belongs to exactly one $\bar S_j$, and all lattice points on the $\bar S_j$ correspond to symmetric $H\in \mathfrak H_p$. In particular $\bar S_j\cap \bar S_k\cap \mathbb N^{p-1}=\emptyset$ for $j\ne k$. According to Proposition 2.2 the generating function for the symmetric semigroups $$H_{\text{sym}}(X_1,\dots,X_{p-1}):=\sum_{\mu\in \cup_{j=1}^t\bar S_j\cap\mathbb N^{p-1}}X^{\mu}=\sum_{j=1}^tH_{\bar S_j}(X_1,\dots,X_{p-1})$$ is a rational function. For pseudo-symmetric semigroups the situation is more complicated. Remember that a numerical semigroup $H$ of genus $g(H)$ and Frobenius number $F(H)$ is [*pseudo-symmetric*]{} if $2g(H)=F(H)+2$. Let $\mathfrak S_{p-1}$ be the permutation group of $\{1,\dots,p-1\}$ and $\mathfrak S_{p-1}^*$ the set of all $\sigma \in \mathfrak S_{p-1}$ such that $$\sigma(i)+\sigma(p-2-i)\equiv \sigma(p-2)\ \text{mod}p\ (i=1,\dots,p-3)$$ $$2\sigma(p-1)\equiv\sigma(p-2)\ \text{mod}p.$$ For $\sigma\in \mathfrak S_{p-1}^*$ let $L_{\sigma}$ be the affine subspace of $\mathbb R^{p-1}$ defined by the linear equations $$X_{\sigma(i)}+X_{\sigma(p-2-i)}=\begin{cases}X_{\sigma(p-2)}\qquad \text{if}\ \sigma(i)+\sigma(p-2-i)=\sigma(p-2)\cr X_{\sigma(p-2)}-1\ \text{if}\ \sigma(i)+\sigma(p-2-i)=\sigma(p-2)+p\end{cases}\leqno (1)$$ $(i=1,\dots,p-3)$ and $$2X_{\sigma(p-1)}=X_{\sigma(p-2)}+\begin{cases}1\quad \text{if}\ 2\sigma(p-1)=\sigma(p-2)\cr 0\quad \text{if}\ 2\sigma(p-1)=\sigma(p-2)+p\end{cases}.$$ This space has dimension $\lfloor\frac{p-1}{2}\rfloor$. The following Proposition is based on \[RG\],4.15. [**3.2. Proposition.**]{} a) For $\sigma\in \mathfrak S_{p-1}^*$ each lattice point $(\mu_{\sigma(1)},\dots,\mu_{\sigma(p-1)})\ne 0$ of $L_{\sigma}\cap C_p$ belongs to a pseudo-symmetric semigroup. b\) For each pseudo-symmetric $H\in \mathfrak H_p$ there exists a $\sigma\in \mathfrak S_{p-1}^*$ such that $\mu(H)\in L_{\sigma}\cap C_p$. c\) For $p>3$ the $L_{\sigma}\cap C_p$ are contained in the boundary of $C_p$. d\) $L_{id}\cap C_p=\bar S\cap H$ where $\bar S$ is the closure of a face $S$ of $C_p$ of dimension $\lfloor\frac{p+1}{2}\rfloor$ containing $0$ and $H$ a hyperplane through $0$ defining a facet of $C_p^*$. [**Proof.**]{} a) Let $h_{\sigma(i)}:=\sigma(i)+\mu_{\sigma(i)}p\ (i=1,\dots,p-1)$. Then $\{h_{\sigma(1)},\dots,h_{\sigma(p-1)}\}$ is the Apéry set of a semigroup $H\in \mathfrak H_p$ and $\sum_{i=1}^{p-1}h_{\sigma(i)}={{p}\choose{2}}+p\sum_{i=1}^{p-1}\mu_{\sigma(i)}={{p}\choose{2}}+g(H)p$. From the equations defining $L_{\sigma}$ we obtain $$h_{\sigma(i)}+h_{\sigma(p-2-i)}=h_{\sigma(p-2)}\ (i=1,\dots,p-3)$$ $$2h_{\sigma(p-1)}=h_{\sigma(p-2)}+p.$$ Obviously $h_{\sigma(i)}<h_{\sigma(p-2)}$ for $i=1,\dots,p-3$, hence $h_{\sigma(p-2)}\ge p-2$, and then by the last equation $h_{\sigma(p-1)}\ge p-1$ and $h_{\sigma(p-2)}-h_{\sigma(p-1)}=h_{\sigma(p-1)}-p\ge -1$. In case $h_{\sigma(p-1)}>h_{\sigma(p-2)}$ we would have $h_{\sigma(p-1)}=p-1, \mu_{\sigma(p-1)}=0$. But then $\mu_i=0$ for $i=1,\dots,p-1$, contrary to the assumption. Thus $h_{\sigma(p-2)}$ is the maximal element of the Apéry set of $H$, i.e. $h_{\sigma(p-2)}=F(H)+p$ with the Frobenius number $F(H)$ of $H$. Adding the above equations gives $$2\sum_{i=1}^{p-1}h_{\sigma(i)}=ph_{\sigma(p-2})+p=p(F(H)+p)+p$$ that is $$2g(H)=F(H)+2$$ which is equivalent to $H$ being pseudo-symmetric. b\) If $H$ is pseudo-symmetric, then $F(H)$ is an even number. By \[RG\], 4.15 there is a permutation $\sigma\in \mathfrak S_{p-1}^*$ such that $\{h_{\sigma(1)},\dots,h_{\sigma(p-1)}\}$ is the Apéry set of $H$ and $$h_{\sigma(i)}+h_{\sigma(p-2-i)}=h_{\sigma(p-2)}\ (i=1,\dots,p-3)$$ $$h_{\sigma(p-2)}=F(H)+p, h_{\sigma(p-1)}=\frac{F(H)}{2}+p$$ in particular $$2h_{\sigma(p-1)}=F(H)+2p=h_{\sigma(p-2)}+p.$$ With $h_{\sigma(i)}:=\sigma(i)+\mu_{\sigma(i)}p\ (i=1,\dots,p-1)$ it follows that $(\mu_{\sigma(1)},\dots,\mu_{\sigma(p-1)})\in L_{\sigma}\cap C_p$. c\) For $p>3$ the equations (1) show that $L_{\sigma}$ is contained in a hyperplane which defines a facet of $C_p$, hence $L_{\sigma}\cap C_p$ belongs to the boundary of $C_p$. d\) $\sigma=id$ belongs to $\mathfrak S_{p-1}^*$ and in this case $L_{\sigma}$ is the intersection of the hyperplanes $$H_i: X_i+X_{p-2-i}=X_{p-2}\ (i=1,\dots,p-3), H: 2X_{p-1}=X_{p-2}.$$ The $H_i$ define facets of $C_p$ containing the origin, and $H$ a facet of $C_p^*$. We have $$L_{id}\cap C_p=\cap_{i=1}^{p-3}(H_i\cap C_p)\cap H$$ and $\bar S:=\cap_{i=1}^{p-3}(H_i\cap C_p)$ is the closure of a face $S$ of $C_p$. The point $P:=(1,\dots,1,2,1)$ is in $L_{id}\cap C_p$ and in the interior of the half-spaces $X_i+X_j\ge X_{i+j} (i+j\ne p-2)$. Therefore an open neighborhood of $P$ in the $\lfloor\frac{p-1}{2}\rfloor$-dimensional affine space $L_{id}$ is contained in $L_{id}\cap C_p$. Hence the face $S$ must have dimension $\lfloor\frac{p+1}{2}\rfloor$. $\Box$ [**3.3. Corollary.**]{} (\[RG\], 4.26). For $p>3$ pseudo-symmetric semigroups of $\mathfrak H_p$ have embedding dimension $<p$. [**Proof.**]{} Otherwise the lattice points of such semigroups would be in the interior of $C_p$ contradicting 3.2c). $\Box$ [**3.4. Example.**]{} In general $L_{id}\cap C_p$ is not a polyhedral cone. If $p=7$, then $L_{id}$ is a 3-space, in which $L_{id}\cap C_7$ is as shown in the next figure: (-3,-.5)(5,5) (0,0)(1.071,0.214) (0,0)(4.16,5.2) (1.071,0.214)(4.7,4.75) (0,0)(1.32,3.96) (0,0)(2.475,1.65) (1.071,0.214)(4.843,1.629) (0,0)(1.071,0.214) (4,5)(1.2,3.6)(2.25,1.5)(4.5,1.5)(4.5,4.5) \[215\](0,0)[$0$]{} Figure 2 [**4. Intersections of $C_p$ and its faces with hyperplanes**]{} Certain classes of numerical semigroups correspond to the lattice points in the intersection of $C_p$ or some of its faces with hyperplanes. [**4.1. Examples.**]{} a\) If $\mu(H)=(\mu_1,\dots,\mu_{p-1})$ for $H\in \mathfrak H_p$, then $g(H)=\sum_{i=1}^{p-1}\mu_i$ is the genus of $H$. Thus if $H_g$ is the hyperplane $\sum_{i=1}^{p-1}X_i=g$, then the lattice points in $C_p\cap H_g$ are in one-to-one correspondence with the semigroups in $\mathfrak H_p$ of genus $g$ and those of $C_p^0\cap H_g$ with the $H\in \mathfrak H_p$ of genus $g$ and maximal embedding dimension $p$. b\) Given $q=i+np\ (i<p,i\ \text{and}\ p$ coprime) the lattice points in the intersection of $C_p$ with the hyperplane $X_i=n$ correspond bijectively to the $H\in \frak H_{pq}$ such that $q\in\text{ Ap}(H,p)$, and the lattice points in the intersection of $C_p$ with the half-space $X_i\le n$ to all $H\in \frak H_{pq}$. In fact, if $\mu(H)=(\mu_1,\dots,\mu_{p-1})$ and $\mu_i\le n$, then $i+\mu_ip\in H$ implies that also $q\in H$. More generally, let $\alpha\in \mathbb N^{p-1}\setminus\{0\}$ be a primitive lattice point, that is, if $\alpha=(\alpha_1,\dots,\alpha_{p-1})$, then $\alpha_1,\dots,\alpha_{p-1}$ are relatively prime. For $x\in \mathbb R^{p-1}$ let $\alpha\cdot x$ denote the scalar product of $\alpha$ and $x$, and $H_n=H_n(\alpha):=\{x\in \mathbb R^{p-1}\vert \alpha\cdot x=n\}$ for $n\in \mathbb N$. We assume that no edge of $C_p^*$ is contained in the hyperplane $H_0$. This condition is satisfied in the Examples 4.1. If in 4.1b) there would be an edge vector $\delta=(\mu_1,\dots,\mu_{i-1},0,\mu_{i+1},\dots,\mu_{p-1})$ there would be infinitely many lattice points on the edge of $C_p^*$ determined by $\delta$. These correspond to semigroups $H\in \frak H$ with $i\in H$. Since $i$ is prime to $p$ there exist only finitely many such $H$, a contradiction. Let $\bar S$ be the topological closure of a $d$-dimensional face $S$ of $C_p$ and let ${\it P}_{\alpha,n}$ resp. $\it P_{\alpha,n}^{\bar S}$ be the intersection of the hyperplane $H_n$ with the cone $C_p$ resp. $\bar S$. For $n\ge 1$ the sets $\it P_{\alpha,n}$ and ${\it P}_{\alpha,n}^{\bar S}$ are rational polytopes of dimension $p-2$ resp. $d-1$. We are interested in the numbers $f_{\alpha}(n)$ resp. $f_{\alpha}^{\bar S}(n)$ of lattice points in $P_{\alpha,n}$ and $P_{\alpha,n}^{\bar S}$ and in the number $f^0_{\alpha}(n)$ of lattice points in $C_p^0\cap P_{\alpha,n}$. [**4.2. Theorem.**]{} Let ${\delta_1,\dots,\delta_t}$ be the canonical system of representatives for the edges of $C_p^*$ and ${\delta_1,\dots,\delta_r}$ the edge vectors contained in $\overline{S^*}$ where $S^*$ is the face of $C_p^*$ parallel to $S$. If $\bar S$ contains a lattice point $u\in \mathbb N^{p-1}$, then $f_{\alpha}^{\bar S}$ is a quasi-polynomial of degree $d-1$ with non-negative leading coefficient. The least common multiple of $\{\alpha\cdot\delta_i\}_{i=1,\dots,r}$ is a quasi-period of $f_{\alpha}^{\bar S}$. In particular $f_{\alpha}$ is a quasi-polynomial of degree $p-2$. Moreover for $n\ge\sum_{i=1}^{p-1}\alpha_i$ $$f^0_{\alpha}(n)=f_{\alpha}(n-\sum_{i=1}^{p-1}\alpha_i).$$ Remember that a function $f: \mathbb N\to \mathbb C, f\not\equiv 0$ with generating function $H_f(T):=\sum_{n=0}^{\infty}f(n)T^n$ is called a [*quasi-polynomial*]{} of degree $d$ and quasi-period $N>0$ if $f$ is of the form $$f(n)=c_d(n)n^d+c_{d-1}(n)n^{d-1}+\dots+c_0(n)\ (n\in \mathbb N)$$ where the $c_i:\mathbb N\to \mathbb C$ are periodic functions with integral period $N$ and $c_d(n)$ does not vanish identically. Equivalently, $f$ is a quasi-polynomial if there exist an integer $N>0$ and polynomials $f_0,\dots,f_{N-1}$ such that $$f(n)=f_i(n)\ \text{if}\ n\equiv i\ \text{mod}\ N.$$ If there exists an integer $N>0$ and polynomials $P(T), Q(T)\in \mathbb C[T]\setminus \{0\}$ with $\text{deg}P(T)<\text{deg}Q(T)$ so that $$H_f(T)=\frac{P(T)}{Q(T)}$$ and $\alpha^N=1$ for each zero $\alpha$ of $Q$, then $f$ is a quasi-polynomial with quasi-period $N$. Its degree is one less than the maximum pole order of the rational function $\frac{P(T)}{Q(T)}$ (\[S2\],4.4.1). $f\equiv 0$ is also considered as a quasi-polynomial. [**Proof of Theorem 4.2.**]{} At first we show that $f_{\alpha}^{\bar S}$ is a quasi-polynomial. Since $H_n(\alpha)$ does not contain an edge of $\bar S^*$ the denominator $\prod_{i=1}^r(1-X^{\delta_i})$ of the generating function $H_{\bar S}$ (see Proposition 2.2) does not vanish if we replace the $X_j$ by $T^{\alpha_j}$ with a variable $T$. Then $$H_{\bar S}(T^{\alpha_1},\dots,T^{\alpha_{p-1}})=\frac{F_{\bar S}(T^{\alpha_1},\dots,T^{\alpha_{p-1}})}{\prod_{i=1}^r(1-T^{\alpha\cdot\delta_i})}=\sum_{n=0}^{\infty}f_{\alpha}^{\bar S}(n)T^n.$$ Set $P(T):=F_{\bar S}(T^{\alpha_1},\dots,T^{\alpha_{p-1}})$ and $Q(T):=\prod_{i=1}^r(1-T^{\alpha\cdot \delta_i})$. Since deg$_{X_j}H_{\bar S}\le -1$ we have deg$P<\ $deg$Q$. Hence $f_{\alpha}^{\bar S}(n)$ is a quasi-polynomial with quasi-period as stated. Its leading coefficient is non-negative since $f_{\alpha}^{\bar S}(n)\in\mathbb N$ for all $n\in \mathbb N$. Now we show that $f_{\alpha}^{\bar S}$ has degree $d-1$. By assumption $\bar S$ contains a lattice point $u\in \mathbb N^{p-1}$. On the line through $u$ and the vertex $v$ of $C_p$ there is a lattice point $w$ not contained in $\bar S$, hence $-w\in \mathbb N^{p-1}$ and $$u+\overline{S^*}\subset\bar S\subset w+\overline{S^*}.\leqno(1)$$ Let $\it P_{\alpha,n}^*:=\bar S^*\cap H_n$ for $n\in \mathbb N_+$. Then $\it P_{\alpha,1}^*$ is a convex polytope of dimension $d-1$ and $\it P_{\alpha,n}^*=n\cdot\it P_{\alpha,1}^*$. To this situation a theorem of Ehrhart (\[S2\],4.6.8) can be applied. It states that if $i(\it P_{\alpha,n}^*)$ is the number of lattice points in $\it P_{\alpha,n}^*$, then this function of $n$ is a quasi-polynomial of degree $d-1$. In order to show this also for $f_{\alpha}^{\bar S}$ observe that (1) implies that for $n\ge l:=\alpha\cdot u$ and $k:=\alpha\cdot w$ we have $$u+\it P_{\alpha,n-l}^*\subset \bar S\cap H_n\subset w+\it P_{\alpha,n-k}^*.$$ Thus for large $n$ the quasi-polynomial $f_{\alpha}^{\bar S}$ is trapped by two quasi-polynomials of degree $d-1$, hence it has also the degree $d-1$. The formula for $f_{\alpha}^0$ follows from Remark 2.1 after substituting $X_i=T^{\alpha_i}\ (i=1,\dots,p-1)$ and expanding into power series in $T$. $\Box$ In the situation of Example 4.1a) let $G(p,g)$ be the number of $H\in \mathfrak H_p$ with genus $g$. Then with $\alpha=(1,\dots,1)$ we have $G(p,g)=f_{\alpha}(n)$, and Theorem 4.2 tells us that $G(p,g)$, as a function of $g$, is a quasi-polynomial of degree $p-2$ with non-negative leading term. Moreover the least common multiple of $\{\alpha\cdot\delta_i\}_{i=1,\dots,r}$ is a quasi-period of $G(p,g)$. $G^0(p.g):=f_{\alpha}^0(g)$ is the number of $H\in \mathfrak H_p$ having maximal embedding dimension $p$ and genus $g$. For $g\ge p-1$ we have $$G^0(p,g)=G(p,g-(p-1)).\leqno(2)$$ If $G_{sym}(p,g)$ is the the number of symmetric $H\in \mathfrak H_p$ with genus $g$, then Proposition 3.1 and Theorem 4.2 imply that $G_{sym}(p,g)$ is a quasi-polynomial of degree $\lfloor\frac{p}{2}\rfloor-1$. In the situation of Example 4.1b) we can apply Theorem 4.2 to the functions $f_{e_i}^{\bar S}$ where $e_i$ is the i-th unit vector and $\bar S$ the closure of a face of $C_p$ containing a lattice point. Let $g_{e_i}(n):=\sum_{j=0}^nf_{e_i}(j)$ and $g_{e_i}^0(n):=\sum_{j=0}^nf^0_{e_i}(j)$, that is, the number of $H\in \mathfrak H_p$ (of maximal embedding dimension $p$) containing also $i+np$. In order to apply Theorem 4.2 also to $g_{e_i}$ and $g^0_{e_i}$ we need the following facts. For functions $f: \mathbb N\to \mathbb C$ we consider the operators $$E: f(n)\to f(n+1)\ \text{(shift})$$ $$\Delta: f(n)\to f(n+1)-f(n)\ \text{(difference})$$ $$\Sigma: f(n)\to \sum_{i=0}^nf(i)\ (\text{sum}).$$ It is easy to see that $f$ is a quasi-polynomial if and only if this is the case for $Ef, \Delta f$ or $\Sigma f$. Moreover deg$Ef$=deg$f$. [**4.4. Lemma.**]{} Let $f: \mathbb N\to \mathbb R, f\not\equiv 0$ be a quasi-polynomial of degree $d$. a\) If $f$ is increasing, then $c_d(n)$ is constant. b\) If $f\ge 0$, then deg$\Sigma f=d+1$. [**Proof.**]{} a) Let $N$ be a quasi-period of $f=\sum_{k=0}^dc_k(n)n^k$ and $f_0,\dots,f_{N-1}$ the polynomials with $f(n)=f_i(n)$ for $i=0,\dots,N-1$. Then $c_d(i)$ is the coefficient of $t^d$ in $f_i(t)$ for $i=0,\dots,N-1$. With $k\in \mathbb N_+$ and $i\in \{0,\dots,N-1\}$ we have $$f_0(kN)=f(kN)\le f(i+kN)=f_i(i+kN)\le f(N+kN)=f_0((k+1)N).$$ If follows that $$c_d(0)=\lim_{k\to\infty}\frac{f_0(kN)}{(kN)^d}\le c_d(i)=\lim_{k\to\infty}\frac{f_i(i+kN)}{(kN)^d}\le\lim_{k\to\infty}\frac{f_0((k+1)N)}{(kN)^d}$$ $$=\lim_{k\to\infty}c_d(0)(\frac{k+1}{k})^d=c_d(0),$$ hence $c_d(0)=c_d(i)$ and $c_d$ is constant. b\) If $c_d(n)=c_d$ is constant, then for $n\equiv i\ \text{mod}\ d$ $$f(n+1)-f(n)=c_d(n+1)^d-c_dn^d+g$$ with a quasi-polynomial $g$ of degree $\le d-1$. Hence deg$(\Delta f)\le d-1$ for an increasing $f$. As $\Sigma f$ is increasing for $f\ge 0$ and $\Delta\Sigma f=Ef$ we have $$\text{deg}\Sigma f\ge\text{deg}\Delta(\Sigma f)+1=\text{deg}Ef+1=d+1.$$ Since $f\ge 0$ there exists $k>0$ so that for $n>k$ all polynomial functions $f_j(n)$ are increasing. Moreover there exists $i\in\{0,\dots,N-1\}$ such that $f_i(n)\ge f_j(n)$ for all $j$ and $n>k$, if $k$ is sufficiently large. Then for $n>k$ $$f(0)+f(1)+\dots+f(n)\le f(0)+\dots+f(k)+(n-k)f_i(n)=:h(n)$$ where $h(n)$ is a polynomial of degree $\le d+1$. Consequently deg$\Sigma f\le d+1$, and b) follows. $\Box$ Theorem 4.2 and Lemma 4.4 imply [**4.5. Proposition.**]{} Let $i\in\{1,\dots,p-1\}$ be prime to $p$ and $N(p,i+np):=g_{e_i}(n)$ resp. Medim$(p,i+np):=g^0_{e_i}(n)$ the number of $H\in \mathfrak H_p$ (of maximal embedding dimension $p$) with $i+np\in H$. Then the functions $N(p,i+np)$ and $\text{Medim}(p,i+np)$ of the variable $n$ can be expressed as quasi-polynomials of degree $p-1$ having the same highest coefficient, which is independent of $n$. More precisely for each $q>p$ which is prime to $p$ $$\text{Medim}(p,q)=N(p,q-p).$$ For the last formula note that $f^0_{e_i}(0)=0, f^0_{e_i}(n)=f_{e_i}(n-1)$ for $n\ge 1$ by 4.2, hence Medim$(p,i+np)=g^0_{e_i}(n)=\sum_{k=0}^nf^0_{e_i}(k)=\sum_{k=0}^{n-1}f_{e_i}(k)= g_{e_i}(n-1)=N(p,i+(n-1)p)$. Similarly as in Proposition 4.5, if $H(p,g)$ denotes the number of $H\in \mathfrak H_p$ of genus $\le g$, then $H(p,g)$ is a quasi-polynomial of $g$ with degree $p-1$ and constant highest coefficient. As in Proposition 3.1 let $\bar S_1,\dots,\bar S_t$ be the closures of the faces of dimension $\lfloor\frac{p}{2}\rfloor$ of $C_p$ whose lattice points correspond to the symmetric $H\in \mathfrak H_p$. Let Sym$(p,i+np)$ be the number of symmetric $H\in \mathfrak H_p$ containing $i+np$. Application of Theorem 4.2 and Lemma 4.4 to $\Sigma f_{e_i}^{\bar S_j}\ (j=1,\dots,t)$ yields [**4.6. Proposition.**]{} $\text{Sym}(p,i+np)$ is, as a function of $n$, a quasi-polynomial of degree $\lfloor\frac{p}{2}\rfloor$ whose leading coefficient is independent of $n$. [**5. Asymptotic estimates and recursion formulas**]{} For an integer $q$ which is prime to $p$ let $\mathfrak H_{pq}$ be the set of all $H\in \mathfrak H_p$ with $q\in H$. We are interested in the functions $N(p,q), \text{Sym}(p,q)$ and Psym$(p,q)$ where $N(p,q)$ is the number of elements of $\mathfrak H_{pq}$, Sym$(p,q)$ resp. Psym$(p,q)$ the number of symmetric resp. pseudo-symmetric $H\in \mathfrak H_{pq}$. As in \[KKW\] and \[KW\] we associate with each $H\in \frak H_{pq}\ (q>p)$ a certain lattice path in the plane which is contained in the triangle $\Delta_0$ bounded by the line $g_0: p(X+1)+q(Y+1)=pq$ and the coordinate axes, starts on the $Y$-axis, ends on the $X$-axis and has only right or downward steps. In the following the word “lattice path” always means such a path. The set of all lattice paths for given $p,q$ is called the $(p,q)$-system. The lattice path belonging to $H\in \frak H_{pq}$ is constructed as follows. A semigroup $H\in \frak H_{pq}$ is obtained from $H_{pq}=<p,q>$ by closing some of its gaps. Each such gap $\gamma$ can be written $$\gamma=pq-(a+1)p-(b+1)q\leqno (1)$$ with a unique $(a,b)\in \Delta_0$. The set $L_H$ of all these $(a,b)$ is bounded by a lattice path and the coordinate axes. It is by definition the path associated to $H$. (-1.1,-1.1)(9,6.1) (-1.1,-1)(9,6.1) (-.5,5.5)(8,-.5) (0,0)(0,7) (0,0)(9,0) (0,3.5)(0,2.5) (0,2.5)(1,2.5) (1,2.5)(1,2) (1,2)(1.5,2) (1.5,2)(1.5,1.5) (1.5,1.5)(3,1.5) (3,1.5)(3,1) (3,1)(4,1) (4,1)(4,0) (-.5,5.5) (0,3.5) (1,2.5) (1.5,2) (3,1.5) (4,1) (4,0) \[45\](1,1)[$L$]{}\[45\](2,2)[$\Delta_0$]{} (8,-.5) \[45\](2.5,3.6)[$g_0$]{} \[180\](0,3.5)[$P_0$]{} \[70\](1,2.5)[$P_1$]{} \[45\](4,1)[$P_{m-1}$]{} \[270\](4,0)[$P_m$]{} Figure 3 We use the notation $(P_0,P_1,\dots,P_m)$ for lattice paths where $P_0$ is the point where it starts, $P_m$ the point where it ends, and the other $P_i$ are the points where after a right step a downward step follows. As is seen from (1) a downward step in the lattice path of $H$ means for the corresponding $\gamma$ an addition of q, a right step a subtraction of $p$. If $(a,b)$ is a point of the lattice path, then the corresponding $\gamma$ is the smallest element of $H$ in the residue class of $-(b+1)q$ modulo $p$, hence an element of Ap($H,p)$, and different $b$ give different elements of the Apéry set. Thus the lattice path is given by the Apéry set and conversely also determines this set. This also indicates the relation to the points of the polyhedral cone $C_p$. To $H=H_{pq}$ we may associate the empty path and the empty set $L_H$. Not every lattice path as above belongs to a semigroup. For an arbitrary lattice path let $L$ be the set of lattice points in the area bounded by the path and the coordinate axes. In order that $L=L_H$ with an $H\in \frak H_{pq}$ the following conditions must be satisfied (\[KKW\]) a\) For $(a,b),(a',b')\in L$ with $a+a'\ge q-1$ also $(a+a'-q+1,b+b'+1)$ must be in $L$. b\) For $(a,b),(a',b')\in L$ with $b+b'\ge p-1$ also $(a+a'+1,b+b'-p+1)$ must be in $L$. Lattice paths whose corresponding set $L$ satisfied a) and b) were called [*admissible*]{}, and their number was denoted by $L(p,q)$. [**5.1. Lemma.**]{} $L(p,q)$ as a function of $p$ and of $q$ is increasing. [**Proof.**]{} Let $(\tilde p,\tilde q)$ be another pair of relatively prime integers with $\tilde p\ge p,\tilde q\ge q$. Let $\tilde\Delta_0$ be the triangle bounded by the line $\tilde p(X+1)+\tilde q(Y+1)=\tilde p\tilde q$ and the axes. Clearly $\Delta_0\subset \tilde\Delta_0$ so that any lattice path in $\Delta_0$ is also one in $\tilde\Delta_0$. Let $L\subset\Delta_0$ be the set of lattice points corresponding to it. If it satisfies the admissibility conditions a) and b) above, then they are also satisfied in the $(\tilde p,\tilde q)$-system: Let $(a,b),(a',b')\in L$ and $a+a'\ge q-1$. If $(a+a'-q+1,b+b'+1)\in L$, then also $(a+a'-\tilde q+1,b+b'+1)=(a+a'-q+1,b+b'+1)-(\tilde q-q,0)\in L$. If $b+b'\ge p-1$ the proof is analogous. $\Box$ If $q=i+np$ and $g_{e_i}$ is the quasi-polynomial studied in Section 4 we have $$N(p,q)=L(p,q)+1=g_{e_i}(n)\leqno (2)$$ where the 1 comes from $H_{pq}$ or the empty lattice path. Proposition 4.5 and Lemma 5.1 imply [**5.2. Proposition.**]{} $\lim_{q\to\infty}\frac{N(p,q)}{q^{p-1}}$ and $\lim_{q\to \infty}\frac{\text{Medim}(p,q)}{q^{p-1}}$ exist and are equal. The above limits give asymptotic estimates of how many $H\in \mathfrak H_{pq}$ (of maximal embedding dimension $p$) exist. With a somewhat different approach it is shown in \[HW\] that also for an arbitrary $q$ which is prime to $p$ the function $N(p,q)$ is a quasi-polynomial in $q$ of degree $p-1$ and estimates of its (constant) highest coefficients are given, i.e. of the above limits. It was shown in \[KW\] that the admissible lattice paths starting at the point $(0,p-2)$ are in one-to-one correspondence with the semigroups of $\frak H_{p,q-p}$ (if $q-p<p$ exchange $p$ and $q-p$). Thus we have the recursion formula $$N(p,q)=N_{pq}+N(p,q-p)+1\qquad (q>p)\leqno (3)$$ where $N_{pq}$ is the number of admissible lattice paths starting at $(0,b)$ with $b\in \mathbb N, b\le p-3$. Also for the numbers $\text{Sym}(p,q)$ resp. $\text{Psym}(p,q)$ of symmetric (pseudo-symmetric) $H\in \frak H_{pq}$ one has recursion formulas $$\text{Sym}(p,q)=S_{pq}+\text{Sym}(p,q-p)+1\qquad(q>p)\leqno(4)$$ $$\text{Psym}(p,q)=P_{pq}+\text{Psym}(p,q-p)\qquad (q>p)\leqno(5)$$ where $S_{pq}$ (resp. $P_{pq}$) is the number of admissible lattice paths starting at a point $(0,b)$ with $b\le p-3$ and defining a symmetric (pseudo-symmetric) semigroup. We shall use the formulas (3)-(5) in the next section to derive explicit recursion formulas for $N(p,q), \text{Sym}(p,q)$ and $\text{Psym}(p,q)$ in the cases $p=3, p=4$. [**6. Examples**]{} The function $G(p,g)$ tells us in how many ways we can remove $g$ numbers from $\mathbb N$, where $0$ and $p$ are not removed, so that the remaining set is additively closed. In \[Ku\], Appendix C various generating functions for $\alpha=(1,\dots,1)$ and $p\le 5$ are listed. The following explicit formulas can be easily derived: $$G(3,g)=\lfloor\frac{g}{3}\rfloor+1$$ $$G(4,g)=\frac{1}{12}g^2+\frac{1}{2}g+\begin{cases}1\quad \text{if}\ g\equiv 0\ \text{mod}\ 6\cr \frac{5}{12}\ \text{if}\ g\equiv 1\ \text{mod}\ 6\cr \frac{2}{3}\quad \text{if}\ g\equiv 2\ \text{mod}\ 6\cr \frac{3}{4}\quad \text{if}\ g\equiv 3\ \text{mod}\ 6\cr \frac{2}{3}\quad \text{if}\ g\equiv 4\ \text{mod}\ 6\cr \frac{5}{12}\ \text{if}\ g\equiv 5\ \text{mod}\ 6\end{cases}= \lfloor\frac{1}{12}g^2+\frac{1}{2}g\rfloor+1$$ $$G_{sym}(3,g)=\begin{cases}1\quad \text{if}\ g\not\equiv 2\ \text{mod}\ 3\cr 0\quad \text{if}\ g\equiv 2\ \text{mod}\ 3\end{cases}\ G_{sym}(4,g)=\lfloor\frac{g}{3}\rfloor+1.$$ For $p=3$ see also Fig.1. For $G^0(p,g)$ see formula (2) of Section 4. $$\begin{matrix} g & G(3,g)&G^0(3,g)&G_{sym}(3,g)&G(4,g)&G^0(4,g)&G_{sym}(4,g)\cr \noalign{\hrule} 0&1&0&1&1&0&1\cr 1&1&0&1&1&0&1\cr 2&1&1&0&2&0&1\cr 3&2&1&1&3&1&2\cr 4&2&1&1&4&1&2\cr 5&2&2&0&5&2&2\cr 6&3&2&1&7&3&3\cr 7&3&2&1&8&4&3\cr 8&3&3&0&10&5&3\cr \end{matrix}$$ The following formulas were communicated to us by H. Knebl: $$G(5,g)=\frac{1}{135}g^3+\frac{4}{45}g^2+R(i)$$ $$G_{sym}(5,g)=\begin{cases}\frac{1}{6}g+S(i)\quad\ g\not\equiv 3\ \text{mod}\ 5\cr \qquad 0\qquad\ \ g\equiv 3\ \text{mod}\ 5\end{cases}$$ with $R(i)$ and $S(i)$, depending on $i\equiv g\ \text{mod}\ 30$, as in the table below. $$\begin{matrix} i & R(i)& S(i)\cr \noalign{\hrule} 0&(7/15)g+1&1\cr 1&(1/3)g+77/135&5/6\cr 2&(19/45)g+20/27&2/3\cr 3&(3/10)g+1/10&-\cr 4&(2/5)g+68/135&4/3\cr 5&(29/90)g+13/54&1/6\cr 6&(13/30)g+3/5&1\cr 7&(11/30)g+29/54&5/6\cr 8&(16/45)g+91/135&-\cr 9&(3/10)g+7/10&1/2\cr 10&(7/15)g+28/27&4/3\cr 11&(13/45)g+28/135&1/6\cr 12&(7/15)g+4/5&1\cr 13&(3/10)g-53/270&-\cr 14&(16/45)g+37/135&2/3\cr 15&(11/30)g+1/2&1/2\cr 16&(13/30)g+131/135&4/3\cr 17&(29/90)g+119/270&1/6\cr 18&(2/5)g+4/5&-\cr 19&(3/10)g+109/270&5/6\cr 20&(19/45)g+20/27&2/3\cr 21&(1/3)g+1/5&1/2\cr 22&(7/15)g+113/135&4/3\cr 23&(23/90)g-7/270&-\cr 24&(2/5)g+4/5&1\cr 25&(11/30)g+29/54&5/6\cr 26&(7/18)g+82/135&2/3\cr 27&(11/30)g+1/2&1/2\cr 28&(2/5)g+68/135&-\cr 29&(23/90)g+47/270&1/6\end{matrix}$$ In \[Ka\], table 1 an extensive list with values of the function $S(m,g)$ counting the semigroups with multiplicity $m$ and genus $g$ is given. If $m$ is a prime number, then $S(m,g)=G(m,g)$ for large $g$, since there are only finitely many $H\in \mathfrak H_m$ of multiplicity $<m$. In the following the recursion formulas (3)-(5) from Section 5 will be used. I\) For $\bf{p=3}$, in order to determine $N_{3q}, S_{3q}$ and $P_{3q}$, we have only to consider admissible lattice paths on the $X$-axis. These are the paths ending at $(j,0)$ with $j\le \lfloor\frac{q}{2}\rfloor-1$. They correspond to the semigroups $H_j=<3,q,2q-3(j+1)>$ where $H_j$ is symmetric if and only if $q$ is even and $j=\frac{q}{2}-1$ and $H_j$ is pseudo-symmetric if and only if $j=0$ or $q$ is odd and $j=\frac{q-1}{2}-1$ (\[KW\], Example 2.5b). Thus $N_{3q}=\lfloor\frac{q}{2}\rfloor$ and $$N(3,q)=N(3,q-3)+\lfloor\frac{q}{2}\rfloor+1\qquad (q>3),\leqno(1)$$ and for $q\ge 4$ $$\text{Sym}(3,q)=\text{Sym}(3,q-3)+\begin{cases}2\quad \text{if $q$ is even}\cr 1\quad\text{if $q$ is odd}\end{cases}\leqno(2)$$ $$\text{Psym}(3,q)=\text{Psym}(3,q-3)+\begin{cases}1\quad\text{if $q$ is even}\cr 2\quad\text{if $q$ is odd}\end{cases}.\leqno(3)$$ Clearly $\text{Sym}(3,1)=1$ and $\text{Sym}(3,2)=2$, further $\text{Psym}(3,1)=\text{Psym}(3,2)=0$. For Medim$(p,q)$, thanks to Proposition 4.5, no further discussion is necessary. $$\begin{matrix} q & N(3,q)& \text{Medim}(3,q)&\text{Sym}(3,q)&\text{Psym}(3,q)\cr \noalign{\hrule} 1&1&0&1&0\cr 2&2&0&2&0\cr 4&4&1&3&1\cr 5&5&2&3&2\cr 7&8&4&4&3\cr 8&10&5&5&3\cr 10&14&8&6&4\cr 11&16&10&6&5\cr 13&21&14&7&6\cr 14&24&16&8&6\end{matrix}$$ The numbers in the columns can also be found by using figure 1. By what was said there we have $N(3,q)=\text{Medim}(3,q)+\text{Sym}(3,q)$, and by Proposition 4.5 Medim$(3,q)=N(3,q-3)$. [**6.1. Proposition.**]{} a) $\lim_{q\to\infty}\frac{N(3,q)}{q^2}=\lim_{q\to \infty}\frac{\text{Medim}(3,q)}{q^2}=\frac{1}{12}$. b\) $\lim_{q\to\infty}\frac{\text{Sym}(3,q)}{q}=\lim_{q\to\infty}\frac{\text{Psym}(3,q)}{q}=\frac{1}{2}$. [**Proof.**]{} a) By \[KW\], Example 3.5 we have $N(3,q)=\lfloor\frac{q^2}{12}+\frac{q}{2}\rfloor+1$ which implies a). Alternately, from the recursion formula (1) follows for $q\ge 7$ that $$N(3,q)-N(3,q-6)=\lfloor\frac{q}{2}\rfloor+\lfloor\frac{q-3}{2}\rfloor+2=q$$ $$N(3,q-6(i-1))-N(3,q-6i)=q-6(i-1)\ (i\ge 1)$$ hence $$N(3,q)-N(3,q-6\lfloor\frac{q}{6}\rfloor)=\lfloor\frac{q}{6}\rfloor q-6\sum_{i=0}^{\lfloor\frac{q}{6}\rfloor-1}i=$$ $$\frac{q^2}{12}+\frac{q}{2}- 3(\frac{q}{6}-\lfloor\frac{q}{6}\rfloor+1)(\frac{q}{6}-\lfloor\frac{q}{6}\rfloor)$$ which also gives a). b\) By (2) and (3) for $q\ge 7$ $$\text{Sym}(3,q)-\text{Sym}(3,q-6)=3=\text{Psym}(3,q)-\text{Psym}(3,q-6),$$ hence $$\text{Sym}(3,q)-\text{Sym}(3,q-6\lfloor\frac{q}{6}\rfloor)=3\lfloor\frac{q}{6}\rfloor= \text{Psym}(3,q)-\text{Psym}(3,q-6\lfloor\frac{q}{6}\rfloor),$$ from which b) follows. $\Box$ II) For $\bf{p=4}$, in order to determine $N_{4q}, S_{4q}$ and $P_{4q}$, we have to investigate the lattice paths starting at $(0,1)$ or $(0,0)$ and the corresponding semigroups. By \[KW\], 2.6 there are ${{2+q'}\choose{2}}-1=\frac{1}{8}(q^2+4q+3)-1\ (q'=\frac{q-1}{2})$ lattice paths in the rectangle $\bf R$ with the corners $(0,0),(0,1),(q'-1,1),(q'-1,0)$ and all are admissible. To find $N_{4q}$ we have also to determine the admissible paths starting at $P_0=(0,1)$ and ending at a point $(q'+i,0)\ (i=0,\dots,q-2-\lfloor\frac{q}{4}\rfloor-q')$. Such a lattice path is admissible if and only if it contains $(2i,1)$ and avoids $(q'-i,1)$. Thus it is of the form $(P_0,P_1,P_2)$ with $P_1=(j,1),P_2=(q'+i,0)\ (2i\le j<q'-i)$ and defines the semigroup $$G_{ij}:=<4,q,2q-4(j+1),3q-4(q'+i+1)>.$$ Given $i$ there are $q'-3i$ such paths, the number of possible downward steps, where $i\le\lfloor\frac{q'}{3}\rfloor=\lfloor\frac{q-1}{6}\rfloor$. Note that the last condition implies that $$i\le q-2-\lfloor\frac{q}{4}\rfloor-q'\leqno(4)$$ so that we stay below the line $g_0$. Altogether there are $$\sum_{i=0}^{\lfloor\frac{q-1}{6}\rfloor}(q'-3i)= (\lfloor\frac{q-1}{6}\rfloor+1)\frac{q-1}{2}-\frac{3}{2}(\lfloor\frac{q-1}{6}\rfloor+1)\lfloor\frac{q-1}{6}\rfloor$$ such lattice paths (semigroups $G_{ij}$), and therefore for $q>4$ $$N(4,q)= \frac{1}{8}(q^2+4q+3)+(\lfloor\frac{q-1}{6}\rfloor+1)(\frac{q-1}{2}-\frac{3}{2}\lfloor\frac{q-1}{6}\rfloor) +N(4,q-4).\leqno(5)$$ Clearly $N(4,1)=1, N(4,3)=4$. In order to determine the numbers $S_{4q}$ and $P_{4q}$ we first check which of the semigroups $G_{ij}$ are symmetric resp. pseudo-symmetric. This is done by considering their genus and Frobenius number. $G_{ij}$ is obtained from $H_{4q}$ by closing $q'+i+j+2$ of its gaps, hence $G_{ij}$ has genus $$g(G_{ij})=\frac{3}{2}(q-1)-(q'+i+j+2)=2q'-(i+j+2).$$ The parallel $4(X+1)+q(Y+1)=3q+4$ to $g_0$ through $(0,2)$ passes also through $(\frac{q}{4},1)$ and cuts the $X$-axis at $(\frac{q}{2},0)$. If $j\ge \lfloor\frac{q}{4}\rfloor$, then by \[KW\], 3.1 the point $(0,2)$ corresponds to the Frobenius number of $G_{ij}$, hence $$F(G_{ij})=3(q-1)-1-2q=q-4=2q'-3$$ and $$2g(G_{ij})-F(G_{ij})-1=2(q'-i-j-1),$$ an even number. Therefore $G_{ij}$ is not pseudo-symmetric, and symmetric if and only if $j=q'-i-1$. Furthermore $q'-i-1\ge\lfloor\frac{q}{4}\rfloor$ by (4). Thus $\lfloor\frac{q-1}{6}\rfloor+1$ of the $G_{ij}$ are symmetric for $q\not\equiv 1\ \text{mod}\ 6$, and $\lfloor\frac{q-1}{6}\rfloor$ are symmetric for $q\equiv 1\ \text{mod}\ 6$. In case $j<\lfloor\frac{q}{4}\rfloor$ the point $(j+1,1)$ corresponds to $F(G_{ij})$, hence $$F(G_{ij})=3(q-1)-1-4(j+1)-q=4q'-4j-6$$ and $$2g(G_{ij})-F(G_{ij})-1=2j-2i+1.$$ In this case $G_{ij}$ is not symmetric, and pseudo-symmetric only for $j=i$. Since $j\ge 2i$ this implies $i=j=0$. According to \[KW\] 2.8, of the semigroups $\tilde G_{ij}$ corresponding to lattice paths in the rectangle $\bf R$ exactly $q'$ are symmetric and just one is pseudo-symmetric. Altogether for $q\ge 7$ $$S_{4q}=\begin{cases}\lfloor\frac{q-1}{6}\rfloor+\frac{q-1}{2}\qquad \text{if}\ q\equiv 1\ \text{mod}\ 6\cr \lfloor\frac{q-1}{6}\rfloor+\frac{q-1}{2}+1\ \text{if}\ q\not\equiv 1\ \text{mod}\ 6 \end{cases}$$ and $P_{4q}=2.$ Thus $$\text{Sym}(4,q)=\text{Sym}(4,q-4)+1+\begin{cases}\lfloor\frac{q-1}{6}\rfloor+ \frac{q-1}{2}\qquad \text{if}\ q\equiv 1\ \text{mod}\ 6\cr \lfloor\frac{q-1}{6}\rfloor+\frac{q-1}{2}+1\ \text{if}\ q\not\equiv 1\ \text{mod}\ 6 \end{cases}$$ $$\text{Psym}(4,q)=\text{Psym}(4,q-4)+2.$$ It is easy to see that $\text{Sym}(4,3)=3, \text{Sym}(4,5)=5$, $\text{Psym}(4,3)=1, \text{Psym}(4,5)=2$. $$\begin{matrix} q & N(4,q)& \text{Medim}(4,q)&\text{Sym}(4,q)&\text{Psym}(4,q)\cr \noalign{\hrule} 1&1&0&1&0\cr 3&4&0&3&1\cr 5&9&1&5&2\cr 7&17&4&8&3\cr 9&29&9&11&4\cr 11&45&17&15&5\cr 13&66&29&19&6\cr 15&93&45&25&7 \end{matrix}$$ [**6.2. Proposition.**]{} $\lim_{q\to\infty}\frac{N(4,q)}{q^3}=\lim_{q\to\infty}\frac{\text{Medim}(4,q)}{q^3}=\frac{1}{72}$. [**Proof.**]{} We know by 5.2 that the limits exists and are equal. Therefore it is enough to consider $q$ of the form $q=4m+1\ (m>0)$. By formula (5) we obtain for $q\ge 5$ $$N(4,q)-N(4,q-4)=\frac{1}{8}(q^2+4q+3)+(\lfloor\frac{q-1}{6}\rfloor+1)(\frac{q-1}{2} -\frac{3}{2}\lfloor\frac{q-1}{6}\rfloor)$$ $$=\frac{8}{3}m^2+g(m)$$ with a quasi-polynomial $g$ of degree $1, g(0)=1$. Therefore with the sum operator $\Sigma$ $$N(4,q)=\frac{8}{3}\sum_{j=1}^m j^2+(\Sigma g)(m).$$ By Fibonacci’s formula (Liber abaci 1202) $\sum_{j=1}^m j^2=\frac{1}{6}m(m+1)(2m+1)$ this implies $$N(4,q)=\frac{8}{9}m^3+h(m)=\frac{1}{72}(q-1)^3+h(m)$$ with a quasi-polynomial $h$ of degree $2$, and the assertion about the limits follows. $\Box$ $N(4,q)$ and $N(5,q)$ have been explicitly computed by H. Knebl (see \[HW\], Example 4.3).
{ "pile_set_name": "ArXiv" }
Andrzej M. Sołtan Nicolaus Copernicus Astronomical Center [Bartycka 18, 00-716 Warsaw, Poland]{} The X-ray background (XRB) is produced by a large number of faint sources distributed over a wide range of redshifts. The XRB carries information on the spatial distribution and evolution of these sources. [The goals of the paper are: 1. to determine the redshift distribution of the soft X-ray background photons produced by all types of extragalactic sources, in order to relate fluctuations of the background to the large scale structures, 2. to determine the redshift distribution of the soft XRB produced by AGN in order to calculate the evolution of the AGN X-ray luminosity density.]{} [A set of major X-ray surveys is used to determine the redshift distributions of the X-ray sources selected at various flux levels. Simple analytic fits to the data allow us to determine the smooth relationship between the redshift distribution and the source flux. The redshift distribution of the integral XRB flux is obtained by averaging the fits over the source counts.]{} [It is shown that the distribution of extragalactic XRB photons in the $0.5-2$keV band is adequately represented by the function: $d\,n_{\rm XRB} / d\log z = 5.24\:z^{1.52}\,\exp(-z/0.63)$. The huge voids postulated to explain the cold spots in the CMB maps create dips in the total XRB flux. However, the expected magnitude of the effect is comparable to the fluctuation amplitude of the XRB generated by the individual sources contributing to the background. The cosmic evolution of the AGN X-ray luminosity density up to redshift of $\sim\!5$ is calculated in an elegant and straightforward way. Systematic uncertainties of the present method are assessed and shown to be small. At redshift greater than one the present results could be compared directly with some recent estimates obtained in a standard way and the agreement between both methods is very good.]{} Keywords[X-rays: diffuse background – intergalactic medium – X-rays: galaxies ]{} Introduction \[intro\] ====================== The X-ray background (XRB) is generated mostly by discrete extragalactic sources (e.g. @lehmann01 [@kim07], and references therein), predominantly by various types of Active Galactic Nuclei (AGN) and cluster of galaxies. A question of flux, luminosity and redshift distributions of these sources has been discussed in a great number of papers for the last $30$ years. One of the major outcome of these investigations is the conclusion that X-ray sources associated with the AGN are subject to strong cosmic evolution (e.g. @miyaji00, @silverman07, and references therein). As a result of the evolution, the redshift distribution of the XRB flux is wide. Thus, the integral XRB comprises the information on the large scale distribution of the X-ray sources over a wide redshift range. In the present paper the redshift distribution of the XRB photons is investigated in detail. The analysis is based on an extensive observational data selected from several published X-ray sky surveys. A convenient analytic approximations are applied to model the observed redshift histograms of the extragalactic X-ray sources selected at several flux levels. These distributions are weighted by the source counts and summed up to obtain the redshift distributions of the integral XRB. Next, the redshift distribution is used to define a relationship between the XRB signal and the large-scale fluctuations of the matter spatial distribution. The investigation has been raised by the recent report on the huge void generating a dip in the surface brightness of the radio background (@rudnick07). This void, responsible for the deficit of the radio surface brightness, allegedly generates also a cold spot in the CMB map via the late-time integrated Sachs-Wolfe effect. Although the careful statistical analysis by [@smith08] has not confirmed the existence of this particular “cold spot” in the radio survey, a relationship between the large scale features of the matter distribution and the integrated sky brightness in various energy bands is a problem deserving some interest. Apart of the question of the XRB fluctuations induced by voids, the XRB redshift distribution is interesting per se, as it allows to assess the evolution of AGN phenomenon. A standard way to estimate a rate and type of this evolution is based on the examination of the X-ray luminosity functions determined at the consecutive redshift bins. Unfortunately, the X-ray surveys produce flux-limited rather than luminosity-limited samples of sources. In effect, luminosity functions at different redshifts cover different luminosity ranges. This in turn severely impedes estimates of the luminosity function over a wide range of luminosities and redshifts. The total level of nuclear activity in galaxies within unit volume is given by the integral of the X-ray luminosity function. The question of the AGN cosmic evolution constitutes one of the central problems of observational cosmology, and has been investigated for the last forty years (this issue was for the first time recognized by @schmidt68). Here a question of the AGN evolution is addressed without the calculations of the X-ray luminosity function. The available observational data on X-ray source counts and redshifts are used to evaluate the redshift distribution as a function of source flux. This relationship and the source counts allow to calculate the redshift distribution of the total XRB and the integral luminosity density generated by the AGN as a function of redshift. The organization of the paper is following. First, I present the formulae used in calculations. Next, in Sec.\[catalogs\], the basic information on the observational material extracted form the various archives is given. Since the comprehensive characteristics of the data and the source catalogs are described in the original papers, only the basic properties of the material are presented here. The numerical fits to the observed distributions are obtained in Sec. \[fits\]. In that section the calculations of the redshift distribution of the XRB photons are described in details. These results are applied in the Sec. \[supervoids\] to quantify the relationship between the voids and the XRB variations. In Sec. \[agn\] the distribution of the XRB flux produced by AGN is used to calculate the evolution of the AGN activity. Finally, potential sources of errors inherent in the present method are discussed in the Sec. \[discussion\]. The ‘canonical’ standard cosmology is assumed throughout, with $H_0 = 70$kms$^{-1}$Mpc$^{-1}$, $\Omega_m = 0.3$, and $\Omega_\Lambda = 0.70$. Basic relationships \[formulae\] ================================ In the present approach the X-ray source catalogs are used to construct redshift distributions of the extragalactic sources as a function of the source flux. At this stage, a question of the source (absolute) luminosities is not addressed. Let $N(S)$ denotes the X-ray source counts, i. e. number of sources brighter than $S$ in a unit solid angle, and $f_S(z) = dn(z\!\mid\!S)/d\log z$ is the redshift distribution of sources with flux $S$. Then, the redshift distribution of the XRB surface brightness, $b(z)$, is equal to: $$b(z) = \frac{1}{b} \int\! dS\,f_S(z)\:S \left|\frac{dN(S)}{dS}\right|\,, \label{bz}$$ where the integration covers the entire “interesting” range of source fluxes $S$ and $b$ denotes the integral background flux: $$b = \int\! b(z)\;d\log z = \int\! dS \,S \left|\frac{dN(S)}{dS}\right|\,. \label{b}$$ It is assumed that the $f_S(z)$ distributions are normalized: $$\int\! f_S(z)\;d\log z = 1\,.$$ Here the integration limits cover the total range of redshifts occupied by X-ray sources. The actual limits of the “interesting” range of fluxes is discussed below. The luminosity density, $\varepsilon(z)$, i. e. a total luminosity $L$ generated in a unit comoving volume, $V$: $$\varepsilon(z) = \frac{dL}{dV}\,,$$ is related to the flux distribution $b(z)$ and the luminosity distance, $D_L(z)$: $$\varepsilon(z)\;\frac{dV}{d\log z} = 4\,\pi\,D_L^2(z)\:b(z)\,. \label{epsilon}$$ The cosmological relationships between the comoving volume and the luminosity distance in a flat space with $\Lambda \neq 0$ is given by [@hogg99]: $$\frac{dV}{dz} = \frac{c}{H_0}\:\frac{D_L^2}{(1+z)^2 \sqrt{\Omega_m (1+z)^3 + \Omega_\Lambda}}\,. \label{dvdz}$$ Combining Eqs. \[epsilon\] and \[dvdz\] we finally get: $$\varepsilon(z) = 4\pi\:\frac{H_0}{c}\;\frac{(1+z)^2 \sqrt{\Omega_m(1+z)^3 + \Omega_\Lambda}}{\ln\!10\;z}\; b(z)\,. \label{lum_dens}$$ Thus, to calculate the distributions $b(z)$ and $\varepsilon(z)$, the source counts $N(S)$ and the functions $n(z\,|\,S)$ have to be determined using the observational material. In the next section the available X-ray surveys are examined from this point of view. Observational material \[catalogs\] =================================== Because the high imaging efficiency of X-ray telescopes in the soft band and numerous extensive identifications programs, we concentrate on the X-ray band of $0.5 - 2.0$keV. At these energies a fraction of the background resolved into discrete sources exceeds $90$% and is higher than in the other bands (e.g. @moretti03, @brandt05). Also a fraction of identified objects with measured spectroscopic or photometric redshifts is relatively high. Equation \[bz\] shows that the $b(z)$ distribution is sensitive to sources which perceptibly contribute to the XRB. Consequently, one needs to calculate the $n(z\,|\,S)$ functions over a quite wide range of fluxes. To achieve this objective I have examined numerous X-ray surveys and selected several major source catalogs for further analysis. The overall characteristics of those catalogs are listed in Table \[surveys\]. A common name of the survey/catalog is given in column 1. From the each catalog, sources for further processing have been extracted within fixed range of fluxes defined on column 2. The numbers of all sources, extragalactic and AGN with known redshifts are given in columns 3, 4, and 5, respectively. Statistical requirements which have to be satisfied by the source samples to properly determine the $n(z\,|\,S)$ distribution are different than those for the luminosity function calculations. The individual sample should contain sources from possible narrow range of fluxes, but the sample has not to be flux limited. The sample provides unbiased estimate of the redshift distribution as long as the process of identification and redshift measurements does not introduce spurious correlation between flux and redshift. [llccc]{} Name &&\ && All &Extragalactic& AGN\ RBS & $1.0\times 10^{-12}-5.0\times 10^{-11}$ & 1764 & 1054 & 681\ NEP & $5.0\times 10^{-14}-1.0\times 10^{-12}$ & 361 & 248 & 192\ RIXOS & $2.5\times 10^{-14}-5.0\times 10^{-13}$ & 393 & 318 & 235\ XMS & $1.0\times 10^{-14}-2.0\times 10^{-13}$ & 275 & 256 & 231\ CDFS & $5.0\times 10^{-17}-1.0\times 10^{-15}$ & 205 & 201 & 197\ CDFN & $1.5\times 10^{-17}-5.0\times 10^{-15}$ & 425 & 412 & 268\ The [*ROSAT*]{} Bright Survey (RBS) ----------------------------------- The identification program of the brightest sources detected in the [*ROSAT*]{} All-Sky Survey, known as [*ROSAT*]{} Bright Survey, resulted in a sample of 2072 sources with the total count rate above $0.2$s$^{-1}$ (@schwope00). More than $99.5$% of sources in the final catalogue is identified. The survey covers high galactic latitudes ($|b|> 30\deg$). After the removal of the Virgo clusters and Magellanic Clouds regions, the catalog contains 2012 sources. In the energy band of $0.5-2.0$keV, 1773 RBS sources generate flux between $1.0\times 10^{-12}$ergcm$^{-2}$s$^{-1}$ (hereafter [[*cgs*]{}]{}) and $5.0\times 10^{-11}$ [[*cgs*]{}]{}. As one might expect, only for a small fraction of the RBS sources the redshifts are undetermined and relatively large number of sources is associated with galactic sources, mostly late type stars and cataclysmic variables. Since the RBS sample covers rather wide range of fluxes, it is useful for our purposes to divide it into several subsamples with narrow flux limits and to estimate the $n(z\,|\,S)$ distribution for the each set separately. We define the bright source sample, labeled RBS(b), which contains sources with $4.0\times 10^{-12} < S < 1.0\times 10^{-11}$ [[*cgs*]{}]{}. Of $365$ RBS sources in this flux range, $129$ is identified with galactic objects and for the other $15$ sources redshifts have not been measured. Eventually, the sample contains $221$ extragalactic objects with known redshifts. Nearly half of the sample, viz. $97$ sources are identified with clusters of galaxies and normal galaxies. The median flux[^1] in the extragalactic subsample $S{\rm_m} = 5.5\times 10^{-12}$ [[*cgs*]{}]{}. ![image](z_hist_all_c.eps){width="90.00000%"} The flux limits of $2.0\times 10^{-12} < S < 4.0\times 10^{-12}$ [[*cgs*]{}]{} have been adopted for the medium flux sample, RBS(m). Within this flux range the RBS comprises of $669$ objects including $272$ galactic stars. Among the $397$ extragalactic sources, $233$ are identified with AGN with known redshifts. The median flux in the RBS(m) extragalactic sample, $S_{\rm m} = 2.7\times 10^{-12}$ [[*cgs*]{}]{}. The faint source sample, RBS(f), contains sources with $1.0\times 10^{-12} < S < 2.0\times 10^{-12}$ [[*cgs*]{}]{}. Of $599$ RBS sources in this flux range $268$ is identified with galactic objects and $24$ has unknown redshifts. Thus, the RBS(f) sample contains $307$ sources extragalactic objects with the median flux of $1.6\times 10^{-12}$ [[*cgs*]{}]{}. The redshift distributions of sources in the RBS (b), (m), and (f) samples are shown in three upper left panels in Fig. \[z\_distr\]. The integrals of the histograms for all the samples in Fig. \[z\_distr\] are normalized to unity. The distributions are plotted using logarithmic redshift bins with $\Delta \log z = 0.125$. Each histogram is labeled with the survey name and the median flux in [[*cgs*]{}]{}. Analytic fits will be discussed in the next section. The [*ROSAT*]{} North Ecliptic Pole Survey (NEP) ------------------------------------------------ The deepest exposure of the [*ROSAT*]{} All-Sky Survey (RASS) is centered at the north ecliptic pole (@voges99). The RASS of this region has been used to construct statistically well defined sample of X-ray sources above a flux limit $\sim\!2\times 10^{-14}$ [[*cgs*]{}]{} (@henry06) which have been followed-up by the optical observations (@gioia03). The identification rate in the final catalog of $443$ sources is very high ($99.6$%). Within the flux limits of $5.0\times 10^{-14}$ and $1.0\times 10^{-12}$ [[*cgs*]{}]{}the NEP survey provided $361$ sources. After excluding $113$ galactic stars, we are left with $248$ extragalactic sources; for $3$ sources the redshift is unknown. The redshift histogram of $245$ sources (including $53$ clusters) is shown in Fig. \[z\_distr\]; the median flux in this sample is equal to $1.2\times 10^{-13}$ [[*cgs*]{}]{}. The [*ROSAT*]{} International X-ray/Optical Survey (RIXOS) ---------------------------------------------------------- This [*ROSAT*]{} medium-sensitivity survey consists of sources found in $82$ PSPC pointing observations at high galactic latitudes ($|b|>28\deg$). A flux limit of $3\times 10^{-14}$ [[*cgs*]{}]{} was adopted in $64$ fields and $8\times 10^{-14}$ [[*cgs*]{}]{} in the remaining $18$ fields. The source selection procedures, optical identifications and the final catalog are given by [@mason00]. For the purpose of the present analysis, $393$ sources with fluxes between $2.5\times 10^{-14}$ and $5\times 10^{-13}$ [[*cgs*]{}]{} have been selected. Within these flux limits $75$ sources are associated with galactic stars. Of the remaining $318$ sources, the redshifts of three objects are unknown, and $49$ sources are still unidentified. The redshift distribution of $266$ sources (including $33$ clusters) is shown in Fig. \[z\_distr\]. The median flux in this subsample $S_{\rm m} = 5.2\times 10^{-14}$ [[*cgs*]{}]{}. The [*XMM-Newton*]{} serendipitous survey (XMS) ----------------------------------------------- The [*XMM-Newton*]{} serendipitous survey (XMS) has been constructed in a similar way as the [*RIXOS*]{}. More than $300$ sources have been isolated in $25$ high galactic latitude ($|b|>22\deg$) pointings covering $\sim\!3$deg$^2$ of the sky (@barcons07). In the $0.5-2.0$keV band the sample is complete above $1.5\times 10^{-14}$ [[*cgs*]{}]{}, and contains many weaker sources. I have extracted from the original catalog $275$ sources with fluxes between $1.0\times 10^{-14}$ and $2.0\times 10^{-13}$ [[*cgs*]{}]{}. The sample is completely identified; it contains $19$ stars and $2$ clusters of galaxies. For $23$ objects the redshifts are unknown. The redshift distribution of $233$ sources is shown in Fig. \[z\_distr\]. The median flux in this subsample $S_{\rm m} = 1.9\times 10^{-14}$ [[*cgs*]{}]{}. The Chandra Deep Field–South (CDFS) ----------------------------------- The $1$Ms [*Chandra*]{} observations known as the [*Chandra*]{} Deep Field South are described by [@giacconi02]. The catalog of sources detected in this field by two independent algorithms contains $304$ objects, of which $275$ have determined fluxes in the $0.5-2.0$keV band. For further processing $205$ sources with fluxes in the range $5.0\times 10^{-17} - 1.0\times 10^{-15}$ [[*cgs*]{}]{} have been selected. Four sources are identified with stars. The redshifts either spectroscopic (@szokoly04, @ravikumar07) or photometric (@zheng04) are known for $200$ sources; one source remains unidentified. The redshift distribution is shown in Fig. \[z\_distr\]. The median flux in this sample $S_{\rm m} = 2.1\times 10^{-16}$ [[*cgs*]{}]{}. The sample contains one galaxy group; two other sources apparently are not associated with the activity in the galactic nuclei (@lehmer06). It is likely, however, that more objects in the CDFS survey should be classified as off-nuclear sources. The available data do not allow for unambiguous separation of AGN and off-nuclear sources at the low flux levels in the CDFS. This reservation holds also for the CDFN samples below. The Chandra Deep Field–North (CDFN) ----------------------------------- The ultra deep Chandra field, $2$Ms exposure, CDFN, resulted in a catalog of $503$ sources detected over $0.12$ sq.deg. (@alexander03). Optical follow-up observations by [@barger03] have rendered a large number of spectroscopic and photometric redshifts. Several more redshifts are taken from [@reddy06], [@donley07] and [@georgakakis07]. In the present investigation, the CDFN catalog has been divided into two samples of bright (b) and faint (f) sources. The (b) sample contains $181$ sources between $2.5\times 10^{-16}$ and $5.0\times 10^{-15}$ [[*cgs*]{}]{}; ten sources have been identified with the galactic stars; at least one has been categorized as ‘starburst’ galaxy (@georgakakis07), for $49$ objects the redshifts have not been measured. The sample consists of $122$ sources, mostly AGN. The median flux in this subsample $S_{\rm m} = 7.3\times 10^{-16}$ [[*cgs*]{}]{}. The faint CDFN sample has been selected between $1.5\times 10^{-17}$ and $2.5\times 10^{-16}$ [[*cgs*]{}]{}. Among $244$ sources satisfying these flux limits, three sources have been identified with stars, $51$ – with the starburst galaxies (@georgakakis07) and for $43$ objects the redshifts have not been measured. The final sample used in the calculations contains $198$ sources with $147$ confirmed AGN. The median flux in the sample $S_{\rm m} = 9.0\times 10^{-17}$ [[*cgs*]{}]{}. The redshift histograms for the (b) and (f) samples are shown in Fig. \[z\_distr\]. The numbers of objects unidentified or without redshift are in some samples quite large. Hence one could expect that the corresponding redshift histograms are not representative for the whole population of sources at given flux. Below this question is de facto worked out where we construct an analytic function which simultaneously fits all the histograms. Approximations and fits \[fits\] ================================ The redshift distribution of sources selected at fixed flux, $n(z|S)$, is a intricate function of a number of parameters, such as the luminosity function, the relationship between the luminosity and observable flux, and the relationship between the volume and redshift. The luminosity function itself depends on redshift and both the latter relationships depend on the cosmological model. However, the existing estimators of the $n(z|S)$ function represented by the nine histograms in Fig. \[z\_distr\] are strongly affected/degraded by the statistical noise. It implies that a simple analytic function with $2-3$ free parameters will provide a statistically satisfactory fit to the observed distributions. It appears that the histograms in Fig. \[z\_distr\] are adequately reproduced by: $$f_S(z) = f_0\;z^\alpha\:e^{-z/z_{\rm c}}\,, \label{fit_1}$$ where $f_0=f_0(S)$, $\alpha = \alpha (S)$, and $z_{\rm c}=z_{\rm c}(S)$ are three parameters fitted to the the histograms $n(z|S_i)$, $i = 1,... ,9$. In Fig. \[z\_distr\] the least square fits for all the distributions are shown with the dotted curves. Apart from a few pronounced features visible in the plots which represent the large scale structures reported in the literature (e.g. @barger02, @gilli03), analytic fits seem to adequately reproduce the observed distributions. It is found that only $z_{\rm c}$ is strongly correlated with $S_{\rm m}$, while the fits do not indicate any statistically significant correlation between $\alpha$ and $S_{\rm m}$. In Fig. \[z\_c\_vs\_s\] the best fit values of $\alpha$ are shown with crosses. The labels and scale on the left-hand ordinate refer to $z_{\rm c}$, and on the right-hand – to $\alpha$. Since the simultaneous fitting of $\alpha$ and $z_{\rm c}$ introduces a spurious correlation between these two parameters, the $\alpha$ parameter has been fixed at the average value found for the $9$ samples, $\bar{\alpha} = 1.934$. Effectively, it means that the shape of the $n(z|S)$ function is fixed and the only dependence on $S$ is limited to the horizontal shift along the $z$ axis. In Fig. \[z\_c\_vs\_s\] the best fit parameters $z_{\rm c}$ found for the fixed $\alpha$ are plotted against the median flux $S_{\rm m}$. In agreement with the expectations, the $z_{\rm c}$ increases with diminishing flux $S_{\rm m}$ over a wide range of fluxes. However, a clear flattening of the relationship is observed below $\sim\!10^{-14}$ [[*cgs*]{}]{}. This apparent absence of correlation between $z_{\rm c}$ and $S$ results from the well-known fact that in the X-ray surveys at low flux levels the maximum detected redshift remains stable while significantly increases fraction of intrinsically weak sources. ![ []{data-label="z_c_vs_s"}](s_med_z_cut_all_c.eps){width="60.00000%"} The points in the $z_{\rm c} - S_{\rm m}$ relationship above $\sim 10^{-14}$ [[*cgs*]{}]{} seem to follow the power law. In the subsequent calculations it is assumed that this relationship is in fact well approximated by the power law in the whole range of fluxes between $10^{-14}$ and $10^{-11}$ [[*cgs*]{}]{}, although the data coverage is rather sparse. Below $S \approx 10^{-14}$ [[*cgs*]{}]{} the data are insufficient to delineate precisely the $z_{\rm c} - S_{\rm m}$ relationship. I have assumed tentatively that $z_{\rm c}$ remains constant and is equal to the average value found for three [*Chandra*]{} samples. The $z_{\rm c}$ in CDFN(b), CDFS and CDFN(f) are equal to $0.69$, $0.69$ and $0.51$, respectively. The average weighted by the uncertainties $\bar{z}_c = 0.61$. In Fig. \[z\_c\_vs\_s\] the model relationship $z_{\rm c} - S$ used in the calculations is shown with the dotted line. The $z_{\rm c} \sim S$ relationship with the fixed $\alpha$ parameter and fixed normalization of the integral: $$\int f_S(z)\;d\log z = 1\,,$$ eliminates formally any free parameters in the fitting the analytic function to the nine histograms in Fig. \[z\_c\_vs\_s\]: $$f_S(z)=\frac{\ln 10}{\Gamma(\alpha)\;z_{\rm c}^\alpha}\;z^\alpha\;e^{-z/z_{\rm c}}\,, \label{fit_2}$$ where $\alpha = 1.934$, $\Gamma(\alpha) = 0.9739$ is the gamma function and $z_{\rm c}$ is specified for each sample by the $z_{\rm c} \sim S$ relationship. Analytic distributions defined in Eq. \[fit\_2\] are shown in Fig. \[z\_distr\] with the solid curves. In all the histograms the model distribution is astonishingly close to the corresponding best three-parameter fit represented by the dotted curves. Most deviations visible in some plots are easily explained by the statistical nature of the problem and/or the large scale structures present in the catalogs based on the localized sky area (@barger02, @gilli03). Systematic shifts between the fits are present in three histograms below $S = 10^{-14}$ [[*cgs*]{}]{}. It is a direct result of the assumption of a single $z_{\rm c} = 0.61$ value for all three [*Chandra*]{} samples. It is noticeable that the constant width (in $\log z$) model fits adequately represent the data over the full range of fluxes. Small differences in the width between the three-parameter fits and the final model which are visible in the NEP, RIXOS and XMS data, apparently do not represent the systematic effects. In the NEP and RIXOS histograms the final model is slightly narrower than the individual fits, while in the XMS sample it is wider. To effectively use the Eq. \[bz\] one needs the representation of the source counts $dN(S)/dS$ over the whole range of fluxes $S$. The parametrization by [@moretti03] adequately suits the present calculations. The smooth functional form for $N(S)$ proposed by Moretti et al. accurately reproduces the observed counts below $10^{-11}$ [[*cgs*]{}]{} down to [*Chandra*]{} threshold of $\sim\!2\times 10^{-17}$ [[*cgs*]{}]{}. Sources within these flux limits generate more than $90$% of the XRB and smooth extrapolation of the [@moretti03] counts down to $\sim\!3\times 10^{-18}$ [[*cgs*]{}]{} is consistent with the entire XRB. Substituting all the components into Eq. \[bz\] we finally get the redshift distribution of the XRB photons. It is shown with the solid curve in the bottom right panel in Fig. \[z\_distr\]. The same normalization has been applied to facilitate comparison with the distributions derived for the individual samples. Points in the plot are discussed below in the Sec. \[discussion\]. A suitable representation of the $b(z)$ distribution has been found using a smooth function of the same form as for the individual redshift histograms. The function: $$b_{\rm fit}(z) = 5.24\;z^{1.52}\;e^{-z/0.63}\,,$$ reproduces the derived distribution of $b(z)$ with the relative error of less than $4$% for $0.06 < z < 6$. It is shown with dotts in the bottom right panel (with normalization rescaled to conform to all the plots in Fig. \[z\_distr\]). XRB and Supervoids \[supervoids\] ================================= The distribution of the XRB photons $b(z)$ peaks at redshift $z\approx 1$ and $50$% of the background originates between the redshifts of $0.4$ and $1.4$ (for $80$% the redshift limits are $0.2$ and $2.1$). Thus, very large structures of the matter distribution at redshift within these limits would generate fluctuations of the integral XRB. As an example I discuss below the X-ray signature of the huge void postulated by [@rudnick07]. The arguments based on the radio survey in favor of the void with a radius of $\sim\!140$Mpc in Eridanus have been questioned (@smith08). Nevertheless, the Integrated Sachs-Wolfe effect operating on extremely large structures of matter remains a valid explanation of the strongest CMB fluctuations. The redshift separation, $\Delta z$, corresponding to the far side and the near side of the void with diameter $R_o$ centered at redshift $z$ is equal to (e.g. @hogg99): $$\Delta z = \frac{2 R_o}{c / H_o} \sqrt{\Omega_m (1+z)^3 + \Omega_\Lambda}\,.$$ Using the $b(z)$ distribution we assess the fractional deficit of the XRB $\delta = \Delta b / b$ created by the completely empty region of size $280$ Mpc. Such void would generate $|\delta| = 5.7$% and $\delta = 4.9$% at redshifts $z=0.5$ and $1$, respectively. Assuming spherical shape of the void, its angular diameter would be $8{\ensuremath{.\hspace{-3pt}^\circ}}4$ and $4{\ensuremath{.\hspace{-3pt}^\circ}}7$ at these redshifts. The XRB depression produced by the void should be compared to the XRB intrinsic fluctuations resulting from the discrete nature of sources generating the background. Assuming purely random distribution of sources, the rms fluctuations of the XRB are defined by the source counts $N(S)$: $$\sigma_b = \left[ \int_{S_{\rm min}}^{S_{\rm max}} \;dS\:S^2\:\omega\; \left|\frac{dN(S)}{dS}\right|\;\right]^{\;1/2}\,,$$ where $\omega$ is the solid angle subtended by the investigated area. Using the the [@moretti03] counts and $S_{\rm max} = 1\cdot 10^{-11}$ [[*cgs*]{}]{} (the amplitude of the XRB fluctuations is dominated by the contribution of sources at the bright end of counts), we get $\delta b/b = 0.035$ and $0.024$ for the circular areas of radius $2^\circ$ and $3^\circ$, respectively. Thus, at $z=0.5$ the signal-to-noise ratio for the void detection amounts to $\sim\!2.4$. In the case of $z=1$ the S/N drops to just $1.4$. To reduce the amplitude of the XRB fluctuations one should remove from the XRB the contribution of bright sources. If the $S_{\rm max}$ is decreased to $1\cdot 10^{-12}$ [[*cgs*]{}]{}, the significance of the void signal reaches $3.5\,\sigma$ at $z=0.5$ and $2.0\,\sigma$ at $z=1$. So, only the low redshift voids would produce the XRB deficits significantly stronger than the statistical fluctuations. Redshift distribution of the XRB and the AGN evolution \[agn\] ============================================================== To assess the distribution of the XRB generated just by the AGN, I have repeated all the procedures described in the previous sections using the samples constructed exclusively from the AGN. The AGN sources are easily separated from the clusters and nearby normal galaxies. However, the distinction between the nuclear activity and stellar emitters in the case of distant and weak sources becomes problematic. Such sources are present in both [*Chandra*]{} surveys. One should keep in mind this limitations in the present investigation. Nevertheless, the AGN are a dominating constituent of all the samples exploited in the paper and even the moderate contamination of the AGN subsamples with the off-nuclear sources would not affect significantly our calculations (see below). The numbers in the AGN samples are smaller than in the full samples and parameter estimates are subject to slightly larger uncertainties. Clusters and normal galaxies populate on the average lower redshift bins and the histograms for the AGN analogous to those in Fig. \[z\_distr\] are shifted towards the higher redshifts. We notice also a weak correlation of the best fit $\alpha$ parameter with the source flux – a shape of of the redshift distribution (in $\log z$ bins) varies with $S_{\rm m}$. In Fig. \[z\_c\_vs\_s\_agn\] the values of $\alpha$ in the nine samples are shown with crosses. The regression line of $\log\alpha$ on $\log S$ is used to fix the value of $\alpha$ for the each sample and to calculate the best fit parameter $z_{\rm c}$. These new $z_{\rm c}$ are shown in Fig. \[z\_c\_vs\_s\_agn\] with the squares. Finally, the best fit line $\log z_{\rm c} \sim \log S_{\rm m}$ is calculated for six brighter samples to obtain $z_{\rm c}$ for $S > 1.9\cdot 10^{-15}$ [[*cgs*]{}]{}. A constant $z_{\rm c}$ is assumed for lower fluxes, and the complete $z_{\rm c} \sim S_{\rm m}$ relationship adopted for further computations is shown in Fig. \[z\_c\_vs\_s\_agn\] with the dotted lines. ![ []{data-label="z_c_vs_s_agn"}](s_med_z_cut_agn_c.eps){width="60.00000%"} The X-ray source counts used in the present case should be limited to the AGN only. Two other major classes of sources contributing to the counts are associated with clusters and normal/starburst galaxies. The analytic formula obtained by [@moretti03] quite accurately represents counts of all the types of extragalactic sources, but the relative contribution of the each class in the total counts is not well established. One should notice, however, that most of the XRB is produced by sources in the middle range of fluxes considered here, while the cluster contribution is significant only at the bright end of counts and the normal and starburst galaxies populate mostly the faint end of counts. Vertical bars at the bottom of Fig. \[z\_c\_vs\_s\_agn\] divide the flux range of $10^{-17} - 10^{-11}$ [[*cgs*]{}]{} into $9$ contiguous bands corresponding approximately to fluxes surveyed by the source samples defined in the paper. The numbers between the bars give the relative contribution of each flux band to the total XRB. To extract the cluster and starburst galaxies contributions we corrected the total counts in the following way. In the RBS(b) sample clusters constitute $40$% of all the extragalactic sources. The slope of the cluster counts at the bright end amounts approximately to $1.3$ (@degrandi99). Substantially flatter slope than that for the total counts reduces the relative cluster contribution at lower fluxes. Although the cluster counts are not well constrained below $\sim\!10^{-12}$ [[*cgs*]{}]{}, their contribution to the total counts drops at the faint end of counts to a negligible level. The source counts attributed to the AGN are assessed by subtracting the cluster counts from the total counts defined by the [@moretti03] formula. The normal and starburst galaxies are relatively abundant in the CDFN(f) sample. Of $241$ extragalactic sources, $147$ have been classified as ‘AGN’, $51$ as ‘starburst’ and for the other $43$ the redshift is unknown. The absolute maximum content of the non-AGN sources in the CDFN(f) sample amounts to $(51+43)/241 \equiv 39$%, assuming that all sources with undetermined redshift are starburst galaxies. The amount of the non-AGN sources at higher flux levels drops quickly. In the XMS sample none extragalactic source with known redshift has been classified as normal or starburst galaxy. The maximum possible contribution of the starburst galaxies at the low flux levels has been accounted for by flattening the slope of the [@moretti03] counts below $10^{-14}$ [[*cgs*]{}]{} to reproduce the reduction of the AGN in the CDFN(f) sample by $39$%. The counts modified this way have been substituted into Eq. \[bz\] to obtain the redshift distribution, $b_{\rm AGN}(z)$. ![ []{data-label="evolution"}](lg_ve_z.eps){width="70.00000%"} One can use the observed distribution of the background flux produced by the AGN to calculate the cosmic history of the luminosity density generated by these objects. The $b_{\rm AGN}(z)$ distribution is inserted into Eq. \[lum\_dens\] which relates the cosmological evolution of the X-ray luminosity density, $\varepsilon(z)$, to the redshift distribution of the background, $b(z)$. Variations of the luminosity density obtained this way are shown in Fig. \[evolution\] with the solid curves. The data are displayed in three panels as a function of redshift, $z$, logarithm of $(1+z)$ and the cosmic time, assuming $t_0 = 13.47\cdot 10^9$ years for the present age of the Universe[^2]. The accuracy of the present $\varepsilon(z)$ estimate depends strongly on a quality of our $b(z)$ fits. Relatively small numbers of sources at redshifts below $\sim\!0.03$ and above $\sim\!3$ generate large statistical fluctuations and weakly constrains the analytic fits $b(z)$ in these redshift ranges. Hence, the present estimates of $\varepsilon(z)$ are also subject to large uncertainties at low and high redshifts. In order to assess the importance of the $N(S)$ uncertainties on the present estimates of $\varepsilon(z)$, I have plotted in Fig. \[lum\_dens\] with the dotted curves the $\varepsilon(z)$ function using the original [@moretti03] formula, i.e. assuming no corrections for clusters and starburst galaxies. The discrepancies between both solutions do not exceed $20$% for redshifts below $\sim\!3$. It implies that our procedure to isolate the contribution of AGN from the total counts, albeit crude, does not contribute significantly to the final errors of $\varepsilon(z)$. Discussion \[discussion\] ========================= The main objectives of the present investigation, viz. estimates of the redshift distribution of the XRB photons, $b(z)$, and the evolution of the AGN luminosity density, $\varepsilon(z)$, have been achieved using the smooth, analytic fits to the observed source redshift histograms. The present method is conceptionally simple and computationally straightforward. Unfortunately, it does not provide error estimates. The major sources of uncertainties have been indicated in the previous section. Here a quantitative estimate of the errors is discussed. The errors of the present measurement of $b(z)$ are generated by the statistical nature of the investigated material and a chain of approximations applied to substitute the observed redshift distributions centered on a selected fluxes by an analytic function $f_S(z)$ continuous in both parameters, $z$ and $S$. In fact, the visual inspection of the analytic fits displayed with the solid curves in Fig. \[z\_distr\] reveals some deviations from the redshift histograms. To estimate the significance of these differences, the calculations have been performed using the actual histograms shown in Fig. \[z\_distr\] with broken solid lines instead of $f_S(z)$. For the each value of flux $S$ in the range $10^{-17} - 10^{-11}$ [[*cgs*]{}]{}, the corresponding redshift distribution has been obtained by the linear interpolation between two histograms from the samples centered on the median fluxes nearest to $S$. The results of this procedure are shown in the lower right panel of Fig. \[z\_distr\] with dots. Generally good agreement between the distribution of points and the solid curve proves that the analytic approximations do not introduce perceptible systematic errors in the present investigation. It appears that the relatively large deviations for three data points (centered at redshifts: $0.087$, $0.65$, and $1.16$) result purely from the large scale structures. This is particularly likely for the first bin ($0.075 < z < 0.1$), where the discrepancy between the fits is produced entirely by the excess of sources in the localized NEP survey. One should also notice, that the uncertainties of our main results are only weakly affected by the limited statistics of the individual samples and histograms. This is because the final distributions are obtained by averaging the individual distributions and this procedure effectively reduces statistical fluctuations. In the upper panel of Fig. \[evolution\] the AGN emissivity calculated by [@hasinger05] is shown. The points with the error bars are redrawn here from their original paper. [@hasinger05] apply more stringent criteria to select sources and use only well defined samples of type-1 AGN. Optically these objects are identified by the broad Balmer emission lines, while using the X-ray criteria, they have unabsorbed spectra indicating low intrinsic column densities. In the present analysis I have included all objects in which the X-ray emission originates in the active nuclei. Thus, our results cannot be compared directly to those by [@hasinger05]. Nevertheless, despite entirely different method applied in the present paper, the distributions show good agreement at redshifts above $\sim\!1$. Although most of the apparent discrepancies, which reach a factor of $3$ at $z\approx 0.5$, are probably due to the distinct selection criteria of both investigations, one cannot exclude that some differences are caused by unrecognized systematic effects inherent in one or both methods. The present method of the luminosity density calculations has also some disadvantages. In our approach the absolute luminosities of the individual objects are not determined. Consequently, only the integral luminosity density is obtained, and the cosmic evolution of any selected AGN luminosity class has to be studied by means of the standard methods. ACKNOWLEDGEMENTS\ This work has been partially supported by the Polish MNiSW grant N N203 395934. Alexander, D. M., Bauer, F. E., Brandt, W. N., Schneider, D. P., Hornschemeier, A. E., et al. 2003, AJ, 126, 539 Barcons, X., Carrera, F. J., Ceballos, M. T., Page, M. J., Bussons-Gordo, J., et al. 2007, AA, 476, 1191 Barger, A. J., Cowie, L. L., Brandt, W. N., Capak, P., Garmire, G. P., et al. 2003, AJ, 124, 1839 Barger, A. J., Cowie, L. L., Capak, P., Alexander, D. M., Bauer, F. E., et al. 2003, AJ, 126, 632 Brandt, W. N. & Hasinger, G. 2005, ARA&A, 43, 827 De Grandi, S., Böhringer, H., Guzzo, L., Molendi, S., Chincarini, G., et al. 1999, ApJ, 514, 148 Donley, J. L., Rieke, G. H., Pérez-González, P. G., Rigby, J. R., & Alonso-Herrero, A. 2007, ApJ, 660, 167 Georgakakis, A., Rowan-Robinson, M., Babbedge, T. S. R., & Georgantopoulos, I. 2007, MNRAS, 377, 203 Giacconi, R., Zirm, A., Wang, J-X., Rosati, P., Nonino, M., et al. 2002, ApJS, 139, 369 Gilli, R., Cimatti, A., Daddi, E., Hasinger, G., Rosati, P., et al. 2003, ApJ, 592, 721 Gioia, I., Henry, J., Mullis, C., Böhringer, H., Briel, U., et al. 2003, ApJS, 149, 29 Hasinger, G., Miyaji, T, & Schmidt, M., 2005, AA, 441, 417 Henry, J., Mullis, C., Voges, W., Böhringer, H., Briel, U., et al. 2006, ApJS, 162, 304 Hogg, D. W. 1999, [*astro-ph/9905116*]{} Kim, M., Wilkes, B. J., Kim, D.-W., et al. 2007, ApJ, 659, 29 Lehmann, I., Hasinger, G., Schmidt, M., et al. 2001, A&A, 371, 833 Lehmer, B. D., Brandt, W. N., Hornschemeier, A. E., Alexander, D. M., Bauer, F. E., et al. 2006, AJ, 131, 2394 Mason, K. O., Carrera, F. J., Hasinger, G., Andernach, H, Aragon-Salamanca, A., et al. MNRAS, 311, 456 Miyaji, T., Hasinger, G., & Schmidt, M. 2000, ApJ, 353, 25 Moretti, A., Campana, S., Lazzati, D., & Tagliaferri, G. 2003, ApJ, 588, 696 Ravikumar, C. D., Puech, M., Flores, H., Proust, D., Hammer, H., et al. 2007, AA, 465, 1099 Reddy, N. A., Steidel, C. C., Erb, D. K., Shapley, A. E., & Pettini, M. 2006, ApJ, 653, 1004 Rudnick, L., Brown, S., & Williams L. R., 2007, ApJ, 671, 40 \[Erratum: ApJ, 678, 1531 (2008)\] Schmidt, M. 1968, ApJ, 151, 398 Schwope, A. D., Hasinger, G., Lehmann, I., Schwarz, R., Brunner, H., et al. 2000, AN, 321, 1 Silverman, J. D., Green, P. J., Barkhouse, W. A., et al. 2007, [*astro-ph/0710.2461*]{} Smith, K, M. & Huterer, D., 2008, [*arXiv:0805.2751*]{} Stocke, J. T., Morris, S. L., Gioia, I. M., Maccacaro, T., Schild, R., et al. 1991, ApJS, 76, 813 Szokoly, G. p., Bergeron, J., Hasinger, G., Lehmann, I., Kewley, L., et al. 2004, ApJS, 155, 271 Voges, W., Aschenbach, B., Boller, Th., Bräuninger, H., Briel, U., et al. 1999, AA, 349, 389 Zheng, W., Mikles, V. J., Mainieri, V., Hasinger, G., Rosati, P., et al. 2004, ApJS, 155, 73 [^1]: The samples extracted from the RBS span over narrow ranges of fluxes and there is no major difference between the mean and median values. In some other samples investigated in this paper the median flux is distinctly smaller than the average. In those cases, the median flux is adopted as the argument in the $n(z\,|\,S)$ functions. [^2]: For the cosmological model defined in Sec. \[intro\] and using the formulae given by [@hogg99].
{ "pile_set_name": "ArXiv" }
--- abstract: 'Coupled multiphysics problems often give rise to interface conditions naturally formulated in fractional Sobolev spaces. Here, both positive- and negative fractionality are common. When designing efficient solvers for discretizations of such problems it would then be useful to have a preconditioner for the fractional Laplacian. In this work, we develop an additive multigrid preconditioner for the fractional Laplacian with positive fractionality, and show a uniform bound on the condition number. For the case of negative fractionality, we re-use the preconditioner developed for the positive fractionality and left-right multiply a regular Laplacian with a preconditioner with positive fractionality to obtain the desired negative fractionality. Implementational issues are outlined in details as the differences between the discrete operators and their corresponding matrices must be addressed when realizing these algorithms in code. We finish with some numerical experiments verifying the theoretical findings.' address: '$^\dagger$Department of Mathematics, University of Oslo, Blindern, Oslo, 0316 Norway' author: - 'Trygve Bærland$^\dagger$' - 'Miroslav Kuchta$^\dagger$' - 'Kent-Andre Mardal$^\dagger$' bibliography: - 'fractional\_mg.bib' title: Multigrid Methods for Discrete Fractional Sobolev Spaces --- [^1] Introduction {#sec:intro} ============ Multiphysics or multiscale problems often involve coupling conditions at interfaces which are manifolds of lower dimensions. The coupling conditions are, because of the lower dimensionality, naturally posed in fractional Sobolev spaces, and this fact seemingly complicates discretization schemes and solution algorithms. Our focus here will be on the development of solution algorithms in terms of multilevel preconditioners that from an implementational point of view only require minor adjustments of standard multilevel algorithms. As simplified examples of problems involving interface conditions, let us consider the following two prototype problems. First an elliptic problem with a trace constraint [$$\label{frac:model1} {\begin{aligned} -\Delta u + T^*\lambda &=& f, \quad x \in \Omega, \\ T u &=& g, \quad x \in \Gamma , \end{aligned}}$$]{} and second an elliptic problem in mixed form with a trace constraint [$$\label{frac:model2} {\begin{aligned} u - \nabla p + T^*\lambda &=& f, \quad x \in \Omega, \\ \nabla \cdot u &=& g, \quad x \in \Omega, \\ T u &=& h, \quad x \in \Gamma . \end{aligned}}$$]{} Here, $\Gamma$ is a sub-manifold either within $\Omega$ or at its boundary, $T$ is a trace operator and $T^*$ its adjoint. Both problems are assumed to be equipped with suitable boundary conditions. We remark that although these problems are single physics problems, they may easily be coupled to other problems through the Lagrange multiplier at the interface. As such, the problems represent well the challenge of handling the interface properly in a multiphysics setting. We may write the above problems as $$\left( \begin{array}{cc} -\Delta & T^* \\ T & 0 \end{array} \right) \left( \begin{array}{c} u \\ \lambda \end{array} \right) = \left( \begin{array}{c} f \\ g \end{array} \right) \quad\mbox{ and }\quad \left( \begin{array}{ccc} I & -\nabla & T^* \\ \nabla\cdot & 0 & 0 \\ T & 0 & 0 \end{array} \right) \left( \begin{array}{c} u \\ p \\ \lambda \end{array} \right) = \left( \begin{array}{c} f \\ g \\ h \end{array} \right) .$$ A crucial challenge is to discretize and solve these problems in a scalable way such that the computations scales linearly with the number of unknowns. Our approach here is to consider iterative methods and develop preconditioners that are both spectrally equivalent with the involved operators and of order-optimal complexity. The main difficulty is the handling of the Lagrange multiplier which falls outside the scope of standard multilevel methods. To provide a general framework, we will consider preconditioners constructed in terms of the so-called operator preconditioning approach [@mardal2011preconditioning] to be used for iterative methods. As will be explained later, the block diagonal preconditioners constructed by this technique will be of the following form: $$\left( \begin{array}{cc} -\Delta^{-1} & 0 \\ 0 & (-\Delta)^{-1/2} \end{array} \right) \quad\mbox{ and }\quad \left( \begin{array}{ccc} (I - \nabla\nabla\cdot)^{-1} & 0 & 0 \\ 0 & I & 0 \\ 0 & 0 & (-\Delta)^{1/2} \end{array} \right),$$ respectively. Multilevel methods spectrally equivalent with both $(-\Delta)^{-1}$ and $(I - \nabla\nabla\cdot)^{-1}$ are well known. The challenging part in both cases is the construction of efficient preconditioning algorithms that approximate the inverse of the fractional Laplace problems on the form $$\label{frac:lap} (-\Delta)^s u = f, \quad x \in \Gamma$$ with $s=1/2$ and $s=-1/2$, equipped with suitable boundary conditions. Furthermore, if $\Gamma$ is of codimension 2, numerical simulations [@kuchta2016preconditioning] indicates that $s\in (-0.2, -0.1)$ gives rise to efficient preconditioners. In this paper we therefore consider methods for $s \in [-1,1]$. There are many examples of applications of fractional Laplacians in the litterature and we mention a few that motivates this work. For non-overlapping domain decomposition preconditioner are studied in [@arioli2012discrete], [@kocvara2016constraint]. Here, they use with $s=\tfrac{1}{2}$ to precondition the interface problem involving the related Steklov-Poincar[' e]{} operator. In [@kuchta2016efficient] the authors use with $s=-\tfrac{1}{2}$ as part of a block diagonal preconditioner for multiphysics problem where the constraint coupling two domains of different topological dimension is enforced by the Lagrange multiplier. Therein the fractionality $s$ is dictated by the mapping properties of the Schur complement operator. Some further examples of coupled systems with domains of different dimensionality include Babu[š]{}ka’s problem for enforcing Dirichlet boundary conditions on an elliptic operator [@babuvska1973finite], flow stabilization by removal of tangential velocity at the boundary through Lagrange multipliers [@BERTOLUZZA201758], the no-slip condition on the surface of a falling solid in the Navier-Stokes fluid [@court2015fictitious], inextensibility constraint in the complex model of vesicle formation [@aland2014diffuse], and the potential jump on a membrane of a cardiac cell [@tveito2017cell]. We note that in these applications the fractional Laplace problem has to be solved with both positive and negative exponent. There are several alternative approaches that have been used in order to approximate fractional Laplacians. Polynomial approximations of discretizations of $A^s$, where $A$ is a discrete Laplacian, can be computed with the standard Krylov subspace methods. However, without any preconditioner a Krylov subspace of large dimension is required for convergence, see e.g. Lanczos method in [@knizhnerman2010new Section 4]. We note that if is used to form a preconditioner the approximation of $A^s$ can be less accurate, and in [@arioli2012discrete] (generalized) Lanczos method is used to construct an efficient preconditioner. Therein the efficiency is due to a small size of the subspace. In [@yang2011novel] Lanczos method with a preconditioner based on the invariant subspace of the $k$ smallest eigenvalues is proposed for solving the fractional heat equation. The method is shown to generate non-polynomial approximations of $A^s$. The contour integral method of [@hale2008computing] and the extended Krylov method of [@knizhnerman2010new] are then related to rational function approximations of $A^s$, while [@harizanov2016optimal] consider the best uniform rational aproximations of the trasformed function $A \mapsto A^{\beta-s}$. In general, the approximation properties of these methods depend on the condition number of $A$ and thus computations of extremal eigenvalues are often part of the algorithm. Further, computational complexity of the methods depends on efficient solvers for auxiliary linear systems, e.g. $(A-q_kI)x=b$ in [@harizanov2016optimal] where $q_k\in{\mathbb{R}}$ is a shift parameter. Almost mesh independent preconditioners for systems arising in [@hale2008computing] and [@knizhnerman2010new] are discussed in [@burrage2012efficient]. An alternative approach to the matrix transfer method is presented in [@bonito2015numerical] where the inverse of the fractional Laplacian is defined via the (integral) Balakrishnan formula [@balakrishnan1960fractional]. Multilevel methods have been considered for fractional Laplacians have been considered in [@bramble2000computational; @oswald1998multilevel; @harizanov2016optimal], but there seems to be a significant untapped potential for advancement. Our work here is closely related to [@bramble2000computational], where order-optimal preconditioners for $A^s$ when $s\in\left(-\frac{3}{2}, \frac{3}{2} \right)$ were constructed using an hierarchical basis approach. The paper did, however, only consider smoothers based on level-dependent scaling and did not put much focus on the actual implementation. Here, we will develop and analyze a multilevel algorithm that is straightforward to implement in a standard multilevel software framework. In fact, the main change required is an adjustment of the smoothers. To illustrate the change, let us assume that we want to solve the system ${\mathsf{A}}{\mathsf{x}} = {\mathsf{b}}$, where ${\mathsf{A}}$ is a stiffness matrix corresponding to a discretized Laplacian. A standard Jacobi algorithm can then be written $${\mathsf{x}}^{n+1}_i = {\mathsf{x}}^{n}_{i} - \frac{1}{{\mathsf{A}}_{i,i}}{\mathsf{r}}_i^n,$$ where ${\mathsf{A}}_{i,i}$ are the diagonal entries of the stiffness matrix for a discretized Laplacian, and ${\mathsf{r}}^n$ is the residual of the $n$’th iterate, ${\mathsf{x}}^n$ . In our case, for ${\mathsf{A}}^s{\mathsf{x}} = {\mathsf{b}}$, the proposed Jacobi smoother may be implemented as $${\mathsf{x}}^{n+1}_i = {\mathsf{x}}^n_{i} - \left(\frac{1}{{\mathsf{M}}_{i,i}^{1-s}{\mathsf{A}}_{i,i}^s}\right) {\mathsf{r}}^n_i, .$$ Here, ${\mathsf{M}}_{i,i}$ are the diagonal entries of the mass matrix. We notice here that for $s=0$ the action is a Jacobi iteration on the mass matrix, for $s=1$ the action is a Jacobi iteration on the stiffness matrix, and for $0<s<1$ the action is an interpolation between these two extremes. From an implementational point of view, the restriction and interpolation operators used are the same as those used in standard multilevel algorithms. However, from a theoretical point of view, the fact that we use standard restriction and interpolation operators, means that the multilevel approach will be non-nested. In fact, the matrices on coarser levels do not correspond to $(-\Delta)^s$-Galerkin projections of the matrix on the finer levels. We therefore employ the framework of non-nested multilevel methods [@bramble1991analysis]. Furthermore, a multiplicative multilevel algorithm would require computing the residual and hence the evaluation of the exact $(-\Delta)^{-s}$ operator on every level. Since the evaluation of the exact $(-\Delta)^{-s}$ is a computationally expensive procedure, we instead rely on the additive multilevel algorithm proposed in [@bramble1990parallel], where the same residual is used on all levels. The additive variant is significantly less efficient than corresponding multiplicative variants in terms of the conditioning in the sense that the conditioning depends on the number of levels. Still, this is a small price to pay (only logarithmic in the number of unknowns) to avoid exact evaluation of the residual. In this paper we will assume quasi-uniform mesh and continuous piecewise linear finite elements. This is mainly for simplicity, and the results can be generalized to higher order discretizations, as well as discontinuous Galerkin methods. The paper is structured as follows. In Section \[sec:preliminaries\], we introduce notation, and some useful operator inequalities related to fractional powers of positive operators. We also give a brief discussion of fractional Sobolev spaces, as well as of some implementational concerns. Section \[sec:abstract-bpx\] is devoted to the analysis of an abstract multilevel framework. In section \[sec:discrete-bpx\] we use this framework to define operators that are spectrally equivalent to the inverse of the fractional Laplacian when the fractionality $s \geq 0$. We discuss some strategies for preconditioning when $s<0$ in section \[sec:preconditioning-negative-s\], and in section \[sec:implementation\] we discuss implementation of the preconditioners developed in the previous sections. Finally, we provide numerical results that verify our theoretical result in section \[sec:numerical-experiments\]. Notation and preliminaries {#sec:preliminaries} ========================== Let $\Omega$ be a bounded, Lipschitz domain in ${\mathbb{R}}^n$, with boundary ${\partial}\Omega$. We denote by $L^2(\Omega)$ the space of square integrable functions over $\Omega$, with inner product ${\left( \cdot, \cdot\right)}$ and norm ${\left\Vert\cdot\right\Vert}$. For $k \in \mathbb{N}$, we denote by $H^k(\Omega)$ the usual Sobolev spaces of functions in $L^2(\Omega)$ with all derivatives up to order $k$ in $L^2(\Omega)$. The norm and inner product in $H^k$ is denoted by $\|\cdot\|_k$ and $(\cdot, \cdot)_k$, respectively. The closure in $H^k$ of smooth functions with compact support in $\Omega$ is denoted as $H^k_0(\Omega)$ and its dual space is $H^{-k}$. In general a Hilbert space $X$ is equipped with a norm $\|\cdot\|_X$ and an inner product $(\cdot, \cdot)_X$ and the dual space is denoted ${X^{'}}$. Let $A$ be a symmetric positive-definite operator on a finite-dimensional Hilbert space $X$ with dimension $N$. Denote by ${\left\{ (\lambda_k,\phi_k) \right\} }_{k=1}^N$ the set of eigenpairs of $A$, normalized so that [$${\left( \phi_k, \phi_l\right)}_X = \delta_{k,l},$$]{} where $\delta_{k,l}$ is the Kronecker delta. Then $\phi_k$, for $k=1,\ldots,N$ forms an orthonormal basis of $X$, and if $u \in X$ has the representation $u = \sum_{k=1}^N c_k \phi_k$, then [$$Au = \sum_{k=1}^N \lambda_k c_k \phi_k.$$]{} For $s\in {\mathbb{R}}$, we define the fractional power $A^s$ of $A$ by [$$A^s u = \sum_{k=1}^N \lambda_k^s c_k \phi_k.$$]{} If $A$ is only positive semi-definite, then we must restrict to $s \geq 0$, and the eigenvectors corresponding to the nullspace of $A$ are left out (also for $s=0$). If $B$ is another symmetric positive semi-definite operator on $X$, we write $A \leq B$ if for every $u \in X$ [$${\left( Au, u\right)}_X \leq {\left( Bu, u\right)}_X$$]{} holds. Note that $A \geq 0$ is equivalent to saying that $A$ is positive semi-definite. A result in operator theory is the Löwner-Heinz inequality, which in our case states that if $A \leq B$, then [$$\label{eq:Loewner-Heinz} A^s \leq B^s, \quad s\in [0,1],$$]{} cf. for instance [@kato1952notes]. Inequality means that the function $x^s$ with $x\in [0,\infty)$ is operator monotone for $s\in[0,1]$. It follows that $-(x)^s$ is operator convex (cf. [@hansen1982monotone]), that is, for any two symmetric positive semi-definite operators $A$ and $B$ on a Hilbert space $X$, the inequality [$$\label{eq:operator-convex-def} \lambda A^s + (1-\lambda)B^s \leq \left(\lambda A + (1-\lambda)B \right)^s$$]{} holds for every $\lambda \in [0,1]$. A key result regarding operator convex functions is the Jensen’s operator inequality (cf. [@hansen2003jensen Theorem 2.1]). The version we will use in the current work states that for any $K\in \mathbb{N}$ and $s\in [0,1]$ [$$\label{eq:jensen-inequality} \sum_{k=1}^K {P_k^*}A_k^s P_k \leq \left(\sum_{k=1}^K {P_k^*}A_k P_k\right)^s,$$]{} where for $k=1,\ldots,K$, $A_k$ are symmetric positive semi-definite operators on $X$, and $P_k$ are linear operators on $X$ so that $\sum_{k=1}^K {P^*}_k P_k \leq 1$. Fractional Sobolev spaces {#sec:interpolation_theory} ------------------------- We consider the interpolation spaces between $H^1(\Omega)$ and $L^2(\Omega)$ as defined in [@lions1972nonhom]. Let the inner product on $H^1(\Omega)$ be realized by the operator $A := I - \Delta $, as [$${\left( u, v\right)}_1 = {\left( Au, v\right)} = {\left( u, v\right)} + {\left( {\nabla}u, {\nabla}v\right)}, \quad u,v \in H^1(\Omega).$$]{} $A$ is unbounded as an operator mapping $L^2(\Omega)$ to $L^2(\Omega)$. However, $A$ is well-defined on the set [$$D(A) = {\left\{ u \in L^2(\Omega) {\,:\,}Au \in L^2(\Omega) \right\} },$$]{} which is a dense subspace of $L^2(\Omega)$. On $D(A)$, $A$ is symmetric and positive-definite, and so the fractional powers of $A$, $A^\theta$ for $\theta\in {\mathbb{R}}$, are well-defined. Note that in the particular case $\theta = \frac{1}{2}$, [$${\left\VertA^{\frac{1}{2}}u\right\Vert}^2 = {\left( Au, u\right)} = {\left\Vertu\right\Vert}_1.$$]{} For $s\in[0,1]$, we define the fractional Sobolev spaces as [$$\label{eq:Frac-Sobolev-space-def} H^s(\Omega) = {\left\{ u \in L^2(\Omega) {\,:\,}A^{\frac{s}{2}}u \in L^2(\Omega) \right\} },$$]{} which is a Hilbert space with inner product given by [$${\left( u, v\right)}_s = {\left( A^s u, v\right)}, \quad u,v\in H^s(\Omega),$$]{} and we denote the corresponding norm by ${\left\Vert\cdot\right\Vert}_s$. We define $H_0^s(\Omega)$ to be the closure of $C_0^\infty(\Omega)$, the space of infinitely smooth functions with compact support in $\Omega$, in the norm of $H^s(\Omega)$. We note that if $s \leq \frac{1}{2}$, the spaces $H_0^s(\Omega)$ and $H^s(\Omega)$ coincide (cf. [@lions1972nonhom Theorem 11.1]). For $s \in [-1,0]$, we define a family of fractional Sobolev spaces using the dual of $H_0^s(\Omega)$. That is, [$$H^s(\Omega) = {\left(H^{-s}_0(\Omega)\right)^{'}}.$$]{} Replacing $H^1(\Omega)$ with $H_0^1(\Omega)$ and setting $A = -\Delta$ in the above construction, will again yield the space $H^s_0(\Omega)$, with equivalent norm, for all $s$ except when $s = \frac{1}{2}$. In this case, interpolation between $H_0^1(\Omega)$ and $L^2(\Omega)$ results in a space that is strictly contained in $H_0^{\frac{1}{2}}(\Omega)$. The subsequent analysis is valid for both $H^s_0(\Omega)$ and $H^s(\Omega)$. We remark that the above defined fractional space $H^s(\Omega)$ is equivalent to the fractional space $\hat{H}^s(\Omega)$ defined in terms of the norm $${\left\Vertu\right\Vert}^2_{\hat{H}^{s}(\Omega)} = {\left\Vertu\right\Vert}^2 + \int_{\Omega \times \Omega} \frac{|u(x) - u(y)|}{|x-y|^{n+s}}\, {\mathrm{d}}x {\mathrm{d}}y.$$ A detailed overview of the various definitions of fractional Sobolev norms and their discretizations can be found in [@lischke2018fractional]. Discrete fractional Sobolev spaces ---------------------------------- We will now consider a discretization of the fractional Sobolev spaces $H_0^s(\Omega)$ and $H^{-s}(\Omega)$ for $s\in[0,1]$. Let $X_h$ be a finite-dimensional subspace of $H_0^1(\Omega)$, with $\dim X_h = N_h$. We define the operator $A_h: X_h \to X_h$ by [$$\label{eq:Ah-def} {\left( A_h u, v\right)} = {\left( {\nabla}u, {\nabla}v\right)}, \quad u,v \in X_h.$$]{} Using the fractional powers of $A_h$, we define for $s\in {\mathbb{R}}$ the discrete fractional inner product on $X_h$ by [$${\left( u, v\right)}_{s,h} = {\left( A^s_h u, v\right)}, \quad u,v \in X_h,$$]{} and denoted the corresponding norm by ${\left\Vert\cdot\right\Vert}_{s,h}$. It is clear that for $s=0$ and $s=1$, the two norms ${\left\Vert\cdot\right\Vert}_{s,h}$ and ${\left\Vert\cdot\right\Vert}_s$ coincide on $X_h$. Therefore, due to [@arioli2009discrete Lemma 2.3], the norms ${\left\Vert\cdot\right\Vert_{s,h}}$ and ${\left\Vert\cdot\right\Vert_{s}}$, when $s\in [0,1]$, are equivalent on $X_h$, with constants of equivalence independent of $N_h$. Let $X_H$ be a subspace of $X_h$, and $A_H: X_H \to X_H$ be defined analogously to $A_h$ in . If $I_H: X_H \to X_h$ is the inclusion map, we see that [$$\label{eq:MhAh-inheritance} A_H = {I^*}_HA_hI_H,$$]{} where ${I^*}_H$ is the adjoint of $I_H$ with respect to the $L^2$ inner product. We may also define $A^s_H: X_H \to X_H$, but generally, $A^s_H \not= {I^*}_H A^s_h I_H$. However, by Jensen’s operator inequality we have the following. \[lem:subspace-As-estimate\] For every $s \in [0,1]$ we have [$$\label{eq:subspace-As-estimate-operator} {I_H^*}A^s_h I_H \leq A^s_H.$$]{} That is, for every $u \in X_H$, [$$\label{eq:subspace-As-estimate} {\left( A^s_h u, u\right)} \leq {\left( A^s_H u, u\right)}.$$]{} For $s=0$ and $s=1$, holds with equality, so let $0<s<1$. We start by noticing that since ${I_H^*}I_H$ is the identity on $X_H$, [$$A^2_H = ({I_H^*}I_HA_H {I_H^*}I_H)^2 = {I_H^*}(I_H A_H {I_H^*})^2 I_H.$$]{} By induction, we find that [$$A^k_H = {I_H^*}(I_H A_H {I_H^*})^k I_H$$]{} for every nonnegative integer $k$. It follows that for any polynomial $p$ [$$\label{eq:subspace-As-estimate-proof1} p(A_H) = {I_H^*}\,p\left( I_H A_H {I_H^*} \right) I_H.$$]{} Take now $\epsilon > 0$. The spectra of both $A_H$ and $I_H A_H {I_H^*}$ are contained in some bounded, nonnegative interval $[0,b]$. By Weierstrass’ approximation Theorem we can thus choose a polynomial $p$ so that [$${\left\Vert\left(I_H A_H {I_H^*} \right)^s - p(I_H A_H {I_H^*} )\right\Vert} < \epsilon, \quad \text{ and } \quad {\left\VertA_H^s - p(A_H)\right\Vert} < \epsilon.$$]{} Using the triangle inequality and we have that [$${\left\VertA_H^s -{I_H^*}\left( I_H A_H {I_H^*} \right)^s I_H\right\Vert} \leq {\left\VertA_H^s - p(A_H)\right\Vert} + {\left\Vert{I_H^*}\left( \left(I_H A_H {I_H^*} \right)^s - p(I_H A_H {I_H^*} )\right) I_H\right\Vert} < 2\epsilon,$$]{} and since $\epsilon$ was arbitrary, this shows that [$$\label{eq:subspace-As-estimate-proof2} A^s_H = {I_H^*}(I_H A_H {I_H^*})^s I_H.$$]{} Using in , we get that [$$\label{eq:subspace-As-estimate-proof3} A^s_H = {I_H^*}(I_H {I_H^*}A_h I_H {I_H^*})^s I_H.$$]{} Finally, $I_H {I_H^*}$ defines a symmetric operator on $X_h$ with $L^2$ operator norm equal to $1$. Since the function $x \mapsto -x^s$ is operator convex on $[0,\infty)$ we can use Jensen’s operator inequality in to get [$$\begin{aligned} A^s_H &\geq {I_H^*}I_H {I_H^*} A^s_h I_H {I_H^*} I_H \\ & = {I_H^*} A^s_h I_H, \end{aligned}$$]{} where we have used that ${I_H^*}I_H$ is the identity on $X_H$. Abstract multilevel theory {#sec:abstract-bpx} ========================== In order to analyze and implement a multigrid preconditioner for the fractional Laplacian there are three main issues that need to be dealt with. First, we need to derive and implement a smoother with the desired properties. As already mentioned in the introduction, this step only requires a minor modification to standard smoothing algorithms. We will discuss the details concerning implementation later. Second, the restriction/interpolation operators do not result in a nested hierarchy of operators in our fractional setting as $A^s_H \not= I^*_H A_h I_H$. For this reason we will employ the framework for non-nested multilevel algorithms developed in  [@bramble1991analysis]. Third, our main motivation for developing fractional multilevel solvers is their application to multiphysics and multiscale problems where the preconditioner for the fractional Laplacian is utilized at the interfaces. As such, the fractional Laplacian operator is not part of the original problem and we may therefore not assume that this operator has been implemented. Furthermore, implementing this operator in an efficient manner is a challenge, as the exact fractional operator involves the solution of a large, global and expensive generalized eigenvalue problem. To avoid the application of the fractional Laplacian on the various levels we employ additive multilevel schemes which enable the residual of the problem to be used at all levels and remove the need for implementing a global fractional Laplacian operator. That said, the theory developed here extends to multiplicative algorithms for problems involving the fractional Laplacian, such as the standard V-cycle. In this section we will address the second and the third issues and outline a theory for an additive multilevel scheme applied to an abstract non-nested problem. As such analysis of this section is the synthesis of the papers [@bramble1990parallel] and [@bramble1991analysis]. Assume that we are given a nested sequence of finite-dimensional function spaces [$$V_1 \subset V_2 \subset \cdots \subset V_J = V, \quad J \geq 2.$$]{} We further assume that $V$, and consequently all subspaces of $V$, is endowed with an inner product ${\left( \cdot, \cdot\right)}$, with corresponding induced norm ${\left\Vert\cdot\right\Vert}$. Moreover, for each $k=1,\ldots, J$, we assume we are given a symmetric positive definite operator $A_k: V_k \to V_k$, and we set $A = A_J$. Note that we do not assume that the $A_k$ operators are nested. For the development and analysis of a multilevel algorithm, it will be useful to define a number of operators on each level $k$. First, we define $P_{k,k-1}: V_k \to V_{k-1}$ by [$${\left( A_{k-1}P_{k,k-1}v, w\right)} = {\left( A_k v, w\right)}, \quad \forall v \in V_k, w \in V_{k-1}.$$]{} We remark that in a nested setting, $P_{k,k-1}$ is the $A$-projection, while since the $A_k$ operators are not nested, the $P_{k,k-1}$ operators are not projections. Next, we define $Q_k: V \to V_k$ by [$${\left( Q_k v, w\right)} = {\left( v, w\right)}, \quad \forall v \in V, w \in V_k.$$]{} It follows by the above definitions that [$$\label{eq:AP_QA} A_{k-1}P_{k,k-1} = Q_{k-1}A_k,$$]{} and $Q_l Q_k = Q_k Q_l = Q_l$ whenever $l \leq k$. For the sake of brevity, it will also be useful to define $P_k: V \to V_k$ by $P_k = P_{k+1,k} P_{k+2,k+1} \cdots P_{J,J-1}$. Using the definition of $P_{j+1,j}$ for $j=k,\ldots,J-1$ we see that [$${\left( A_kP_kv, w\right)} = {\left( Av, w\right)}, \quad \forall v \in V, w\in V_k.$$]{} Furthermore, applying repeatedly, we find that [$$\label{eq:AP_QA-all-levels} A_k P_k = Q_k A.$$]{} Finally, suppose we are given for each $k$ a smoother, which is a symmetric positive definite operator $R_k : V_k \to V_k$, which in some sense should approximate ${A_k^{-1}}$ on $V_k \backslash V_{k-1}$. We can now define a additive multilevel operator $B: V \to V$ by [$$\label{eq:BPX-def} B = \sum_{k=1}^J R_k Q_k.$$]{} As remarked in [@bramble1990parallel], $B$ can viewed as a additive version of the standard multiplicative V-cycle multigrid algorithm, where $R_k$ plays the role of smoother. Because of this, it is reasonable that the assumptions we need to make to establish spectral equivalence between ${A^{-1}}$ and $B$ are similar to those made for standard multigrid algorithms. We assume that for $k=2,\ldots,J$ [$$\tag{A.1} \label{ass:noninheritance} {\left( A_k v, v\right)} \leq {\left( A_{k-1}v, v\right)}, \quad \forall v \in V_{k-1}.$$]{} Under assumption and the definition of $P_{k,k-1}$ we see that for any $v \in V_k$ [$$\begin{aligned} {\left( A_{k-1}P_{k,k-1}v, P_{k,k-1}v\right)} &= {\left( A_k v, P_{k,k-1}v\right)} \\ &\leq {\left( A_k P_{k,k-1}v, P_{k,k-1}v\right)}^{\frac{1}{2}}{\left( A_k v, v\right)}^{\frac{1}{2}} \\ & \leq {\left( A_{k-1} P_{k,k-1}v, P_{k,k-1}v\right)}^{\frac{1}{2}}{\left( A_k v, v\right)}^{\frac{1}{2}}. \end{aligned}$$]{} Thus, we see that implies [$$\label{eq:noninheritance-adjoint} {\left( A_{k-1}P_{k,k-1}v, P_{k,k-1}v\right)} \leq {\left( A_k v, v\right)}, \quad \forall v \in V_k.$$]{} Conversely, assume . Then, for any $v \in V_{k-1}$ and the definition of $P_{k,k-1}$ [$$\begin{aligned} {\left( A_k v, v\right)} &= {\left( A_{k-1}P_{k,k-1}v, v\right)} \\ &\leq {\left( A_{k-1}P_{k,k-1}v, P_{k,k-1}v\right)}^{\frac{1}{2}} {\left( A_{k-1}v, v\right)}^{\frac{1}{2}} \\ &\leq {\left( A_{k}v, v\right)}^{\frac{1}{2}} {\left( A_{k-1}v, v\right)}^{\frac{1}{2}}, \end{aligned}$$]{} which implies . Thus, and are equivalent. Notice that a similar inequality to would also hold for $P_k$, namely [$$\label{eq:noninheritance-adjoint-all-levels} {\left( A_k P_k v, P_k v\right)} \leq {\left( A v, v\right)}, \quad \forall v \in V.$$]{} For the operators $R_k$, we assume there are constants $C_1, C_2 > 0$, independent of $k$ so that [$$\tag{A.2} \label{ass:smoother-stability} C_1 \frac{{\left\Vertv\right\Vert}^2}{\lambda_k} \leq {\left( R_k v, v\right)} \leq C_2 {\left( {A_k^{-1}}v, v\right)}, \quad \forall v \in V_k,$$]{} where $\lambda_k$ is the largest eigenvalue of $A_k$. Lastly, as is common in multigrid theory, we will use an approximation assumption to establish spectral equivalence between $B$ and ${A^{-1}}$. In this work, we assume the following approximation property: That there is an $\alpha \in (0,1]$ and constant $C_3 > 0$, independent of $k$, so that [$$\tag{A.3} \label{ass:approximation-property} {\left( A_k(I-P_{k,k-1})v, v\right)} \leq C_3^\alpha \left( \frac{{\left\VertA_k v\right\Vert}^2}{\lambda_k} \right)^\alpha {\left( A_k v, v\right)}^{1-\alpha}, \quad \forall v \in V_k.$$]{} We are now in a position to state and prove the main theorem of this section: \[thm:bpx-spectral-equivalence\] Assume that , , and hold. Then, with $B$ given in , [$$\label{eq:bpx-spectral-equivalence} C_1 {C_3^{-1}}J^{1-\frac{1}{\alpha}}{\left( A v, v\right)} \leq {\left( B Av, Av\right)} \leq C_2J{\left( A v, v\right)}.$$]{} holds for every $v \in V$. Fix $v \in V$. Using the definition of $B$ together with we find that [$${\left( BA v, Av\right)} = \sum_{k=1}^J {\left( R_k Q_k A v, Q_k A v\right)} = \sum_{k=1}^J {\left( R_k A_k P_k v, A_k P_k\right)}.$$]{} Thus, applying the second inequality of and gives [$${\left( B Av, Av\right)} \leq C_2\sum_{k=1}^J {\left( A_k P_kv, P_k v\right)} \leq C_2 J {\left( A v, v\right)},$$]{} which proves the second inequality of . For the first inequality of we write [$$v = \sum_{k=1}^J (P_k - P_{k-1})v,$$]{} where we interpret $P_0 = 0$ and $P_J =I$. By the definition of $P_k$, we have that $P_{k-1} = P_{k,k-1}P_k$, and so [$$v = \sum_{k=1}^J (I - P_{k,k-1})P_k v.$$]{} It follows that [$${\left( A v, v\right)} = \sum_{k=1}^J {\left( A_k(I-P_{k,k-1})P_k v, P_k v\right)}.$$]{} Using and , gives [$$\begin{aligned} {\left( Av, v\right)} &\leq C_3^\alpha \sum_{k=1}^J\left({\lambda_k^{-1}}{\left\VertA_kP_kv\right\Vert}^2\right)^\alpha {\left( A_k P_k v, P_k v\right)}^{1-\alpha} \\ &\leq C_3^\alpha {\left( A v, v\right)}^{1-\alpha}\sum_{k=1}^J\left({\lambda_k^{-1}}{\left\VertA_kP_kv\right\Vert}^2\right)^\alpha. \end{aligned}$$]{} The first inequality of then implies that [$$\begin{aligned} {\left( A v, v\right)} &\leq ({C_1^{-1}}C_3)^\alpha {\left( A v, v\right)}^{1-\alpha}\sum_{k=1}^J{\left( R_kA_k P_k v, A_k P_k v\right)}^\alpha \\ &\leq ({C_1^{-1}}C_3)^\alpha J^{1-\alpha} {\left( A v, v\right)}^{1-\alpha}\left(\sum_{k=1}^J{\left( R_kA_k P_k v, A_k P_k v\right)}\right)^\alpha \\ &\leq ({C_1^{-1}}C_3)^\alpha J^{1-\alpha} {\left( A v, v\right)}^{1-\alpha}{\left( B A v, A v\right)}^\alpha, \end{aligned}$$]{} where the second step follows by Hölder’s inequality. The last step follows by the definition of and . Dividing by $({C_1^{-1}}C_3)^\alpha {\left( A v, v\right)}^{1-\alpha}J^{1-\alpha}$ on both sides and raising to the power $\frac{1}{\alpha}$ gives the first inequality of . \[rem:bpx-weak-spectral-equivalence\] Analagously to what was done in [@bramble1990parallel], we can replace assumption with an assumption on the projections $Q_k$. In particular, if instead of , we assume that there is a constant $C_4 > 0$, independent of $k$ so that [$${\left\Vert(I-Q_{k-1})v\right\Vert}^2 \leq C_4 {\lambda_k^{-1}} {\left( A v, v\right)}, \quad \forall v \in V,$$]{} then we can use an argument like what was made in [@bramble1990parallel Theorem 1 and Corollary 1] to show that [$$\label{eq:bpx-weak-spectral-equivalence} {C_4^{-1}}C_1{J^{-1}}{\left( A v, v\right)} \leq {\left( B Av, Av\right)} \leq C_2 J {\left( A v, v\right)}$$]{} for every $v \in V$. Preconditioner for discrete fractional Laplacian {#sec:discrete-bpx} ================================================ In this section we use the abstract theory developed in Section \[sec:abstract-bpx\] to derive an order optimal preconditioner for the discrete fractional Laplacian $A^s_h$, described in Section \[sec:preliminaries\], when $s \in [0,1]$. Let $\Omega$ be a bounded, polygonal domain in ${\mathbb{R}}^n$ and suppose we are given a quasi-uniform triangulation of $\Omega$, denoted by ${\mathcal{T}}_h$, where $h$ denotes the characteristic mesh size. We restrict our discussion to the case when $V_h$ is the space of continuous, piecewise linear functions relative to the triangulation ${\mathcal{T}}_h$ which vanish on ${\partial}\Omega$. To define a nested sequence of subspaces, we suppose that ${\mathcal{T}}_h$ is constructed by successive refinements. That is, we are given a sequence, [$${\mathcal{T}}_1 \subset \cdots {\mathcal{T}}_J ={\mathcal{T}}_h,$$]{} of quasi-uniform triangulations, and ${\mathcal{T}}_k$ has characteristic mesh size $h_k$ for $k=1,\ldots,J$. In the following, we will assume the bounded refinement hypothesis, that is, $h_{k-1} \leq \gamma h_k$ for $k=2,\ldots,J$, where $\gamma \geq 1$ is a constant. In practice, $\gamma$ is around $2$. For each $k$ we define $V_k$ as the space of continuous, piecewise linear functions relative to ${\mathcal{T}}_k$ that vanish on ${\partial}\Omega$. Further, we define $A_k: V_k \to V_k$ by [$${\left( A_k v, w\right)} = {\left( {\nabla}v, {\nabla}w\right)}, \quad v,w\in V_k.$$]{} We now fix $s \in [0,1]$. Since $A_k$ is symmetric positive definite, we can define SPD operators $A^s_k$ and corresponding norms [$${\left\Vertv\right\Vert}_{s,k}^2 := {\left( A^s_k v, v\right)}, \quad v\in V_k.$$]{} Note that if $s=0$ or $s=1$, the norm ${\left\Vert\cdot\right\Vert}_{s,k}$ coincides with the $L^2$- and $H^1_0$-norm, respectively. That is, ${\left\Vert\cdot\right\Vert}_{0,k} = {\left\Vert\cdot\right\Vert}$, and ${\left\Vert\cdot\right\Vert}_{1,k} = {\left\Vert\cdot\right\Vert}_{1}$. Analogous to the discussion in Section \[sec:abstract-bpx\] we also define operators $P^s_{k,k-1}: V_k \to V_{k-1}$ by [$$\label{eq:Ps_k-definition} {\left( A^s_{k-1}P^s_{k,k-1}v, w\right)} = {\left( A^s_k v, w\right)}, \quad \forall v \in V_k, w \in V_{k-1}.$$]{} $Q_{k}: V_J \to V_k$ as the $L^2$-projection, and $P^s_k := P^s_{k+1,k} P^s_{k+2,k+1}\cdots P^s_{J,J-1}$. To complete the description of a multilevel preconditioner, we still need to define smoothers, $R^s_k$, for each $k$ and $s$. In this work, we will define additive smoothers based on domain decomposition. To that end, let ${\mathcal{N}}_k$ be the set of vertices in the triangulation ${\mathcal{T}}_k$, and for each $\nu \in {\mathcal{N}}_k$, let ${\mathcal{T}}_{k,\nu}$ be the set of triangles meeting at the vertex $\nu$. Then ${\mathcal{T}}_{k,\nu}$ forms a triangulation of a small subdomain $\Omega_{k,\nu}$. We define $V_{k,\nu}$ to be the subspace of functions in $V_k$ with support contained in $\bar{\Omega}_{k,\nu}$. Analogously to what we did for $V_k$, we may define for each $\nu \in {\mathcal{N}}_k$ operators $A^s_{k,\nu}: V_{k,\nu} \to V_{k,\nu}$, and $Q_{k,\nu}: V_k \to V_{k,\nu}$. For $k=2,\ldots,J$, we define [$$\label{eq:Rk-definition} R^s_k := \sum_{\nu \in {\mathcal{N}}_k} A^{-s}_{k,\nu}Q_{k,\nu},$$]{} while on the coarsest level we set $R^s_1 = A^{-s}_1$. We note that the smoothers are symmetric positive-definite, and their inverse satisfy [$$\label{eq:smoother-inverse-norm} {\left( {(R^s_k)^{-1}}v, v\right)} = \inf_{\underset{v_\nu \in V_{k,\nu}}{v = \sum_{\nu} v_\nu}} \sum_{\nu \in {\mathcal{N}}_k} {\left( A^s_{k,\nu}v_\nu, v_\nu\right)}, \quad v \in V_k.$$]{} Our preconditioner now reads [$$\label{eq:BPXs-preconditioner-definition} B_h^s := \sum_{k=1}^J R^s_k Q_k.$$]{} We want to apply Theorem \[thm:bpx-spectral-equivalence\] to the preconditioner defined by and , so we need to verify assumptions -. Using Lemma \[lem:subspace-As-estimate\], we immediately find that for every $k$, [$${\left( A^s_k v, v\right)} \leq {\left( A^s_{k-1} v, v\right)}, \quad \forall v \in V_{k-1},$$]{} which verifies in the current context. We present the verification of in the following Lemma. \[lem:smoother-stability-verification\] For $k=1,\ldots,J$, let $R^s_k: V_k \to V_k$ be defined by . Assume that there exists a constant $K_0$, so that for every $k$ and $v \in V_k$ there exists a decomposition $\sum_{\nu \in {\mathcal{N}}_k}v_\nu = v$, with $v_\nu \in V_{k,\nu}$ so that [$$\label{eq:L2-stable-decomposition} \sum_{\nu \in {\mathcal{N}}_k}{\left\Vertv_\nu\right\Vert}^2 \leq K_0 {\left\Vertv\right\Vert}^2.$$]{} Then there are constants $C_1, C_2 > 0$, so that for every $k$, [$$\label{eq:smoother-stability-BPXs} C_1 \frac{{\left\Vertv\right\Vert}^2}{\lambda^s_k} \leq {\left( R^s_k v, v\right)} \leq C_2 {\left( A^{-s}_k v, v\right)}, \quad \forall v \in V_k,$$]{} where $\lambda_k^s$ is the largest eigenvalue of $A^s_k$. It is evident that holds at the coarsest level, i.e., for $k=1$ is satisfied with $C_1 = C_2 = 1$. So let $k \geq 2$, and fix $v \in V_k$. For $\nu \in {\mathcal{N}}_k$, let $\lambda_{k,\nu}^s$ denote the largest eigenvalue of $A^s_{k,\nu}$. To prove the first inequality in , we begin by noting that for every $\nu \in {\mathcal{N}}_k$ [$$\lambda_k^1 = \sup_{w \in V_k}\frac{{\left( A^1_k w, w\right)}}{{\left( w, w\right)}} \geq \sup_{w \in V_{k,\nu}}\frac{{\left( A^1_{k,\nu} w, w\right)}}{{\left( w, w\right)}} = \lambda_{k,\nu}^1.$$]{} Thus, since $\lambda_k^s = (\lambda_k^1)^s$, we have that [$$\label{eq:smoother-stability-proof1} \lambda_k^s \geq \lambda_{k,\nu}^s.$$]{} Now, using and assumption , together with the definition of $Q_{k,\nu}$ and $R^s_k$ yields [$$\begin{aligned} \frac{{\left( v, v\right)}}{\lambda_k^s} &= \frac{1}{\lambda_k^s}\sum_{\nu \in {\mathcal{N}}_k}{\left( v, v_\nu\right)} \\ &= \frac{1}{\lambda_k^s}\sum_{\nu \in {\mathcal{N}}_k}{\left( Q_{k,\nu}v, v_\nu\right)} \\ &\leq \left(\frac{1}{\lambda_k^s}\sum_{\nu \in {\mathcal{N}}_k}{\left( Q_{k,\nu}v, Q_{k,\nu}v\right)}\right)^{\frac{1}{2}} \left(\frac{1}{\lambda_k^s}\sum_{\nu \in {\mathcal{N}}_k}{\left\Vertv_\nu\right\Vert}^2\right)^{\frac{1}{2}} \\ &\leq \left(\sum_{\nu \in {\mathcal{N}}_k}\frac{1}{\lambda_{k,\nu}^s}{\left( Q_{k,\nu}v, Q_{k,\nu}v\right)}\right)^{\frac{1}{2}}\left(\frac{K_0}{\lambda_k^s}{\left\Vertv\right\Vert}^2\right)^{\frac{1}{2}} \\ &\leq \left(\sum_{\nu \in {\mathcal{N}}_k}{\left( A^{-s}_{k,\nu}Q_{k,\nu}v, Q_{k,\nu}v\right)}\right)^{\frac{1}{2}}\left(\frac{K_0}{\lambda_k^s}{\left\Vertv\right\Vert}^2\right)^{\frac{1}{2}} \\ &\leq {\left( R^s_k v, v\right)}^{\frac{1}{2}} \left(\frac{K_0}{\lambda_k^s}{\left\Vertv\right\Vert}^2\right)^{\frac{1}{2}}, \end{aligned}$$]{} which proves the first inequality of with $C_1 = {K_0^{-1}}$. For the second inequality, we begin by noting that for $s=1$, it was proven in [@xu1992iterative Lemma 7.2] that there is a constant $C$, independent of $k$ so that [$${\left( R^1_k v, v\right)} \leq C{\left( {A_k^{-1}}v, v\right)}, \quad \forall v \in V_k.$$]{} Since $s\in [0,1]$, it follows by the Löwner-Heinz inequality that [$$\label{eq:smoother-stability-proof3} {\left( (R^1_k)^s v, v\right)} \leq C^s{\left( A^{-s}_k v, v\right)}, \quad \forall v \in V_k.$$]{} Thus, if we can show that [$$\label{eq:smoother-stability-proof4} {\left( R^s_k v, v\right)} \leq C {\left( (R^1_k)^s v, v\right)},$$]{} for some constant $C$, which is independent of $k$, then we can use together with to prove the second inequality of . We aim to prove using Jensen’s operator inequality. To that end, we need to scale $R^s_k$, so that is applicable. Using the assumed stable decomposition of $v$, we note that [$$\begin{aligned} \sum_{\nu \in {\mathcal{N}}_k}{\left( Q_{k,\nu}v, Q_{k,\nu} v\right)} &= \sum_{\nu \in {\mathcal{N}}_k} {\left( Q_{k,\nu} v, v\right)} \\ &= \sum_{\nu, \eta \in {\mathcal{N}}_k}{\left( Q_{k,\nu} v, v_\eta\right)} \\ &\leq \sum_{\nu, \eta \in {\mathcal{N}}_k}{\left\VertQ_{k,\nu} v\right\Vert}{\left\Vertv_\eta\right\Vert}_{\Omega_{k,\nu}} \\ &\leq \sum_{\eta \in {\mathcal{N}}_k}\left( \sum_{\nu \in {\mathcal{N}}_k} {\left\VertQ_{k,\nu}v\right\Vert}^2 \right)^\frac{1}{2}\left(\sum_{\nu \in {\mathcal{N}}_k}{\left\Vertv_\eta\right\Vert}_{\Omega_{k,\nu}}^2\right)^\frac{1}{2}. \end{aligned}$$]{} Because of the shape-regularity of ${\mathcal{T}}_k$, we can bound [$$\sum_{\nu \in {\mathcal{N}}_k}{\left\Vertv_\eta\right\Vert}_{\Omega_{k,\nu}}^2 \leq K_1 {\left\Vertv_\eta\right\Vert}^2,$$]{} for some $K_1$ that is independent of $\eta$ and $k$. Thus, [$$\begin{aligned} \sum_{\nu \in {\mathcal{N}}_k}{\left( Q_{k,\nu}v, Q_{k,\nu} v\right)} &\leq K_1 \sum_{\eta \in {\mathcal{N}}_k} {\left\Vertv_\eta\right\Vert}^2 \\ &\leq K_1 K_0 {\left\Vertv\right\Vert}^2. \end{aligned}$$]{} If we now define $\tilde{Q}_{k,\nu} = (K_0K_1)^{-\frac{1}{2}}Q_{k,\nu}$, and $\tilde{R}^s_k = {(K_0 K_1)^{-1}}R^s_k$, we have that [$$\sum_{\nu \in {\mathcal{N}}_k}{\left( \tilde{Q}_{k,\nu}v, \tilde{Q}_{k,\nu} v\right)} \leq {\left\Vertv\right\Vert}^2 \text{ and } {\left( \tilde{R}^s_k v, v\right)} = \sum_{\nu \in {\mathcal{N}}_k} {\left( A^{-s}_{k,\nu}\tilde{Q}_{k,\nu} v, \tilde{Q}_{k,\nu} v\right)}.$$]{} We can now use Jensen’s operator inequality , together with an argument analogous to that in the proof of Lemma \[lem:subspace-As-estimate\] to get [$$\begin{aligned} R^s_k &= K_0K_1 \tilde{R^s_k} \\ &\leq K_0 K_1 (\tilde{R}^1_k)^s \\ &= (K_0 K_1)^{1-s}(R^1_k)^s \end{aligned}$$]{} This, together with , proves the second inequality of with $C_2 = (K_0K_1)^{1-s}C^s$. We observe that the proof of Lemma \[lem:smoother-stability-verification\], shows that if the decomposition $V_k = \sum_{\eta \in {\mathcal{N}}_k}V_{k,\nu}$ is stable in both $L^2$- and $H_0^1$-norms, then it is also stable in the fractional norm ${\left\Vert\cdot\right\Vert}_{s,k}$. That is, if there are constants $c_0, c_1 > 0$ so that [$${\left( A^s_k v, v\right)} \leq c_s{\left( {(R^s_k)^{-1}}v, v\right)}, \quad \forall v\in V_k,$$]{} with $s=0$ and $s=1$, then the same holds for every $s\in [0,1]$, with $c_s = c_0^{1-s}c_1^s$. In this way, the smoother defined by is the natural interpolation between the corresponding smoothers for $s=0$ and $s=1$. In our current context, where $V_k$ is the space of continuous, piecewise linear functions relative to ${\mathcal{T}}_k$, the assumption in Lemma \[lem:smoother-stability-verification\] can be verified in the following manner. Fix $k$ and $v \in V_k$. Let ${\left\{ \theta_\nu \right\} }_{\nu \in {\mathcal{N}}_k}$ be a partition of unity subordinate to ${\left\{ \Omega_{k,\nu} \right\} }_{\nu \in {\mathcal{N}}_k}$, and let $\pi_k$ denote the nodal interpolant on $V_k$. We then set $v_\nu = \pi_k \theta_\nu v \in V_{k,\nu}$, in which case we see that $v = \sum_{\nu \in {\mathcal{N}}_k} v_\nu$. Furthermore, we have that [$$\label{eq:CG1-L2-stable-decomposition} {\begin{aligned} \sum_{\nu \in {\mathcal{N}}_k}{\left\Vertv_\nu\right\Vert}^2 &=\sum_{\nu \in {\mathcal{N}}_k} \int_{\Omega_{k,\nu}} |v_\nu|^2 {\mathrm{d}}x \\ &\leq \sum_{\nu \in {\mathcal{N}}_k} {\left\Vertv\right\Vert}^2_{\Omega_{k,\nu}} \\ &\leq (n+1){\left\Vertv\right\Vert}^2, \end{aligned}}$$]{} where the last inequality follows by the fact that no point in $\Omega$ is contained in more than $n+1$ subdomains $\Omega_{k,\nu}$. Thus, holds with $K_0 = n+1$. As noted in [@bramble1987newconvergence Remark 5.1] the $\alpha$ in the approximation and regularity assumption is closely related to the elliptic regularity of the continuous problem. Therefore, we make the following assumption: \[ass:elliptic-regularity\] There is an $\alpha \in (0,1]$ so that $A$ is a bounded operator from $H^1_0(\Omega) \bigcap H^{1+\alpha}(\Omega)$ to $H^{-1+\alpha}(\Omega)$, and $A^{-1}$ is a bounded operator from $H^{-1+\alpha}(\Omega)$ to $H^1_0(\Omega) \bigcap H^{1+\alpha}(\Omega)$. Assumption \[ass:elliptic-regularity\] is standard for proving condition in the case of $s=1$ (cf. for instance [@bramble1987newconvergence]). In [@bonito2015numerical Thm 4.3, and Rem. 4.1] Bonito et al. used Assumption \[ass:elliptic-regularity\] to prove the error estimate [$$\label{eq:bonito-error-estimate} {\left\Vert(A^{-s}-A_k^{-s}Q_k)f\right\Vert} \leq C h_k^{2s}{\left\Vertf\right\Vert}, \quad \forall f \in L^2(\Omega),$$]{} when $\alpha > s$. By the triangle inequality and the bounded refinement hypothesis it then follows that [$$\label{eq:bonito-error-estimate2} {\left\Vert(A^{-s}_{k}-A_{k-1}^{-s}Q_{k-1})f\right\Vert} \leq C h_k^{2s}{\left\Vertf\right\Vert}, \quad \forall f \in V_k,$$]{} for each $k$. This estimate is sufficient to verify in our current context. The result is stated in the following Lemma. \[lem:approximation-property\] Assume that Assumption \[ass:elliptic-regularity\] is satisfied with $\alpha > s$. Then there is a constant $C_3 > 0$, so that for every $k$ [$$\label{eq:approximation-property-BPXs} {\left( A^s_k(I-P^s_{k,k-1})v, v\right)} \leq C_{3} \frac{{\left\VertA^s_k v\right\Vert}^2}{\lambda^s_k}.$$]{} From the definition of $P^s_{k,k-1}$ in , [$$I - P^s_{k,k-1} = I - A^{-s}_{k-1}Q_{k-1}A^{s}_k = (A^{-s}_k - A^{-s}_{k-1}Q_{k-1})A^s_k,$$]{} and so, for any $v \in V_k$ [$${\left( A^s_k(I-P^s_{k,k-1})v, v\right)} \leq {\left( (A^{-s}_k - A^{-s}_{k-1}Q_{k-1})A^s_k v, A^s_k\right)}.$$]{} Using Cauchy-Schwarz inequality together with the error estimate , we get [$$\label{eq:approximation-property-BPXs-proof1} {\left( A^s_k(I-P^s_{k,k-1})v, v\right)} \leq {\left\Vert(A^{-s}_k - A^{-s}_{k-1}Q_{k-1})A^s_k v\right\Vert} {\left\VertA^s_k v\right\Vert} \leq C h_k^{2s} {\left\VertA^s_k v\right\Vert}^2.$$]{} By the quasi-uniformity of the mesh and $h_k^2 \leq C \lambda_k^{-1}$ it follows that $h_k^{2s} \leq C\lambda_k^{-s}$ . Using this in completes the proof. We are finally in a position to prove the main theorem of this section. \[thm:BPXs-spectral-equivalence\] Let Assumption \[ass:elliptic-regularity\] be satisfied with $\alpha > s$. Then, for $s \in [0,1]$ with $B_h^s$ defined by and satisfies [$$\label{eq:BPXs-spectral-equivalence} C_1 {C_3^{-1}}{\left( A_h^s v, v\right)} \leq {\left( B_h^s A_h^s v, A_h^s v\right)} \leq C_2J{\left( A_h^s v, v\right)}, \quad \forall v \in V,$$]{} where $C_1$, $C_2$, and $C_3$ are the same as in Lemmas \[lem:smoother-stability-verification\] and \[lem:approximation-property\]. This result is a straightforward application of Theorem \[thm:bpx-spectral-equivalence\] together with Lemmas \[lem:smoother-stability-verification\] and \[lem:approximation-property\]. Theorem \[thm:BPXs-spectral-equivalence\] shows that the condition number $K(B_h^s A_h^s) \leq C J$, and so increases linearly with the number of mesh levels, but is independent of $h$. With less regularity of the domain, we can still prove a slightly weaker form of spectral equivalence. By the assumed quasi-uniformity of ${\mathcal{T}}_k$, we have for $k=2,\ldots,J$ that [$${\left\Vert(I-Q_{k-1})v\right\Vert}^2 \leq C h_k^2 {\left\Vertv\right\Vert}_{1}^2, \quad \forall v \in V_k.$$]{} This, together with the boundedness of $I-Q_{k-1}$ and interpolation theory, yields [$${\left\Vert(I-Q_{k-1})v\right\Vert}^2 \leq C h_k^{2s}{\left\Vertv\right\Vert}_{s,k}^2 \leq C_4 \lambda_k^{-s}{\left\Vertv\right\Vert}_{s,k}^2,$$]{} for some constant $C_4$, independent of $k$. By the discussion in Remark \[rem:bpx-weak-spectral-equivalence\], we get that [$$\label{eq:BPXs-weak-spectral-equivalence} {C_4^{-1}}C_1 {J^{-1}}{\left( A^s_h v, v\right)} \leq {\left( B^s_h A^s_h v, A^s_h v\right)} \leq C_2 J{\left( A^s_h v, v\right)}, \quad \forall v \in V_h,$$]{} and the condition number is bounded by $K(B^s_h A^s_h) \leq C J^2$. Preconditioner when $s \in [-1,0]$ {#sec:preconditioning-negative-s} ================================== For $s \in [-1,0]$, the large eigenvalues of $A^s_h$ correspond to smooth functions. In a multilevel setting this means that neither relaxation nor coarse grid correction will reduce the oscillatory components of the error. As such, we cannot expect a direct multigrid approach to work. Moreover, when $s <0$ the Löwner-Heinz’ and Jensen’s operator inequalities in and fail to hold, and the argument of Section \[sec:discrete-bpx\] is no longer valid. In this section, we will therefore investigate an alternative approach for constructing preconditioners. We will base the preconditioner for $A^s_h$ when $s$ is negative on our previously defined preconditioners $B^t_h$ for $t \in [0,1]$ together with the multiplicative decomposition of $A_h$ [$$\label{eq:As-neg-decomp} A^{-s}_h = A_h^{-\frac{1+s}{2}}A_h A_h^{-\frac{1+s}{2}}.$$]{} We have for every $u \in V_h$ and $t \in {\mathbb{R}}$ that [$${\left\Vertu\right\Vert}_{-s,h} = {\left\VertA^{-\frac{t+s}{2}}_hu\right\Vert}_{t,h}.$$]{} The specific form we will use below is [$$\label{eq:As-neg-mapping-property} {\left\Vertu\right\Vert}_{-\frac{1+s}{2}+\beta,h} = {\left\VertA^{-\frac{1+s}{2}}_h u\right\Vert}_{\frac{1+s}{2}+\beta,h},$$]{} which is valid for any $\beta \in {\mathbb{R}}$. Replacing the left- and rightmost factor of the right hand side in with a spectrally equivalent preconditioner $B^{\frac{1+s}{2}}_h$, yields a symmetric positive definite operator [$$\label{eq:tildeBs-def} \tilde{B}^{s}_h := B^{\frac{1+s}{2}}_h A_h B^{\frac{1+s}{2}}_h.$$]{} We want $\tilde{B}^s_h$ to be spectrally equivalent to $A^{-s}_h$. That is, there exist constants $C_1, C_2$ so that for every $u \in V_h$, [$$\label{eq:tildeBs-spectral-equivalence} C_1{\left( A^s_h u, u\right)} \leq {\left( \tilde{B}^s_h A^s_h u, A^s_h u\right)} \leq C_2 {\left( A^s_h u, u\right)}$$]{} holds. By the definition of $\tilde{B}^s_h$, [$${\left( \tilde{B}^s_h A^s_h u, A^s_h u\right)} = {\left( A_hB^{\frac{1+s}{2}}_hA^s_h u, B^{\frac{1+s}{2}}_hA^s_h u\right)} = {\left\VertB^{\frac{1+s}{2}}_hA^s_h u\right\Vert_{1}}^2,$$]{} and since ${\left( A^s_h u, u\right)} = {\left\Vertu\right\Vert_{s,h}}^2$, we see that the spectral equivalence in is equivalent to [$$\label{eq:tildeBs-norm-equivalence} C_1^{\frac{1}{2}}{\left\Vertu\right\Vert}_{s,h} \leq {\left\VertB_h^{\frac{1+s}{2}}A^s_h u\right\Vert}_1 \leq C_2^{\frac{1}{2}} {\left\Vertu\right\Vert}_{s,h}, \quad \forall u \in V_h.$$]{} Using the preconditioner described in section \[sec:discrete-bpx\], we have by the spectral equivalence established in Theorem \[thm:BPXs-spectral-equivalence\] that there are constant $C_1, C_2 > 0$ so that [$$\label{eq:Bs-norm-spectral-equivalence} C_1 {\left\Vertu\right\Vert}_{-\frac{1+s}{2},h} \leq {\left\VertB_h^{\frac{1+s}{2}}u\right\Vert}_{\frac{1+s}{2},h} \leq C_2 J {\left\Vertu\right\Vert}_{-\frac{1+s}{2},h}, \quad u \in V_h.$$]{} We assume now some additional regularity on $B_h^{\frac{1+s}{2}}$, similar to . That is, for some $\beta$, we have the norm equivalence [$$\label{eq:tildeBs-norm-equivalence-regularity} C_1 {\left\Vertu\right\Vert}_{-\frac{1+s}{2}+\beta,h} \leq {\left\VertB_h^{\frac{1+s}{2}}u\right\Vert}_{\frac{1+s}{2}+\beta,h} \leq C_2 J {\left\Vertu\right\Vert}_{-\frac{1+s}{2}+\beta,h}.$$]{} In particular, with $\beta = \frac{1-s}{2} \in \left[\frac{1}{2}, 1 \right]$, and replacing $u$ by $A^s_h u$ in we recover and the spectral equivalence . We note also that if we assume the additional regularity of , we can bound the condition number of $\tilde{B}^s_h A^s_h$, $K(\tilde{B}^s_hA^s_h)$, by [$$\label{eq:tildeBs-condition-number-bound} K(\tilde{B}^s_h A^s_h) \leq K(B_h^{\frac{1+s}{2}}A^{\frac{1+s}{2}}_h)^2.$$]{} Implementational concerns {#sec:implementation} ========================= The discrete operators discussed so far are related to, but are not the same as the matrices used in the implementation. In this section we will discuss how to implement these operators. We begin by discussing the matrix representation of the discrete fractional operators. We refer also to [@mardal2011preconditioning] for more details. While the discrete fractional operators satisify the group property $A_h^s A_h^t = A_h^{s+t}$, their matrix representations do not. In particular, for $t=-s$, $A_h^s A_h^{-s} = I_h$ and the finite element matrix representation of the identity is the mass matrix. Hence, in order to provide a precise description of the interpolation of the involved matrices, we let ${\left\{ \phi_h^i \right\} }_{i=1}^{N_h}$ be the standard nodal basis for $V_h$, and we introduce the operators $\pi_h, \mu_h: V_h\to {\mathbb{R}}^{N_h}$, defined by [$$\label{eq:matvec-representations} {\begin{aligned} v &= \sum_{i=1}^{N_h}\left(\pi_h v\right)_i \phi^i_h, \quad \text{ and } \\ (\mu_h v)_i &= {\left( v, \phi^i_h\right)}, \quad i=1,\ldots,N_h. \end{aligned}}$$]{} Subsequently, we will refer to $\pi_h v$ and $\mu_h v$ as the primal- and dual vector representation of $v$, respectively. The primal representation is sometimes called the nodal representation. We then have that [$$\label{eq:vector-representation-adjoints} \mu_h^* = \pi_h^{-1}, \quad \text{ and } \quad \pi_h^* = \mu_h^{-1}$$]{} To see this, take ${\mathsf{v}} \in {\mathbb{R}}^{N_h}$, and $u \in V_h$. Then, [$${\begin{aligned} {\left( \mu_h^* {\mathsf{v}}, u\right)} &= {\left( {\mathsf{v}}, \mu_h u\right)}_{l^2} \\ &= \sum_{i=1}^{N_h} {\mathsf{v}}_i {\left( u, \phi_h^i\right)} \\ &= {\left( u, \sum_{i=1}^{N_h}{\mathsf{v}}_i\phi_h^i\right)} \\ &= {\left( u, \pi_h^{-1}{\mathsf{v}}\right)}, \end{aligned}}$$]{} where ${\left( \cdot, \cdot\right)}_{l^2}$ is the standard Euclidean inner product on ${\mathbb{R}}^{N_h}$. This proves the first identity in . The second identity is proven similarly. Using these operators, the stiffness matrix can then be expressed as $${\mathsf{A}}_h = \mu_h A_h \pi_h^{-1}, \quad \mbox{and} \quad ({\mathsf{A}}_h)_{i,j} = (A_h \phi_h^j, \phi_h^i ), \quad 1\le i, j \le N_h,$$ and the mass matrix is $${\mathsf{M}}_h = \mu_h I_h \pi_h^{-1} = \mu_h \pi_h^{-1}, \quad \mbox{and} \quad ({\mathsf{M}}_h)_{i,j} = (M_h \phi_h^j, \phi_h^i ), \quad 1\le i, j \le N_h .$$ We see that for both the stiffness- and mass matrix, a matrix-vector product takes primal vectors as input and returns dual vectors. For the matrix realization of $A^s_h$, let ${\left\{ (\lambda_i, {\mathsf{u}}_i) \right\} }_{i=1}^{N_h} \subset {\mathbb{R}}\times {\mathbb{R}}^{N_h}$ be the eigenpairs of the generalized eigenvalue problem [$$\label{eq:generalized-eigenvalue-problem-matrices} {\mathsf{A}}_h {\mathsf{u}}_i = \lambda_i {\mathsf{M}}_h {\mathsf{u}}_i,$$]{} normalized so that ${\mathsf{u}}_j^{\top} {\mathsf{M}}_h {\mathsf{u}}_i = \delta_{i,j}$. Setting ${\mathsf{\Lambda}}_h = \operatorname{diag} (\lambda_1,\ldots,\lambda_{N_h})$, and ${\mathsf{U}} = [{\mathsf{u}}_1,\ldots, {\mathsf{u}}_{N_h}]$, we have that [$$\label{eq:orthonormal-eigenvectors} {\mathsf{U}}^{\top} {\mathsf{M}}_h {\mathsf{U}} = I, \quad \text{ and } \quad {\mathsf{U}}^{\top}{\mathsf{A}}_h {\mathsf{U}} = {\mathsf{\Lambda}}_h.$$]{} We then define [$$\label{eq:Ash-matrix-realization} {\mathsf{A}}_h^s = \left({\mathsf{M}}_h {\mathsf{U}} \right){\mathsf{\Lambda}}_h^s \left({\mathsf{M}}_h {\mathsf{U}} \right)^{\top}.$$]{} One can verify that the entries of ${\mathsf{A}}^s_h$ satisfy [$$({\mathsf{A}}_h^s)_{i,j} = {\left( A^s_h \phi_h^j, \phi_h^i\right)},$$]{} in which case ${\mathsf{A}}^s_h = \mu_h A^s_h \pi_h^{-1}$. Using we are also able to see that [$$\label{eq:Ash-matrix-inverse} {({\mathsf{A}}_h^s)^{-1}} = {\mathsf{U}}{\mathsf{\Lambda}}_h^{-s} {\mathsf{U}}^{\top} = \pi_h A^{-s}_h \mu_h^{-1},$$]{} making it the matrix realization of $A^s_h$ viewed as an operator from $X_h$ to ${X_h^{'}}$. However, the group properties mentioned above only make sense when we consider $A^s_h$ as operators on $X_h$. Thus, we see that for the matrices [$$\pi_h A^s_h {\pi_h^{-1}} = (\pi_h {\mu_h^{-1}}) \mu_h A^s_h {\pi_h^{-1}} = {{\mathsf{M}}_h^{-1}} {\mathsf{A}}_h^s,$$]{} the group properties are satisfied. This can also be verified using the definition of ${\mathsf{A}}^s_h$ in . Since matrix-vector products involving ${\mathsf{A}}^s_h$ take primal vectors as input and return dual vectors, the matrix realization of $B^s_h$ should take dual vectors as input and return primal vectors. Then the product ${\mathsf{B}}^s_h {\mathsf{A}}^s_h$ acts on primal vectors, and is thus suitable for a Krylov subspace method. See also [@mardal2011preconditioning Section 6] and [@bramble1993multigrid Section 15]. Therefore, we define [$$\label{eq:matBs-def} {\mathsf{B}}^s_h = \pi_h B^s_h \mu_h^{-1}.$$]{} To see how ${\mathsf{B}}^s_h$ is implemented, we begin by supposing that $\dim V_k = N_k$ for $k=1,\ldots,J$. Let ${\left\{ \phi_k^i \right\} }_{i=1}^{N_k}$ be bases for $V_k$, and we define operators $\pi_k, \mu_k : V_k \to {\mathbb{R}}^{N_k}$ analogously to . We then define mass- and stiffness matrices on level $k$ as ${\mathsf{M}}_k = \mu_k \pi_k^{-1}$ and ${\mathsf{A}}_k = \mu_k A_k \pi_k^{-1}$, respectively. By assumption, for every $k$, $V_k \subset V_h$, and so there are matrices ${\mathsf{I}}_k: {\mathbb{R}}^{N_k} \to {\mathbb{R}}^{N_h}$ so that [$$\phi_k^i = \sum_{j=1}^{N_h} ({\mathsf{I}}_k)_{i,j}\phi_h^j, \quad i=1,\ldots,N_k.$$]{} In fact, ${\mathsf{I}}_k$ is the matrix realization of the inclusion operator $I_k: V_k \to V_h$, i.e. ${\mathsf{I}}_k = \pi_h I_k \pi_k^{-1}$. Using that $Q_k = I_k^*$ and we have that the transpose of ${\mathsf{I}}_k$ satisfies [$$\begin{aligned} {\mathsf{I}}_k^{\top} &= \left(\pi_h I_k \pi_k^{-1} \right)^* \\ &= (\pi_k^{-1})^* Q_k \pi_h^* \\ &= \mu_k Q_k \mu_h^{-1} \\ &=: {\mathsf{Q}}_k, \end{aligned}$$]{} which is the matrix realization of $Q_k$ in dual representation. Thus, for the matrix ${\mathsf{B}}^s_h$ we have that [$$\label{eq:matBs-impl-step1} {\begin{aligned} {\mathsf{B}}^s_h &= \pi_h B_h^s \mu_h^{-1} \\ &= \sum_{k=1}^J \pi_h I_k R_{k}^{s} Q_k \mu_h^{-1} \\ &= \sum_{k=1}^J (\pi_h I_k \pi_k^{-1}) (\pi_k R_k^s \mu_k^{-1}) (\mu_k Q_k \mu_h^{-1}) \\ &= \sum_{k=1}^J {\mathsf{Q}}_k^{\top} {\mathsf{R}}^s_k {\mathsf{Q}}_k, \end{aligned}}$$]{} where we define ${\mathsf{R}}^s_k = \pi_k R_k^s \mu_k^{-1}$ as the matrix realization of $R^s_k$. We see that due to ${\mathsf{R}}^s_1 = ({\mathsf{A}}^s_1)^{-1}$. For $k\geq2$ we define for $\nu \in {\mathcal{N}}_k$ operators $\pi_{k,\nu}, \mu_{k,\nu}: V_{k,\nu} \to {\mathbb{R}}^{\dim V_{k,\nu}}$ and matrices ${\mathsf{Q}}_{k,\nu}: {\mathbb{R}}^{N_k} \to {\mathbb{R}}^{\dim V_{k,\nu}}$, similarly to the above. The matrix realization of $R^s_k$ then becomes [$$\label{eq:matRs-impl} {\mathsf{R}}^s_k = \sum_{\nu \in {\mathcal{N}}_k} {\mathsf{Q}}_{k,\nu}^{\top} ({\mathsf{A}}^s_{k,\nu})^{-1} {\mathsf{Q}}_{k,\nu}.$$]{} Here, ${\mathsf{A}}^s_{k,\nu} = \mu_{k,\nu} A^s_{k,\nu} \pi_{k,\nu}^{-1}$. By , the implementation of ${\mathsf{R}}^s_k$ will require solving many small eigenvalue problems. In the particular case of continuous, piecewise linear finite element functions, and subdomains $\Omega_{k,\nu}$ as described in Section \[sec:discrete-bpx\], the subspaces $V_{k,\nu}$ are one-dimensional. The matrix ${\mathsf{R}}^s_k$ is then diagonal, with entries [$$({\mathsf{R}}^s_k)_{i,i} = \frac{1}{({\mathsf{M}}_k)_{i,i}^{1-s}({\mathsf{A}}_k)_{i,i}^s}, \quad i=1,\ldots,N_k.$$]{} That is, this is the smoother mentioned in the introduction. Inserting into we get [$$\label{eq:matBs-impl-final} {\mathsf{B}}^s_h = {\mathsf{Q}}_1^{\top} ({\mathsf{A}}^s_1)^{-1} {\mathsf{Q}}_1 + \sum_{k=2}^{J}{\mathsf{Q}}_k^{\top}\left(\sum_{\nu \in {\mathcal{N}}_k} {\mathsf{Q}}_{k,\nu}^{\top} ({\mathsf{A}}^s_{k,\nu})^{-1} {\mathsf{Q}}_{k,\nu} \right){\mathsf{Q}}_k.$$]{} We end this section by showing how to implement $\tilde{B}^s_h$, when $s\in[-1,0]$. In this case, the matrix realization of $\tilde{B}^s_h$ can be found from ${\mathsf{B}}^{\frac{1+s}{2}}_h$ and ${\mathsf{A}}_h$ by [$$\label{eq:mattildeBs-impl} {\begin{aligned} {\mathsf{\tilde{B}}}^s_h &:= \pi_h \tilde{B}^s_h \mu_h^{-1} \\ &= (\pi_h B^{\frac{1+s}{2}}_h \mu_h^{-1}) (\mu_h A_h \pi_h^{-1})(\pi_h B^{\frac{1+s}{2}}_h \mu_h^{-1}) \\ &= {\mathsf{B}}^{\frac{1+s}{2}}_h {\mathsf{A}}_h {\mathsf{B}}^{\frac{1+s}{2}}_h. \end{aligned}}$$]{} That is, $\tilde{B}^s_h$ is implemented as an application of $B^{\frac{1+s}{2}}_h$, followed by a multplication of the stiffness matrix and a second application of $B^{\frac{1+s}{2}}_h$. Numerical experiments {#sec:numerical-experiments} ===================== In this section we present a series of numerical experiments that aim to validate the theoretical results we established in previous sections. We also present numerical results for the case when $s < 0$, using $\tilde{B}^s_h$, defined in , as preconditioner. Specifically, in Section \[sec:numerical-experiments-fracLap\] we solve [$$A^s_h u = f,$$]{} using preconditioned conjugate gradient method with $B^s_h$ defined in as preconditioner. In Section \[sec:numerical-experiments-EMI\], we consider a coupled multidomain problem, where the weakly imposed continuity on the interface leads to a Lagrange multiplier in $H^{\pm\frac{1}{2}}$. The numerical experiments are conducted using random initial guess. Convergence in the iterative methods used is reached when the relative preconditioned residual, i.e. $\frac{{\left( Br_k, r_k\right)}}{{\left( Br_0, r_0\right)}}$, where $r_k$ is the residual at the $k$-th iteration and $B$ is the preconditioner, is below a given tolerance. Preconditioning the fractional Laplacian {#sec:numerical-experiments-fracLap} ---------------------------------------- In the first set of numerical experiments, we show the performance of the preconditioners $B^s_h$ and $\tilde{B}^s_h$, defined in and , respectively, depending on the sign of $s$ for the $A^s_h$ inner product. That is, for a given $f_h \in V_h$, we solve: Find $u_h \in V_h$ such that [$$\label{eq:numerex1} {\left( A^s_h u_h, v\right)} = {\left( f_h, v\right)}, \quad \forall v \in V_h,$$]{} where $s \in [-1,1]$. We take $\Omega = [0,1] \subset {\mathbb{R}}$, and ${\mathcal{T}}_h$ is a uniform partition of $\Omega$ consisting of $N = \frac{1}{h}$ elements. $V_h$ is then the space of continuous, piecewise linear functions relative to ${\mathcal{T}}_h$ that vanish on ${\partial}\Omega$. We solve the linear system arising from using preconditioned conjugate gradient, with $B^s_h$ as preconditioner if $s \geq 0$, and $\tilde{B}^s_h$ if $s < 0$. For $s \geq 0$, iteration counts and estimated condition numbers can be viewed in Table \[tab:fraclap\_positive\_s\_bpx\]. From these results we see that both the iteration counts and condition numbers stay uniformly bounded for each $s$. The analogous results for $s \leq 0$ can be seen in Table \[tab:fraclap\_negative\_s\_bpx\]. Here the situation is slightly more complicated. For each $s$, the iteration counts and condition numbers seem to increase for small $N$ (large $h$), but ultimately stay bounded when $N$ is increased. Worth noting is that the bound is relatively sharp. For instance, for $s = -1$, the preconditioner $\tilde{B}^s_h$ does two applications of $B^0_h$, and has estimated condition numbers around $193$. By Table \[tab:fraclap\_positive\_s\_bpx\], $K(B^0_hA^0_h) \approx 13.8$, and so $K(\tilde{B}^{-1}A^{-1}_h) \approx K(B^0_hA^0_h)^2$. Similar relations holds for other values of $s \leq 0$. 32 64 128 256 512 ------- ------------ ------------ ------------ ------------ ------------ $0.0$ $20(13.5)$ $25(13.6)$ $28(13.8)$ $29(13.8)$ $29(13.9)$ $0.1$ $18(8.7)$ $21(8.9)$ $23(8.9)$ $24(8.9)$ $24(8.9)$ $0.2$ $16(5.8)$ $18(6.4)$ $19(6.5)$ $21(6.5)$ $21(6.6)$ $0.3$ $14(4.2)$ $15(4.7)$ $17(4.9)$ $18(5.0)$ $18(5.0)$ $0.4$ $12(3.4)$ $14(3.7)$ $15(3.8)$ $15(3.9)$ $16(3.9)$ $0.5$ $11(2.9)$ $12(3.0)$ $13(3.1)$ $13(3.1)$ $14(3.2)$ $0.6$ $12(2.9)$ $13(3.0)$ $13(3.0)$ $14(3.1)$ $14(3.0)$ $0.7$ $12(3.0)$ $13(3.0)$ $14(3.1)$ $14(3.1)$ $14(3.1)$ $0.8$ $13(3.2)$ $14(3.3)$ $14(3.3)$ $14(3.3)$ $14(3.3)$ $0.9$ $14(3.5)$ $15(3.6)$ $15(3.6)$ $15(3.6)$ $15(3.6)$ $1.0$ $14(4.0)$ $16(4.1)$ $16(4.1)$ $16(4.1)$ $16(4.1)$ : Numerical results for $(-\Delta)^s$ with nonnegative $s$. Table shows the number of preconditioned conjugate gradient iterations until reaching error tolerance $10^{-15}$. Estimated condition numbers are shown inside parentheses. $N$ is number of elements on the finest mesh. $J = 5$ in all tests.[]{data-label="tab:fraclap_positive_s_bpx"} 32 64 128 256 512 -------- ------------- ------------- ------------- ------------- ------------- $-1.0$ $32(184.4)$ $47(192.4)$ $56(192.7)$ $64(193.8)$ $62(191.2)$ $-0.9$ $28(119.0)$ $43(118.9)$ $50(120.5)$ $54(120.7)$ $55(119.9)$ $-0.8$ $26(78.3)$ $37(82.6)$ $46(84.5)$ $48(83.8)$ $49(83.9)$ $-0.7$ $25(53.0)$ $33(60.1)$ $40(61.9)$ $42(62.1)$ $45(61.5)$ $-0.6$ $24(36.9)$ $31(43.8)$ $35(45.8)$ $38(46.2)$ $41(46.2)$ $-0.5$ $22(26.8)$ $25(31.9)$ $30(34.3)$ $34(34.9)$ $38(35.1)$ $-0.4$ $20(20.4)$ $24(24.8)$ $28(26.5)$ $32(27.0)$ $37(27.1)$ $-0.3$ $17(16.1)$ $21(19.3)$ $27(20.7)$ $30(21.1)$ $34(21.1)$ $-0.2$ $17(13.1)$ $21(15.3)$ $25(16.4)$ $29(16.7)$ $32(16.7)$ $-0.1$ $16(11.0)$ $20(12.4)$ $23(13.2)$ $27(13.5)$ $29(13.5)$ $0.0$ $14(9.4)$ $17(10.4)$ $20(11.0)$ $24(11.2)$ $27(11.1)$ : Numerical results for $(-\Delta)^s$ with negative $s$. Table shows the number of preconditioned conjugate gradient iterations until reaching error tolerance $10^{-15}$. Estimated condition numbers are shown inside parentheses. $N$ is number of elements on the finest mesh. $J = 5$ in all tests.[]{data-label="tab:fraclap_negative_s_bpx"} Multidomain preconditioning {#sec:numerical-experiments-EMI} --------------------------- In this section we apply the multilevel algorithm to construct mesh independent preconditioners for a coupled multidomain problem originating from a geometrically accurate model of electric signal propagation in cardiac tissue, the EMI model, [@tveito2017cell]. We remark that the EMI model is simple in a sense that it is a single-physics problem where two elliptic equations are coupled. However, the interface problems encountered here are identical to those found in multiphysics applications, e.g. the couled Darcy-Stokes system [@layton2002coupling] or the Stokes-Biot system [@Ambartsumyan2018]. Let $\Omega\subset\mathbb{R}^2$ be a bounded domain decomposed into two non-overlapping subdomains $\Omega_1$, $\Omega_2$ with a common interface $\Gamma=\partial\Omega_1\bigcap\partial\Omega_2$ forming a closed curve. Motivated by the application the subdomain $\Omega_1$ is designated as the *exterior* domain, i.e. $\partial\Omega_2\bigcap\partial\Omega=\emptyset$. With $\epsilon > 0$ and $n$ the outer normal of the exterior domain we now aim to solve $$\label{eq:emi_system} \begin{aligned} u_1-\Delta u_1 &= f_1, & x\in \Omega_1,\\ u_2 -\Delta u_2&= f_2, & x\in \Omega_2,\\ n\cdot \nabla u_1 - n\cdot\nabla u_2 &= 0, & x\in \Gamma,\\ \epsilon (u_1 - u_2) + n\cdot \nabla u_1 &= g, & x\in\Gamma. \end{aligned}$$ The choice of boundary conditions for shall be discussed shortly. We remark that in the EMI model the parameter $\epsilon$ plays a role of inverse time step and thus algorithms robust with respect to the parameter are of interest. However, here the system will be considered only for a fixed choice of the parameter. Considering with homogeneous Neumann boundary conditions $n\cdot\nabla u_1=0$ on $\partial\Omega$ and letting $W_1=H^1(\Omega_1)\times H^1(\Omega_2)\times (H^{-1/2}(\Gamma)\bigcap \epsilon^{-1/2}L^2(\Gamma))$ the variational formulation of defines an operator $\mathcal{A}_1: W_1\rightarrow W'_1$ $$\label{eq:mortarA} \mathcal{A}_1 = \begin{pmatrix} I-\Delta & 0 & T_{1}^{*}\\ 0 & I-\Delta & -T_{2}^{*}\\ T_{1} & -T_{2} & -\epsilon^{-1}I_{} \end{pmatrix},$$ where $T_i$, $T_i v=v|_{\Gamma}$ for $v\in C(\bar{\Omega}_i)$, $i=1, 2$ are the trace operators on $H^1(\Omega_1)$ and $H^1(\Omega_2)$, respectively. Tveito et al. [@tveito2017cell] further discuss a mixed formulation of the system where additional unknowns $\sigma_i=-\nabla u_i$, $i=1,2$ are introduced. If homogeneous Dirichlet boundary conditions $u_1=0$ on $\partial\Omega$ are assumed the mixed formulation leads to operator $\mathcal{A}_2: W_2\rightarrow W'_2$ $$\label{eq:hdivA} \mathcal{A}_2 = \begin{pmatrix} I & \nabla & T^{*}\\ -\nabla\cdot& -I & 0\\ T & 0 & -\epsilon I \end{pmatrix},$$ with $W_2=H(\text{div}, \Omega)\times L^2(\Omega) \times (H^{1/2}(\Gamma)\bigcap \epsilon^{1/2} L^2(\Gamma))$ and $T$ the normal trace operator $T\sigma= \sigma|_{\Gamma}\cdot n$ for $v\in \left[C(\Omega)\right]^2$. We remark that operators $\mathcal{A}_1$ and $\mathcal{A}_2$ also arise naturally in the analysis of non-overlapping domain decomposition methods for second order elliptic problems in the primal [@wohlmuth2000mortar] and mixed formulation [@cowsar1995balancing] respectively. Assuming that the operators $\mathcal{A}_1$ and $\mathcal{A}_2$ are isomorphisms on their respective spaces[^2] the preconditioners can be established within the framework of operator preconditioning [@mardal2011preconditioning]. In particular, the Riesz map preconditioner for is $$\label{eq:mortarB} \mathcal{B}_1=\begin{pmatrix} I-\Delta & &\\ & I-\Delta & \\ & & \epsilon^{-1}I + (-\Delta + I)^{-1/2} \end{pmatrix}^{-1},$$ while will be preconditioned by $$\label{eq:hdivB} \mathcal{B}_2=\begin{pmatrix} I - \nabla \nabla\cdot & &\\ & I & \\ & & \epsilon I + (-\Delta + I)^{1/2} \end{pmatrix}^{-1}.$$ Note that the operator sums in $\mathcal{B}_1$, $\mathcal{B}_2$ are due to the fact that the interface spaces are intersection spaces [@bergh]. In order to simplify the setting and focus only on the fractional operators in the preconditioners we remove the parameter dependence from the problems by settting $\epsilon=\infty$ in , and similarly $\epsilon=0$ for , . In turn, the interface spaces reduce to $H^{-1/2}(\Gamma)$ and $H^{1/2}(\Gamma)$ respectively and the multilevel algorithm is directly applicable for the related interface problems which now involve the operator $I-\Delta$, cf. the Laplacian operator in the previous sections. Robustness of $\mathcal{B}_1$, $\mathcal{B}_2$, and in particular the fractional Sobolev space preconditioner, are finally demonstrated by observing the iteration counts of the preconditioned MinRes method. In the experiments we let $\Omega=\left[0, 1\right]^2$ and $\Omega_2=\left[0.25, 0.75\right]^2$. The finite element discretization of $W_1$ uses continuous linear Lagrange elements. For $W_2$ the first subspace is constructed from the $H(\text{div})$-conforming lowest order Raviart-Thomas elements and the remaining subspaces use piecewise constant elements. The discrete preconditioners shall use off-the-shelve methods for the first two blocks. More specifically, a single $V$ cycle of algebraic multigrid is used for $\mathcal{B}_1$ while for $\mathcal{B}_2$ the action is computed exactly by a direct solver. The final block of the preconditioners is realized by the proposed multilevel preconditioner with different number of levels $J=2, 3, 4$. We note that with the choice of discretization for the space $W_2$ the fractional multigrid algorithm is applied outside of the setting used in the analysis of Section \[sec:discrete-bpx\]. The number of MinRes iterations is shown in Table \[tab:hdiv\] and Table \[tab:mortar\]. Here, the iterations were started from a random initial vector and terminated once the relative preconditioned residual norm was less then $10^{-8}$ in magnitude. For both $\mathcal{B}_1$ and $\mathcal{B}_2$ the iterations are bounded in the discretization parameter. The linear dependence on the number of levels is nicely visible in the results of $\mathcal{B}_2$. The tables further list iteration counts for preconditioners where the fractional operators were realized in terms of spectral decomposition. As expected from the theory and experiments for the Laplace problem the difference in iteration counts between the multilevel realization and specral realization is larger for $\mathcal{B}_1$ then it is for $\mathcal{B}_2$. ---------- --------- ------ ------- ------- ------- ---- $J=2$ $J=3$ $J=4$ 2.21E-02 20736 128 27 28 27 22 1.10E-02 82432 256 27 32 32 22 5.52E-03 328704 512 27 33 36 22 2.76E-03 1312768 1024 27 33 40 22 1.38E-03 5246976 2048 25 35 40 22 ---------- --------- ------ ------- ------- ------- ---- : Number of MinRes itarations for the operator $B_2\mathcal{A}_2$ and $\epsilon=10^{-15}$ using multilevel algorithm with $J$ levels as a preconditioner for $(-\Delta + I)^{1/2}$. Realizing the fractional operator with spectral decomposition leads to iteration counts in the last column. []{data-label="tab:hdiv"} ---------- --------- ------ ------- ------- ------- ---- $J=2$ $J=3$ $J=4$ 2.21E-02 4481 128 67 93 103 36 1.10E-02 17153 256 68 92 111 35 5.52E-03 67073 512 66 90 112 35 2.76E-03 265217 1024 64 90 112 34 1.38E-03 1054721 2048 64 88 108 33 ---------- --------- ------ ------- ------- ------- ---- : Number of MinRes itarations for the operator $B_1\mathcal{A}_1$ and $\epsilon=10^{15}$ using multilevel algorithm with $J$ levels as a preconditioner for $(-\Delta + I)^{-1/2}$. Realizing the fractional operator with spectral decomposition leads to iteration counts in the last column. []{data-label="tab:mortar"} [^1]: Thanks [^2]: The proof of this result as well as stable finite element discretization of the problem are subject of current work and will be reported elsewhere. We remark that operator $\mathcal{A}_1$ in the limit case $\epsilon=\infty$ has been studied in [@lamichhane2004mortar] in the context of mortar finite element method.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We introduce an asymptotically unbiased estimator for the full high-dimensional parameter vector in linear regression models where the number of variables exceeds the number of available observations. The estimator is accompanied by a closed-form expression for the covariance matrix of the estimates that is free of tuning parameters. This enables the construction of confidence intervals that are valid uniformly over the parameter vector. Estimates are obtained by using a scaled Moore-Penrose pseudoinverse as an approximate inverse of the singular empirical covariance matrix of the regressors. The approximation induces a bias, which is then corrected for using the lasso. Regularization of the pseudoinverse is shown to yield narrower confidence intervals under a suitable choice of the regularization parameter. The methods are illustrated in Monte Carlo experiments and in an empirical example where gross domestic product is explained by a large number of macroeconomic and financial indicators. **Keywords:** high-dimensional regression, confidence intervals, Moore-Penrose pseudoinverse, random projection, ridge regression' author: - | Tom Boot[^1]\ University of Groningen - | Didier Nibbering[^2]\ Erasmus University Rotterdam\ Tinbergen Institute bibliography: - 'literature.bib' title: | Inference in high-dimensional\ linear regression models --- #### Notation We use the following notation throughout the paper: For any $n\times 1$ vector $a=(a_1,\dots,a_n)'$, the $l_{q}$-norm is defined as $||a||_q:=(\sum^n_{i=1} |a_i|^q)^{1/q}$ for $q>0$ and $||a||_{0}$ denotes the number of nonzero elements of $a$. The maximum norm is written as $||a||_{\infty} = \max(|a_1|,\dots,|a_n|)$. For a $p\times n$ matrix $A$, the $l_{q}$-norm is defined as $||A||_q:=\sup_{x, ||x||_{q}=1}\left\{||Ax||_{q}\right\}$ and the maximum norm is written as $||A||_{\max} = \max_{i=1,\dots,n,j=1,\dots,p}|A_{ij}|$. The $n\times n$ identity matrix is denoted by $I_{n}$. The vector $e_{i}$ has its $i$-th entry equal to 1 and zeros everywhere else. For the regressor matrix $X$, we index the rows with the subscript $i=1,\ldots,n$ and the columns with the subscript $j=1,\ldots,p$. If $U$ is a $p\times p$ orthogonal matrix, we write $U\in \mathcal{O}(p)$. When two random variables $X$ and $Y$ follow the same distribution, this is denoted as $X\overset{(d)}{=}Y$. High-dimensional linear regression {#sec: gf} ================================== Consider the data generating process $$\label{eq:dgp} y = X\beta + \varepsilon, \quad \varepsilon \sim N(0,\sigma^2 I_{n}),$$ where $y$ is an $n \times 1$ response vector, $X$ an $n \times p$ regressor matrix, $\beta=(\beta_1,\dots,\beta_p)'$ a $p \times 1 $ vector of unknown regressor coefficients, and $\varepsilon$ an $n \times 1$ vector of errors which are independent and normally distributed with variance $\sigma^2$. The empirical covariance matrix of $X$ is denoted by $\hat{\Sigma} = \frac{1}{n}X'X$. We will show how the normality assumption on the errors can be relaxed. Approximate inverse and bias correction {#sec: approxinv} --------------------------------------- Define $M$ as a $p\times n$ matrix for which $MX$ is close to the $p\times p$ identity matrix $I_{p}$, in a sense that will be made precise below. We refer to $M$ as an approximate inverse for $X$. We start by considering estimators for $\beta$ of the form $$\label{eq:start} \begin{split} \hat{\beta} &= My\\ &=MX\beta +M\varepsilon\\ & = \beta +\left(MX - I_{p}\right)\beta + M\varepsilon.\\ \end{split}$$ The second term of represents a bias which depends on the accuracy of the approximate inverse $M$. When $p\leq n$, ordinary least squares yields unbiased estimates by choosing $M = (X'X)^{-1}X'$. When $p>n$, the matrix $X'X$ is singular, and we have to resort to an expression for $M$ for which the bias is not equal to zero. Suppose we have an accurate initial estimator $\hat{\beta}^{\text{init}}$, then we can reduce the bias in by applying a correction $$\label{eq:betac} \begin{split} \hat{\beta}^{c} &= My - \left(MX - I_{p}\right)\hat{\beta}^{\text{init}} \\ & = \beta +\left(MX - I_{p}\right)\left(\beta-\hat{\beta}^{\text{init}}\right) + M\varepsilon.\\ \end{split}$$ For the initial estimator $\hat{\beta}^{\text{init}}$ we take the lasso estimator of @tibshirani1996regression. Alternative initial estimators can be used, as long as they satisfy a sufficiently tight accuracy bound on the $l_{1}$ norm of $\beta-\hat{\beta}^{\text{init}}$. The goal of this paper is to introduce choices of $M$ such that the bias of the estimator $\hat{\beta}^{c}$ is of lower order than the variance. Anticipating the usual $\sqrt{n}$ rate of convergence, we rescale the estimator in as $$\label{eq: corr} \begin{split} \sqrt{n}\left(\hat{\beta}^{c}-\beta\right) &= \Delta+Z\\ \Delta &= \sqrt{n}\left(MX-I_{p}\right)\left(\beta-\hat{\beta}^{\text{init}}\right)\\ Z&=\sqrt{n}M\varepsilon \end{split}$$ The term $\Delta$ reflects the bias of the corrected estimator. To ensure asymptotic unbiasedness, $\Delta$ should be of lower order than the noise term $Z$. We propose specifications for the approximate inverse $M$ for which $Z|X\sim N(0,\sigma^2\Omega)$ with $\Omega = nMM'$ and the variance $\Omega_{jj} = O_{p}(1)$. This shows that the standard errors of the estimator $\hat{\beta}^{c}$ decrease at the familiar $n^{-1/2}$ rate. In order for the bias to vanish compared to the variance term, given that $\Omega_{jj}= O_{p}(1)$, we now need $||\Delta||_{\infty} = o_{p}(1)$. Under a sparsity assumption on $\beta$, we show that this is indeed the case, which implies that $\hat{\beta}^{c}$ is an asymptotically unbiased estimator. Combined with a closed-form expression for the covariance matrix $\Omega$, confidence intervals can be constructed for the $j$-th parameter as $$\label{eq: ci} \left[\hat{\beta}_{j}^{c} - z_{\alpha/2}\sqrt{\sigma^2\Omega_{jj}},\quad\hat{\beta}_{j}^{c} + z_{\alpha/2}\sqrt{\sigma^2\Omega_{jj}}\right],$$ where $z_{\alpha/2}$ is the $\alpha/2$ critical value for the standard normal distribution. We discuss estimation of $\sigma$ in Section \[sec: noise\]. The estimator defined in occurs in a different form in @zhang2014confidence, @van2014asymptotically and @javanmard2014confidence, who consider $ \hat{\beta}^{c} = \hat{\beta}^{\text{lasso}} + \frac{1}{n}\bar{M}X'(y-X\hat{\beta}^{\text{lasso}}). $ This leads to an interpretation of $\hat{\beta}^{c}$ as a ‘desparsified’ version of the lasso estimator. An alternative to the standard lasso estimator is put forward by @caner2014asymptotically. The matrix $\bar{M}$ serves as an approximate inverse to the empirical covariance matrix $\frac{1}{n} X' X$, which is found by a series of lasso regressions in @zhang2014confidence and @van2014asymptotically, or direct numerical optimization in @javanmard2014confidence. As a consequence of the complex estimation procedures, standard errors are not available in closed form, and their validity depends on the appropriate selection of one or more tuning parameters. Choosing the approximate inverse $M$ {#sec: chooseM} ------------------------------------ This section proposes specifications of $M$ for which the bias $||\Delta||_{\infty}$ in is small. We ensure that the diagonal terms of $MX - I_{p}$ are identically equal to zero by introducing a $p\times p$ diagonal matrix $D$, with diagonal elements $d_{j}$, and taking $$\label{eq:diagonalscaling} M = D\tilde{M}, \qquad d_{j} = (\tilde{m}_{j}'x_{j})^{-1},$$ with $\tilde{m}_{j}'$ the $j$-th row of $\tilde{M}$. We first choose $M$ in the form defined in , with $\tilde{M}$ specified as the Moore-Penrose pseudoinverse of $X$. Subsequently, we consider regularized alternatives obtained by random least squares and ridge regression. ### The Moore-Penrose pseudoinverse A tuning parameter free choice for $\tilde{M}$ in is the Moore-Penrose pseudoinverse. When $p\leq n$, and the columns of $X$ are linearly independent, $\tilde{M} = (X'X)^{-1}X'$. In the high-dimensional setting where $p>n$, the matrix $X$ has linearly dependent columns by default. In this case the pseudoinverse equals $X'(XX')^{-1}$, and $$\begin{aligned} \label{eq: mppi} M^{\text{MPI}} = D^{\text{MPI}}X'(XX')^{-1}.\end{aligned}$$ The diagonal elements $d_{j}^{\text{MPI}}$ of the diagonal scaling matrix $D^{\text{MPI}}$ equal $$d_{j}^{\text{MPI}}= \left[x_{j}'(X X')^{-1}x_{j}\right]^{-1}.$$ This provides a closed-form expression for the approximate inverse. In addition, since the bias term of the estimator is of lower order compared to the variance, the covariance of $\hat{\beta}^{c}$ is available in closed form as well, $$\label{eq:varbeta} V(\hat{\beta}^{c}) = D^{\text{MPI}}X'(XX')^{-2}XD^{\text{MPI}}.$$ ### Regularizing the Moore-Penrose pseudoinverse The accuracy of the Moore-Penrose pseudoinverse depends on the concentration of the eigenvalues of the matrix $XX'$, which can be weak when $p$ is close to $n$. Regularizing the approximate inverse can improve in accuracy, with smaller standard errors as a result. This section introduces two regularization techniques, for which Section \[sec: theory\] shows the appropriate choice for the regularization strength. #### Random Least Squares This method is based on projecting the highdimensional regressor matrix $X$ onto a $k<n$ dimensional subspace by post-multiplying with a $p\times k$ matrix $R$ with independently standard normally distributed elements, $$\label{eq: R} R_{jl} \sim N(0,1),\quad j={1,\ldots p}, \quad l={1,\ldots ,k}.$$ The multiplication yields a low-dimensional analogue to , $$y = XR\gamma_R + u.$$ Least squares estimation of $\gamma_R$ is straightforward as $$\hat{\gamma}_{R} = (R'X'XR)^{-1}R'X'y,$$ from which an estimator for $\beta$ can be constructed by $\hat{\beta}_R = R\hat{\gamma}_{R}$. Since $R$ is random, Jensen’s inequality can be used to show that the accuracy of this estimator can be improved by averaging over different realizations of $R$. We then arrive at the following estimator of $\beta$, $$\begin{aligned} \label{eq:rls} \hat{\beta}_{\bar{R}} = \text{E}_{R}[R\hat{\gamma}_{R}] = \text{E}_{R}[R(R'X'XR)^{-1}R']X'y. \end{aligned}$$ From equation , we recognize that random least squares yields an approximate inverse covariance matrix of $X$. Defining $\tilde{M}= \text{E}_{R}[R(R'X'XR)^{-1}R']X'$ in yields $$\begin{aligned} \label{eq:randproj} M^{\text{RLS}} = D^{\text{RLS}}\text{E}_{R}\left[R(R'X'XR)^{-1}R'\right]X',\end{aligned}$$ with $$\label{eq: drls} d^{\text{RLS}}_{j}=\left\{\text{E}_{R}[r_{j}'(R'X'XR)^{-1}R']X' x_{j}\right\}^{-1}.$$ #### Ridge regression An alternative regularization strategy is to use a ridge adjustment, $$\begin{aligned} \label{eq: ridge} M^{\text{RID}} = D^{\text{RID}}(X'X+\gamma I_{p})^{-1}X',\end{aligned}$$ where $\gamma$ denotes the ridge penalty and the elements of the diagonal scaling matrix $D^{\text{RID}}$ equal $$d^{\text{RID}}_{j}=\left(v_{j}'X'x_{j}\right)^{-1},$$ with $v_{j}$ the $j$-th row of $(X'X+\gamma I_{p})^{-1}$. The regularization in can be related to the Moore-Penrose pseudoinverse, since the latter is defined as $$\label{eq:ridgeMP} \begin{split} X'(XX')^{-1} &= \lim_{\gamma\rightarrow 0}\left(X'X + \gamma I_{p}\right)^{-1}X'\\ & = \lim_{\gamma\rightarrow 0}X'\left(XX' + \gamma I_{n}\right)^{-1}. \end{split}$$ which can be shown using the singular value decomposition of $X$ as in @albert1972regression. Estimation of the noise level {#sec: noise} ----------------------------- A consistent estimator of the noise level $\sigma^2$ is crucial to construct valid confidence intervals. Existing methods, such as @van2014asymptotically and @javanmard2014confidence rely on the scaled lasso developed by @scaledlassoSun, for which holds that $\left|\frac{\hat{\sigma}}{\sigma}-1\right| = o_{p}(1)$ under Assumption A\[ass:sparsity\] and Assumption A\[ass:regressor\] discussed in Section \[sec: ass\]. However, in the Monte Carlo simulations in Section \[sec: mc\], and in line with findings by @reid2013study, we find the scaled lasso to be unreliable in many settings. An alternative is to use $$\label{eq: lassosigma} \hat{\sigma}^{2}_{\text{lasso}} = \frac{1}{n-\hat{s}}\hat{\varepsilon}'\hat{\varepsilon},$$ with $\hat{s}$ the number of non-zero coefficients retained by the lasso, and $\hat{\varepsilon}$ the $n\times 1$ vector of lasso regression errors. Corresponding to the results in @reid2013study, we find that this leads to more robust estimation of the noise level. Theoretical results {#sec: theory} =================== This section provides the main results of the paper. Proofs for the theorems in this section are given in Appendix \[app:proofs\]. Assumptions {#sec: ass} ----------- Performing inference in a linear regression model with more variables than observations requires additional assumptions over its low-dimensional counterpart. Our assumptions parallel @fan2008sure and @wang2015high. We will provide a discussion below. \[ass:sparsity\] The sparsity $s_{0} = ||\beta||_{0}$ satisfies $s_{0} = o\left(\frac{\sqrt{n}}{\log p}\right)$. \[ass:regressor\] The regressor matrix $X$ is generated from an elliptical distribution, i.e. $$\label{eq:svdofX} X = \Sigma_{1}^{1/2}Z\Sigma_{2}^{1/2} = \Sigma_{1}^{1/2}VSU'\Sigma_{2}^{1/2},$$ where the $n \times n$ population covariance matrix $\Sigma_1$ and the $p \times p$ population covariance matrix $\Sigma_2$ determine the dependence between the rows and columns of $X$, respectively. The elements of the $n \times p$ matrix $Z$ are generated independently from a spherically symmetric distribution, $V\in\mathcal{O}(n)$, $S$ is an $n\times p$ matrix of singular values, and $U\in\mathcal{O}(p)$. Furthermore, $$\label{eq:subgaussian} P\left(\lambda_{\max}(p^{-1}ZZ')\geq c_{Z}, \quad \lambda_{\min}(p^{-1}ZZ')\leq c_{Z}^{-1}\right)\leq e^{-C_{Z} n},$$ where $\lambda_{\max}(.)$ and $\lambda_{\min}(.)$ are the largest and smallest eigenvalues of a matrix respectively, and $c_{Z},C_{Z}$ are positive constants. \[ass:condnumb\] For both the population covariance matrices $\Sigma_{1}$ and $\Sigma_{2}$, the eigenvalues are bounded by a constant, i.e. for $i = 1,2$, $$0<c_{i,1}\leq \lambda_{\min}(\Sigma_{i})\leq \lambda_{\max}(\Sigma_{i})\leq c_{i,2}<\infty.$$ Assumption A\[ass:sparsity\] imposes a sparsity constraint which restricts the number of non-zero coefficients in $\beta$ by $ s_{0}=||\beta||_{0}$. For lasso consistency, it is required that $s_{0}^2 = o\left(n/\log p\right)$. As noted in @van2014asymptotically and @javanmard2014confidence, a slightly stronger assumption is needed when constructing confidence intervals. In recent work, for example by @chernozhukov2015valid, assumption A\[ass:sparsity\] is relaxed to allow for approximate sparsity, arguably a more realistic assumption in practical applications. This restricts only the number of large non-zero coefficients, and allows the remaining coefficients to be sufficiently small. Since our results only depend on the $l_{1}$ norm of the lasso estimation error, which does not change under approximate sparsity, they remain valid under approximate sparsity. Assumption A\[ass:regressor\] requires that the regressors are generated from an elliptical distribution. The class of elliptical distributions includes the multivariate normal distribution, but also allows for heavier tailed distributions such as the power exponential distribution and the multivariate $t$ distribution [@serfling2006multivariate; @dasgupta2012concentration]. This class precludes $X$ to consist of binomial variables. However, our results rely on the distribution of the elements of $X'(XX')^{-1}X$, which consist of sums of binomial variables. It is possible that one can use the convergence of these sums towards a normal distribution to extend the results towards binomial regressors. The matrices $\Sigma_{1}$ and $\Sigma_{2}$ in Assumption A\[ass:regressor\] allows for dependence between the rows and the columns of $X$, respectively. Assumption A\[ass:condnumb\] states that the eigenvalues of these population covariance matrices are finite and independent of the dimensions $n$ and $p$. This assumption can be relaxed by replacing $c_{i,2}$ with $c_{i,2}n^{\alpha}$. The standard errors then decrease at the rate of $1/\sqrt{n^{1-\alpha}}$ instead of $1/\sqrt{n}$. Asymptotic unbiasedness and normality using the\ Moore-Penrose pseudoinverse {#sec: distrest} ------------------------------------------------ To prove that $\hat{\beta}^{c}$ in based on the Moore-Penrose pseudoinverse is an asymptotically unbiased estimator, we show that with high probability the bias term in is small and of lower order than the noise. Moreover, the construction of confidence intervals as in requires $Z|X$ to follow a normal distribution. Efficiency of the estimator is ensured by showing that the standard errors decrease at the usual $n^{-1/2}$ rate. The first requirement follows from bounding the bias term of the estimator in by a norm inequality, $$\label{eq:splitbias} ||\Delta||_{\infty}\leq \sqrt{n}\left|\left|MX-I_{p}\right|\right|_{\max}||\beta-\hat{\beta}^{\text{init}}||_{1},$$ which is an element-wise bound on $MX - I_{p}$ together with an $l_{1}$ accuracy bound on $\beta-\hat{\beta}^{\text{init}}$. The following lemma bounds on the first term in probability. \[theorem:biasvanish\] Suppose Assumption A\[ass:regressor\] and A\[ass:condnumb\] hold. Define $M^{MPI} = D^{MPI}X'(XX')^{-1}$ with $D^{MPI}$ a diagonal matrix with elements $d^{\text{MPI}}_{j}=(x_{j}'(XX')^{-1}x_{j})^{-1} $, then we have $$P\left(\left|\left|M^{MPI}X-I_{p}\right|\right|_{\max}\geq a\sqrt{\frac{\log p}{n}}\right) = O(p^{-\tilde{c}}),$$ with $\tilde{c}=\frac{c}{2c_{s}}a^2-2$ where $a,c,c_{s}>0$. A proof is presented in Appendix \[app:prooftheorembiasvanish\]. Note that the diagonal elements of $M^{\text{MPI}}X-I_{p}$ are identically zero, due to the diagonal scaling with $D^{\text{MPI}}$. Lemma \[theorem:biasvanish\] is therefore a statement on the off-diagonal elements of $M^{\text{MPI}}X-I_{p}$. Next we show that the $l_{1}$ norm of the initial estimation error, in the second term in the bound for $||\Delta||_{\infty}$ in , is bounded with high probability. As the initial estimator we use the lasso estimator by @tibshirani1996regression, which is defined as $$\label{eq:lasso} \hat{\beta}^{\text{lasso}} = \arg\min_{b}\left[\frac{1}{n}(y-Xb)'(y-Xb) + \lambda||b||_{1}\right].$$ The following bound applies to the $l_{1}$-error of the lasso estimator. \[lem:acclasso\] Suppose Assumption A\[ass:sparsity\] and Assumption A\[ass:regressor\] hold. Consider the lasso estimator with $\lambda\geq 8\sigma\sqrt{\frac{\log p}{n}}$, then with probability exceeding $1-2p^{-1}$ we have $$\label{eq:accguar} \left|\left|\beta-\hat{\beta}^{\text{lasso}}\right|\right|_{1} = O_{p}\left(s_{0}\sqrt{\frac{\log p}{n}}\right).$$ A proof is presented in Appendix \[proof:compcond\]. As shown in @buhlmann2011statistics, this bound applies under a so-called compatibility condition on $X$. The proof amounts to showing that the compatibility condition is indeed satisfied under Assumption A\[ass:sparsity\] and Assumption A\[ass:regressor\]. Combining Assumption A\[ass:sparsity\], Lemma \[theorem:biasvanish\], and Lemma \[lem:acclasso\], we see that the bias can be bounded by $$\label{eq:biasvanish2} ||\Delta||_{\infty} = O_{p}\left(s_{0}\frac{\log p}{\sqrt{n}}\right) = o_{p}(1).$$ In order for the estimator to be asymptotically unbiased, it is necessary that the bias in is of lower order than the noise term of the estimator, given by $Z$ in . The following lemma states that this is indeed the case. \[theorem:finitevar\] Suppose Assumption A\[ass:regressor\] and A\[ass:condnumb\] hold. For $j=1,\ldots,p$ we have $$\begin{split} Z_{j} &= \sqrt{n}d^{\text{MPI}}_{j}x_{j}'(XX')^{-1}\varepsilon,\\ Z_{j}|X &\sim N(0,\sigma^2\Omega_{j}),\\ ||\Omega_{jj}||_{2} &= O_{p}(1), \end{split}$$ where $\Omega_{jj} = nm_{j}'m_{j}$ with $m_{j}'$ the $j$-th row of $M^{\text{MPI}} = D^{\text{MPI}}X'(XX')^{-1}$ and $D^{\text{MPI}}$ a diagonal matrix with $d_{j}^{\text{MPI}} = [x_{j}'(XX')^{-1}x_{j}]^{-1}$. A proof is presented in Appendix \[app:prooftheoremfinitevar\]. Appendix \[app:nongaus\] shows that under additional assumptions this result also holds for independent and identically distributed errors $\varepsilon_{i}$. Combining Lemma \[theorem:finitevar\] with yields the central theorem of this paper. \[theorem:main\] Suppose A\[ass:sparsity\]-A\[ass:condnumb\] hold. Let $\hat{\beta}^{c}=My - \left(MX-I_p\right)\hat{\beta}^{\text{init}}$, with $\hat{\beta}^{\text{init}}$ such that $||\hat{\beta}^{\text{init}}-\beta||_{1} = O_{p}\left(s_{0}\sqrt{\log(p)/n}\right)$, and take $M$ as $$M^{\text{MPI}} = D^{\text{MPI}}X'(XX')^{-1},$$ where $D^{\text{MPI}}$ is a diagonal matrix with elements $d^{\text{MPI}}_{j}=n\left[x_{j}'(X X')^{-1}x_{j}\right]^{-1}$. Then, $$\begin{aligned} \sqrt{n}(\hat{\beta}^{c}-\beta) & = Z + o_{p}(1),\\ Z|X &\sim N\left(0,\sigma^2 \Omega\right),\end{aligned}$$ where $\Omega= nM^{MPI}M^{MPI'}$ and $\Omega_{jj} = O_{p}(1)$. This theorem shows that the estimator $\hat{\beta}^{c}$ in is asymptotically unbiased with covariance matrix $\Omega$, and standard errors that decrease at the usual $n^{-1/2}$ rate. Theorem \[theorem:main\] allows for the construction of confidence intervals that are uniformly valid over $j$. Uniformity is guaranteed since the bound on the lasso estimator given in Lemma \[lem:acclasso\] holds uniformly over all sets $S_{0}$ of size $s_{0} = o(\sqrt{n}/\log p)$, see @van2014asymptotically for a discussion. Since the resulting covariance matrix of the estimator is available in closed form, efficient multiple testing procedures as in @buhlmann2013statistical can be employed, together with joint tests on estimated coefficients, as well as confidence intervals around predictions for future values of the dependent variable. Regularized approximate inverse {#sec: distrest_reg} ------------------------------- When the number of variables is of the same order as the number of observations, the concentration of the eigenvalues in Assumption A\[ass:regressor\] might not be very tight. In this case, regularization of the pseudoinverse can increase the accuracy. We therefore analyze two regularization approaches. #### Random least squares {#sec: theoryrls} The key to the behavior of the regularized covariance matrix in repeated least squares, is the projection dimension $k$. The following lemma parallels Lemma \[theorem:biasvanish\] and Lemma \[theorem:finitevar\] for an appropriate choice of the projection dimension. \[theorem:RLS\] Define $M^{\text{RLS}} = D^{\text{RLS}}\text{E}_{R}\left[R(R'X'XR)^{-1}R'\right]X'$ where $D^{\text{RLS}}$ is a diagonal matrix with diagonal elements $ d^{\text{RLS}}_{j}=\left\{\text{E}_{R}[r_{j}'(R'X'XR)^{-1}R']X' x_{j}\right\}^{-1} $, and $R$ a $p\times k$ matrix with normally and independently distributed entries. Choose the projection dimension $k$ as $$k = \left(1-c_{\kappa}\sqrt{(\log p)/n}\right)(n-1),$$ where $c_{k}$ is a positive constant. Then we have $$P\left(||M^{RLS}X-I_{p}||_{\max}\geq a\sqrt{\frac{\log p}{n}}\right) = O\left(p^{-\tilde{c}}\right),$$ with $\tilde{c}$ as in Lemma \[theorem:biasvanish\] with $a$ replaced by $\tilde{a}<a$. Furthermore, for $Z = \sqrt{n}d^{\text{RLS}}_{j}\text{E}[r_{j}(R'X'XR)^{-1}R']X'\varepsilon$, we have $$\begin{split} Z|X&\sim N(0,\sigma^2 \Omega^{RLS}),\\ \Omega^{RLS} &= nM^{RLS}M^{RLS'},\\ \Omega^{RLS}_{jj}& = O_{p}(1). \end{split}$$ The proof of Lemma \[theorem:RLS\] given in Appendix \[app:prooftheoremRLS\] relies on showing that when $k$ is sufficiently close to $n$, the regularized inverse approximates the Moore-Penrose inverse. The results from Section \[sec: distrest\] can then be used to show that regularizing using random least squares does not adversely affect the bias. The proof of Lemma \[theorem:RLS\] also elicits that random least squares is equivalent to a generalized form of ridge regression, where the regularization strength is dependent on the eigenvalues of the regressor matrix $X$. Details on the constant $c_{k}$ are provided in the proof. #### Ridge regularization Because of the relation between the Moore-Penrose pseudoinverse and ridge regularized covariance matrices displayed in , intuition suggests that for a sufficiently small penalty parameter $\lambda$, the results under a Moore-Penrose inverse carry over to a ridge adjusted estimator. The following lemma formalizes this intuition. \[theorem:ridge\] Define $ M^{\text{RID}} = D^{\text{RID}}(X'X+\gamma I_{p})^{-1}X'$, with the elements of the diagonal scaling matrix $D^{\text{RID}}$ equal to $ d^{\text{RID}}_{j}=\left(e_{j}'(X'X + \gamma I_{p})^{-1}X'x_{j}\right)^{-1} $. If the ridge penalty parameter satisfies $\gamma \leq c_{\gamma}p\sqrt{\frac{\log p}{n}}$, where $c_{\gamma}$ is a positive constant, then we have $$P\left(||M^{\text{RID}}X-I_{p}||_{\infty}\geq a\sqrt{\frac{\log p}{n}}\right) = O\left(p^{-\tilde{c}}\right),$$ with $\tilde{a}$ and $\tilde{c}$ as in Lemma \[theorem:RLS\]. Furthermore, for $Z = \sqrt{n}d^{\text{RID}}_{j}(X'X + \gamma I_{p})^{-1}X'\varepsilon$, we have $$\begin{split} Z|X&\sim N(0,\sigma^2 \Omega^{RID}),\\ \Omega^{RID} &= nM^{RID}M^{RID'},\\ \Omega^{RID}_{jj}& = O_{p}(1). \end{split}$$ A proof is provided in Appendix \[app:prooftheoremridge\], which also gives a more detailed description of the constant $c_{\gamma}$. #### Inference using a regularized approximate inverse Using Lemma \[theorem:RLS\] and Lemma \[theorem:ridge\], we arrive at the following theorem for the regularized estimators. \[theorem:main2\] Suppose A\[ass:sparsity\]-A\[ass:condnumb\] hold. Let $\hat{\beta}^{c}=My - \left(MX-I_p\right)\hat{\beta}^{\text{init}}$, with $\hat{\beta}^{\text{init}}$ such that $||\hat{\beta}^{\text{init}}-\beta||_{1} = O_{p}\left(s_{0}\sqrt{\log(p)/n}\right)$, and take $M$ as either $ M^{\text{RLS}} = D^{\text{RLS}}\text{E}_R\left[R(R'X'XR)^{-1}R'\right]X'$ or $M^{\text{RID}} = D^{\text{RID}}(X'X+\gamma^{*} I_p)^{-1}X'$, where the elements of the diagonal matrices $D$ are defined in Lemma \[theorem:RLS\] and Lemma \[theorem:ridge\], $R$ is a $p\times k^{*}$ matrix with independent standard normal entries, $k^{*}=k$ as in Lemma \[theorem:RLS\], and $\gamma^{*}=\gamma$ as in Lemma \[theorem:ridge\]. Then, $$\begin{aligned} \sqrt{n}(\hat{\beta}^{c}-\beta) & = Z + o_{p}(1),\\ Z|X &\sim N\left(0,\sigma^2 \Omega\right), \\ \Omega &= nMM',\\ \Omega_{jj}&= O_{p}(1).\end{aligned}$$ This theorem follows directly from Lemma \[theorem:RLS\] and Lemma \[theorem:ridge\]. It confirms that when $k$ is close to $n$ and $\gamma$ is sufficiently small, the estimator in is asymptotically unbiased with covariance matrix $\Omega$, and standard errors that decrease at the usual $n^{-1/2}$ rate. The reason one would opt for the regularized variants despite the additional tuning parameters is provided by the following theorem. Here we compare the variance of $Z$ in equation for the different estimators. \[theorem:power\] Denote the variance of the estimator $\hat{\beta}^{c}_{j}$ under a diagonal scaling matrix $D$ by $\Omega_{jj}(D)$. For the choice of $k$ as in Lemma \[theorem:RLS\], or $\gamma$ as in Lemma \[theorem:ridge\], we have $$\label{eq:th4} \Omega_{jj}(D)^{RLS} - \Omega_{jj}(D)^{MPI}\leq 0, \qquad \Omega_{jj}(D)^{RID}-\Omega_{jj}(D)^{MPI}\leq 0.$$ The proof is given in Appendix \[app:prooftheorempower\]. Note that Theorem \[theorem:power\] requires the regularized estimator and the estimator based on the Moore-Penrose pseudoinverse to use the same diagonal scaling matrix. Using $D^{\text{MPI}}$ for the Moore-Penrose inverse, $D^{\text{RLS}}$ for the repeated least squares estimator, and $D^{\text{RID}}$ for the ridge regularized inverse, does not yield an ordering in terms of power. However, in all cases we have encountered, the inequality in Theorem \[theorem:power\] is satisfied when using the diagonal matrix specific to the estimator under consideration. This is also evident from the Monte Carlo results in Section \[sec: mc\]. Consistency {#sec: consist} ----------- Although our focus in this paper is on the construction of confidence intervals, the estimator $\hat{\beta}^{c}$ can be shown to be consistent when we restrict the growth rate of the number of variables relative to the number of observations. \[ass:consistency\] The number of variables grows near exponentially with the number of observations, i.e. $$\frac{\log p}{n} = o(1).$$ Since $Z_{i}$ is (asymptotically) normal, we have that $\max_{i=1,\ldots,j}|Z_{i}| = O_{p}(\sqrt{\log p})$. Since $\hat{\beta}^{c} = \beta + \frac{1}{\sqrt{n}}(\Delta + Z)$, Assumption A\[ass:consistency\] then guarantees that $\lim_{n\rightarrow\infty}\hat{\beta}^c= \beta$. If one is only interested in consistency, then Assumption A\[ass:regressor\] can potentially be relaxed. In that case the bias is not required to be of lower order compared to the variance. [^1]: Department of Economics, Econometrics and Finance, University of Groningen, Nettelbosje 2, 9747 AE Groningen, The Netherlands, e-mail: [^2]: Econometric Institute, Erasmus University Rotterdam, P.O. Box 1738, NL-3000 DR Rotterdam, The Netherlands, e-mail: We would like to thank Paul Bekker, Patrick Groenen, Trevor Hastie, Christiaan Heij, Richard Paap, Andreas Pick, and participants of the workshop on statistical learning and econometrics at the Erasmus University Rotterdam for helpful comments.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Polynomial chaos is a powerful technique for propagating uncertainty through ordinary and partial differential equations. Random variables are expanded in terms of orthogonal polynomials and differential equations are derived for the expansion coefficients. Here we study the structure and dynamics of these differential equations when the original system has Hamiltonian structure, has multiple time scales, or displays chaotic dynamics. In particular, we prove that the differential equations for the expansion coefficients in generalized polynomial chaos expansions of Hamiltonian systems retain the Hamiltonian structure relative to the ensemble average Hamiltonian. We connect this with the volume-preserving property of Hamiltonian flows to show that, for an oscillator with uncertain frequency, a finite expansion must fail at long times, regardless of the order of the expansion. Also, using a two-time scale forced nonlinear oscillator, we show that a polynomial chaos expansion of the time-averaged equations captures uncertainty in the slow evolution of the Poincar[é]{} section of the system and that, as the time scale separation increases, the computational advantage of this procedure increases. Finally, using the forced Duffing oscillator as an example, we demonstrate that when the original dynamical system displays chaotic dynamics, the resulting dynamical system from polynomial chaos also displays chaotic dynamics, limiting its applicability.' author: - 'Jos[é]{} Miguel Pasini$^\dagger$' - 'Tuhin Sahai$^\dagger$' bibliography: - 'Hamiltonian.bib' date: | $^\dagger$United Technologies Research Center\ 411 Silver Lane, East Hartford, CT 06108, USA\ {pasinijm, sahait}@utrc.utc.com title: 'Polynomial chaos based uncertainty quantification in Hamiltonian, multi-time scale, and chaotic systems' --- Introduction ============ Uncertainty quantification techniques allow one to quantify output variability in the presence of parametric uncertainty. Typically, the moments of the output distributions are computed using sampling methods such as Monte Carlo [@McQMc], Quasi-Monte Carlo [@QMC], and importance sampling [@Importance_Sampling]. Non-sampling approaches include response surface [@Response1:book; @Response2:book] and polynomial chaos based methods [@Wiener]. Depending on the problem, different methods are applicable/appropriate in different scenarios. Polynomial chaos based techniques for propagating uncertainty have been used on a multitude of applications such as aeroelastic modeling [@Allen2009], transport in heterogeneous media [@Ghanem1998], Ising models [@TuhinPoly], switching systems [@hybrid_uq], combustion [@Najm2009], fluid flow [@Xiu2003], and materials models [@Cit:Materials], to name a few. Here we study the properties and utility of using polynomial chaos expansions to propagate uncertainty through systems that have either Hamiltonian structure, multiple scales, or display chaos. We point out that polynomial chaos [@Wiener] and chaos theory [@Cit:Stro] are unrelated areas. Originally proposed by Nobert Wiener [@Wiener] in 1938 (prior to the development of chaos theory—hence the unfortunate usage of the term *chaos*), polynomial chaos expansions are a popular method for propagating uncertainty through low dimensional systems with smooth dynamics. They rely on expanding random variables in terms of orthogonal basis functions [@BeyondWienerAskey]. Note that the orthogonal polynomials are chosen such that they are orthogonal to one another with respect to the prior distribution on the uncertain parameters [@BeyondWienerAskey]. For example, if the underlying distribution on the uncertain parameters is Gaussian, then the associated orthogonal polynomials are Hermite polynomials [@Orthopoly:book]. Similarly, if the underlying prior distribution is uniform, the associated orthogonal polynomials are Legendre [@Orthopoly:book]. In general, one can construct orthogonal polynomials for arbitrary distributions [@BeyondWienerAskey]. The advantage of polynomial chaos based techniques is that they provide exponential convergence for smooth processes with finite variance [@cameron1947orthogonal]. Chaos, on the other hand, refers to “aperiodic long-term behavior in *deterministic* systems that exhibits sensitive dependence on initial conditions” [@Cit:Stro]. Chaos theory has been applied to a wide variety of applications such as fluid turbulence [@Cit:turb], celestial dynamics [@Cit:solar], and weather modeling [@Cit:Lorenz]. It is important to point out that although the dynamics has sensitive dependence on initial conditions, it is inherently deterministic. In other words, no associated parametric uncertainty is required to observe chaos. In this work, we present three new results. In the first part, we show that the dynamical systems that one gets by applying polynomial chaos expansions to Hamiltonian systems with uncertain parameters are also Hamiltonian. To do this, we first perform a polynomial chaos expansion of the generalized coordinates and conjugate momenta and find the evolution equations for the coefficients. We consider the expansion coefficients of the generalized coordinates as a new, larger set of “uncertain” generalized coordinates. By considering the averaged Hamiltonian (over parameter space) as a function of the expansion coefficients, we show that, for each of these new generalized coordinates, the coefficient in the expansion of the corresponding conjugate momentum and to the corresponding order is itself the conjugate momentum relative to the average Hamiltonian, thus demonstrating that the Hamiltonian structure in the derived differential equations is preserved. We then connect this result with the volume-preserving property of Hamiltonian flows to show that any finite polynomial chaos expansion of a harmonic oscillator with uncertain frequency must fail at long times, regardless of how many terms are kept. In the second part, we demonstrate the application of polynomial chaos to systems with multiple time scales using perturbation theory [@Cit:Rand-notes]. We demonstrate how the uncertain parameters influence the averaged dynamics of the dynamical system. In particular, we show that uncertainty can be propagated through the averaged equations instead of through the original equations, thus avoiding the computational burden of simulating a stiff system. In the third part, we apply polynomial chaos to a chaotic dynamical system (forced Duffing oscillator) [@Gucken:book] and demonstrate that the resulting equations for the coefficients are also chaotic. We then show that chaotic dynamics significantly reduce the efficacy of polynomial chaos expansions at propagating uncertainty. Introduction to polynomial chaos ================================ Starting with a complete probability space $\Gamma$ given by $(\Omega, \mathcal{F},\mathbb{P})$, where $\Omega$ is the sample space, $\mathcal{F}$ is the $\sigma$-algebra on $\Omega$ and $\mathbb{P}$ is a probability measure, let $L_{2}(\Gamma,X)$ denote the Hilbert space of square-integrable, $\mathcal{F}$-measurable, $X$-valued random elements. Then one can, in general, define a polynomial chaos basis $\{\psi_k(\lambda(\omega))\}$, where $\lambda(\omega)$ is a random vector, $\omega \in \Omega$, and $k = (k_1,k_2,\dots)$ is a vector of non-negative indices. We denote the probability density function of the random vector $\lambda$ by $\rho(\lambda)$. Generalized polynomial chaos (gPC) [@BeyondWienerAskey] provides a framework for representing second-order stochastic processes $\kappa\in L_{2}(\Gamma,X)$ for arbitrary distributions of $\lambda$ by the following expansion: $$\kappa(\lambda) = \displaystyle\sum_{|k|=0}^{\infty}a_k \psi_k (\lambda), \label{eq:expan1}$$ where $|k| = \sum_{i} k_{i}$ is the sum of the indices of $k$ and $\psi_k (\lambda)$ are orthonormal polynomials on $\Gamma$ with respect to $\rho(\lambda)$. Restricting our formalism to $\mathbb{R}^{n}$ (relevant for this work) we get the orthonormality is given by $$\displaystyle\int_{\mathbb{R}^n} \rho(\lambda)\psi_i(\lambda)\psi_k(\lambda)d\lambda = \delta_{ik}, \label{eq:ortho}$$ where $\delta_{ik}$ is the Kronecker delta product. Depending on $\rho(\lambda)$ one can generate an appropriate orthogonal basis for representing $\kappa(\lambda)$. As mentioned earlier, if $\rho$ is Gaussian, then the appropriate polynomial chaos basis is the set of Hermite polynomials; if $\rho$ is the uniform distribution, then the basis is the set of Legendre polynomials. For details on the correspondence between distributions and polynomials see [@PolyReview; @Ogura]. A framework to generate polynomials for arbitrary distributions has been developed in [@BeyondWienerAskey]. The advantage of using polynomial chaos is that it provides exponential convergence in smooth processes [@cameron1947orthogonal]. However, the approach suffers from the curse of dimensionality, making them infeasible for problems with more than a handful of parameters. To mitigate the curse of dimensionality, sparse grid techniques [@Webster2007; @Nobile2008; @Zabaras2008], iterative methods [@surana_uq; @Sahai2012; @Klus2011], regression based algorithms [@blatman2010adaptive; @blatman2011adaptive], hierarchical methods [@ma2009adaptive], dimensionality reduction based techniques [@ma2011kernel; @marzouk2009dimensionality] have been developed. In practice, the expansion in Eqn. \[eq:expan1\] is truncated at a particular order, say, $r$. One can then use Galerkin projections to obtain a set of differential equations for the coefficients $a_k$ in Eqn. \[eq:expan1\] [@BeyondWienerAskey]. Typically, low order truncations are found to capture the uncertainty in smooth systems [@cameron1947orthogonal]. Polynomial chaos based uncertainty quantification in Hamiltonian systems ======================================================================== Consider a system described by the Hamiltonian $H(q,p;\lambda)$, where $\lambda$ is a vector of uncertain parameters with probability density $\rho(\lambda)$. The generalized coordinates and momenta $q_i$ and $p_i$ ($i=1,\ldots,N$) satisfy Hamilton’s equations. $$\begin{aligned} \dot{q}_i &= \frac{\partial H}{\partial p_i}, \\ \dot{p}_i &= -\frac{\partial H}{\partial q_i}.\end{aligned}$$ The generalized polynomial chaos (gPC) expansion of the coordinates and momenta is, $$\begin{aligned} q_i(t;\lambda) &= \sum_k Q_{ik}(t) \psi_k(\lambda), \\ p_i(t;\lambda) &= \sum_k P_{ik}(t) \psi_k(\lambda),\end{aligned}$$ where the $\psi_k$ form an orthonormal basis with respect to the density $\rho$ (see Eq. \[eq:ortho\]). The gPC coefficients $Q_{ik}$ and $P_{ik}$ follow deterministic equations, obtained by projecting the equations of motion along $\psi_s$ $$\begin{aligned} \int \dot{q}_i \psi_s \rho d\lambda &= \int \frac{\partial H}{\partial p_i} \psi_s \rho d\lambda, \\ \int \dot{p}_i \psi_s \rho d\lambda &= -\int \frac{\partial H}{\partial q_i} \psi_s \rho d\lambda.\end{aligned}$$ Inserting the gPC expansions and using the orthonormality condition  we obtain, $$\begin{aligned} \dot{Q}_{is} &= \int \frac{\partial H}{\partial p_i} \psi_s \rho d\lambda, \\ \dot{P}_{is} &= -\int \frac{\partial H}{\partial q_i} \psi_s \rho d\lambda.\end{aligned}$$ Let us define the average Hamiltonian $\hat{H}$ $$\hat{H} = \int H \rho d\lambda.$$ By using the gPC expansion of $q$ and $p$ we can consider $\hat{H}$ as a function of $Q$ and $P$, where $Q$ and $P$ denote the sets of coefficients $Q_{ik}$ and $P_{ik}$, respectively. \[thm:hamiltonian\_structure\] The gPC expansion coefficients $\{Q,P\}$ together with $\hat{H}(Q,P)$ form a Hamiltonian system, with the corresponding expansion coefficients $P_{ik}$ as conjugate momenta to $Q_{ik}$. In other words, $$\begin{aligned} \dot{Q}_{ik} &= \frac{\partial \hat{H}}{\partial P_{ik}}, \label{eq:Qdot} \\ \dot{P}_{ik} &= -\frac{\partial \hat{H}}{\partial Q_{ik}}. \label{eq:Pdot}\end{aligned}$$ We start with the right-hand side of Eq. : $$\begin{aligned} \frac{\partial \hat{H}}{\partial P_{ik}} &= \int \sum_s \left( \frac{\partial H}{\partial q_s} \frac{\partial q_s}{\partial P_{ik}} + \frac{\partial H}{\partial p_s} \frac{\partial p_s}{\partial P_{ik}} \right) \rho d\lambda, \\ &= \int \sum_s \sum_r \frac{\partial H}{\partial p_s} \delta_{is} \delta_{kr} \psi_r \rho d\lambda, \\ &= \int \frac{\partial H}{\partial p_i} \psi_k \rho d\lambda, \\ &= \dot{Q}_{ik}.\end{aligned}$$ Similarly for Eq. : $$\begin{aligned} \frac{\partial \hat{H}}{\partial Q_{ik}} &= \int \sum_s \left( \frac{\partial H}{\partial q_s} \frac{\partial q_s}{\partial Q_{ik}} + \frac{\partial H}{\partial p_s} \frac{\partial p_s}{\partial Q_{ik}} \right) \rho d\lambda, \\ &= \int \sum_s \sum_r \frac{\partial H}{\partial q_s} \delta_{is} \delta_{kr} \psi_r \rho d\lambda, \\ &= \int \frac{\partial H}{\partial q_i} \psi_k \rho d\lambda, \\ &= -\dot{P}_{ik}.\end{aligned}$$ Note that the proof depends only on the *form* of the expansion and does not require that the expansion be complete. In other words, the coefficients of a *truncated* expansion will also form a (finite) Hamiltonian system relative to the average Hamiltonian when expressed as a function of the truncated expansion coefficients. Hence, polynomial chaos expansions when applied to Hamiltonian systems are also Hamiltonian. This result is not only interesting but also has practical implications. In particular, if the underlying system is Hamiltonian and one desires to propagate uncertainty using polynomial chaos, symplectic integrators [@Cit:geom] will be needed to maintain numerical accuracy for long times. We now illustrate the preservation of Hamiltonian structure on the Duffing oscillator with parametric uncertainty. Example: Duffing oscillator --------------------------- To provide an example of a Hamiltonian system with uncertainty, we consider the Duffing oscillator, $$\begin{aligned} \ddot{q} + \lambda q + q^3 = 0, \label{eq:duffing_nf}\end{aligned}$$ where $\lambda$ is a normally distributed uncertain parameter with mean $\mu(\lambda) = \lambda_0 = -1.0$ and standard deviation $\sigma(\lambda) = 0.1$. This system has the following Hamiltonian: $$\begin{aligned} H = \frac{1}{2} p^2 + \frac{\lambda}{2}q^2 + \frac{1}{4} q^4, \label{eq:hamilton}\end{aligned}$$ where $p=\dot q$. The phase portrait of the undamped Duffing oscillator (in Eq. \[eq:duffing\_nf\]) is shown in Fig. \[Fig:hamiltphase\]. One can observe the Hamiltonian structure evident in phase space. In particular, the system has two centers at located at $(-1,0)$ and $(1,0)$. The equilibrium at $(0,0)$ is a saddle point. For a detailed discussion on the characteristics of the Duffing oscillator and its volume preserving flow we point the reader to [@Gucken:book]. The resulting dynamical system is of the form, $$\begin{pmatrix} \dot q\\ \dot p \end{pmatrix}= \begin{pmatrix} p\\ - \lambda q - q^3 \end{pmatrix}. \label{eq:duffingsimple}$$ Assuming that $\lambda$ is an uncertain parameter, we now perform a polynomial chaos expansion [@BeyondWienerAskey] given by, $$\begin{split} q(t;\lambda) &= \displaystyle\sum_{i=0}^{r}Q_{i}(t)\psi_{i}(\lambda),\\ p(t;\lambda) &= \displaystyle\sum_{i=0}^{r}P_{i}(t)\psi_{i}(\lambda). \end{split} \label{eq:expx}$$ By substituting the above expansion, for $r=1$, into Eq. \[eq:duffingsimple\] and imposing orthogonality constraints we get the following set of equations, $$\begin{pmatrix} \dot Q_{0}\\ \dot P_{0}\\ \dot Q_{1}\\ \dot P_{1} \end{pmatrix}=\begin{pmatrix} P_{0}\\ - \lambda_{0}Q_{0} - \sigma Q_{1} - (Q_{0}^3 + 3Q_{0}Q_{1}^2)\\ P_{1}\\ - \lambda_{0}Q_{1} - \sigma Q_{0} - 3(Q_{1}^3 + Q_{0}^2Q_{1}) \end{pmatrix}. \label{eq:coeffs}$$ It is easy to check that the Hamiltonian for the above system of equations is given by $$\begin{split} H_{pc} &= \frac{1}{2}P_{0}^{2} + \frac{1}{2}P_{1}^{2} + \frac{\lambda_0}{2}(Q_{0}^{2} + Q_{1}^{2}) + \sigma Q_{0}Q_{1} \\ &\quad + \frac{3}{2}Q_{0}^{2}Q_{1}^2 + \frac{1}{4} Q_{0}^4 + \frac{3}{4}Q_{1}^4. \end{split} \label{eq:Hpc}$$ Similar Hamiltonians can be constructed for higher order expansions (arbitrary $r$) in the Duffing oscillator as well as for other Hamiltonian systems, such as the double pendulum and $N$ bodies interacting through Newton’s law of gravitation. ![Phase portrait of the Duffing oscillator.\[Fig:hamiltphase\]](Hamiltonian_struct.eps) Harmonic oscillator with uncertain frequency -------------------------------------------- We have proved that the PC equations have Hamiltonian structure. We now combine this with other results in Hamiltonian theory to show that, for certain problems, the uncertainty cannot be captured by a finite PC expansion for long times, *regardless of the order of the expansion*. Specifically, we focus on a harmonic oscillator with uncertain frequency and certain initial conditions: $$\ddot{q} + \omega^2 q = 0 \qquad q(0) = 1 \qquad \dot{q}(0) = 0. \label{eq:harmonic_oscillator}$$ We choose $\omega$ uniformly distributed in $(\omega_1, \omega_2)$ and define $\omega \equiv \omega_0 + \alpha \lambda$, with $\lambda \sim U(-1,1)$, so $\alpha = (\omega_2 - \omega_1)/2$ is a measure of the magnitude of the uncertainty in frequency. The coefficients of a finite PC expansion of the true solution for this system converge to zero at long times. The solution for this system is $$q(t;\lambda) = \cos (\omega_0 + \alpha \lambda) t \label{eq:harmonic_solution}$$ and the PC expansion in this case is $$\begin{aligned} q(t;\lambda) &= \sum_{k=0}^r Q_k(t) \psi_k(\lambda) \\ p(t;\lambda) &= \sum_{k=0}^r P_k(t) \psi_k(\lambda),\end{aligned}$$ where $\psi_k(\lambda) = \sqrt{2k+1} \,{\cal P}_k(\lambda)$, and ${\cal P}_k$ is the usual Legendre polynomial of order $k$. $\{\psi_k\}$ is a set of orthonormal polynomials in $[-1,1]$ with respect to the density $\rho(\lambda) = 1/2$. Explicitly, $$\begin{aligned} \psi_k(\lambda) &= \sum_{\ell=0}^k B_{k\ell} \lambda^\ell \\ B_{k\ell} &= \sqrt{2k+1} \, 2^k \binom{k}{\ell} \binom{(k+\ell-1)/2}{k}.\end{aligned}$$ We project the PC expansion onto this basis: $$\begin{aligned} \int_{-1}^1 q(t;\lambda) \rho(\lambda) d\lambda &= \int_{-1}^1 \sum_{k=0}^r Q_k(t) \psi_k(\lambda) \rho(\lambda) d\lambda \\ \int_{-1}^1 p(t;\lambda) \rho(\lambda) d\lambda &= \int_{-1}^1 \sum_{k=0}^r P_k(t) \psi_k(\lambda) \rho(\lambda) d\lambda.\end{aligned}$$ If we had an infinite-order expansion ($r=\infty$), the sum would be a series, and to exchange the integral and the series we would require uniform convergence of the sum. For any *finite* sum, however, we can do the exchange and obtain these equations for the coefficients: $$\begin{aligned} Q_k(t) &= \int_{-1}^1 \cos[(\omega_0 + \alpha \lambda) t] \psi_k(\lambda) \rho(\lambda) d\lambda \\ P_k(t) &= \int_{-1}^1 (-\alpha \lambda) \sin[(\omega_0 + \alpha \lambda) t] \psi_k(\lambda) \rho(\lambda) d\lambda.\end{aligned}$$ We now show that these coefficients go to zero as $t\to\infty$. Indeed, $$Q_k(t) = \frac{1}{2} \sum_{\ell=0}^k B_{k\ell} I_\ell,$$ where we can integrate by parts twice to obtain a recurrence formula: $$\begin{aligned} I_\ell &= \int_{-1}^1 \lambda^\ell cos(\omega_0 + \alpha \lambda)t \, d\lambda \\ &= \frac{1}{\alpha t} [\sin\omega_2 t - (-1)^\ell \sin \omega_1 t] \\ &+ \frac{\ell}{(\alpha t)^2} [\cos \omega_2 t - (-1)^{\ell-1} \cos \omega_1 t ] - \frac{\ell(\ell-1)}{(\alpha t)^2} I_{\ell-2}.\end{aligned}$$ Since both $I_0$ and $I_1$ go to zero as $t\to\infty$, all $Q_k(t) \to 0$ as $t\to\infty$. Similarly, we can prove that all $P_k(t) \to 0$ as $t\to\infty$. With all the coefficients in any finite expansion of the true solution converging to zero as $t\to\infty$, the volume of this flow decreases. On the other hand, the solution of the PC equations obtained must preserve volume, so those coefficients cannot go to zero to match the behavior of the true solution without violating Liouville’s theorem [@Liboff1998]. In other words, Liouville’s theorem prevents any finite PC expansion from representing the true solution of this system at long times. Polynomial chaos based uncertainty quantification in systems with multiple time scales ====================================================================================== Systems with multiple time scales are prevalent in a wide variety of applications related to smart grids [@He2011; @Parpas2011; @Susuki2011], building systems [@surana_uq], and micromechanical oscillators [@Tuhin_mems1; @Tuhin_mems2; @Tuhin_mems3], to name a few. Simulating these systems is challenging due to their inherent stiffness [@Gear1971]. The method of multiple scales (or averaging) is a very popular approach for simulating such systems. The approach typically involves perturbing off a dynamical system whose solution can be computed in closed form [@Cit:Rand-notes]. Note that this approach is applicable only in the scenario that the perturbation is small $O(\epsilon)$. The method of multiple scales captures the dynamics of the system on an $n-1$ dimensional section transversal to the flow, also known as the Poincaré section [@Gucken:book]. For a detailed discussion on the method of multiple scales or averaging theory, see [@Cit:Rand-notes; @Gucken:book]. To the best of our knowledge, no attempt has been made to extend polynomial chaos based methods to systems with multiple time scales using the method of multiple scales. Here we apply polynomial chaos to the two-time scale system given below, $$\begin{aligned} \ddot q + q +\epsilon\delta\dot q + \epsilon\beta q^3 = \epsilon\gamma \cos \omega t, \label{eq:twotime_eq}\end{aligned}$$ where $\epsilon\delta$ is the system damping, $1$ and $\epsilon\beta$ are the linear and nonlinear stiffnesses respectively, and $\epsilon\gamma$ and $\omega$ are the forcing amplitude and frequency respectively. Note that in the above system, we assume that $\epsilon$ is a small parameter (i.e. $\epsilon\ll 1$). We assume that $\gamma = \gamma_{0} + \sigma(\gamma)\eta$ is an uncertain parameter , where $\gamma_{0} = 1.0$ is the mean of $\gamma$ and $\sigma(\gamma) = 0.1$ is its standard deviation ($\eta$ is a normal random variable with zero mean and unit variance). Using the two time scales as $\xi=\omega t$ and $\chi = \epsilon t$, one can derive the averaged equations for the system [@Cit:Rand-notes; @Gucken:book]. This is done by substituting $\frac{d}{dt} = \omega\frac{\partial}{\partial\xi} + \epsilon\frac{\partial}{\partial\chi}$, $\frac{d^{2}}{dt^{2}} = \omega^{2}\frac{\partial^{2}}{\partial\xi^{2}} + 2\omega\epsilon\frac{\partial^{2}}{\partial\xi\partial\chi} + \epsilon^{2}\frac{\partial^{2}}{\partial\chi^{2}}$, and $q(\xi,\chi) = q_{0}(\xi,\chi) + \epsilon q_{1}(\xi,\chi) + \hdots$ in Eqn. \[eq:twotime\_eq\]. Collecting terms, we obtain $$\begin{aligned} O(1): \frac{\partial^{2}q_{0}}{\partial\xi^{2}} + q_{0} &=0,\label{eq:o1}\\ O(\epsilon): \frac{\partial^{2}q_{1}}{\partial\xi^{2}} + q_{1} &= -2\frac{\partial^{2}q_{0}}{\partial\xi\partial\chi} - \delta\frac{\partial q_{0}}{\partial\xi} - \beta q_{0}^{3} + \gamma\cos\xi \label{eq:o2}.\end{aligned}$$ The solution to Eqn. \[eq:o1\], is $q_{0}(\xi,\chi) = A(\chi)\cos\xi + B(\chi)\sin\xi$. Substituting the solution into Eqn. \[eq:o2\] and imposing that there be no secular terms [@Cit:Rand-notes; @Gucken:book] yields the averaged equations $$\begin{split} 2\frac{\partial A}{\partial\chi} + \delta A -\frac{3}{4}\beta B(A^{2} + B^{2}) &= 0, \\ 2\frac{\partial B}{\partial\chi} + \delta B +\frac{3}{4}\beta A(A^{2} + B^{2}) &=\gamma. \end{split} \label{eq:twotime_averaged}$$ Note that the above dynamical system captures the dynamics on the Poincaré section of the original system [@Gucken:book]. From here on we take $\beta = 1$. We will also focus on the deterministic initial condition $q=2$, $\dot{q} = 0$, so $A(0) = 2$ and $B(0) = 0$. We now apply a polynomial chaos expansion to Eq. \[eq:twotime\_averaged\], to first order: $$\begin{split} A(\chi, \eta) &= a_0(\chi) H_0(\eta) + a_1(\chi) H_1(\eta) \\ B(\chi, \eta) &= b_0(\chi) H_0(\eta) + b_1(\chi) H_1(\eta), \end{split} \label{eq:PC_expansion_of_slow_equations}$$ where $H_0(\eta) = 1$ and $H_1(\eta) = \eta$ are the first two probabilist’s Hermite polynomials. Substituting Eq. \[eq:PC\_expansion\_of\_slow\_equations\] into Eq. \[eq:twotime\_averaged\] gives $$\begin{split} 2a_0' &= - \delta a_0 + \frac{3}{4} b_0 (a_0^2 + b_0^2) \\ 2b_0' &= - \delta b_0 - \frac{3}{4} a_0 (a_0^2 + b_0^2) + \gamma_0 \\ 2a_1' &= - \delta a_1 + \frac{3}{4} (2a_0 b_0 a_1 + a_0^2 b_1 + 3b_0^2 b_1) \\ 2b_1' &= - \delta b_1 - \frac{3}{4} (3a_0^2 a_1 + b_0^2 a_1 + 2a_0 b_0 b_1) + \sigma, \end{split} \label{eq:twotime_averaged_PC}$$ with initial condition $(a_0,b_0,a_1,b_1) = (2,0,0,0)$. For comparison purposes, we also do an equivalent polynomial chaos expansion of the original two-time equations, defining $x = q$, $y = \dot{q}$, and doing a first order expansion $$\begin{aligned} x(t, \eta) &= x_0(t) H_0(\eta) + x_1(t) H_1(\eta) \\ y(t, \eta) &= y_0(t) H_0(\eta) + y_1(t) H_1(\eta).\end{aligned}$$ The resulting system is $$\begin{split} \dot{x}_0 &= y_0 \\ \dot{y}_0 &= -x_0 - \epsilon \delta y_0 - \epsilon x_0^3 + \epsilon \gamma_0 \cos \omega t \\ \dot{x}_1 &= y_1 \\ \dot{y}_1 &= -x_1 - \epsilon \delta y_1 - 3\epsilon x_0^2 x_1 + \epsilon \sigma \cos \omega t, \end{split} \label{eq:twotime_PC}$$ with initial condition $(x_0,y_0,x_1,y_1) = (2,0,0,0)$. In what follows, we choose $\omega=1$ and, more importantly, $\delta = 0$. This is done in order to avoid having a system in which the dissipation artificially reduces the overall error. Figure \[Fig:TwoTime\_comparison\] shows the error in mean and standard deviation of both PC expansions compared with Monte Carlo simulations with $10^3$ samples of the original two-time system, evaluated at the Poincaré sections where the forcing is maximal: $t = 2\pi n$ ($n = 0, 1, 2, \ldots$). Solutions of both PC expansions and the Monte Carlo trajectories of the original system were obtained using Matlab’s ode45 solver with relative tolerance of $10^{-6}$. The error has two sources: the time averaging and the truncation of the polynomial chaos expansion. As we decrease $\epsilon$, the error from time averaging decreases and the main source of error becomes the truncation of the polynomial expansion. As $\epsilon$ decreases, the two expansions yield increasingly similar results, but the PC expansion of the original equation becomes more expensive to compute, scaling as $1/\epsilon$, because the solver needs to trace each fast oscillation, even if we’re only interested in the slow evolution of the Poincaré section. Figure \[Fig:TwoTime\_function\_evaluations\] shows the number of function evaluations required by this expansion as $\epsilon$ decreases (solid line) compared with the $\epsilon$-independent behavior of the averaged PC. ![Absolute deviation of the polynomial chaos expansions on the averaged equations (solid line) and on the original equations (dashed line) of the two-time oscillator. The reference is a Monte Carlo simulation with 1000 samples.\[Fig:TwoTime\_comparison\]](Twotime_PC_error_delta0_epsilon0.1_Nsamples1000.eps "fig:"){width="50.00000%"}![Absolute deviation of the polynomial chaos expansions on the averaged equations (solid line) and on the original equations (dashed line) of the two-time oscillator. The reference is a Monte Carlo simulation with 1000 samples.\[Fig:TwoTime\_comparison\]](Twotime_PC_error_delta0_epsilon0.01_Nsamples1000.eps "fig:"){width="50.00000%"} ![Function evaluations required to solve the polynomial chaos expansion of the original two-time system as a function of the time scale separation parameter $\epsilon$. The solution was obtained using Matlab’s ode45 with relative tolerance $10^{-6}$. The dashed line shows the number of function evaluations required by the polynomial expansion on the averaged equations, which is independent of $\epsilon$. \[Fig:TwoTime\_function\_evaluations\]](Twotime_PC_function_evaluations_vs_epsilon_delta0.eps){width="50.00000%"} Polynomial chaos based uncertainty quantification in chaotic systems ==================================================================== We now demonstrate that dynamics of the coefficients of the polynomial chaos expansions can be chaotic, if the underlying dynamical system is chaotic. We will then show that if the underlying system is chaotic, the applicability of polynomial chaos is significantly reduced. While this is a natural result, it is not obvious. PC aims to capture the moments of the output distribution and not individual trajectories of the underlying dynamical system. Since mixing introduces averaging that could, in principle, smooth out the moments and allow them to be captured accurately, it is not obvious that chaos would necessarily make predictions worse. For this demonstration, we pick the forced Duffing oscillator [@Gucken:book] given by $$\ddot q + \delta \dot q + \lambda q + q^3 = \gamma \cos \omega t, \label{eq:duffing}$$ where $\delta = 0.2$, $\gamma = 0.3$, $\omega = 1.0$, and $\lambda = -1.0$. Note that the above equation (Eq. \[eq:duffing\]) is the same as Eq. \[eq:duffing\_nf\] with the addition of damping and forcing terms. We can write Eq. \[eq:duffing\] in the form $$\begin{pmatrix} \dot q\\ \dot p \end{pmatrix}= \begin{pmatrix} p\\ - \delta p - \lambda q - q^3 + \gamma \cos\omega t \end{pmatrix}. \label{eq:duffing2d}$$ The dynamics of the above system have been studied extensively (see, e.g., [@Gucken:book]). For the forced Duffing oscillator, the Poincaré section is given by taking “snapshots” of the system at phase $\phi = 0$, where $\phi = (\omega t\, \mbox{mod}\, 2\pi)$. The intersection of a single trajectory with the Poincaré section can be seen in Fig. \[Fig:ForcedDuffing\], starting from the initial condition $(q,p)=(1,0)$. The dynamics of the forced Duffing oscillator (at the parameter values given above) is well known to be chaotic [@Gucken:book]. In fact, one can numerically compute Lyapunov exponents ($\Xi$) [@Cit:Stro] for the above system and show that they are positive. Note that $\Xi>0$ is considered to be the signature of a chaotic system since it implies that the system response is sensitive to initial conditions. We find that the nominal system gives $\Xi \approx 0.93$, hence (numerically) implying the existence of chaos. ![Poincaré section of the forced Duffing oscillator with damping at phase $\phi=0$. The oscillator displays chaotic dynamics and the attractor above displays the stretching and folding properties of chaos [@Gucken:book]. \[Fig:ForcedDuffing\]](Dynamics_Forced_Duffing.eps) Let us now assume that $\lambda$ is normally distributed. Let $\lambda = \lambda_{0} + \sigma\eta$, where $\lambda_{0} = -1.0$ is the mean of $\lambda$ and $\sigma = 0.1$ is its standard deviation. Is is easy to see that $\eta$ will now become a normally distributed random variable with zero mean and unit standard deviation. Since $\eta$ is normally distributed, we use Hermite polynomials in our expansion [@Wiener]. In Eq. \[eq:duffing2d\] we use the expansion in Eq. \[eq:expx\]. Truncating the expansion at $r=1$ gives the following set of differential equations: $$\begin{pmatrix} \dot Q_{0}\\ \dot P_{0}\\ \dot Q_{1}\\ \dot P_{1} \end{pmatrix}=\begin{pmatrix} P_{0}\\ - \delta P_{0} - \lambda_{0}Q_{0} - \sigma Q_{1} - (Q_{0}^3 + 3Q_{0}Q_{1}^2) + \gamma \cos\omega t\\ P_{1}\\ - \delta P_{1} - \lambda_{0}Q_{1} - \sigma Q_{0} - 3(Q_{1}^3 + Q_{0}^2Q_{1}). \end{pmatrix}, \label{eq:a}$$ Note that there is nothing special about order $r=1$, and the same procedure can be repeated for any $r$. The initial condition on the generalized coordinates $q$ and conjugate momenta $p$ gets incorporated into the initial conditions on the coefficients of expansion: $Q_{i}$ and $P_{i}$. The Poincaré section for Eqs. \[eq:a\] are shown in Fig. \[Fig:Coeff\_certain\]. In this case, the stretching and folding structure of the Duffing oscillator is not as evident as in Fig. \[Fig:ForcedDuffing\]. However, the resulting dynamical system in Eqs. \[eq:a\] has a Lyapunov exponent of $\Xi \approx 0.73$, implying the persistence of sensitive dependence to initial conditions. Note that the route to chaos [@Gucken:book] for the the original Duffing oscillator is well known. In [@Cit:DuffingR1; @Cit:DuffingR2], the authors numerically demonstrate that the forced Duffing oscillator becomes chaotic due to a sequence of period doubling bifurcations. Due to the onset of chaos, the solution becomes increasingly difficult for polynomial chaos to capture. We point out that polynomial chaos is known to suffer from an inability to track output distributions for long term simulations [@Cit:long-term]. The reason for the inability of polynomial chaos to track the output distribution lies in the increasingly oscillatory nature of the solution $q(t;\lambda)$ in terms of the uncertain parameter $\lambda$. In other words, any finite expansion in Eqn. \[eq:expx\] will fail at some $t$, since $q(t,\lambda)$ is too oscillatory in terms of $\lambda$. The greater the oscillatory nature of the output in terms of $\lambda$, the worse polynomial chaos performs. In [@hybrid_uq], the oscillatory nature of the output is again found to adversely impact the propagation of uncertainty through hybrid dynamical systems. However, we find that chaotic dynamics exacerbates this phenomenon. In particular, due to the coexistence of periodic orbits of different periods along with the chaotic attractor, the solution is found to rapidly become oscillatory with respect to $\lambda$ (depicted in Fig. \[Fig:Duffing\_osc\]). Polynomial chaos is unable to track the first moment (mean) of $q(t;\lambda)$ beyond certain time (10 secs in Fig. \[Fig:Duffing\_tracking\], 25 secs in Fig. \[Fig:Duffing\_tracking\_IC4\]). In contrast to the forced Duffing oscillator, polynomial chaos is able to accurately track the mean of $q$ in the undamped and unforced Duffing oscillator with an order of expansion of $r=1$ (see Figs. \[Fig:Duffing\_simple\_tracking\] and \[Fig:Duffing\_simple\_tracking\_IC4\]). Note that all parameters and initial conditions are held constant here (except for the removal of the forcing and damping terms). Hence, one needs to be careful when applying polynomial chaos to systems that are chaotic. We point to a caveat that if the initial condition is chosen close to the stable and unstable manifolds of the saddle equilibrium $(0,0)$, polynomial chaos performs poorly on the undamped, unforced oscillator case due to the discontinuity associated with the basin boundary [@hybrid_uq]. ![$q(t;\lambda)$ as a function of $\lambda$ for the Duffing oscillator at $t\approx 15$ sec. The solution is already too oscillatory in terms of $\lambda$ for an expansion to $r=1$.\[Fig:Duffing\_osc\]](oscillatory_T_15.eps) ![Comparison of Monte Carlo with polynomial chaos for the mean of $q$ as a function of time in the Duffing oscillator with initial condition $(q,p)=(1,0)$. After $t\approx 10$s, polynomial chaos (expansion to $r=1$) is unable to accurately track the first moment of the output distribution.\[Fig:Duffing\_tracking\]](Duffing_tracking.eps) ![Comparison of Monte Carlo with polynomial chaos for the mean of $q$ as a function of time in the undamped, unforced Duffing oscillator with initial condition $(q,p)=(1,0)$. Polynomial chaos (expansion to $r=1$) is able to accurately track the first moment of the output distribution.\[Fig:Duffing\_simple\_tracking\]](Duffing_simple_tracking.eps) ![Comparison of Monte Carlo with polynomial chaos for the mean of $q$ as a function of time in the Duffing oscillator with initial condition $(q,p)=(4,0)$. After $\approx t=25$s, polynomial chaos (expansion to $r=1$) is unable to accurately track the first moment of the output distribution.\[Fig:Duffing\_tracking\_IC4\]](Duffing_tracking_IC4.eps) ![Comparison of Monte Carlo with polynomial chaos for the mean of $q$ as a function of time in the undamped, unforced Duffing oscillator with initial condition $(q,p)=(4,0)$. Polynomial chaos (expansion to $r=1$) is able to accurately track the first moment of the output distribution.\[Fig:Duffing\_simple\_tracking\_IC4\]](Duffing_simple_tracking_IC4.eps) When propagating uncertainty through chaotic systems with uncertain initial conditions, polynomial chaos again will need to be used carefully. Assume that $\lambda$ is not uncertain anymore, but instead the *initial conditions* are normally distributed as $(q,p) = (1,0) + (\sigma\eta,0)$, where $\eta$ is a Gaussian variable with zero mean and unit variance. The first order expansion yields the following system: $$\begin{pmatrix} \dot Q_{0}\\ \dot P_{0}\\ \dot Q_{1}\\ \dot P_{1} \end{pmatrix}=\begin{pmatrix} P_{0}\\ - \delta P_{0} - \lambda Q_{0} - (Q_{0}^3 + 3Q_{0}Q_{1}^2) + \gamma \cos \omega t\\ P_{1}\\ - \delta P_{1} - \lambda Q_{1} - 3(Q_{1}^3 + Q_{0}^2Q_{1}). \end{pmatrix}, \label{eq:ic}$$ Note that the uncertainty in initial conditions does not appear explicitly in these equations, but rather enters through the initial condition $Q_1(0) = \sigma$. For the purpose of our simulations we take $\sigma=0.1$. The resulting Poincaré sections are shown in Fig. \[Fig:Coeff\_uncertain\]. The Lyapunov exponent is numerically found to be $\approx 0.85$, suggesting the persistence of chaos in the resulting polynomial chaos equations. This implies that any long term simulation that aims to track the output distribution will also suffer from problems of round-off in the initial conditions (given that the distribution on the initial condition will require computation of the initial conditions of the coefficients). Conclusions =========== Polynomial chaos is slowly becoming an established and popular approach for propagating uncertainty through smooth systems. Every year researchers use the approach to propagate uncertainty through a wide variety of engineering [@Allen2009; @Ghanem1998; @Najm2009; @Cit:Materials] and biological systems [@Cit:polybio]. A systematic study on the properties and applicability of polynomial chaos to systems based on their structure and dynamics appears to be lacking. In this work, we presented three main results. In the first part, we proved that when polynomial chaos is applied to Hamiltonian systems, the resulting equations are also Hamiltonian, even when the expansion is truncated. This is important, as it implies that structure in Hamiltonian systems is not only inherited by the new equations but also require the use of structure-preserving integrators [@Cit:geom] to accurately propagate uncertainty. We also used the volume-preserving property of Hamiltonian systems to show that, on a particular example, a finite expansion must fail at long times, regardless of the order of the expansion. In the second part, we show that polynomial chaos may be applied to the averaged equations of a forced two-time system, allowing much faster uncertainty propagation than polynomial chaos on the original system. As the time scale separation increases, both the computational advantage as well as the quality of the approximation improves. In the third part, we demonstrate that polynomial chaos also inherits chaotic dynamics from underlying systems. The presence of chaos is shown to negatively influence the applicability of polynomial chaos. It reduces the length of time that polynomial chaos accurately tracks the output distributions and complicates computations when there is uncertainty in initial conditions. Acknowledgments {#acknowledgments .unnumbered} =============== The authors would like to thank Amit Surana, as well as the anonymous reviewers for constructive comments.
{ "pile_set_name": "ArXiv" }
--- abstract: 'For probability vectors $x$ and $y$, the catalytic majorization relation $x \prec_T y$ is defined to hold when there exists a probability vector $z$ such that $x \otimes z$ is majorized by $y \otimes z$. In this paper, an infinite family of functions is given such that, subject to some trivial restrictions, $x \prec_T y$ if and only if $f_r(x)<f_r(y)$ for all functions $f_r$ in the family. An outline of a proof of this result is provided. The catalytic majorization relation is known to provide a determination of which transformations of jointly held pure quantum states are possible using local operations and classical communication when an additional jointly held state may be specified to facilitate the transformation without being consumed.' author: - Matthew Klimesh bibliography: - 'maj2007.bib' date: 'September 23, 2007' title: | Inequalities that Collectively Completely Characterize\ the Catalytic Majorization Relation --- Introduction {#sec:intro} ============ Let $x=(x_1,\ldots,x_d)$ and $y=(y_1,\ldots,y_d)$ be $d$-dimensional vectors with real components. Let $x^{\downarrow}$ denote the vector obtained by arranging the components of $x$ in decreasing order: $x^{\downarrow} = (x_1^{\downarrow},\ldots,x_d^{\downarrow})$ where $x_1^{\downarrow} \geq \cdots \geq x_d^{\downarrow}$. Then $x$ is said to be *majorized* by $y$, written $x \prec y$, if $$\sum_{i=1}^k x_i^{\downarrow} \leq \sum_{i=1}^k y_i^{\downarrow}$$ when $1 \leq k < d$ and $\sum_{i=1}^d x_i = \sum_{i=1}^d y_i$. The majorization relation has been well studied; useful literature on the subject includes [@Mar79] and [@Bha97]. For a $d$-dimensional vector $x$ and an $\ell$-dimensional vector $z$ the notation $x \otimes z$ denotes the $d\ell$-dimensional vector tensor product, which is a vector whose components are all of the terms of the form $x_i z_j$. For our purposes, the order of the components is irrelevant. Our interest is in the following question: given two $d$-dimensional probability vectors $x$ and $y$, does there exist a (finite dimensional) probability vector $z$ such that $x \otimes z \prec y \otimes z$? This question motivates the definition of the *catalytic majorization* relation: we say $x$ is *catalytically majorized* by $y$, written $x \prec_T y$, when such a $z$ exists.[^1] Probability vectors of different dimensions can be compared with this relation by appending zeros to the shorter vector, and in fact we find it convenient to consider catalytic majorization as a relation among probability vectors of all finite dimensions. In this paper, an infinite family of functions is given with the property that, subject to some trivial restrictions, $x \prec_T y$ if and only if $f_r(x)<f_r(y)$ for all functions $f_r$ in the family. This result provides an answer to open problem 4 of [@Kru05] (also posed in [@Nie99a]), which asks to determine relatively simple conditions to decide whether or not $x \prec_T y$. This result also essentially resolves a conjecture of Nielsen (as stated in [@Daf04]) on that problem; with some minor adjustments, that conjecture would hypothesize the present main theorem. We regard the main mathematical significance of our result to be in clarifying the nature of the catalytic majorization relation, as well as providing a formulation that may be more useful mathematically than the definition. However, we also note that our result should provide a means of determining whether $x \prec_T y$ for a given $x$ and $y$; the definition does not suggest a general practical method for making this determination. The catalytic majorization relation has its origins in the field of quantum information. In recent years in this field has developed rapidly and much research has been directed toward understanding what sort of manipulations of quantum-mechanically entangled states are possible. This research has in part been motivated by applications of entanglement including quantum teleportation, quantum dense coding, quantum cryptography, and quantum computation. Both majorization and catalytic majorization have been shown to arise in the study of transformations of entangled bipartite pure quantum states. Nielsen obtained the following result [@Nie99b]: \[thm:nielsen\] Suppose Alice and Bob are in joint possession of an entangled pure quantum state $|\psi_1\rangle$ that they wish to transform into another bipartite entangled pure state $|\psi_2\rangle$. Let $|\psi_1\rangle = \sum_{i=1}^d \sqrt{\alpha_i}|i_A\rangle |i_B\rangle$ be a Schmidt decomposition of $|\psi_1\rangle$ and let $|\psi_2\rangle = \sum_{i=1}^d \sqrt{\beta_i}|i'_A\rangle |i'_B\rangle$ be a Schmidt decomposition of $|\psi_2\rangle$. Then $|\psi_1\rangle$ can be converted to $|\psi_2\rangle$ (with success guaranteed) using only local operations and classical communication (LOCC) if and only if the vector $\alpha = (\alpha_1,\ldots,\alpha_d)$ is majorized by $\beta = (\beta_1,\ldots,\beta_d)$. Jonathan and Plenio [@Jon99] extended this result by showing that even if it is not possible to convert $|\psi_1\rangle$ to $|\psi_2\rangle$ directly (according to Theorem \[thm:nielsen\]), it may be possible to convert $|\psi_1\rangle |\phi\rangle$ to $|\psi_2\rangle |\phi\rangle$, where $|\phi\rangle$ is an additional bipartite state shared by Alice and Bob. If $x$, $y$, and $z$ are the vectors of (squared) Schmidt coefficients of $|\psi_1\rangle$, $|\psi_2\rangle$, and $|\phi\rangle$ respectively, then the Schmidt coefficients of $|\psi_1\rangle |\phi\rangle$ are the components of $x \otimes z$ and the Schmidt coefficients of $|\psi_2\rangle |\phi\rangle$ are the components of $y \otimes z$; thus Nielsen’s Theorem implies that $|\psi_1\rangle |\phi\rangle$ can be converted to $|\psi_2\rangle |\phi\rangle$ when $x \otimes z \prec y \otimes z$. The state $|\phi\rangle$ is not consumed by this transformation; for this reason $|\phi\rangle$ is referred to as a catalyst and we say $|\phi\rangle$ catalyzes the transformation from $|\psi_1\rangle$ to $|\psi_2\rangle$. A state exists that can catalyze the transformation from $|\psi_1\rangle$ to $|\psi_2\rangle$ if and only if $x \prec_T y$. Main Result =========== To state our main theorem we specify a family of functions, indexed by a real number $r$. For a $d$-dimensional probability vector $x$, let $$f_r(x) = \begin{cases} \ln \sum_{i=1}^d x_i^r & (r>1); \\ \sum_{i=1}^d x_i \ln x_i & (r = 1); \\ -\ln \sum_{i=1}^d x_i^r & (0<r<1); \\ -\sum_{i=1}^d \ln x_i & (r = 0); \\ \ln \sum_{i=1}^d x_i^r & (r<0). \end{cases}$$ If any of the components of $x$ are $0$, we take $f_r(x) = \infty$ for $r \leq 0$. Our main theorem is the following: \[thm:main\] Let $x=(x_1,\ldots,x_d)$ and $y=(y_1,\ldots,y_d)$ be $d$-dimensional probability vectors. Suppose that $x$ and $y$ do not both contain components equal to $0$ and that $x \neq y$. Then $x \prec_T y$ if and only if $f_r(x) < f_r(y)$ for all $r \in {\mathbb{R}}$. We refer to the inequalities $f_r(x) < f_r(y)$ as the $f_r$ inequalities. Note that the restrictions on $x$ and $y$ in Theorem \[thm:main\] do not limit the scope of the theorem in any essential way; for example, it is clear that adding or removing components that are $0$ from $x$ and $y$ does not affect the status of the catalytic majorization relation between them. We previously reported the statement of Theorem \[thm:main\] in [@Kli04]. It should be clear that there are many equivalent choices for the $f_r$ functions, since for any particular $r$ the composition of any increasing function with $f_r$ will give the same inequality in Theorem \[thm:main\]. For example, $\sum_{i=1}^d x_i^r$ could be used in place of $\ln \sum_{i=1}^d x_i^r$ when $r>1$. With our definition, for $r>0$ the function $f_r$ is, up to a positive constant factor that depends on $r$, the negative of the Rényi entropy $H_r$. In particular, the function $f_1$ is the negative of the Shannon entropy. Related Results in the Literature ================================= There are many related results in the literature; we attempt to list the most relevant here. Recently, Aubrun and Nechita [@Aub07] obtained a result that has many aspects of our Theorem \[thm:main\]. To describe their result, it is convenient to consider catalytic majorization to be a relation among infinite-dimensional probability vectors with a finite number of nonzero components (i.e., with finite support). Aubrun and Nechita considered the set $T_c(y)$, defined to be the set of all $x$ such that $x \prec_T y$. In terms of the $f_r$ functions, their main result implies that $x \in \overline{T_c(y)}$ if and only if $f_r(x) \leq f_r(y)$ for $r > 1$, where $\overline{T_c(y)}$ is defined to be the $\ell_1$ closure of $T_c(y)$ within the space of infinite probability vectors with finite support. Stated another way, their result is that if $x$ and $y$ are infinite probability vectors with finite support, then the following are equivalent: - For any $\epsilon>0$ there exists an $x'$ with finite support such that $\|x-x'\|_1 < \epsilon$ and $x' \prec_T y$. - The inequalities $f_r(x) \leq f_r(y)$ hold for $r > 1$. We point out a subtlety that arises when comparing this result to ours. Let $P_d$ be the set of all infinite probability vectors with at most $d$ nonzero components, and let $T(y) = \{ x \in P_d : x \prec_T y \}$. Then, as noted in [@Aub07], even though $T(y) = T_c(y) \cap P_d$, the $\ell_1$ closure $\overline{T(y)}$ is generally a strict subset of $\overline{T_c(y)} \cap P_d$. Essentially, for some $d$-dimensional vectors $x$, one cannot find an $x'$ that is “close” to $x$ for which $x' \prec_T y$ unless $x'$ can have more nonzero components than $x$. Alternatively, this can be regarded as an indication of the importance of the $f_r$ inequalities with $r<1$ for characterizing $T(y)$. The result of Aubrun and Nechita that we have stated above can be obtained as a consequence of our Theorem \[thm:main\]. However, their main result also applies to multiple-copy transformations, which are not addressed by the present work. Transformations of states that use catalyst states as described in Section \[sec:intro\] are said to be transformations that use entanglement-assisted LOCC (ELOCC). Thus Theorem \[thm:main\] can be regarded as a characterization of transformations that are possible under ELOCC. If for some positive integer $k$, the state $|\psi_1\rangle^{\otimes k}$ (i.e., a state that is $k$ copies of $|\psi_1\rangle$) can be transformed to $|\psi_2\rangle^{\otimes k}$ under LOCC, then we say that the state $|\psi_1\rangle$ can be transformed to $|\psi_2\rangle$ under multiple-copy LOCC (MLOCC), and the transformation is called a multiple-copy transformation. It was observed in [@Ban02] that MLOCC allows transformations that are not possible under LOCC. Duan et al. [@Dua05] showed that any transformation that is possible under MLOCC is also possible under ELOCC. Additional results on MLOCC and its close relation to ELOCC can be found in [@Fen06; @Dua05]. The main result of Aubrun and Nechita [@Aub07] implies a characterization of MLOCC that is identical to their characterization of ELOCC. Vidal [@Vid99] showed that Nielsen’s theorem ([@Nie99b], stated above as Theorem \[thm:nielsen\]) can be generalized to provide the optimum probability of transforming a state $|\psi_1\rangle$ to another state $|\psi_2\rangle$ under LOCC. (Nielsen’s theorem covers the case where the transformation can be made with certainty.) It was noted by Jonathan and Plenio [@Jon99] that catalyst states can be useful in probabilistic entanglement transformations. Properties of catalyst-assisted probabilistic entanglement transformations were investigated in [@Daf04; @Fen05]. We do not consider MLOCC or probabilistic entanglement transformations in this paper. Additive Schur-Convex Functions {#sec:asc} =============================== A useful tool in the study of majorization is the notion of a *Schur-convex* function. A function $f: {\mathbb{R}}^d \rightarrow {\mathbb{R}}$ is Schur-convex if $f(x) \leq f(y)$ whenever $x \prec y$ [@Mar79]. Nielsen [@Nie99a] has introduced the notion of an *additive Schur-convex* function: A function $f$ from probability vectors (of any dimension) to ${\mathbb{R}}$ is additive Schur-convex if $f$ is Schur-convex (when restricted to probability vectors of a given dimension), and $f(x \otimes y) = f(x) + f(y)$. If $x \prec_T y$ then we must have $f(x) \leq f(y)$ for any additive Schur-convex function $f$, since if $x \otimes z \prec y \otimes z$ then $f(x)+f(z) = f(x \otimes z) \leq f(y \otimes z) = f(y)+f(z)$. Clearly, this observation is a motivating factor for the conjecture of Nielsen (mentioned in Section \[sec:intro\]) regarding conditions for catalytic majorization. Our functions $f_r$ are known to be additive Schur-convex functions. This can be verified with the aid of the following fact (from, e.g., [@Mar79]): A differentiable function $f(x)$ is Schur-convex if and only if it is invariant to permutations of the components of $x$ and $$(x_i-x_j)\left( \frac{\partial f}{\partial x_i} - \frac{\partial f}{\partial x_j} \right) \geq 0$$ for any pair of indices $i,j$. (For our application, the definition of additive Schur-convexity must be modified slightly to allow for the fact that $f_r$ can be $\infty$ when $r \leq 0$, but this does not cause much difficulty.) Clearly we have that $f_r(x)$ is continuous in $r$ when $r \neq 0,1$. However we are really more interested in the inequality $f_r(x) < f_r(y)$ than the actual values of the functions. It turns out that this inequality is continuous in $r$ in a certain sense for all $r \in {\mathbb{R}}$. Specifically, if for any neighborhood of $r$ there exists an $r'$ in the neighborhood such that $f_{r'}(x) \leq f_{r'}(y)$, then $f_r(x) \leq f_r(y)$. This continuity of the $f_r$ inequalities in $r$ can be exhibited as follows. Let $\tilde{f}_r(x) = (1/r(r-1)) \ln \sum_{i=1}^d x_i^r$, for $r \neq 0,1$. Clearly $\tilde{f}_r(x) < \tilde{f}_r(y)$ is equivalent to $f_r(x) < f_r(y)$ for $r \neq 0,1$. For the case $r=1$, we note that the limit of $\tilde{f}_r(x)$ as $r$ approaches $1$ is equal to $f_1(x)$. On the other hand, in the limit as $r \rightarrow 0$, $\tilde{f}_r(x)$ goes to $\pm \infty$ (depending on which direction the limit is from). However, the difference $\tilde{f}_r(y)- \tilde{f}_r(x)$ converges to $(f_0(y)-f_0(x))/d$ if neither $x$ nor $y$ contains components equal to zero. Thus, if we use the limiting values when $r \neq 0,1$ then the difference $\tilde{f}_r(y)- \tilde{f}_r(x)$ is continuous in $r$ over all of ${\mathbb{R}}$ when neither $x$ nor $y$ contains components equal to zero. Given this fact and our main theorem, it may be convenient in some cases to assess whether $x \prec_T y$ by examining a plot of $\tilde{f}_r(y)- \tilde{f}_r(x)$ as a function of $r$. There are other additive Schur-convex functions that are not members of our $f_r$ family of functions; examples are given in, e.g., [@Nie99a]. We note in particular the functions $$x \mapsto \ln x_1^{\downarrow},$$ $$x \mapsto -\ln x_d^{\downarrow},$$ and $$x \mapsto -\ln \left|\operatorname{supp}x \right|,$$ where $\left|\operatorname{supp}x \right|$ is the number of nonzero components of $x$. From our main theorem and the above discussion, it must be the case that if for some $x$ and $y$ we have $f_r(x) \leq f_r(y)$ for all $r \in {\mathbb{R}}$, then $f(x) \leq f(y)$ for any additive Schur-convex function $f$. This can be easily verified directly for the examples above. Order-Free Characterization of Majorization {#sec:order-free} =========================================== The usual definition of majorization involves arranging the components of the vectors in decreasing order. This definition appears to be inconvenient to work with when considering catalytic majorization. Thus we rely on the following well-known equivalence (see, e.g., [@Bha97]), where the notation $(c)^{+}$ means the positive portion of $c$; that is, $(c)^{+} =\max(c,0)$. For $d$-dimensional vectors $x$ and $y$ the following are equivalent: - $x \prec y$; - $\sum_{i=1}^d x_i = \sum_{i=1}^d y_i$ and for all $t \in {\mathbb{R}}$, $$\sum_{i=1}^d (x_i-t)^{+} \leq \sum_{i=1}^d (y_i-t)^{+};$$ - $\sum_{i=1}^d x_i = \sum_{i=1}^d y_i$ and for all $t \in {\mathbb{R}}$, $$\sum_{i=1}^d (t-x_i)^{+} \leq \sum_{i=1}^d (t-y_i)^{+}.$$ In the sufficiency portion of the proof of our main result, we use the following formulation. If $x$ and $y$ are $d$-dimensional probability vectors and $z$ is an $\ell$-dimensional vector, then $x \otimes z \prec y \otimes z$ if and only if for all $t \in {\mathbb{R}}$ we have $$\label{eq:order_free_cm} \sum_{i=1}^d \sum_{j=1}^{\ell} (x_i z_j-t)^{+} \leq \sum_{i=1}^d \sum_{j=1}^{\ell} (y_i z_j-t)^{+}.$$ Necessity ========= We first establish the necessity direction of our main theorem: under the hypothesis of Theorem \[thm:main\], $x \prec_T y$ implies $f_r(x) < f_r(y)$ for all $r \in {\mathbb{R}}$. This almost follows from the discussion of Schur-convex functions in Section \[sec:asc\]. However, to show that the inequalities must all be strict, we use the following result of Daftuar and the author [@Daf01]: \[cor:key\] Let $x =(x_1,\ldots, x_d)$ and $y=(y_1,\ldots, y_d)$ be $d$-dimensional probability vectors. Let $T(y)$ denote the set of all $d$-dimensional probability vectors $x'$ such that $x' \prec_T y$. Suppose that $x \prec_T y$ and $y_1^{\downarrow} > x_1^{\downarrow}$ and $y_d^{\downarrow} < x_d^{\downarrow}$. Then $x$ is in the interior of $T(y)$ (relative to the space of probability vectors). *Sketch of proof of necessity portion of Theorem \[thm:main\]:* Suppose $x \prec_T y$. Because each $f_r$ is additive Schur-convex, we must have $f_r(x) \leq f_r(y)$ for all $r \in {\mathbb{R}}$. However, we must show that equality cannot hold. We assume that the components of both $x$ and $y$ are in decreasing order. Without loss of generality, we may assume that no component of $x$ is equal to any component of $y$, as removing such a component from both $x$ and $y$ does not change the status of the catalytic majorization relation between them, nor does it affect the $f_r$ inequalities. Using this assumption and the fact that $f_r(x) \leq f_r(y)$ as $r$ approaches $+\infty$ and $-\infty$, we find that we may assume $x_1 < y_1$ and $x_d > y_d$. Let $w$ be the $d$-dimensional vector $(1,0,\ldots,0,-1)$. From Corollary \[cor:key\], $x$ is not on the boundary of $T(y)$, so for a sufficiently small $\epsilon>0$, we have $x+\epsilon w \prec_T y$. It is straightforward to verify that $f_r(x) < f_r(x+\epsilon w)$ with this $\epsilon$ for all $r \in {\mathbb{R}}$. But $f_r(x+\epsilon w) \leq f_r(y)$ since $x+\epsilon w \prec_T y$, so necessity is established. Sufficiency Outline {#sec:sufficiency_outline} =================== We now outline how the sufficiency portion of Theorem \[thm:main\] can be proved. That is, under the hypothesis of Theorem \[thm:main\], we outline how to show that if $f_r(x) < f_r(y)$ for all $r \in {\mathbb{R}}$ then there exists a catalyst vector $z$ such that $x \otimes z \prec y \otimes z$. For convenience, we assume that the components of $x$ and $y$ are arranged in decreasing order. As in the necessity proof, we may assume without loss of generality that $x_1 < y_1$ and $x_d > y_d$. We allow catalyst to vectors to be unnormalized. We argue that we need only consider vectors $x$ and $y$ whose components are all nonzero. The constraint $x_d > y_d$ implies that the components of $x$ are all nonzero. For $y$, we use the following claim. If $x$ and $y$ satisfy the $f_r$ inequalities, and $x_1 < y_1$ and $x_d > y_d$, then there exists a $y'$ (with components assumed to be in decreasing order) whose components are all nonzero, with $x_1 < y'_1$ and $x_d > y'_d$, such that $y' \prec y$, and the $f_r$ inequalities are satisfied with $y$ replaced by $y'$. This claim is straightforward (but slightly tedious) to verify; we provide a proof in the Appendix. Suppose the sufficiency portion of Theorem \[thm:main\] is shown to hold for probability vectors whose components are all nonzero. Then given $x$ and $y$ satisfying the $f_r$ inequalities, we can find $y'$ with the properties stated in the claim and conclude that there exists a $z$ such that $x \otimes z \prec y' \otimes z$. But $y' \prec y$ implies $y' \otimes z \prec y \otimes z$, so we have $x \otimes z \prec y \otimes z$ as desired. Thus we henceforth assume that all components of $y$ (and of $x$) are nonzero. The sufficiency proof can be divided into steps as follows. In these steps, we assume $x$ and $y$ are $d$-dimensional probability vectors with all nonzero components, the components of $x$ and $y$ are arranged in decreasing order, and $x_1 < y_1$ and $x_d > y_d$. **Step 1.** Show that if the $f_r$ inequalities hold for all $r \geq 0$, then there exists a continuous, decreasing function $z_{+}: [0, \infty) \rightarrow (0,1]$ with $z_{+}(0)=1$, along with a constant $s_{+}$, such that 1. for all $t \in (0,y_1)$, $$\label{eq:1ci} \sum_{i=1}^d \int_0^{\infty} (x_i z_{+}(s)-t)^{+}\,ds < \sum_{i=1}^d \int_0^{\infty} (y_i z_{+}(s)-t)^{+}\,ds;$$ 2. if $s \geq s_{+}$ then $z_{+}(s) = z_{+}(s_{+}) e^{-(s-s_{+})}$; and 3. the function $z_{+}$ satisfies a Lipschitz condition; that is, there exists a $K$ such that if $s_1,s_2 \geq 0$ then $|z_{+}(s_1)-z_{+}(s_2)| \leq K|s_1-s_2|$. **Step 2.** Show that if the $f_r$ inequalities hold for all $r \leq 0$, then there exists a continuous, increasing function $z_{-}: [0, \infty) \rightarrow [1,\infty)$ with $z_{-}(0)=1$, along with a constant $s_{-}$, such that 1. for all $t \in (y_d,\infty)$, $$\sum_{i=1}^d \int_0^{\infty} (t-x_i z_{-}(s))^{+}\,ds < \sum_{i=1}^d \int_0^{\infty} (t-y_i z_{-}(s))^{+}\,ds;$$ 2. if $s \geq s_{-}$ then $z_{-}(s) = z_{-}(s_{-}) e^{s-s_{-}}$; and 3. the function $z_{-}$ satisfies a Lipschitz condition on the interval $[0, s_{-}]$. **Step 3.** Show that if $z_{+}$ and $z_{-}$ satisfy the conditions given in Steps 1 and 2, then there exists a continuous, decreasing, positive function $z_{*}(s)$ on an interval $[0,a]$, with $z_{*}(0)=1$, such that 1. for all $t \in (y_d z_{*}(a),y_1)$, $$\label{eq:step3} \sum_{i=1}^d \int_0^a (x_i z_{*}(s)-t)^{+}\,ds < \sum_{i=1}^d \int_0^a (y_i z_{*}(s)-t)^{+}\,ds;$$ and 2. the function $z_{*}$ satisfies a Lipschitz condition on $[0,a]$. **Step 4.** Show that if $z_{*}$ satisfies the conditions given in Step 3, then there exists a finite-dimensional $z$ for which $x \otimes z \prec y \otimes z$. Clearly, the results to be proved in Steps 1 through 4 together imply the sufficiency portion of Theorem \[thm:main\]. Overview of Steps 1 and 2 ========================= Here we present an outline of how Steps 1 and 2 can be completed. The presentation here is somewhat less detailed than that of the other portions of the proof, but we hope that this outline provides a reasonable indication of our general strategy and of the methods involved. First we provide some motivation for why the results in these steps are useful. The basic idea is that conditions (i) of Steps 1 and 2 are reminiscent of the order-free conditions for majorization discussed in Section \[sec:order-free\]. In fact one could regard condition (i) of Step 1 as implying that $x$ is catalytically majorized by $y$ with catalyst $z_{+}$, in an appropriately generalized sense. However, we note that this condition by itself does not imply that $x \prec_T y$. (A similar observation is made in [@Daf04; @Aub07].) Conditions (ii) of Steps 1 and 2 together provide a way to combine the results of Steps 1 and 2, as will be clear in carrying out Step 3. We provide further thoughts related to conditions (i) of Steps 1 and 2 in Section \[sec:discussion\]. In Step 1, we first find a function $\tilde{z}_{+}$ that satisfies condition (i) without regard for condition (ii). A key idea is to consider functions of the form $$\tilde{z}_{+}(s) = e^{-s^{1/n}},$$ where $n$ is a positive integer. It turns out that if for a given $x$ and $y$ the $f_r$ inequalities are satisfied for $r \geq 0$, then for sufficiently large $n$ condition (i) holds with this $\tilde{z}_{+}$ substituted for $z_{+}$. To establish this result, we parameterize $t$ as $t = e^{-n/r}$; the range $r \in (0,\infty)$ corresponds to the range $t \in (0,1)$. For a given $r$ there is a correspondence between the inequality (\[eq:1ci\]) for $t = e^{-n/r}$ and the $f_r$ inequality. Specifically, the following holds: If for some $r>0$ the $f_r$ inequality is satisfied, then for all sufficiently large $n$ the inequality (\[eq:1ci\]) holds for $t = e^{-n/r}$ and $\tilde{z}_{+}(s) = e^{-s^{1/n}}$. In other words, given $r>0$, if $f_r(x) < f_r(y)$ then for all sufficiently large $n$ we have $$\label{eq:cond1a} \sum_{i=1}^d \int_0^{\infty} (x_i e^{-s^{1/n}} - e^{-n/r})^{+}\,ds < \sum_{i=1}^d \int_0^{\infty} (y_i e^{-s^{1/n}} - e^{-n/r})^{+}\,ds.$$ We require the existence of an $n$ such that (\[eq:cond1a\]) holds for all $r>0$, so it is not enough just to show that the above implication holds for each $r>0$. Our strategy in showing such an $n$ exists involves dividing the values of $r$ into five regions: regions 1, 3, and 5 corresponds conceptually to neighborhoods of $0$, $1$, and $\infty$, respectively, while regions 2 and 4 fill in the gaps. More precisely, the steps in the strategy are as follows. **Region 5.** Show that the inequality $y_1 > x_1$ implies that there is an interval $R_5 = [r_4,\infty)$ for some $r_4 \in (1,\infty)$, along with a positive integer $n_5$, such that if $r \in R_5$ and $n \geq n_5$, then (\[eq:cond1a\]) holds. **Region 3.** Show that the $f_r$ inequality with $r=1$ implies that there is an interval $R_3 = [r_2,r_3]$, where $r_2 \in (0,1)$ and $r_3 \in (1,r_4)$, along with a positive integer $n_3$, such that if $r \in R_3$ and $n \geq n_3$, then (\[eq:cond1a\]) holds. **Region 1.** Show that the $f_r$ inequality with $r=0$ implies that there is an interval $R_1 = (0,r_1]$ for some $r_1 \in (0,r_2)$, such that if $r \in R_1$ and $n \geq 1$, then (\[eq:cond1a\]) holds. Note the condition $n \geq 1$; unlike the other regions, $n$ does not need to be large here. **Region 2.** The interval $R_2$ is $[r_1,r_2]$. The task for this step is to show that the $f_r$ inequalities for $r \in R_2$ imply that there exists a positive integer $n_2$ such that if $r \in R_2$ and $n \geq n_2$, then (\[eq:cond1a\]) holds. **Region 4.** Similarly, the interval $R_4$ is $[r_3,r_4]$. The task for this step is to show that the $f_r$ inequalities for $r \in R_4$ imply that there exists a positive integer $n_4$ such that if $r \in R_4$ and $n \geq n_4$, then (\[eq:cond1a\]) holds. We proceed with some mathematical preliminaries that we found useful for completing the region steps above. For convenience we let $$\alpha(c,t,n)= \int_0^{\infty} (ce^{-s^{1/n}} - t)^{+} \,ds,$$ where $c \in (0,1)$ and $n \geq 1$ and $t>0$. We first observe that for $n \geq 1$ and $s \geq 0$, $$\int e^{-s^{1/n}}\,ds = -n!\,e^{-s^{1/n}} \sum_{j=0}^{n-1} \frac{1}{j!} s^{j/n}.$$ This can be verified by taking the derivative of the right side with respect to $s$. Suppose $c \geq t$. Then $$\begin{aligned} \alpha(c,t,n) & = \int_0^{(\ln \frac{c}{t})^n} (ce^{-s^{1/n}} - t) \,ds \nonumber \\ & = -cn! \left[ e^{-s^{1/n}} \sum_{j=0}^{n-1} \frac{1}{j!} s^{j/n} \right]_0^{(\ln \frac{c}{t})^n} - \left( \ln \frac{c}{t} \right)^n t \nonumber \\ & = -cn! \left(\frac{t}{c} \sum_{j=0}^{n-1} \frac{1}{j!} \left( \ln \frac{c}{t} \right)^j - 1 \right) - \left( \ln \frac{c}{t} \right)^n t \nonumber \\ & = cn! - tn! \left(\sum_{j=0}^{n-1} \frac{1}{j!} \left( \ln \frac{c}{t} \right)^j + \frac{1}{n!} \left( \ln \frac{c}{t} \right)^n \right) \nonumber \\ & = cn! - tn! \sum_{j=0}^n \frac{1}{j!} \left( \ln \frac{c}{t} \right)^j. \label{eq:alpha_expr1}\end{aligned}$$ Noting that $\sum_{j=0}^{\infty} \frac{1}{j!} (\ln \frac{c}{t})^j = c/t$, it follows from (\[eq:alpha\_expr1\]) that $$\alpha(c,t,n) = tn! \sum_{j=n+1}^{\infty} \frac{1}{j!} \left( \ln \frac{c}{t} \right)^j.$$ Substituting $t=e^{-n/r}$ gives $$\alpha(c,e^{-n/r},n) = e^{-n/r} n! \sum_{j=n+1}^{\infty} \frac{1}{j!} \left( \ln c + \frac{n}{r} \right)^j.$$ We found this form of $\alpha$ to be a useful starting point for completing the region 5 step. Returning to (\[eq:alpha\_expr1\]), we substitute $t=e^{-n/r}$ and find that $$\begin{aligned} \alpha(c,e^{-n/r},n) & = cn! - e^{-n/r} n! \sum_{j=0}^n \frac{1}{j!} \left( \ln c + \frac{n}{r} \right)^j \nonumber \\ & = cn! - e^{-n/r} n! \sum_{j=0}^n \frac{1}{j!} \sum_{k=0}^j {j \choose k} (\ln c)^k \left(\frac{n}{r} \right)^{j-k} \nonumber \\ & = cn! - e^{-n/r} n! \sum_{k=0}^n (\ln c)^k \sum_{j=k}^n \frac{1}{j!} {j \choose k} \left(\frac{n}{r} \right)^{j-k} \nonumber \\ & = cn! - e^{-n/r} n! \sum_{k=0}^n \frac{1}{k!} (\ln c)^k \sum_{j=k}^n \frac{1}{(j-k)!} \left(\frac{n}{r} \right)^{j-k} \nonumber \\ & = cn! - e^{-n/r} n! \sum_{k=0}^n \frac{1}{k!} (\ln c)^k \sum_{j=0}^{n-k} \frac{1}{j!} \left(\frac{n}{r} \right)^j, \label{eq:alpha_region12}\end{aligned}$$ where in the last step we have replaced $j-k$ with $j$. We found the form of $\alpha$ given by (\[eq:alpha\_region12\]) to be a useful starting point for completing the region 1 and region 2 steps. Now note that $\sum_{k=0}^{\infty} \frac{1}{k!} (\ln c)^k \sum_{j=0}^{\infty} \frac{1}{j!} (\frac{n}{r})^j = ce^{n/r}$. Thus we can conclude from (\[eq:alpha\_region12\]) that $$\label{eq:alpha_region34} \alpha(c,e^{-n/r},n) = e^{-n/r} n! \sum_{k=0}^{\infty} \frac{1}{k!} (\ln c)^k \sum_{j=\max(n+1-k,0)}^{\infty} \frac{1}{j!} \left(\frac{n}{r} \right)^j.$$ We found this form of $\alpha$ to be a useful starting point for completing the region 3 and region 4 steps. As an example we discuss our approach for the region 4 step in some detail. We start with (\[eq:alpha\_region34\]) where we replace $j$ with $j+n+1-k$: $$\begin{aligned} \alpha(c,e^{-n/r},n) & = e^{-n/r} n! \sum_{k=0}^{\infty} \frac{1}{k!} (\ln c)^k \sum_{j=\max(0,k-n-1)}^{\infty} \frac{1}{(n+1+j-k)!} \left(\frac{n}{r} \right)^{n+1+j-k} \nonumber \\ & = e^{-n/r} n! \sum_{k=0}^{\infty} \frac{1}{k!} r^k (\ln c)^k \sum_{j=\max(0,k-n-1)}^{\infty} \frac{n^{n+1+j-k}}{(n+1+j-k)!} r^{-n-1-j} \nonumber \\ & = e^{-n/r} n! \frac{n^{n+1}}{r^{n+1}(n+1)!} \sum_{k=0}^{\infty} \frac{1}{k!} (\ln c^r)^k \sum_{j=\max(0,k-n-1)}^{\infty} \frac{(n+1)!\,n^{j-k}}{(n+1+j-k)!} r^{-j} \label{eq:alpha_region4}\end{aligned}$$ We present a sequence of successively refined results regarding this expression. It is not difficult to show that for fixed $j$ and $k$, $$\lim_{n \rightarrow \infty} \frac{(n+1)!\,n^{j-k}}{(n+1+j-k)!} = 1.$$ By crudely bounding the difference between $\frac{(n+1)!\,n^{j-k}}{(n+1+j-k)!}$ and $1$ for large $n$, it can be shown that for fixed $r>1$ and fixed $k$, $$\lim_{n \rightarrow \infty} \sum_{j=\max(0,k-n-1)}^{\infty} \frac{(n+1)!\,n^{j-k}}{(n+1+j-k)!}r^{-j} = \sum_{j=0}^{\infty} r^{-j} = \frac{r}{r-1}.$$ By instead carefully bounding the quantities involved in the double sum in (\[eq:alpha\_region4\]) it can be shown that for a fixed $c \in (0,1)$ and a fixed $r>1$, $$\lim_{n \rightarrow \infty} \sum_{k=0}^{\infty} \frac{1}{k!} (\ln c^r)^k \sum_{j=\max(0,k-n-1)}^{\infty} \frac{(n+1)!\,n^{j-k}}{(n+1+j-k)!} r^{-j} = \frac{r}{r-1} \sum_{k=0}^{\infty} \frac{1}{k!} (\ln c^r)^k = \frac{r}{r-1} c^r.$$ Continuing along these lines, it can be shown that for a fixed $r>1$, if $\sum_{i=1}^d x_i^r < \sum_{i=1}^d y_i^r$ then for $n$ sufficiently large (\[eq:cond1a\]) holds. Finally, given $r_3$ and $r_4$ with $1<r_3<r_4$, it can be shown that if the $f_r$ inequality holds for all $r \in [r_3,r_4]$, then there exists an $n_4$ such that if $n \geq n_4$ then (\[eq:cond1a\]) holds for all $r \in [r_3,r_4]$. Note that only this last result is needed to complete the region 4 step; the sequence of results is given to provides some intuition as to why the last result holds, as well as suggesting how to establish it. After all of the region steps are completed, it follows quickly that condition (i) of Step 1 is satisfied by $\tilde{z}_{+}(s) = e^{-s^{1/n}}$ for sufficiently large $n$. We now explain how a $z_{+}$ can be constructed that also satisfies condition (ii). We start with a straightforward result that shows that whether (\[eq:1ci\]) holds for a given $t$ (or range of $t$) depends only on a portion of the function $z_{+}$. In the context of Step 1, let $z_{+}^{(0)}$ and $z_{+}^{(1)}$ be continuous, positive, decreasing functions on $[0,\infty)$ with $z_{+}^{(0)}(0) = z_{+}^{(1)}(0) = 1$. - Suppose $z_{+}^{(0)}(s) = z_{+}^{(1)}(s)$ for all $s \in [0,s_0]$ for some $s_0$, and that (\[eq:1ci\]) holds for $z_{+}^{(0)}$ when $t \in [y_1 z_{+}^{(0)}(s_0), y_1)$. Then (\[eq:1ci\]) holds for $z_{+}^{(1)}$ when $t$ is in this same interval. - Fix $t$ and let $[s_1,s_2]$ be the interval for which $z_{+}^{(0)}(s)$ is in $[t/y_1, t/y_d]$. Suppose for a given $\Delta$ we have $z_{+}^{(1)}(s+\Delta) = z_{+}^{(0)}(s)$ for all $s \in [s_1,s_2]$ (where necessarily $s_1+\Delta \geq 0$), and that (\[eq:1ci\]) holds for $z_{+}^{(0)}$ with our specific $t$. Then (\[eq:1ci\]) holds for $z_{+}^{(1)}$ and $t$. *Sketch of proof.* We outline a proof of part (ii); part (i) can be proved by similar means. Observe that if $c$ is a component of $x$ or $y$, then $c z_{+}^{(1)}(s) \geq t$ when $s \leq s_1+\Delta$, and $c z_{+}^{(1)}(s) \leq t$ when $s \geq s_2+\Delta$. We therefore have $$\begin{aligned} \sum_{i=1}^d \int_0^{\infty} (x_i z_{+}^{(1)}&(s) - t)^{+} \,ds \nonumber \\ & = \sum_{i=1}^d \left( \int_0^{s_1+\Delta} (x_i z_{+}^{(1)}(s) - t) \,ds + \int_{s_1+\Delta}^{s_2+\Delta} (x_i z_{+}^{(1)}(s) - t)^{+} \,ds \right) \nonumber \\ & = \sum_{i=1}^d \left(-(s_1+\Delta)t + x_i \int_0^{s_1+\Delta} z_{+}^{(1)}(s)\,ds + \int_{s_1}^{s_2} (x_i z_{+}^{(0)}(s) - t)^{+} \,ds \right) \nonumber \\ & = -(s_1+\Delta)td + \int_0^{s_1+\Delta} z_{+}^{(1)}(s)\,ds + \sum_{i=1}^d \int_{s_1}^{s_2} (x_i z_{+}^{(0)}(s) - t)^{+} \,ds \label{eq:intsplit1}\end{aligned}$$ and similarly $$\label{eq:intsplit2} \sum_{i=1}^d \int_0^{\infty} (x_i z_{+}^{(0)}(s) - t)^{+} \,ds = \int_0^{s_1} z_{+}^{(0)}(s) \,ds - s_1 td + \sum_{i=1}^d \int_{s_1}^{s_2} (x_i z_{+}^{(0)}(s) - t)^{+} \,ds.$$ We see from (\[eq:intsplit1\]) and (\[eq:intsplit2\]) that the difference $$\sum_{i=1}^d \int_0^{\infty} (x_i z_{+}^{(1)}(s) - t)^{+} \,ds - \sum_{i=1}^d \int_0^{\infty} (x_i z_{+}^{(0)}(s) - t)^{+} \,ds$$ does not depend on $x$. The analogous expression with $x$ replaced by $y$ has the same value. Thus (with our specific $t$), condition (\[eq:1ci\]) holds for $z_{+}^{(0)}$ if and only if (\[eq:1ci\]) holds for $z_{+}^{(1)}$. To produce the desired $z_{+}$, we start with $\tilde{z}_{+}(s) = e^{-s^{1/n_0}}$, where $n_0$ is chosen to be large enough that $\tilde{z}_{+}$ satisfies condition (i) of Step 1. Let $z_{+}(s) = \tilde{z}_{+}(s)$ when $s \in [0,s_0]$, where $s_0$ is chosen to be large enough that (\[eq:1ci\]) holds for all $t \in [e^{-n_0/r_1},y_1)$ regardless of how we specify the rest of $z_{+}$; the existence of such an $s_0$ is implied by part (i) of the preceding claim. The remaining $t$ fall in the interval $(0, e^{-n_0/r_1})$ which, with $t$ parameterized as $t=e^{-n_0/r}$, corresponds to $r \in R_1$. Recall that in this region it was not necessary for $n$ to be large for (\[eq:1ci\]) to be satisfied by $\tilde{z}_{+}(s) = e^{-s^{1/n}}$. The idea is to extend $z_{+}$ beyond $s_0$ in such a way that $z_{+}$ gradually changes from the form $e^{-s^{1/n_0}}$ to the form $e^{-s}$. We outline how this can be accomplished. We extend $z_{+}$ with sections from functions of the form $e^{-s^{1/n}}$ (with offsets chosen to make $z_{+}$ continuous), where $n$ decreases from section to section and is now no longer necessarily an integer. We make each such section long enough that the value of $z_{+}$ decreases by at least a factor of $y_1/y_d$ within the section. Part (ii) of our claim then implies that to verify (\[eq:1ci\]) for a given $t$ we need only consider the value of $z_{+}$ in at most two sections. The starting point for doing this is to show that (\[eq:cond1a\]) holds for noninteger $n \geq 1$ for $t$ in region 1. Then it must be checked that when the decrement to $n$ is small enough, (\[eq:1ci\]) continues to hold for the relevant values of $t$. Finally, it must be shown that for sufficiently many decrements we can reach a segment with $n=1$. The segment with $n=1$ is used for the remaining $s$, and part (ii) of our claim is again used to show that (\[eq:1ci\]) holds. At this point $z_{+}$ can be verified to satisfy all parts required by Step 1. Step 2 is largely similar to Step 1. A key idea in Step 2 is to consider functions of the form $\tilde{z}_{-}(s) = e^{s^{1/n}}$ where $n$ is a positive integer. If the $f_r$ inequalities are satisfied for all $r \leq 0$, then for sufficiently large $n$ condition (i) in Step 2 holds with this $\tilde{z}_{-}$. As in Step 1, our strategy for showing this involves five regions of values of $r$, in this case with regions $1'$, $3'$, and $5'$ corresponding conceptually to neighborhoods of $0$, $-1$, and $-\infty$. A $z_{-}$ satisfying condition (ii) of Step 2 is constructed analogously to the construction of the $z_{+}$ satisfying condition (ii) of Step 1. Step 3 ====== The function $z_{*}$ in this step may be regarded as a continuous precursor to a finite-dimensional catalyst vector $z$. The construction of $z_{*}$ is roughly as follows. The function $z_{+}$ from Step 1 is used for the left side of the interval on which $z_{*}$ is defined, and the function $z_{-}$ from Step 2 is reflected horizontally, scaled appropriately, and used for the right side of the interval on which $z_{*}$ is defined. There is a region in the center, in which $z_{*}$ is equal to a decaying exponential function, where the two halves coincide. The center region allows the transition from the submajorization-like condition involving $z_{+}$ to the supermajorization-like condition involving $z_{-}$ to be made. We proceed with the details. Let $a=s_{+}+s_{-}+ \ln(y_1/y_d)$. The function $z_{*}$ is defined on the interval $[0,a]$. We let $$\label{eq:zstar_defn} z_{*}(s) = \begin{cases} z_{+}(s) & \text{if $s \in [0, a-s_{-}]$;} \\ \frac{y_d}{y_1} \frac{z_{+}(s_{+})}{z_{-}(s_{-})} z_{-}(a-s) & \text{if $s \in (a-s_{-},a]$.} \end{cases}$$ Observe that if $s \in [s_{+}, a-s_{-}]$ then $$\begin{aligned} \frac{y_d}{y_1} \frac{z_{+}(s_{+})}{z_{-}(s_{-})} z_{-}(a-s) & = \frac{y_d}{y_1} \frac{z_{+}(s_{+})}{z_{-}(s_{-})} z_{-}(s_{-}) e^{a-s_{-}-s} \\ & = \frac{y_d}{y_1} z_{+}(s_{+}) e^{s_{+} + \ln(y_1/y_d) - s} \\ & = z_{+}(s_{+}) e^{-(s-s_{+})} \\ & = z_{+}(s),\end{aligned}$$ and therefore when $s \in [s_{+},a-s_{-}]$, either part of (\[eq:zstar\_defn\]) can be used. This is the aforementioned region where the two parts of $z_{*}(s)$ coincide. Also observe that if $s \in [s_{+},a]$ then $z_{*}(s) = z_{*}(a)z_{-}(a-s)$. Note that $z_{*}$ is continuous and decreasing and satisfies a Lipschitz condition. Let $t^{*} = y_1 z_{*}(a-s_{-})$. Observe that also $$t^{*} = y_1 z_{*}(a-s_{-}) = y_1 z_{+}(s_{+}) e^{-(a-s_{-}-s_{+})} = y_1 z_{+}(s_{+}) e^{- \ln(y_1/y_d)} = y_d z_{+}(s_{+}).$$ Suppose $c$ is a component of $x$ or $y$, so that in particular $0 < c \leq y_1$. If $s \in [a-s_{-},a]$ and $t \geq t^{*}$ then $$c z_{*}(s)-t \leq y_1 z_{*}(a-s_{-}) - t^{*} = 0,$$ and similarly if $s \geq a-s_{-}$ and $t \geq t^{*}$ then $c z_{+}(s)-t \leq 0$. Since $z_{*}(s)=z_{+}(s)$ when $s \in [0,a-s_{-}]$, for all $t \geq t^{*}$ we have $$\int_0^a (cz_{*}(s)-t)^{+} \,ds = \int_0^{\infty} (cz_{+}(s)-t)^{+} \,ds.$$ Now the result from Step 1 implies that if $t \in [t^{*},y_1)$ then $$\label{eq:cmaj1} \sum_{i=1}^d \int_0^a (x_i z_{*}(s)-t)^{+} \,ds < \sum_{i=1}^d \int_0^a (y_i z_{*}(s)-t)^{+} \,ds.$$ Next we find an analogous result that uses the result from Step 2. The derivation is slightly more involved due to the fact that $z_{-}$ is reflected and scaled to form the right half of $z_{*}$. Suppose again that $c$ is a component of $x$ or $y$, so that in particular $c \geq y_d$. If $s \in [0,s_{+}]$ and $t \leq t^{*}$ then $$t-cz_{*}(s) \leq t^{*}-y_d z_{*}(s_{+}) = 0.$$ Thus we have $$\begin{aligned} \int_0^a (t-cz_{*}(s))^{+} \,ds & = \int_{s_{+}}^a (t-cz_{*}(s))^{+} \,ds \\ & = \int_{s_{+}}^a (t-cz_{*}(a)z_{-}(a-s))^{+} \,ds \\ & = z_{*}(a) \int_{s_{+}}^a \left(\frac{t}{z_{*}(a)} - cz_{-}(a-s) \right)^{+} ds \\ & = z_{*}(a) \int_0^{a-s_{+}} \left(\frac{t}{z_{*}(a)} - cz_{-}(u) \right)^{+} du,\end{aligned}$$ where we have made the substitution $u=a-s$. For $u \geq a-s_{+}$ and $t \leq t^{*}$ we have $$\frac{t}{z_{*}(a)} - c z_{-}(u) \leq \frac{t^{*}}{z_{*}(a)} - y_d z_{-}(a-s_{+}) = \frac{y_d z_{+}(s_{+})}{z_{*}(a)} - y_d \frac{z_{*}(s_{+})}{z_{*}(a)} = 0.$$ Therefore $$z_{*}(a) \int_0^{a-s_{+}} \left(\frac{t}{z_{*}(a)} - cz_{-}(u) \right)^{+} du = z_{*}(a) \int_0^{\infty} \left(\frac{t}{z_{*}(a)} - cz_{-}(u) \right)^{+} du$$ and so $$\int_0^a (t-cz_{*}(s))^{+} \,ds = z_{*}(a) \int_0^{\infty} \left(\frac{t}{z_{*}(a)} - cz_{-}(u) \right)^{+} du.$$ Thus the result from Step 2 implies that if $t \in (y_d z_{*}(a), t^{*}]$ then $$\label{eq:cmaj2} \sum_{i=1}^d \int_0^a (t-x_i z_{*}(s))^{+} \,ds < \sum_{i=1}^d \int_0^a (t-y_i z_{*}(s))^{+} \,ds.$$ Finally we show that for any given $t$, conditions (\[eq:cmaj1\]) and (\[eq:cmaj2\]) are equivalent. Observe that $c = (c)^{+} - (-c)^{+}$ for any real $c$; thus $$\begin{aligned} \sum_{i=1}^d \int_0^a (x_i z_{*}(s)-t)^{+} \,ds & = \sum_{i=1}^d \int_0^a (x_i z_{*}(s)-t + (t-x_i z_{*}(s))^{+}) \,ds \\ & = \int_0^a z_{*}(s)\,ds - tad + \sum_{i=1}^d \int_0^a (t-x_i z_{*}(s))^{+} \,ds.\end{aligned}$$ Subtracting this equation from the analogous equation for $y$ yields $$\begin{aligned} \sum_{i=1}^d \int_0^a (y_i z_{*}(s)-&t)^{+}\,ds - \sum_{i=1}^d \int_0^a (x_i z_{*}(s)-t)^{+}\,ds \\ & = \sum_{i=1}^d \int_0^a (t-y_i z_{*}(s))^{+}\,ds - \sum_{i=1}^d \int_0^a (t-x_i z_{*}(s))^{+}\,ds,\end{aligned}$$ from which the equivalence of (\[eq:cmaj1\]) and (\[eq:cmaj2\]) is clear. Thus in particular (\[eq:cmaj1\]) holds for all $t \in (y_d z_{*}(a), y_1)$, so Step 3 is complete. Step 4 ====== Step 3 has provided a function $z_{*}$ that can be regarded as a continuous catalyst, as the condition (\[eq:step3\]) is roughly analogous to the condition (\[eq:order\_free\_cm\]) for a finite-dimensional vector $z$. It is fairly straightforward to produce a suitable catalyst vector $z$ from $z_{*}$. We provide the details in this section. From the result of Step 3 we have that (\[eq:step3\]) holds for all $t \in (y_d z_{*}(a), y_1)$; therefore in particular (\[eq:step3\]) holds for all $t \in [x_d z_{*}(a), x_1]$. This latter interval is compact, and $\sum_{i=1}^d \int_0^a (x_i z_{*}(s)-t)^{+}\,ds$ and $\sum_{i=1}^d \int_0^a (y_i z_{*}(s)-t)^{+}\,ds$ are continuous functions of $t$, so we may pick $\delta>0$ so that $$\label{eq:sumintdiff} \sum_{i=1}^d \int_0^a (y_i z_{*}(s)-t)^{+}\,ds - \sum_{i=1}^d \int_0^a (x_i z_{*}(s)-t)^{+}\,ds \geq \delta$$ when $t \in [x_d z_{*}(a),x_1]$. Now let $\ell$ be an integer that is large enough to imply that $|z_{*}(s_1)-z_{*}(s_2)| \leq \delta/2y_1 ad$ whenever $|s_1-s_2| \leq a/\ell$; such an $\ell$ must exist from the Lipschitz condition on $z_{*}(s)$. This $\ell$ will be the dimension of our catalyst $z$. For each $j \in \{1,\ldots, \ell \}$ pick $z_j$ as any value in the interval $[z_{*}((j/\ell)a),z_{*}(((j-1)/ \ell)a)]$. Our catalyst $z$ is the vector $(z_1,\ldots,z_{\ell})$. We now verify that $x \otimes z \prec y \otimes z$. Since the sum of all components of $x \otimes z$ is equal to the sum of all components of $y \otimes z$, we need only verify that for all $t$, $$\label{eq:step4ineq} \sum_{i=1}^d \sum_{j=1}^{\ell} (y_i z_j-t)^{+} - \sum_{i=1}^d \sum_{j=1}^{\ell} (x_i z_j-t)^{+} \geq 0.$$ For $t \geq x_1$, the second sum is zero, so (\[eq:step4ineq\]) must hold. For $t \leq x_d z_{*}(a)$ we have $$\sum_{i=1}^d \sum_{j=1}^{\ell} (y_i z_j-t)^{+} \geq \sum_{i=1}^d \sum_{j=1}^{\ell} (y_i z_j-t) = \sum_{i=1}^d \sum_{j=1}^{\ell} (x_i z_j-t) = \sum_{i=1}^d \sum_{j=1}^{\ell} (x_i z_j-t)^{+},$$ so again (\[eq:step4ineq\]) holds. For the remaining values of $t$, namely $t \in (x_d z_{*}(a), x_1)$, we show that the (appropriately scaled) values of the double sums in (\[eq:step4ineq\]) are close to the values of the corresponding sum-integrals, then we use (\[eq:sumintdiff\]) to establish (\[eq:step4ineq\]). We have $$\begin{aligned} \Biggl| \frac{a}{\ell} \sum_{i=1}^d \sum_{j=1}^{\ell} (x_i &z_j-t)^{+} - \sum_{i=1}^d \int_0^a (x_i z_{*}(s)-t)^{+}\,ds \Biggr| \\ & = \left| \sum_{i=1}^d \sum_{j=1}^{\ell} \int_{\frac{j-1}{\ell}a}^{\frac{j}{\ell}a} ((x_i z_j-t)^{+} - (x_i z_{*}(s)-t)^{+})\,ds \right| \\ & \leq \sum_{i=1}^d \sum_{j=1}^{\ell} \int_{\frac{j-1}{\ell}a}^{\frac{j}{\ell}a} \left| (x_i z_j-t)^{+} - (x_i z_{*}(s)-t)^{+} \right| ds \\ & \leq \sum_{i=1}^d \sum_{j=1}^{\ell} \int_{\frac{j-1}{\ell}a}^{\frac{j}{\ell}a} | x_i z_j - x_i z_{*}(s)| \,ds \\ & \leq \sum_{i=1}^d \sum_{j=1}^{\ell} \int_{\frac{j-1}{\ell}a}^{\frac{j}{\ell}a} \frac{x_i \delta}{2y_1 ad} \,ds \\ & \leq \sum_{i=1}^d \sum_{j=1}^{\ell} \int_{\frac{j-1}{\ell}a}^{\frac{j}{\ell}a} \frac{\delta}{2ad} \,ds \\ & = \delta/2.\end{aligned}$$ Similarly, $$\left| \frac{a}{\ell} \sum_{i=1}^d \sum_{j=1}^{\ell} (y_i z_j-t)^{+} - \sum_{i=1}^d \int_0^a (y_i z_{*}(s)-t)^{+}\,ds \right| \leq \delta/2.$$ Thus $$\begin{aligned} \frac{a}{\ell} \Biggl( \sum_{i=1}^d \sum_{j=1}^{\ell} (y_i z_j-t)^{+} &- \sum_{i=1}^d \sum_{j=1}^{\ell} (x_i z_j-t)^{+} \Biggr) \\ & \geq -\delta + \sum_{i=1}^d \int_0^a (y_i z_{*}(s)-t)^{+} \,ds - \int_0^a (x_i z_{*}(s)-t)^{+} \,ds \\ & \geq 0,\end{aligned}$$ so (\[eq:step4ineq\]) holds. Thus we have shown that (\[eq:step4ineq\]) holds for all $t$, so $x \otimes z \prec y \otimes z$ as desired. Discussion {#sec:discussion} ========== A proof of Theorem \[thm:main\] that follows along the lines outlined here appears to be constructive enough that one could produce a crude upper bound to the minimum dimension of the catalyst, as a function of $x$ and $y$. It may take a substantial effort to do this. It may be of interest to determine an asymptotic growth rate of the minimum dimension of the catalyst as, say, $x$ approaches the boundary of $T(y)$. The characterization of catalytic majorization provided by Theorem \[thm:main\] requires verifying a continuum of inequalities before concluding that $x \prec_T y$. In practice, for a particular $x$ and $y$, one can generally use the properties of the $f_r$ to check all $f_r$ inequalities with a finite amount of work. Given this, one might speculate that some of the inequalities are redundant, but we think that this is not the case: we conjecture that for any $r_0 \in {\mathbb{R}}$, there exists an $x$ and $y$ such that $f_r(x) < f_r(y)$ holds for all $r$ except $r = r_0$ (note that this would imply that $f_{r_0}(x) = f_{r_0}(y)$). The methods used here suggest several possible additional mathematical investigations. We mention one direction that seems intriguing. The idea is that it may be fruitful to define weak catalytic majorization. As background, we note that the usual weak majorizations (submajorization and supermajorization) can be naturally defined for infinite vectors $x$ and $y$ as follows: An infinite vector $x=(x_1,x_2,\ldots)$ can be defined to be submajorized by an infinite vector $y$ if for all $k \geq 1$ we have $\sum_{i=1}^k x_i^{\downarrow} \leq \sum_{i=1}^k y_i^{\downarrow}$. This definition makes sense only if for any component of $x$, there are a finite number of larger components of $x$, and similarly for $y$. If the components of $x$ and $y$ are nonnegative and have finite sums, then this definition is equivalent to the order-free formulation that requires $\sum_{i=1}^{\infty} (x_i-t)^{+} \leq \sum_{i=1}^{\infty} (y_i-t)^{+}$ for all $t \geq 0$. Similarly, $x$ can be defined to be supermajorized by $y$ if for all $k \geq 1$ we have $\sum_{i=1}^k x_i^{\uparrow} \leq \sum_{i=1}^k y_i^{\uparrow}$. This definition makes sense only if for any component of $x$, there are a finite number of smaller components of $x$, and similarly for $y$. If the components of $x$ and $y$ are nonnegative and unbounded, then this definition is equivalent to the order-free formulation that requires $\sum_{i=1}^{\infty} (t-x_i)^{+} \leq \sum_{i=1}^{\infty} (t-y_i)^{+}$ for all $t \geq 0$. With these notions in place, we can state a possibility for weak catalytic majorization. Suppose $x$ and $y$ are $d$-dimensional vectors with nonnegative components. It is not assumed that the sum of the components of $x$ equals the sum of the components of $y$. We consider infinite catalyst vectors $z = (z_1, z_2, \ldots)$ with nonnegative components (and at least one nonzero component). A possible definition is that $x$ is catalytically submajorized by $y$ if there exists a catalyst $z$ such that $x \otimes z$ is submajorized by $y \otimes z$ and $\sum_{j=1}^{\infty} z_j^r$ is finite for all $r>0$. Also $x$ is catalytically supermajorized by $y$ if there exists a catalyst $z$ such that $x \otimes z$ is supermajorized by $y \otimes z$ and $\sum_{j=1}^{\infty} z_j^r$ is finite for all $r<0$. We think that if the condition on $z$ is chosen properly in both cases (we are not sure if we have it right as written), then $x$ being both catalytically submajorized and catalytically supermajorized by $y$ will imply $x \prec_T y$. Note that with these definitions it is possible that, say, $x$ is catalytically submajorized by $y$ and the sum of the components of $x$ is equal to the sum of the components of $y$, but $x \not\prec_T y$. On a related note, we ask: if for some $r_0 \geq 0$ we only require the sum $\sum_{j=1}^{\infty} z_j^r$ to be finite when $r>r_0$, then are the $f_r$ inequalities for $r \geq r_0$ sufficient to imply the existence of a $z$ for which $x \otimes z$ is submajorized by $y \otimes z$? Analogously, if for some $r_0 \leq 0$ we only require the sum $\sum_{j=1}^{\infty} z_j^r$ to be finite when $r<r_0$, then are the $f_r$ inequalities for $r \leq r_0$ sufficient to imply the existence of a $z$ for which $x \otimes z$ is supermajorized by $y \otimes z$? Here we prove a claim that was used in Section \[sec:sufficiency\_outline\] to show that in our sufficiency proof we need only consider those $y$ whose components are all nonzero. We restate the claim here. Suppose $x$ and $y$ are $d$-dimensional probability vectors with components in decreasing order, and such that $x_1 < y_1$ and $x_d > y_d$ and for all $r \in {\mathbb{R}}$ the inequality $f_r(x) < f_r(y)$ holds. Then there exists a $d$-dimensional probability vector $y'$ whose components are all nonzero, with $x_1 < y'_1$ and $x_d > y'_d$ (we assume the components of $y'$ are in decreasing order), such that $f_r(x) < f_r(y')$ for all $r \in {\mathbb{R}}$, and such that $y' \prec y$. *Proof.* If all components of $y$ are nonzero, we can simply take $y'=y$, and we are done. Therefore we suppose $k$ components of $y$ are $0$, where $k \geq 1$. Let $w$ be the $d$-dimensional vector for which the first $d-k$ components are $0$, and the remaining $k$ components are $1$. For $n \geq 1$ let $y^{(n)} = \frac{n-1}{n}y + \frac{1}{nk} w$. Clearly, for sufficiently large $n$, say $n \geq N_0$, the components of $y^{(n)}$ are in decreasing order and $y^{(n)} \prec y$. Note also that $y^{(n)} \prec y^{(n+1)}$ for $n \geq N_0$. Choose $N_1$ to be large enough that $1/N_1 k < x_d$. Choose $r_1<0$ so that if $r<r_1$ then $$\left(\frac{1}{N_1 k} \right)^r > x_d^r d.$$ Choose $N_2$ to be large enough that $\frac{N_2-1}{N_2}y_1 > x_1$. Choose $r_2>1$ so that if $r>r_2$ then $$\left(\frac{N_2-1}{N_2} y_1 \right)^r > x_1^r d.$$ For general $x$, $y$, and $r$ we define $$F(x,y,r) = \begin{cases} \frac{1}{r(r-1)}( \ln \sum_{i=1}^d y_i^r - \ln \sum_{i=1}^d x_i^r), & \text{if $r \neq 0,1$;} \\ -\frac{1}{d} ( \sum_{i=1}^d \ln y_i - \sum_{i=1}^d \ln x_i ), & \text{if $r=0$;} \\ \sum_{i=1}^d y_i \ln y_i - \sum_{i=1}^d x_i \ln x_i, & \text{if $r=1$.} \end{cases}$$ If all components of $x$ are nonzero then $F(x,y,r) > 0$ is equivalent to $f_r(x) < f_r(y)$. It is straightforward to verify that for fixed $x$ and $y$, both with all components nonzero, $F(x,y,r)$ is continuous in $r$ over all of ${\mathbb{R}}$. For fixed $x$ and $r$ the function $F(x,y,r)$ is Schur-convex in $y$. For our specific $x$ and $y$ (with $y^{(n)}$ constructed from $y$) let $g_n(r) = F(x,y^{(n)},r)$. We then have that for all $n \geq N_0$ the function $g_n(r)$ is continuous in $r$. Also if $n \geq N_0$ then $y^{(n)} \prec y^{(n+1)}$ and so $g_n(r) \leq g_{n+1}(r)$ for all $r \in {\mathbb{R}}$. Finally, for all $r \in {\mathbb{R}}$ it can be verified that $\lim_{n \rightarrow \infty} g_n(r) > 0$ (specifically, the limit is $F(x,y,r)$, and in particular it is $+\infty$ when $r \leq 0$). From these properties of $g_n(r)$, a subclaim that we state below implies that there exists an $N$ such that for all $r \in [r_1,r_2]$ we have $g_n(r)>0$ when $n \geq N$. Let $y' = y^{(n)}$ for $n = \max(N_0,N_1,N_2,N)$. Now if $r \in [r_1,r_2]$ then $g_n(r)>0$ so that $f_r(x) < f_r(y')$. If $r<r_1$ then we have $$\sum_{i=1}^d x_i^r \leq x_d^r d < \left(\frac{1}{N_1 k}\right)^r \leq \left(\frac{1}{n k}\right)^r \leq \sum_{i=1}^d (y'_i)^r,$$ which implies $f_r(x) < f_r(y')$. If $r>r_2$ then we have $$\sum_{i=1}^d x_i^r \leq x_1^r d < \left(\frac{N_2-1}{N_2}y_1 \right)^r \leq \left(\frac{n-1}{n}y_1 \right)^r \leq \sum_{i=1}^d (y'_i)^r,$$ and again we have $f_r(x) < f_r(y')$. Thus $f_r(x) < f_r(y')$ for all $r \in {\mathbb{R}}$, so the proof is complete except for establishing the subclaim. Suppose $[a,b]$ is an arbitrary closed interval, and $g_n(r)$ satisfies the following: - for all $n \geq N_0$ the function $g_n(r)$ is continuous in $r$ over all of ${\mathbb{R}}$; - $g_n(r)$ is increasing in $n$ when $n \geq N_0$ and $r \in {\mathbb{R}}$; and - $\lim_{n \rightarrow \infty} g_n(r) > 0$ for all $r \in {\mathbb{R}}$. Then there exists an $N$ such that $g_n(r)>0$ for all $r \in [a,b]$ and $n \geq N$. *Proof of subclaim.* For all $n \geq N_0$ let $E_n = \{\, r \mid g_n(r)>0 \,\}$. Then from (i) each $E_n$ is an open set, and it follows from (ii) that $E_n \subset E_{n+1}$. From (iii) we have $\bigcup_{n \geq N_0} E_n = {\mathbb{R}}$, so in particular $[a,b] \subset \bigcup_{n \geq N_0} E_n$. Because the interval $[a,b]$ is compact, there exists a finite subset of $\{E_n\}_{n \geq N_0}$ that covers $[a,b]$. In view of the fact that $E_n \subset E_{n+1}$, there is thus an $N$ for which $[a,b] \subset E_N$, which implies that $[a,b] \subset E_n$ for $n \geq N$. Thus the claim holds using this choice of $N$. *Remarks.* Our subclaim appears to be a simpler relative of Dini’s theorem, and our proof of the subclaim is similar to a standard proof of Dini’s theorem. The fact that when $n$ is large enough $y^{(n)}$ has the desired properties can also be proved without the subclaim using the definition of the $f_r$ directly. The author is indebted to Sumit Daftuar and Michael Nielsen for introducing him to (what we now call) catalytic majorization, and for useful discussions and encouragement when he started investigating the subject. [^1]: The catalytic majorization relation is called the trumping relation in [@Daf01; @Daf04; @Nie99a].
{ "pile_set_name": "ArXiv" }
--- abstract: 'A better grasp of the physical foundations of life is necessary before we can understand the processes occurring inside a living cell. In his physical theory of the cell, American physiologist Gilbert Ling introduced an important notion of the resting state of the cell. He describes this state as an independent stable thermodynamic state of a living substance in which it has stored all the energy it needs to perform all kinds of biological work. This state is characterised by lower entropy of the system than in an active state. The main contribution to this reduction in entropy is made by the cellular water (the dominant component with a concentration of 14 M) which remains in a bound quasi-crystallised state in a resting cell. When the cell becomes active the water gets desorbed and the system’s entropy goes up sharply while the free energy of the system decreases as it is used up for biological work. However, Ling’s approach is primarily qualitative in terms of thermodynamics and it needs to be characterised more specifically. To this end, we propose a new thermodynamic approach to studying Ling’s model of the living cell (Ling’s cell), the centrepiece of which is the non-ergodicity property which has recently been proved for a wide range of systems in statistical mechanics \[7\]. In many ways this new thermodynamics overlaps with the standard quasi-stationary thermodynamics and is therefore compatible with the principles of the Ling cell, however a number of new specific results take into account the existence of several non-trivial motion integrals communicating with each other, whose existence follows from the non-ergodicity of the system (Ling’s cell). These results allowed us to develop general thermodynamic approaches to explaining some of the well-known physiological phenomena, which can be used for further physical analysis of these phenomena using specific physical models.' author: - 'D.V. Prokhorenko [^1], V.V. Matveev [^2]' title: 'The Significance of Non-ergodicity Property of Statistical Mechanics Systems for Understanding Resting State of a Living Cell' --- Introduction ============ The living state of a substance has always attracted the attention of physicists. And indeed, only a clear understanding of the thermodynamic characteristics of the living matter can give us insight into the processes that occur in living organisms. However, despite the fact that there is a lot of interest in this problem, this field can’t be said to be developing by leaps and bounds. Before proceeding with our analysis we have to establish its boundaries and conditions. There are two very different approaches to the thermodynamics of living systems: one based on thermodynamics of equilibrium process \[1,2,3,4\] and the other on thermodynamics of non-equilibrium processes \[5\]. Shroedinger’s research \[6\] differs from both of these. An obscure Russian scientist of Hungarian descent Ervin Bauer \[4\] was probably the first to suggest that the living state should be regarded as an unstable equilibrium. Treating the physical state of living substance in this way allowed him to draw a number of interesting conclusions and generalizations, but on the whole the most part of research was purely theoretical. According to Ling \[1,2,3\], whose position is of special interest to us, the minimal cell in the physical sense is a complex comprising protein in an unfolded configuration and water with ions , water and ions. The most significant characteristic of this complex is the state of water in it; it is absorbed by the protein in the form of a multi-layered structure that surrounds it along the entire length of the polypeptide. This ’coat’ consisting of water molecules is stabilized by hydrogen links that are stronger than the hydrogen links in volumetric water. This increase in strength is a result of an increase in the dipole moment of the water molecules under the influence of other dipoles that are stronger than water such as the functional groups in the peptide link bound (NH and CO). The polarization of water molecules explains both their strong binding by the polypeptide frame of the protein and the multi-layered absorption of water on the surface of the unfolded protein. According to Ling, almost all the water in a the cell is in a bound state. Because, in terms of the number of molecules, water is the most abundant compound inside the cell, its transition into a quasi-crystal state results in a significant fall in the entropy of the cell. This lower entropy is what brings about the rise in the amount of free energy of the resting living cell. According to Ling \[1,2,3\], whose position is of special interest to us, the minimal cell in the physical sense is a complex comprising protein in an unfolded configuration, water and ions. The most significant characteristic of this complex is the state of water in it; it is absorbed by the protein in the form of a multi-layered structure that surrounds it along the entire length of the polypeptide. This ’coat’ consisting of water molecules is stabilized by hydrogen links that are stronger than the hydrogen links in volumetric water. This increase in strength is a result of an increase in the dipole moment of the water molecules under the influence of other dipoles that are stronger than water such as the functional groups in the peptide bound (NH and CO). The polarization of water molecules explains both their strong binding by the polypeptide frame of the protein and the multi-layered absorption of water on the surface of the unfolded protein. According to Ling, almost all water in the cell is in a bound state. Because, in terms of the number of molecules, water is the most abundant compound inside the cell, its transition into a quasi-crystal state results in a significant fall in the entropy of the cell. This lower entropy is what brings about the rise in the amount of free energy of the resting living cell. The introduction of the concept of a resting state is one of Ling’s achievements. It’s this state that is used as the reference point for all the physical and chemical processes that take place inside the cell. When a cell is activated by an external stimulant or some other signal, it changes its state from resting to active. The active state is characterized by the disintegration of the water-protein-ions complex. The bound water breaks free and the system’s entropy increases. The free energy of the resting state is released and is used up for all kinds of biological work. This is the Ling model of the living cell (Ling’s cell) that will be the focus of our analysis. According to Ling, a cell can remain resting without exchanging energy or substances with the external environment. The cell just maintains diffusion equilibrium with the. This view directly contradicts Prigozhin’s approach according to which a living cell can only maintain its organization as long as it keeps exchanging substance and energy with the environment on a continuous basis. In other words, a cell can be likened to a burning candle flame; the flame will remain ’alive’ only as long as there is sufficient supply of fuel and oxidizer. Ling believes that such understanding of the thermodynamics of life is completely inadequate in the case of a living cell. His calculations demonstrate that if a continuous inflow of energy was really necessary for the experimentally observed exchange of Na+ ions between a resting cell and the environment (as is postulated in the traditional mechanism), the cell would simply be incapable to produce the necessary amounts of energy \[3\] and therefore the universally adopted model of ion transport contradicts the energy preservation law. Ling’s other argument proceeds as follows; if a living cell and a burning candle were to be frozen to the temperature of liquid nitrogen, both the flame and the life in the cell will ’go out’, but if they’re heated back to room temperature, the flame won’t start burning again, but the life processes in the cell will resume. The contradictions between the thermodynamic approaches to the phenomenon of life are so pronounced that the need for further research in this field is self-evident. The purpose of this paper is the demonstrate that the property of non-ergodicity that has recently been proved for a large number of systems in statistical mechanics can help better understand the nature of the resting state of a Ling’s cell and supports his understanding of the living cell’s thermodynamics. One feature of this approach is that it suggests that the resting state of a Ling’s cell should be considered to be a non-equilibrium stationary state, whose existence is a direct consequence of the non-ergodicity property that we postulate for Ling’s cells. In this approach thermodynamics of non-equilibrium stationary states must be constructed (analogous to the standard quasi-stationary thermodynamics) to explain why biological work becomes possible in the context of the proposed approach (biological work here means any changes in the cell that have a biological significance and that use up energy, for instance muscle contraction). There is no real contradiction between Ling’s stationary resting state and the obviously continuous metabolism necessary to maintain life, because a real cell constantly changes its state from active to resting and back. Metabolism and energy are needed to go back to a resting state rather than maintain it. Non-ergodicity means that there exist non-trivial first integrals of the system (i.e. there are values that are invariant under the Heisenberg and Hamilton motion equations). Because by definition we consider these first integrals to be real, then, in the case of quantum mechanics, they must be represented by self-adjoint operators and be experimentally observable values. It’s then only natural to ask how come they can only be observed in biological systems (as is shown below) as well as in liquid helium and superconductors, systems that are as far from biological as can be? When answering this question we come across a certain mathematical similarity between the super-fluid state of helium and the resting state of the Ling cell, and this similarity helps us better understand the physics of the living state. The main result of this paper is that by looking at the Ling cell as a non-ergodic system we were able to propose a common physical mechanism for various physiological phenomena, which were previously explained with the help of separate mechanisms barely related to each other. The main goal of physics in physiology must be to find out the thermodynamic nature of an active living cell, i.e. the source of all the manifestations of life. With this goal in mind we only discuss some of the characteristics of the Ling cell and provide only a most general physical description for them. We demonstrate, for example, that when a Ling’s cell is activated it emits heat rather than absorbs it. When we consider the properties of the physical model we use, it becomes perfectly clear why it is that unfolded proteins that make up the structural foundation of a resting Ling’s cell, begin to fold when it goes active, why potassium ions exit the cell into the environment and why an active cell changes its size (usually it shrinks) and what makes a cell dead. It is not our goal in this paper to compare the results we obtain for the Ling’s cell with the properties of a real living cell. Non-ergodicity of statistical mechanics systems. ================================================ In this section we first give the definition of ergodicity in statistical mechanics. We then formulate the main result of \[7\] and its classical analogue. After that we demonstrate how non-ergodicity follows from this result on the classical level. **Definition**. Suppose that a quantum system is described by Hamiltonian $H$ and some set of self-adjoint integrals of motion $K_1$,...,$K_l$ commuting to each other. We say that the system is ergodic with respect to the integrals $K_i$,...,$K_l$, if each dynamical variable commuting with $H$, $K_1$,...,$K_l$, is their function, or in other words the joint spectrum of operators $H$, $K_1$,..., $K_l$ is simple. The classical version of this definition can be easily obtained by replacing the word commutator with a Poisson bracket. In our case of Bose gas, apart from the Hamiltonian, there are the following trivial commuting first integrals: impulse $\vec{P}$ and the number of particles $N$. We will be looking at ergodicity relative to this set of first integrals. Let us now formulate the main result of our work \[7\]. Let $\Psi(x)$ and $\Psi^+(x)$ be secondary quantized wave functions of Bose gas that satisfy the canonical commutative relations $$\begin{aligned} {[\Psi(x),\Psi(x')]=[\Psi^+(x),\Psi^+(x')]=0},\nonumber\\ {[}\Psi(x),\Psi^+(x'){]}=\delta(x-x'),\end{aligned}$$ where the brackets designate a commutator and $\delta(x-x')$- means Dirak $\delta$ — function of a vector argument. An algebra generated by secondary-quantized wave functions is called the algebra of canonical commutative relations. The main result of \[7\] can be formulated as the following theorem. **Theorem 1.** There exist a linear functional $\langle\cdot\rangle$ on the algebra of canonical commutative relations (within a formal perturbation theory applied to an interaction constant) such that: a\) It is stationary, i.e. $\langle[H,v]\rangle=0$, where $H$ is a Hamiltonian of the system and $v$ is an arbitrary element of the canonical commutative relations algebra. b)$\langle\cdot\rangle$ is translation-invariant. c))$\langle\cdot\rangle$ Commutates with the operator of the number of particles $N$, i.e. ${([N, v]}) = 0$, where $N$ is the operator of the number of particles in the system and $v$ is an arbitrary element of the algebra of commutative relations. d\) Satisfy to the week cluster property, i.e. $$\begin{aligned} \lim_{|{a}|\rightarrow\infty}\int \langle\Psi^\pm(t,{x}_1+\delta_1 e_1{a}) ...\Psi^\pm(t,{x}_n+\delta_n e_1 {a} )\rangle f({x}_1,...,{x}_n)d^3x_1...d^3x_n\nonumber\\ =\int\langle\Psi^\pm(t,{x}_{i_1})...\Psi^\pm(t,{x}_{i_k})\rangle \langle\Psi^\pm(t,{x}_{i_k})...\Psi^\pm(t,{x}_{i_n})\rangle \nonumber\\ \times f({x}_1,...,{x}_n)d^3x_1...d^3x_n,\end{aligned}$$ where $\delta_i\in\{1,0\},\;i=1,2...n$ and $$\begin{aligned} i_1<i_2<...<i_k,\nonumber\\ i_{k+1}<i_{k+2}<...<i_n,\nonumber\\ \{i_1,i_2,...,i_k\}=\{i=1,2...n|\delta_i=0\}\neq\emptyset,\nonumber\\ \{i_{k+1},i_{k+2},...,i_n\}=\{i=1,2...n|\delta_i=1\}\neq\emptyset.\end{aligned}$$ $f(x_1,...,x_n)$ - is a test function (i.e. it is sufficiently smooth and it decreases sufficiently fast at infinity with all its derivatives), $e_1$ is a unit vector parallel to the x axis. e)$\langle\cdot\rangle$ is not describe the Gibbs distribution. Before proving the non-ergodicity of the system from this result, we will formulate its classical analogue. In classical statistical mechanics a system containing $N$ particles $N\rightarrow +\infty$, is described by a distribution function of the form $\rho((x_1,p_1),...,(x_N,p_N))$, where $(x_i,p_i)$ are the coordinates and impulse of the $i$-th particle $i = 1,...,N$. The distribution function $\rho$ is assumed to be symmetrical relative to the permutation of its arguments $(x_1,p_1),..., (x_N,p_N)$. Also partial distribution functions are introduced: $$\begin{aligned} \rho_0=1,\nonumber\\ \rho_1(x_1)=V \int dx_2dp_2...dx_Ndp_N \rho((x_1,p_1)...(x_N,p_N)),\nonumber\\ \rho_2((x_1,p_1),(x_2,p_2))=V^2 \int dx_3dp_3...dx_Ndp_N \rho((x_1,p_1)...(x_N,p_N)),\nonumber\\ ...................................................................................................... \label{DISTR}\end{aligned}$$ $V$ is a volume of the system. Partial distribution functions $\rho_0=1,\rho_1,...,\rho_n...$ satisfy the following consistency condition: $$\begin{aligned} \rho_n((x_1,p_1),...,(x_n,p_n))=\nonumber\\ =V^{-m} \int dx_{m+1}dp_{m+1}...dx_{n+m}dp_{n+m}\rho_n((x_1,p_1),...,(x_{n+m},p_{n+m})).\end{aligned}$$ If the number of particles in the system is variable, then in a similar way to (\[DISTR\]), we can also define partial distribution fucntions. They will also satisfy the consistency conditions if $N$ approaches infinity. If we take the main result of \[7\] and pass it through the limit we get the following theorem. **Theorem 1’**. There exists such a state $\rho$ (not necessarily positively defined and possibly with a varied number of particles) of a gas consisting of particles with weak interaction, that a\) This state is stationary, i.e. it doesn’t change with time due to its motion equations. b\) This state is translation-invariant. c\) This state satisfies the weak cluster property, i.e. $\forall n=1,...,\infty$ and $\forall k=1,...,n$ $$\begin{aligned} \lim \limits_{|l|\rightarrow\infty}\rho_n((x_1,p_1),...,(x_k,p_k),(x_{k+1}+le_1,p_{k+1}),...,(x_{n}+le_1,p_{n}))=\nonumber\\ =\rho_k((x_1,p_1),...,(x_k,p_k))\rho_{n-k}((x_{k+1},p_{k+1}),...,(x_{n},p_{n}))\end{aligned}$$ d\) The state $\rho$ cannot be described with a Gibbs distribution. Let us finally demonstrate how this result leads to the non-ergodicity of our system. To simply things we’ll only be considering the case of classical mechanics. For a detailed discussion of the quantum version of this result see \[7\]. Let us suppose that our system is ergodic, i.e. that its energy and impulse form a maximal set of first integrals in involution. Two dynamic variables are said to be in involution if their Poisson bracket equals zero. From here on for the sake of brevity, we’ll be simply talking about energy rather than energy, impulse and the number of particles. Let $f$ be a distribution function for our system in a phase space that corresponds to the state we talked about before. Then because of the system’s assumed ergodicity it must be a function of only energy $f=f(E_\Gamma)$ ($\Gamma$ - is a point in the phase space) and can be presented as a superposition of micro-canonical distributions: $$\begin{aligned} f(E_\Gamma)=\sum \limits_\alpha c_\alpha \delta(E_\Gamma-E_\alpha),\end{aligned}$$ The summation in this context is understood in the broadest possible sense and can be infinite (in other words it can be replaced with integration). Let 1 be a sufficiently large but finite subsystem of our system and let 2 - be a subsystem obtained from 1 by shifting it along the $x$ axis by a distance of $l$. Let 12 - be a union of systems 1 and 2. Let $\Gamma_1$, $\Gamma_2$ and $\Gamma_{12}$ designate points in the phase space for our systems 1, 2 and 12 respectively. Let $f_1(\Gamma_1)$, $f_2(\Gamma_2)$ and $f_{12}(\Gamma_{12})$ be distribution functions for systems 1, 2 and 12 respectively. Then using the same method that is used for obtaining a canonical distribution from a microcanonical one, for $|l|=\infty$ we can find that: $$\begin{aligned} f_{12}=\sum c_\alpha d_\alpha\frac{e^{-\frac{E_{\Gamma_1}}{T_\alpha}}}{Z_\alpha}\frac{e^{-\frac{E_{\Gamma_2}}{T_\alpha}}}{Z_\alpha}.\label{I}\end{aligned}$$ Here $d_\alpha>0$ - is a weight multiplier, $E_{\Gamma_1}$ and $E_{\Gamma_2}$ - are the energies of subsystems 1 and 2, $T_\alpha$ - is the temperature corresponding to energy $E_\alpha$, and $Z_\alpha$ - is the statistical sum: $$\begin{aligned} Z_\alpha=\int d \Gamma_1 e^{-\frac{E_{\Gamma_1}}{T_\alpha}}.\end{aligned}$$ But from the weak cluster property it follows that: $$\begin{aligned} f_{12}=f_1f_2.\end{aligned}$$ Therefore it is follows from (\[I\]) that all $c_\alpha=0$ except one and: $$\begin{aligned} f(E)=c\delta(E-E_0),\end{aligned}$$ (for some $c$ and $E_0$), i.e. the whole system is described by a microcanonical distribution while any of its sufficiently large but finite subsystems is described by a canonical Gibbs distribution, which contradicts the fact that our state is not a Gibbs state. This contradiction proves that our system is non-ergodic. Discussion of Non-ergodicity ============================ Because not enough research has bee conducted into the specific features of the dynamics of statistical mechanics systems connected with non-ergodic property, the discussion in this section will be informal and hypothetical in places. **Boltzmann’s hypothesis.** Boltzmann hypothesis (1871) postulates that in statistical mechanical systems the dynamic variables that are average in time are equal to the dynamic variables that are average across the (micro) canonical ensemble. We will only consider the quantum mechanics case. The discussion presented here was taken from \[7\]. Before we can formulate the Boltzmann hypothesis for a Bose gas with weak pair interaction, some designations need to be introduced. Let $\langle a\rangle_{\beta,\vec{v},\mu}$ be the Gibbs state corresponding to the inverse temperature $\beta$, the system velocity $\vec{v}$ and its chemical potential $\mu$, i.e. $$\begin{aligned} \langle a\rangle_{\beta,\vec{v},\mu}=\frac{1}{Z_{\beta,\vec{v},\mu}} {\rm tr \mit} (a e^{-\beta(H-\mu N+\vec{v}\vec{P})}),\end{aligned}$$ where $a$ is an arbitrary element from the canonical commutative relations algebra, $H$ is the Hamiltonian of the system, $N$ is the operator of the number of particles in the system, $\vec{P}$ is the operator of the system impulse and $Z$ is the so called large statistical sum, i.e. $$\begin{aligned} Z_{\beta,\vec{v},\mu}={\rm tr \mit} (e^{-\beta(H-\mu N+\vec{v}\vec{P})}). \end{aligned}$$ Let $V'_G$ be a space of all linear functionals on the algebra of canonical commutative relations, each of which is a superposition of Gibbs states. Let $\langle\cdot\rangle$ be is a translation-invariant functional on the algebra of canonical commutative relations such that $\forall t \in \mathbb{R}$ $\langle e^{itH}(\cdot) e^{-itH}\rangle$ is a well defined linear functional. The Boltzmann hypothesis states that in this case there will be a linear functional $\langle\cdot\rangle' \in V'_G$ such that for any element $a$ from the algebra of commutative relations: $$\begin{aligned} \lim \limits_{T\rightarrow +\infty}\frac{1}{T} \int \limits_{0}^{T}\langle e^{itH} a e^{-itH}\rangle dt=\langle a\rangle'. \label{Bol}\end{aligned}$$ Let us demonstrate that if formulated in this way the Boltzmann hypothesis is wrong. Theorem 1 states that there exists a translation invariant stationary linear functional on the algebra of canonical commutative relations and that said functional is not a superposition of Gibbs functions. However, this functional is built only in the form of a formal power series on the interaction constant and nothing is known about whether this series has a finite sum or not. \[7\], however, demonstrates that there is a way around the issue of whether this series converges or not and that there exists a stationary translation invariant functional on the algebra of canonical commutative relations $\langle\cdot\rangle_s$ that does not belong to $V'_G$. If this composite function is substituted for $\langle\cdot\rangle$ in the left side of equation (\[Bol\]), then the left side itself will equal $\langle\cdot\rangle_s$ and from equation (\[Bol\]) it will then follow that $\langle\cdot\rangle_s$ is superposition of Gibbs states. This contradiction shows that the Boltzmann hypothesis does not hold. **Discussion of the relation between a system’s behaviour on its boundaries and its tendency towards thermodynamic equilibrium.** Non-ergodicity means that Bose gas (with weak pair interaction) will not tend towards thermodynamic equilibrium in an infinite volume. Consequently to prove the tendency towards thermodynamic equilibrium we have to take into account its behaviour on the boundaries. In fact if the role played by the boundaries is disregarded we may end up with an infinite system. In \[7\] the role of boundaries is illustrated with Bogolubov’s derivation of Boltzmann’s kinetic equation. Bogolubov’s programme for deriving equations considers an infinite strengthening chain of equations for partial distribution functions (BBGKI chain) equivalent to Luiville’s equations, and then closed kinetic equations are obtained by selecting the suitable conditions for breaking the chain (correlation breaking conditions ). The first member of this chain has the form: $$\begin{aligned} \frac{\partial}{\partial t}\rho_1((x_1p_1)|t)+\frac{{p}}{m}{\nabla}\rho_1((x_1,p_1)|t) +\int dx_2dp_2 \frac{{p_2}}{m}\frac{\partial}{\partial {x}_2}\rho_2((x_1,p_1),(x_2,p_2)|t)=\nonumber\\ =\int dx_2dp_2\frac{\partial V(x_1-x_2)}{\partial {x}_1}\frac{\partial \rho_2((x_1,p_1),(x_2,p_2)|t)}{\partial p_1}. \label{II}\end{aligned}$$ Here $t$ is time, $m$ is the mass of the particles, and $V$ is the potential of pair interaction. An integral over the coordinates in the last term in the equation on the left (\[II\]) may be transformed into an integral along the boundary of a three-dimensional space using the Gauss theorem, and usually this fact is used as the basis for assuming that this integral equals to zero. However in \[7\] it is demonstrated if the last term on the left side of (\[II\]) is not discarded and Bogolubov’s strategy for deriving a Boltzmann equation is followed (if translation-invariant distributions are considered from the start), then this term will eventually cancel out with the collision integral and the system will have no kinetic evolution. Even though, it is possible that this effect is purely mathematical, it does make one wonder about the role that the system’s behaviour on its boundaries plays in proving its tendency towards thermodynamic equilibrium. Although it is now clear that the system’s behaviour on the boundary must be taken into account when proving its tendency towards thermodynamic equilibrium, the following question now arises: if the linear sizes of the system equal $\sim L$ and tend to infinity, then the volume of the system will b ehave like $L^3$ while the surface area of its boundary will equal $L^2$ which makes one think that the influence of the boundaries must be very insignificant. A similar situation can be observed in the theory of second order phase transitions, for instance in the theory of ferromagnetism. If, for instance, you take a two dimensional Izing model at a temperature below the critical level, and assume that all the spins are directed upwards along the boundaries of the crystal, then Peierls estimates (see \[8\]) show that in the thermodynamic limit there will be a non-zero magnetization. Second order phase transitions are often related to breaking of some symmetry. In the case of the Izing model, the symmetry that is broken is the symmetry relative to the change in the direction of all the spins in the grid. In the case of kinetic evolution there is also as symmetry that gets broken, it’s the symmetry relative to the time sign (Boltzmann’s H-theorem). We will have a more detailed discussion of this similarity later on. The necessity to take into account a system’s behaviour on its boundaries when proving the system’s tendency towards thermodynamic equilibrium raises an interesting philosophical question. If a system’s behaviour on its boundaries must be taken into account when proving the system’s tendency towards thermodynamic equilibrium, then it means we have to know about the behaviour of the particles outside the system’s boundaries because those influence the behaviour of the particles immediately inside the system’s boundaries, in other words we have to know how the environment that the system is a part of behaves. This in turn means that we have to know the behaviour of the larger environment around the immediate environment of the system and so on and so forth. So how do we escape from this infinite regress? To answer this question we will consider the following example. It is a known fact that there are black holes in our universe. Black holes are known to emit energy but for this phenomenon to be explained certain boundary conditions have to be assumed, which essentially are based on the fact that a black hole is formed as a result of the collapse of a mass body. According to one theorem of Penrose, this means that at one time in the past there existed a certain feature of the space time. The surface of a black hole, the so called event horizon, is the boundary at which the dependence of the particles inside the system on particles outside the system is broken. In fact due to the structure of light cones, information on the event horizon can only fall into the black hole, while the Hawking radiation is in equilibrium and therefore carries only the bare minimum of information. On the other hand one of Penrose-Hawking’s theorems \[9\] (whose conditions are confirmed by reliable astronomical observations) postulates that at some time in the past there was a singularity of space-time which can be equated to the Big Bang. At the same time Boltzmann also posited (on the basis of his $H$-theorem) that there had to have been a moment in time which can be equated to the beginning of the universe. In the theorems of Penrose and Hawking there appears entropy (equal to the surface area of the event horizon), which increase and is related to the temperature of the Hawking radiation through a known thermodynamic relation. It follows from all of this that there is a much deeper link between the ’proof’ of Boltzmann’s Big Bang and Penrose — Hawking’s singularity than a simple mathematical similarity. It is also worth noting that due to the arguments presented above, the non-ergodicity of Bose gas with weak pair interaction can be interpreted as the impossibility of deriving macroscopic dynamics from microscopic dynamics. The non-ergodicity theorem, as follows from its proof (see \[7\]), is applicable to a broad range of systems in statistical mechanics and we make the assumption that it can also be applied to the Ling’s cell. Therefore, the general results obtained for all non-ergodic systems can also be applied in this case. A more detailed study of how the properties of a system’s surface influence the process whereby it achieves thermodynamic equilibrium can be helpful in improving our understanding of living cells and the role of the cell’s surface in the physiological processes inside it. **Analogy with second order phase transitions.** Let’s discuss in more detailed the aforementioned analogy between the broking of the symmetry relative to the time sign and second order phase transitions. We’ll do this by discussing the Green-Cubo formula that describes the linear response of a system to small perturbations. Let there be $n$ fluctuating variables in the system $x_1,...,x_n$ and let $\hat{x}_1,...,\hat{x}_n$ be corresponding quantum mechanics operators. Let us assume that the fluctuations are sufficiently small and that the system’s entropy can be expressed with the formula $$\begin{aligned} S=S_0- \beta_{ij}x_ix_j,\end{aligned}$$ where $\beta_{ij}$ is some positively defined matrix (here and through the end of this section indexes repeated twice mean summation). Let $X_1,...,X_n$ be thermodynamically conjugate values to $x_1,...,x_n$ i.e. $$\begin{aligned} X_k=-\frac{\partial S}{\partial x_k}.\end{aligned}$$ The equation describing the relaxation of this system look like this: $$\begin{aligned} \dot{x}_i=-\gamma_{ik}X_k,\label{Lars}\end{aligned}$$ where $\gamma_{ik}$ - are the so called kinetic coefficients. This coefficients satisfy Onsanger’s symmetry principle: $$\begin{aligned} \gamma_{ik}=\gamma_{ki}.\end{aligned}$$ If the system is affected by external forces $f_k\;k=1,...,n$, in other words if the system’s Hamiltonian looks like this $$\begin{aligned} H=H_0+f_k\hat{x}_k,\end{aligned}$$ ($H_0$ is the Hamiltonian of an unperturbed system), then we can determine the generalized sensitivity of the system $\alpha_{ik}(\tau)$ in such a way that $$\begin{aligned} x_k(t)=\int \limits_{-\infty}^{+\infty} \alpha_{ki}(t-\tau)f_i(\tau)d\tau.\end{aligned}$$ Let’s use $\tilde{\alpha}_{ik}(\omega)$ to designate the Fourier transform of the generalized sensitivity $$\begin{aligned} \tilde{\alpha}_{ik}(\omega)=\int e^{i\omega t} \alpha_{ik}(t) dt.\end{aligned}$$ The generalized sensitivity can be found from the Green — Cubo formula: $$\begin{aligned} \alpha_{ik}(t)=-i\theta(t)\langle[\hat{x}_i(t),\hat{x}_k(0)]\rangle, \label{Vos1}\end{aligned}$$ where $\theta(t)$ is a Heaviside step function that equals 1 when its argument is positive and zero if its negative and $\langle\cdot\rangle$ is the averaging over the equilibrium state of an unperturbed system. It has to be noted, however, that when the Green— Cubo formula is applied the system is assumed to be in equilibrium with $t=-\infty$. If we assume, on the contrary, that the system is in equilibrium with $t=\infty$, then by transforming the time in the Green Cubo — formula, we will get the following formula for the generalised sensitivity $$\begin{aligned} \alpha'_{ik}(t)=i\theta(-t)\langle[\hat{x}_i(t),\hat{x}_k(0)]\rangle.\end{aligned}$$ It can be easily seen that $$\begin{aligned} \tilde{\alpha}'_{ik}(\omega)=(\tilde{\alpha}_{ik}(\omega))^\star,\end{aligned}$$ where the asterisk means complex conjugation. If, however, we start with an approximation in which relaxation equations (\[Lars\]) hold, then it can be shown that there exists the following relations between the matrices $\alpha,\beta$ and $\gamma$: $$\begin{aligned} \tilde{\alpha}_{ik}(\omega)=\frac{1}{T}(\beta_{ik}-i\omega\gamma_{ik}^{-1})^{-1}.\label{Vos2}\end{aligned}$$ It can be seen from this formula that if $t$ is replaced with $-t$ and vice versa in the Green – Cubo formula, then $\gamma_{ik}$ will become $-\gamma_{ik}$ and will no longer describe kinetic evolution. So, how can we choose between the requirement for equilibrium with $t=-\infty$ and with $t=+\infty$, in other words how can we determine whether the equation with a retarded commutator (\[Vos1\]) or the equation with a advanced commutator (\[Vos2\]) should be used for the generalized sensitivity of the system? It should be noted that a similar problem arises in classical electrodynamics when working out the formula for retarded potentials. A vector potential $A^\mu,\;\mu=0,1,...,3$ in classical electrodynamics satisfies the heterogeneous wave equation: $$\begin{aligned} \Box A^\mu=J^\mu,\end{aligned}$$ where $$\begin{aligned} \Box=\frac{\partial^2}{(\partial x^0)^2}-\frac{\partial^2}{(\partial x^1)^2}-\frac{\partial^2}{(\partial x^2)^2}-\frac{\partial^2}{(\partial x^2)^3},\label{Vawe}\end{aligned}$$ and $J^\mu$ is a four-vector of current, satisfying the continuity equation: $$\begin{aligned} \partial_\mu J^\mu=\frac{\partial}{\partial x^0}J^0+\frac{\partial}{\partial x^1}J^1+\frac{\partial}{\partial x^2}J^2+\frac{\partial}{\partial x^3}J^3=0.\end{aligned}$$ Where $x_0$ is time, and $x_1,x_2,x_3$ are the spacial Euclidean coordinates. The solution to equation (\[Vawe\]) usually has the form: $$\begin{aligned} A^\mu(x)=\int D(x-y)J^\mu(x)d^4x, \label{Zap}\end{aligned}$$ where $D(x)$ is the fundamental solution to the wave equation: $$\begin{aligned} \Box D(x)=\delta(x),\end{aligned}$$ with a support in the upper light cone: $$\begin{aligned} {\rm supp \mit} D(x)\subset \bar{V}^+= \{x|x_0\geq 0,\;x^2\geq 0\}.\end{aligned}$$ There exists a single fundamental solution that satisfies this property. Formula (\[Zap\]) is called the formula of retarded potentials. The similarity to the Green—Cubo formula becomes even closer if we remember that the vacuum mean of the commutators of the secondary quantum vector potential in electrodynamics is presented in the following form $$\begin{aligned} \langle 0|[A^\mu(x),A^\nu(y)]|0\rangle\theta(x^0-y^0)=ig^{\mu\nu}D(x-y),\end{aligned}$$ where $g^{\mu\nu}$ is a metric tensor: $$\begin{aligned} g^{\mu\nu}=0,\;{\rm if \mit}\; \mu\neq\nu,\nonumber\\ g^{00}=1,\;g^{ii}=-1,\;i=1,2,3.\end{aligned}$$ If in (\[Vawe\]) we switch to the Fourier transform in time, then we will get: $$\begin{aligned} \{\omega^2+\frac{\partial^2}{(\partial x^1)^2}+\frac{\partial^2}{(\partial x^2)^2}+\frac{\partial^2}{(\partial x^2)^3}\}A^\mu(\omega,\vec{x})=-J^\mu(\omega,\vec{x}).\end{aligned}$$ The solution to this equation, corresponding to the formula of delayed potentials, will have the form: $$\begin{aligned} A^\mu(\omega,\vec{x})=-\frac{1}{4\pi}\int \frac{e^{i\omega|x-y|}}{|x-y|}J^\mu(\omega, \vec{y})d^3\vec{y}.\end{aligned}$$ If $S_R$ is a sphere with a radius $R$ and its center at zero and $R\rightarrow \infty$, then $A^\mu(\omega,\vec{x})$ on this sphere will satisfy the Sommerfeld radiation condition: $$\begin{aligned} \frac{\partial}{\partial\vec{n}}A^\mu(\omega,\vec{x})-i\omega A^\mu(\omega,\vec{x})=O(\frac{1}{R^2}),\end{aligned}$$ where $\frac{\partial}{\partial\vec{n}}$ means differentiation along the external normal to sphere $S_R$. By applying a inverse Fourier transform to this formula, we will find the condition that must be met by $S_R$, to ensure that a retarded commutator is selected in the formula for retarded potentials: $$\begin{aligned} \frac{\partial}{\partial\vec{n}}A^\mu(x)+\frac{\partial}{\partial x^0} A^\mu(x)=O(\frac{1}{R^2}). \label{Zom}\end{aligned}$$ The selection of a advanced commutator would result in the following formula $$\begin{aligned} \frac{\partial}{\partial\vec{n}}A^\mu(x)-\frac{\partial}{\partial x^0} A^\mu(x)=O(\frac{1}{R^2}).\end{aligned}$$ Thus we can see that a advanced or a retarded commutator in Green—Cubo formulas can be selected (which is necessary to specify the direction of time), at least in electrodynamics, by applying appropriate boundary conditions on the infinitely removed boundary of a three dimensional space. This approach is in full accord with our ideas about the role that the boundaries of a system play in proving its tendency towards thermodynamic equilibrium. It should also be noted that boundary condition (\[Zom\]) imposes a limit on the flow of information (this interpretation is also possible): information can only go out of the system (flow outside). This restriction on the direction of the flow of information is realized, for instance, on the event horizon of a black hole due to the structure of the light cones on it, as has already been noted above. Let us go back to our analogy with ferromagnetism (Izing model). We said earlier that non-zero magnetization can be obtained by changing the direction of the spins on the boundary of the grid. However, non-zero magnetization can also be achieved in another way. We can introduce a fictitious infinitesimally small external field, which is expressed by adding the following expression to the system’s Hamiltonian $$\begin{aligned} h\sum \limits_{p} \sigma_p,\end{aligned}$$ where the summation is done for all the nodes $p$ of the grid, and $\sigma_p$ is the spin in the node of the grid with $\sigma_p\in \{-1,1\}$. Then the system passes to an unlimited volume and after that $h$ starts approaching zero, however it always remains positive or negative. The means calculated in this way are called quasi-means (Bogolubov’s quasi-means). It turns out that non-zero magnetization occurs in this case as well (and it’s equal to the magnetization that results when the spins on the boundaries are given a certain direction). It begs the question whether it is possible to make the right choice between a advanced and retarded commutator method when working out a Green—Cubo formula, by analogy with Bogolubov’s quasi-means? The answer to this question is positive and here is why. Let $H(t)$ be a Hamiltonian of the system dependent on time: $$\begin{aligned} \hat{H}(t)=\hat{H}_0+\hat{V}f(t),\end{aligned}$$ Let $f(t)$ be an external and infinitesimally small classical force changing with time. The von Neumann equation for the density matrix will look like this : $$\begin{aligned} \frac{d}{dt}\rho(t)=-i[\hat{H}(t),\rho(t)]=-i[\hat{H}_0,\rho(t)]-if(t)[\hat{V},\rho(t)].\end{aligned}$$ If $f(t)\equiv 0$, $\rho=\rho_0$ is the density matrix of the equilibrium of an unperturbed system. Consequently if we’re looking for $\rho$ in the first order over $f(t)$, then the following equation can be used for $\rho$: $$\begin{aligned} \frac{d}{dt}\rho(t)=-i[\hat{H}_0,\rho(t)]-if(t)[\hat{V},\rho_0]. \label{vn}\end{aligned}$$ The standard method for solving this equation when working out a Green — Cubo formula involves choosing a boundary condition $\rho(t)\rightarrow \rho_0$ with $t\rightarrow -\infty$ and solving (\[vn\]) by using the constant variation method. What we get then is: $$\begin{aligned} \rho=\rho_0-i\int \limits_{-\infty}^t e^{-i\hat{H}_0(t-\tau)}[\hat{V},\rho_0]e^{+i\hat{H}_0(t-\tau)} f(\tau)d\tau. \label{GK1}\end{aligned}$$ This expression for the density matrix immediately leads to a Green—Cubo formula with a retarded commutator. Instead of imposing initial conditions on the density matrix, we will introduce an infinitesimally small term $\varepsilon\rho(t)$, $\varepsilon>0$, $\varepsilon\rightarrow 0$ into the left side of (\[vn\]). In other words instead of (\[vn\]) we will be solving the following equation for the density matrix: $$\begin{aligned} \frac{d}{dt}\rho(t)+\varepsilon\rho+i[\hat{H}_0,\rho(t)]=-if(t)[\hat{V},\rho_0].\label{vn1}\end{aligned}$$ Applying a Fourier transform to this equation we get: $$\begin{aligned} (-i\omega+\varepsilon+i[\hat{H}_0])\tilde{\rho}(\omega)=-i\tilde{f}(\omega)[\hat{V},\rho_0],\label{v1}\end{aligned}$$ where $[\hat{H}_0]$ designates the following super-operator: $$\begin{aligned} [\hat{H}_0]\rho:=[\hat{H}_0,\rho].\end{aligned}$$ Therefore it follows from (\[v1\]) that: $$\begin{aligned} \rho(\omega)=\frac{-i}{-i\omega+\varepsilon+i[\hat{H}_0]}[\hat{V},\rho_0]f(\omega).\end{aligned}$$ Since a Fourier transform transforms convolution into product we can find: $$\begin{aligned} \rho(t)=\int \limits_{-\infty}^{+\infty}\alpha(t-\tau)f(\tau)d\tau,\label{v2}\end{aligned}$$ where $$\begin{aligned} \alpha(t)=\frac{1}{2\pi}\int \limits_{-\infty}^{+\infty} e^{-i\omega t}\frac{-i}{-i\omega+\varepsilon+i[\hat{H}_0]}[\hat{V},\rho_0]d\omega.\label{v3}\end{aligned}$$ The super-operator $[\hat{H}_0]$ is the Liouvillian of an unperturbed system and is therefore self-adjoint (if we define the scalar product of the operators applicable in the space of states as $\langle\hat{A},\hat{B}\rangle={\rm tr \mit} (\hat{A}^\star\hat{B})$). Then, calculating the integral in (\[v3\]) through residue we will get: (\[v3\])$$\begin{aligned} \alpha(t)=-i\theta(t)e^{(-i[\hat{H}_0]-\varepsilon)t}[\hat{V},\rho_0] =\theta(t)e^{-\varepsilon t}e^{-i\hat{H}_0t}[\hat{V},\rho_0]e^{i\hat{H}_0t}\end{aligned}$$ Substituting the expression we just found in (\[v2\]) for $\alpha(t)$ and assuming that $\varepsilon\rightarrow 0$ we will find: $$\begin{aligned} \rho=\rho_0-i\int \limits_{-\infty}^t e^{-i\hat{H}_0(t-\tau)}[\hat{V},\rho_0]e^{i\hat{H}_0(t-\tau)} f(\tau)d\tau,\end{aligned}$$ Thus we’ve once again got expression (\[GK1\]) for the density matrix. Thus we see that boundary conditions for a density matrix can be chosen using a method similar to Bogolubov’s method of quasi means, i.e. by introducing an appropriate infinitesimally small term in the equations describing the evolution of the density matrix through time. This should explain the similarity between spontaneous breaking of the symmetry in second order phase transitions and the breaking of the time symmetry in kinetic equations. Thermodynamics of stationary non-equilibrium states =================================================== In section 1 we said that according to the results obtained in \[7\], even the most realistic systems in statistical mechanics are non-ergodic. It should also be reminded at this stage, that a system is ergodic if its trivial first integrals (Hamiltonian, impulse, number of particles) form the maximum set of first integrals in involution. Hereinafter for the sake of brevity we will use the term energy to mean the energy and the impulse of a system. From the non-ergodicity of a system it follows that it must have stationary states that are not super-positions of Gibbs states. We will call the states of a system that are super-positions of Gibbs states equilibrium states. It therefore makes sense to build thermodynamics of non-equilibrium stationary states. This thermodynamics will really come in handy in the next section when we formulate the principle that the substance inside a living cell is in a non-equilibrium stationary state, which is the goal of this paper. This section describes how we build this thermodynamics. The equations we derive in this section will be applicable to any system in statistical mechanics, including the Ling’s cell. For the sake of simplicity we will only consider the case of classical mechanics. Some basic concepts of classical (Hamiltonian) mechanics should be remembered here. The phase space $\Gamma$ of a Hamiltonian system is a $2n$-dimensional $n = 1,2,...$ smooth manifold in which for any two (good) functions f, g there is a function $(f,g)$ defined on their Poisson brackets, which satisfies the following conditions: 1\) $(fg, h) = f(g, h) + g(f, h)$ (Leibniz rule) 2\) $(f,g) = -(g,f)$ (Anti-symmetry) 3\) $(f, (g, h) + (g, (h, f)) + (h, (f,g)) = 0$ (Jacobi identity) 4\) For any point $x \in \Gamma$, and for any function $f$: ${\rm grad \mit} f(x)\neq 0$, there is a function $g$ on $\Gamma$ such that $(f,g)(x)\neq 0$ (non-degeneracy). $n$ is called the number of degrees of freedom of the system. An example of a phase space would be $\Gamma=\{(p_1,...,p_n,q_1,...,q_n)\mid p_i,q_i \in \mathbb{R}\}$ in which the Poisson bracket is defined as: $$\begin{aligned} (f,g)=\sum \limits_{i=1}^n(\frac{\partial f}{\partial p_i}\frac{\partial g}{\partial q_i}-\frac{\partial f}{\partial q_i}\frac{\partial g}{\partial p_i}). \label{Kan}\end{aligned}$$ According to the Darboux theorem (see \[10\]) local coordinates $p_1,...,p_n,q_1,...,q_n$ can always be introduces in which the Poisson bracket will look like (\[Kan\]). Such coordinates are called canonical or symplectic. The evolution of a Hamiltonian system is described by a one-parameter group of diffeomorphisms (i.e. homeomorphisms which are smooth with their inverse), $x\mapsto G_\tau(x),\;\tau \in \mathbb{R}$ preserving the Poisson bracket ($\tau$ is time). This one-parameter group of diffeomorphisms is called a phase flow. It can be shown that for any phase flow $G_\tau$ (locally) there is a function $H$ on $\Gamma$, called a Hamiltonian, such that $$\begin{aligned} \frac{d}{d\tau}f_\tau=(H,f_\tau),\end{aligned}$$ Where $f_\tau(x):=f(G_\tau(x))$. For more about Hamiltonian dynamics see (\[10\]). Thus, let’s assume we have a Hamiltonian system in which in addition to Hamiltonian $H$ there is also , $k = 1,2,3...$ first integrals $K_1,...K_k$ in involution, i.e. all the matching Poisson brackets equal zero: $\forall i,j=1,...,k\;(K_i,K_j)=0$. We want to describe a state of our system in which it is in equilibrium with its environment and in which its Hamiltonian $H$ and the first integrals $K_1,...,K_k$ take on some specific values $E,\;K'_1,...,K'_k$ respectively. The requirement that the first integrals $K_1,..., K_k$ must be in involution can be explained as follows: we want to describe states in which the integrals of motion $K_1,...,K_k$ take on specific values: $E,\;K'_1,...,K'_k$. But in quantum mechanics (see \[11\]) simultaneous measurability of the observables means that they are commutative. Now it becomes clear why $K_1,...,K_k$ must be in involution, if we take into account the fact that the commutator is the quantum mechanics counterpart of the Poisson bracket (see \[12\]). Let us first consider the way the above task is solved with $k = 0$. The state of a system in statistical mechanics is described by a distribution function $\rho(x)$ on the phase space $\Gamma$, i.e.by a function that satisfies the following conditions a\) $\forall x \in \Gamma \rho(x)\geq 0$ (Positive definiteness), b\) $\int \rho(x)d \Gamma_x=1$ (Normalization). Here the phase volume is determined in the following way: in canonical coordinates it looks like $d\Gamma=dp_1...dp_ndq_1...dq_n$. It can be demonstrated that this definition is correct. The observed macroscopic value $\bar{F}$ of dynamic variable $F (x)$ has the form: $$\begin{aligned} \bar{F}=\int F(x)\rho(x)d \Gamma_x. \label{SR}\end{aligned}$$ The role of the distribution function $\rho(x)$ is assigned to the so called micro-canonical Gibbs distribution: $$\begin{aligned} \rho(x)=c\delta(H(x)-E),\end{aligned}$$ In which the constant $c$ is chosen on the basis of the normalization condition. It’s important to note here that we assumed that the system under consideration is in thermodynamic equilibrium with its environment. For this reason $H(x)$ takes into account interaction with the particles in the environment and $H(x)$ is a function of time and time is included in $H(x)$ through the coordinates and impulses of the particles in the environment. This will be important further down the road. It has also be remembered, however, that in the thermodynamic limit these extra terms in $H(x)$ that account for the interaction with the environment, can be disregarded. If the system’s Hamiltonian is a function of time, $H(t)$, then the system’s evolution will be described by a co-cycle, i.e. $\forall t_1,t_2$ there are canonical transformations (i.e. transformation that preserves the Poisson bracket) $x\mapsto G_{t_1,t_2}(x)$ meets the co-cycle property: $$\begin{aligned} G_{t_1,t_2}G_{t_2,t_3}=G_{t_1,t_3}.\end{aligned}$$ Any dynamic variable $f$ is a function of time by definition $f_t(x):=f(G_{t,0}(x))$. Now it has to be said that in actuality, the motion of the system is described by some trajectory through the phase space and for this reason the left side of (\[SR\]) has to be interpreted as a mean over time, or $$\begin{aligned} \lim \limits_{T\rightarrow 0}\frac{1}{T} \int \limits_{0}^{T}F(G_{\tau,0}(x))d\tau=c\int F(x)\delta(H(x)-E)d\Gamma_x. \label{SR1}\end{aligned}$$ Equation (\[SR1\]) does not in any way mean that the system is ergodic, because from the very beginning we said that our system is a subsystem of a large system. A phase flow preserves its phase volume(see \[10\]). Therefore if is the system’s distribution function, then $\rho$ Consequently, over time the distribution function will evolve into: $$\begin{aligned} \rho_t(x)=\rho(G_{0,t}(x)).\end{aligned}$$ Function pt meats the conditions of the following differential equation: $$\begin{aligned} \frac{d}{dt}\rho_t=(\rho_t,H(t)).\end{aligned}$$ Now let us assume that the system’s Hamiltonian $H_\lambda$ is a function of some parameter $\lambda$. We will demonstrate that if $\lambda$ undergoes adiabatic changes the micro-canonical distribution remains the same. Let $\lambda(t)$ be some function of time and let $\varepsilon$ be some small positive number. Let $\lambda_\varepsilon(t):=\lambda(\varepsilon t)$. Let $x$ be a point in the phase space such that $H_{\lambda(\varepsilon t)}(x)=E$. Let $[0,\Delta]$ be a sufficiently long period of time such that a mean over this period is the same as the mean over the micro-canonical ensemble. Let us now select such a small $\varepsilon$ that will not change very much in this time period. Let’s calculate $$\begin{aligned} H_{\lambda_\varepsilon(t)}(G_{t,0}(x))|_{t=\Delta}.\end{aligned}$$ What we get is: $$\begin{aligned} H_{\lambda_\varepsilon(t)}(G_{t,0}(x))|_{t=\Delta}=\nonumber\\ =H_{0}((x))+\int \limits_{0}^\Delta dt \{\frac{d\lambda}{dt}\frac{\partial H_{\lambda_\varepsilon(t)}}{\partial\lambda}+(H_{\lambda_\varepsilon(t)}, H_{\lambda_\varepsilon(t)})\}(G_{t,0}(x))=\nonumber\\ =H_{0}((x))+\int \limits_{0}^\Delta dt\frac{d\lambda}{dt}\frac{\partial H_{\lambda_\varepsilon(t)}}{\partial\lambda}(G_{t,0}(x))=\nonumber\\ =H_{0}((x))+\Delta \frac{d\lambda}{dt}\overline{\frac{\partial H_{\lambda_\varepsilon(t)}}{\partial\lambda}}+O(\varepsilon^2)\end{aligned}$$ Thus with a precision of up to $\varepsilon^2$ canonical transformation $G_{t,0}$ transforms function $H_{\lambda_\varepsilon(t)}(x)-E-\Delta \frac{d\lambda}{dt}\overline{\frac{\partial H_{\lambda_\varepsilon(t)}}{\partial\lambda}}$ into $H_{\lambda_\varepsilon(0)}(x)-E$, i.e. transformation $G_{t,0}$ transforms the micro-canonical distribution $c\delta(H_{\lambda_\varepsilon(0)}(x)-E)$ into $c\delta(H_{\lambda_\varepsilon(t)}(x)-E-\Delta \frac{d\lambda}{dt}\overline{\frac{\partial H_{\lambda_\varepsilon(t)}}{\partial\lambda}})$ with a precision of up to $\frac{1}{\varepsilon}$. The function $\lambda_\varepsilon(t)$ changes by a finite amount within an interval of order $\frac{1}{\varepsilon}$. Therefore, if we break down this time period into sub-intervals of duration $\Delta$ we will find that after time period $\frac{1}{\varepsilon}$ elapses the form of the micro-canonical distribution will change by a value of order $\varepsilon$. If we have an $\varepsilon$ approaching zero, we can prove the above contention. Now we will consider the case when the system has $k$- first integrals in involution. By analogy with the micro-canonical distribution we will assume that in this case if the values of the parameters $E,\;K'_1,...,K'_k$ are fixed the system’s distribution function will have the form: $$\begin{aligned} \rho(x)=c\delta(H(x)-E)\prod \limits_{i=1}^k\delta(K_i(x)-K'_i), \label{MKdis}\end{aligned}$$ where$c$ is a normalizing multiplier. First of all let us note that this definition is correct, i.e. the form of the distribution will not change if the integrals $K_1,..., K_k$ are replaced with the functions $\hat{K}_1,...,\hat{K}_k$ of these integrals and the Hamiltonian. In fact: $$\begin{aligned} c\delta(H-E)\prod \limits_{i=1}^k\delta(\hat{K}_i-\hat{K}'_i)=\nonumber\\ =c\delta(H-E)\prod \limits_{i=1}^k\delta(K_i-K'_i)\frac{D(H, K_1,...K_k)}{D(H, \hat{K}_1,...\hat{K}_k)},\end{aligned}$$ where $$\begin{aligned} \frac{D(H, K_1,...K_k)}{D(H, \hat{K}_1,...\hat{K}_k)}\end{aligned}$$ - is the Jacobian of the transformation of the variables $H,\; K_1,...K_k$ into the variables $H,\; K_1,...K_k$. This Jacobian is constant on the shared level surface of the functions $H,\; K_1,...K_k$, which proves the correctness of the definition. We will call distribution (\[MKdis\]) the generalized micro-canonical distribution. It has to be said that the generalized micro-canonical distribution can be derived from the requirement that the Von Neumann entropy must achieve its maximum if the energy value and the values of the first integrals $K_1,...,K_k$, are fixed. In other words, the generalized micro-canonical distribution corresponds to a state of the system in which the information that is available about it (with the fixed values of the Hamiltonian and the first integrals $K_i,...,K_k$ is minimal. Using a method similar to that we used above, we can demonstrate that the micro-canonical distribution does not change its form as a result of quasi-static processes. The following formula has to be noted here: if the Hamiltonian $H$ and the first integrals $K_1,...,K_k$ are functions of parameter $\lambda$, then in quasi stationary processes: $$\begin{aligned} (\frac{\partial K'_i}{\partial\lambda})_{qs}=\overline{\frac{\partial K_i}{\partial\lambda}},\end{aligned}$$ where $qs$ means that the derivative is calculated along the quasi-static process. We will now define the entropy corresponding to the micro-canonical distribution as: $$\begin{aligned} S(E,K'_1,...,K'_k)=\rm ln \mit W(E,K'_1,...,K'_k),\end{aligned}$$ Where $W(E,K'_1,...,K'_n)$ is the so called statistical weight: $$\begin{aligned} W(E,K'_1,...,K'_k)=\int d\Gamma_x \delta(H(x)-E)\prod \limits_{i=1}^k\delta(K_i(x)-K'_i).\end{aligned}$$ Let’s found out how correct this definition is, i.e. how will the entropy change if we move from the integrals $H,\;K_1,...,K_k$ to the integrals $H,\;\hat{K}_1,...,\hat{K}_k$. What we get is: $$\begin{aligned} W(E,K'_1,...,K'_n)=c\int d\Gamma_x \delta(H(x)-E)\prod \limits_{i=1}^n\delta(\hat{K}_i(x)-\hat{K}'_i),\end{aligned}$$ where $c$ is the transformation Jacobian. The appearance of this multiplier next to the statistical weight will create an additional term in the entropy $\rm ln \mit\, c$. It has to be noted here, that a similar problem also arises in standard thermodynamics ($k = 0$), where this parameter is simply discarded because it is assumed that it’s influence in the thermodynamic limit is infinitesimally small. We will do the same here. It also has to be noted that instead of (\[MKdis\]) the following formula can be used for the micro-canonical distribution: $$\begin{aligned} \rho(x)=c\Delta(H(x)-E)\prod \limits_{i=1}^k\Delta(K_i(x)-K'_i), \label{Delta}\end{aligned}$$ Where function $\Delta$ equals one in the small neighborhood of zero and equals zero outside this neighborhood. In the thermodynamic limit, this neighborhood will be so narrow that $\Delta$-function can be replace with $\delta$-function, and we will get back to (\[MKdis\]). In light of the above notes, the statistical weight can be defined as: $$\begin{aligned} W(E,K'_1,...,K'_n)=\int d\Gamma_x \Delta(H(x)-E)\prod \limits_{i=1}^k\Delta({K}_i(x)-{K}'_i).\end{aligned}$$ It can be seen from this formula that the statistical weight can be interpreted as the number of microscopic states compatible with a given macroscopic state, or as the thermodynamic probability. Now let’s assume our system is in equilibrium with the environment. Let’s demonstrate that we can choose the first integrals $\hat{K}_1,...,\hat{K}_n$ in such a way that their values will remain the same as the system heats up. And indeed, the energy level E corresponds to the integral values $K'_1(E),...,K'_n(E)$. Let’s replace the integrals $K_1,...,K_n$ with the integrals $\hat{K}_1,...,\hat{K}_n$ using the formula $\hat{K}_i=K_i-K'_i(H)$. It can be seen that these integrals remain constant as the system heats up. We then assume that these integrals also remain constant when the system temperature increases. What we then have is $$\begin{aligned} (\frac{\partial K'_i}{\partial\lambda})_{qs}=\overline{\frac{\partial K'_i}{\partial\lambda}}.\end{aligned}$$ From which we can derive: But because: $$\begin{aligned} (\frac{\partial K'_i}{\partial\lambda})_{qs}=(\frac{\partial K'_i}{\partial\lambda})_{E}+(\frac{\partial K'_i}{\partial E})_\lambda (\frac{\partial E}{\partial\lambda})_{qs},\end{aligned}$$ and $(\frac{\partial K'_i}{\partial E})_\lambda=0$ we have $$\begin{aligned} (\frac{\partial K'_i}{\partial\lambda})_{E}=(\frac{\partial K'_i}{\partial\lambda})_{qs}=\overline{\frac{\partial K_i}{\partial\lambda}}. \label{74}\end{aligned}$$ It has to be noted that the entropy remains constant in a quasi-static process. Indeed, in a quasi-static process, the (generalized) micro-canonical distributions corresponding to the various parameters turn out to be bound, as was demonstrated above, by a canonical transformation and their entropy is the same because canonical transformations preserve the phase volume. One important result of this section is the conclusion that in the thermodynamics of stationary non-equilibrium states, the standard thermodynamic equations hold true. Let’s derive the standard thermodynamic equation: $$\begin{aligned} dE=TdS-PdV, \label{TD}\end{aligned}$$ where $T$ is the temperature, $P$ is the pressure, and $V$ is the volume of the system. Put by definition that $$\begin{aligned} \frac{1}{T'}=\frac{dS}{dE}\end{aligned}$$ and demonstrate that $T' = T$, i.e. that it is the same as the ordinary temperature. Let $E'$ be the energy of the environment around the system and $S(E')$ its entropy. We will assume that our system is in equilibrium with its environment (but it’s not in equilibrium with itself, being in a stationary non-equilibrium state.) The probability $P(E, E')$ that the system’s energy equals E, and the energy of the environment equals $\sim e^{S(E)+S'(E')}$ because we earlier defined the system’s entropy $S(E)$ as the logarithm of thermodynamic probability. If the system is in equilibrium with its environment, then $P(E, E')$ reaches its maximum (with an additional condition that $E+E'=\rm const \mit$), because equilibrium is the most probable state in this case. Consequently: $$\begin{aligned} (\frac{dS(E)}{dE}-\frac{dS'(E')}{dE'})dE=0,\end{aligned}$$ i.e. $$\begin{aligned} T'=T,\end{aligned}$$ and that is exactly what we’ve been trying to prove. Let us remind you that $S = \ln W$, where $$\begin{aligned} W=\int d\Gamma_x\delta(E-H(x))\prod \limits_{i=1}^n\delta(K'_i(V)-K_i(x)).\end{aligned}$$ Here we directly specify the dependence of $K_i$ on $V$. We just demonstrated that $(\frac{\partial S}{\partial E})_V=\frac{1}{T}$. Now all we’ve got left to do is to demonstrate that $(\frac{\partial S}{\partial V})_E=\frac{1}{T}P$. Hereinafter derivatives are taken with $V = V_0$ for some $V_0$. What we get is: $$\begin{aligned} (\frac{\partial S}{\partial V})_E=\nonumber\\ =\frac{1}{W} \int d \Gamma_x \{-\frac{\partial H(x)}{\partial V}\delta'(E-H(x))\prod \limits_{i=1}^{n}\delta(K'_i(V)-K(x))+\nonumber\\ +\delta(E-H(x))\sum \limits_{j=1}^n [\prod \limits_{i=1,\;i\neq j}\delta(K'_i-K_i(x))\delta'(K'_j-K_j(x))(\frac{\partial K'_j(V)}{\partial V}-\frac{\partial K_j(x)}{\partial V})]=\nonumber\\ =\frac{1}{W} \int d \Gamma_x \{-\frac{\partial H(x)}{\partial V}\delta'(E-H(x))\prod \limits_{i=1}^{n}\delta(K'_i-K(x))+\nonumber\\ +\sum \limits_{j=1}^n \frac{1}{W}\frac{\partial}{\partial K'_j}[W((\frac{\partial K'_j}{\partial V})_E-\overline{\frac{\partial K_j(x)}{\partial V}})]\}-\frac{\partial}{\partial V}(\frac{\partial K'_j(V)}{\partial K'_j(V_0)})\label{80}.\end{aligned}$$ Because of equation (\[74\]) the second to last term on the right of the last equation (\[80\]) equals zero. As far as the last term is concerned, the most reasonable assumption would be that $\frac{\partial K'_j(V)}{\partial K'_j(V_0)}\approx 1$ or $\frac{\partial K'_j(V)}{\partial K'_j(V_0)}=1-\rm const \mit \frac{V-V_0}{V_0}$, with $V\approx V_0$. Therefore the last term in (81) disappears in the thermodynamic limit. Thus the final result is: $$\begin{aligned} (\frac{\partial S}{\partial V})_E=-\frac{1}{W} \int d \Gamma_x \frac{\partial H(x)}{\partial V}\delta'(E-H(x))\prod \limits_{i=1}^{n}\delta(K'_i-K(x))=\nonumber\\ =-\frac{1}{W} \frac{\partial}{\partial E} \int d \Gamma_x \frac{\partial H(x)}{\partial V}\delta(E-H(x))\prod \limits_{i=1}^{n}\delta(K'_i-K(x))=\nonumber\\ =-\frac{1}{W}\frac{\partial}{\partial E} (\overline{(\frac{\partial H(x)}{\partial V})}W).\end{aligned}$$ But $-\overline{(\frac{\partial H(x)}{\partial V})}$, as was already demonstrated earlier, is the speed at which the energy changes in an adiabatic process, in which the changing parameter is the volume, i.e. pressure. Thus $$\begin{aligned} (\frac{\partial S}{\partial V})_E=\frac{1}{W}\frac{\partial}{\partial E}(PW)=\nonumber\\ =P\frac{\partial S}{\partial E}+\frac{\partial P}{\partial E}.\end{aligned}$$ But energy is an extensive value while pressure is an intensive one. For this reason in the thermodynamic limit $\frac{\partial P}{\partial E}=0$. Finally, remember that $\frac{\partial S}{\partial E}=\frac{1}{T}$ we get: $$\begin{aligned} (\frac{\partial S}{\partial V})_E=\frac{1}{T}P,\end{aligned}$$ This is what we’ve been trying to prove. U sing the standard method and applying the Legendre transformation we can introduce thermodynamic functions. The free energy F is defined as: $$\begin{aligned} F:=E-TS.\end{aligned}$$ Using (\[TD\]) we can find: $$\begin{aligned} dF=-SdT-PdV.\end{aligned}$$ Similarly we can define the thermodynamic potential $\Phi$: $$\begin{aligned} \Phi:=F+PV.\end{aligned}$$ What we get is: $$\begin{aligned} d\Phi=-SdT+VdP.\end{aligned}$$ If the number of particles in the system $N$ is variable then (\[TD\]) should be changed to: $$\begin{aligned} dE=TdS-PdV+\mu dN.\end{aligned}$$ Using the theorem of small additions we can find that: $$\begin{aligned} dF=-SdT-PdV+\mu dN,\nonumber\\ d\Phi=-SdT+VdP+\mu dN.\end{aligned}$$ Integrating the last equation with constant $V$ and $T$ between zero and $N$, we can find that: $$\begin{aligned} \Phi=\mu N.\end{aligned}$$ If the system has particles of various sorts numbered with $i = 1,..., M$ and $N_i$ is the number of particles of sort $i$, then in all the previous formulas the term $\mu d N$ needs to be replaced with $\sum \limits_{i=1}^M \mu_i dN_i$. We also have: $$\begin{aligned} \Phi=\sum \limits_{i=1}^M\mu_i N_i.\end{aligned}$$ In all of the above all the thermodynamic values, including the entropy, proved to be dependent on the values $K'_1,...,K'_k$ of the motion integrals $K_1,...,K_k$. Now we’re going to formulate the following **Equivalence principle.** The entropy $S(E,K'_1,...,K'_k)$ does not depend on $K'_1,...,K'_k$ (if all the other parameters of the system are constant). Here is how it can be substantiated. We assigned some specific value to the energy $E$, and for the sake of simplicity assume that $n=1$. We normalize $K':=K'_1$ in such a way as to ensure that $K'$ changes from $-1$ to $1$ and that at $0$ the entropy reaches its maximum. If $S(K')$ depends on $K'$ and our system is in thermodynamic equilibrium with its environment then it will be established such a value of $K'$ that corresponds to maximum entropy, i.e. $K' = 0$. In a sufficiently small neighborhood of $K' = 0$, $S(K')$ will look like this: $$\begin{aligned} S(K')=S_0-c{K'}^2,\end{aligned}$$ $c> 0$. Because our system is macroscopic and $c$ must be comparable to $S_0$, we have to assume that $c$ is very great. In this case the thermodynamic probability will look like this: $$\begin{aligned} W(K')=\rm const \mit e^{-c{K'}^2},\end{aligned}$$ i.e. $W(K')$ will have a pronounced peak at $K' = 0$, while all the other values of $K'$ will be extremely improbable. This means that, in other words, that it doesn’t really matter whether we’re averaging using the micro-canonical distribution $$\begin{aligned} \rho(x)=c\delta(E-H(x))\end{aligned}$$ or using the generalized micro-canonical distribution: $$\begin{aligned} \rho(x)=c\delta(E-H(x))\delta(K'-K(x)).\end{aligned}$$ Thus if $S(E, K')$ depends on $K'$ then we don’t have to use the micro-canonical distribution with a fixed value of the integral $K$. Thus the most general case which we can come across is where the entropy $S(H,K'_1,...,K'_k)$ depends only on the values of the integrals $K_1,...,K_l$, $l=1,...,k$ and does not at all depend on the values of the integrals $K_{l+1},...,K_k$. The integrals $K_1,...,K_l$ are assigned values corresponding to maximum entropy and the mean for the generalized micro-canonical distribution, corresponding to the values of the integrals $K_1=K'_1$,..$K_l=K'_l$, $K_{l+1}=K'_{l+1}$,...,$K_k=K'_k$ will equal the meant for the distribution: $$\begin{aligned} c\delta(H-E)\prod \limits_{i=l+1}^k\delta(K_i-K'_i).\end{aligned}$$ The observed values of the integrals $K_{l+1},...,K_k$ can be arbitrary. Now let us consider an example from the condensed matter physics where the general case above is realized. The case in hand is the superfluid helium. Let us consider a system of $N$ particles enclosed in a certain macroscopic volume $V$ and obeying the Bose statistics. The Hamiltonian of such system takes the form: $$\begin{aligned} H=\sum \limits_{i=1}^N \frac{p_i^2}{2m}+\sum \limits_{i=1}^N \sum \limits_{j=i+1}^N \Phi(|x_i-x_j|),\end{aligned}$$ where $\Phi(|x_i-x_j|)$ is the potential energy of the $i$-th and $j$-th particles. Let us assume that $$\begin{aligned} \nu(p)=\int \Phi(|x|)e^{ipx}d^3x.\end{aligned}$$ Then, in secondary quantization representation the Hamiltonian will take the form: $$\begin{aligned} \Gamma:=H-\mu N=\sum \limits_p (\frac{p^2}{2m}-\mu)b_p^+b_p+\nonumber\\ +\frac{1}{2V} \sum \limits_{p_1,p_2,p'_1,p'_2}\Delta(p_1+p_2-p'_1-p'_2)\nu(p_1-p_1')b^+_{p_1}b^+_{p_2}b_{p'_1}b_{p'_2}.\end{aligned}$$ Here $\Delta(x)$ is a the function equal to 1 when $x=0$ and equal to zero in all the other cases. $b_p^+$, $b_p$ are particle creation-annihilation operators in a state with the momentum of $p$. With low interaction between the particles almost all of them are in condensed state. And as $b_0,\;b_0^+\sim \sqrt{N}$, in commuting relation we can neglect 1 and consider $b_0$, $b_0^+$ commutation variables. Let us denote the number of particles in the condensate by $N_0$. Keeping only the terms that are quadratic by the creation-annihilation operators of the super-condensate particles in the Hamiltonian, we derive: $$\begin{aligned} \Gamma=-\frac{1}{2}\frac{N_0^2}{V} \nu(0)+\sum \limits_{p\neq0} \{\frac{p^2}{2m}+\frac{N_0}{V} \nu(p)\}b_p^+b_p+\nonumber\\ +\frac{N_0}{2V}\sum \limits_{p\neq0}\nu(p)(b^+_pb^+_{-p}+b_{-p}b_p).\end{aligned}$$ This Hamiltonian can be diagonalized by the canonical Bogoliubov transformation \[13\]: $$\begin{aligned} \xi_p:=\frac{b_p-A_pb^+_{-p}}{\sqrt{1-A_p^2}},\nonumber\\ \xi_p^+:=\frac{b_p^+-A_pb_{-p}}{\sqrt{1-A_p^2}},\end{aligned}$$ where $$\begin{aligned} A_p=\frac{V}{N_0\nu(p)}\{E(p)-\frac{p^2}{2m}-\frac{N_0}{V}\nu(p)\},\nonumber\\ E(p)=\sqrt{\frac{N_0}{V}\frac{p^2\nu(p)}{m}+\frac{p^4}{4m}}.\end{aligned}$$ In new variables the Hamiltonian of the system will take the form: $$\begin{aligned} H=H_0+\sum \limits_{p\neq0} E_p\xi^+_p\xi_p.\end{aligned}$$ If the condensate moves at non-zero rate $u$, then the Hamiltonian of the systems takes the form: $$\begin{aligned} H=H_0+\sum \limits_{p\neq0} (E_p-up)\xi^+_p\xi_p.\end{aligned}$$ This Hamiltonian has the evident motion integrals — the occupation numbers. Or rather let us separate the set of all non-zero momentums $\{p\}$ to disjoint subsets $S_i$, containing $G_i$ elements, corresponding to the close values of the momentums. We assume that $G_i\rightarrow \infty$ when $V\rightarrow \infty$, but, together with that, $\frac{G_i}{V}\rightarrow 0$ when $V\rightarrow \infty$. Then the motion integrals will be the occupation numbers $N_i$ of the cells $S_i$. Let us assume $n_i=\frac{N_i}{G_i}$. The entropy of the system will take form: $$\begin{aligned} S=\sum \limits_i G_i[(1+n_i)\ln (1+n_i)-n_i \ln n_i].\end{aligned}$$ This value reaches the maximum under the fixed energy and number of particles on Bose distribution. On the other side, except the occupation numbers, Bose gas has one more first integral — the condensate motion velocity. Let us analyze the dependence of the thermodynamic values of our system on the value of this integral. The state of the liquid helium is convenient to be described via the condensate wave function $\Xi$ \[14\]. It satisfies the following differential equation: $$\begin{aligned} i\frac{\partial}{\partial t}\Xi(x,t)=-\frac{1}{2m}\nabla^2\Xi(x,t)+U_0\Xi(x,t)|\Xi(x,t)|^2-c\Xi(x,t),\end{aligned}$$ where $U_0,\;c >0$. If the system is kept at a constant temperature and its state is described by the number of $\lambda_1,...,\lambda_l$, then in equilibrium the free energy $F(T,\lambda_1,...,\lambda_l)$ must reach the minimum. In our case the condensate wave function $\Xi(x,t)$ performs the role of these parameters. But the condensate wave function can vary with time and in meaning, $\Xi(x,t),\;\Xi^+(x,t)$ represent canonically conjugate variables. That is why the requirement of free energy minimality should be replaced by the requirement of stationary action: $$\begin{aligned} A[\Xi,\Xi^+]=\int \limits_{t_1}^{t_2}[i \int d^3 x\Xi^+\frac{\partial}{\partial t}\Xi-F[\Xi,\Xi^+]]dt.\end{aligned}$$ As $F$ is defined accurate to constant, we will find: $$\begin{aligned} F[\Xi,\Xi^+]=\int d^3x \{-\frac{1}{2m} \Xi^+(x)\nabla^2\Xi(x)+\frac{U_0}{2}|\Xi(x)|^4-c|\Xi(x)|^2\}.\end{aligned}$$ To show that the thermodynamic characteristics do not depend on the velocity of the condensate motion it is sufficient to show that the action $A[\Xi,\Xi^+]$ is invariant under Galilean transformation: $$\begin{aligned} \Xi(x,t)\mapsto e^{i(\frac{mv^2}{2}t+mvx)}\Xi(x-vt,t),\end{aligned}$$ $v$ is relative motion velocity. But the invariance of the action $A[\Xi,\Xi^+]$ in relation to the Galilean transformation is easy to find out by direct calculation. So, the action $A[\Xi,\Xi^+]$ does not depend on the condensate flow velocity and this fact provides the stability of the condensate flow. We needed this example to illustrate the stability of the generalized microcanonical distributions. Besides we obtained the answer to the question of why not all the first integrals residing in the system are observed. Only those integrals are observed that have the entropy unchanged while the values of these integrals are changing (at constant values of other integrals). In the next section when we analyze the Ling cell, it will be essential that some of the first integrals of the Ling cell satisfy the equivalence principle. In addition to that, these first integrals have never been observed directly the way that the velocity of the Bose condensate through the capillary can be observed. These first integrals could only be observed indirectly by analyzing physiological phenomena testifying their presence. For this reason there would be a lot of doubts about or theory if no inorganic systems could be found for which there exist non-trivial first integrals satisfying the equivalence principle and which can therefore be observed. The example cited above allows us to dispense with such doubts. Let us discuss the equivalence principle more detailed. For the sake of simplicity we will assume that the entropy of the system $S(E,K')$ depends on the value of $K'$ of one integral $K$, however, the reasoning given below is directly generalized for the case when the entropy depends on the values of the two and more first integrals in involution. According to the equivalence principle the entropy $S(E,K')$ does not depend on $K'$. But it would be more precise to say that the entropy reaches its maximum on the certain interval of the values of the integral $K'$ $[\alpha(\lambda),\beta(\lambda)]$ and steeply decreases out of this interval. Here $\lambda$ — the parameter on which the Hamiltonian of the system depends. Here the energy of the system depending on the parameter $\lambda$ is chosen the way that at any value of $\lambda$ the entropies of the system were always the same. At the adiabatic infinitesimal change of the parameter $\lambda$, $\lambda_0\mapsto\lambda'=\lambda_0+\delta\lambda$ the value of the integral $K'$ is transformed by the formula $K'\mapsto K'+\overline{\frac{\partial K}{\partial \lambda}}\delta\lambda.$ Let us show that at this process the plateau $[\alpha(\lambda_0),\beta(\lambda_0)]$ transforms into the plateau $[\alpha(\lambda'),\beta(\lambda')]$. Our system is in equilibrium with the thermostat and let us assume that $K'_e \in [\alpha(\lambda_0),\beta(\lambda_0)]$ any equilibrium value of the integral $K'$ ($\lambda=\lambda_0$). Let us show that when the system’s energy is fixed at $E$ and $K' \in [\alpha(\lambda_0),\beta(\lambda_0)]$ the temperature of the system does not depend on $K'$. Let us assume that $T$ — the temperature corresponding to the equilibrium value of $K'_e$ of the integral $K$, and equal to the temperature of the thermostat. Let us assume that $K'_1$ — any value of the integral $K$ from the interval $[\alpha(\lambda_0),\beta(\lambda_0)]$. Let us assume that $T'$ — The temperature corresponding to it, and $T'\neq T$. Let us surround our system by an adiabatic membrane and change $K'_e\mapsto K'_1$. When this happens neither the system energy + thermostat nor the system entropy + thermostat are not changing. Now let us make our system contact the thermostat again which will lead to the equalization of the temperatures of the system and the thermostat. At this stage we assume that system + thermostat is an isolated system, particularly, its full energy is constant. As at the process of temperature equalization a self-induced heat flow takes place, the entropy of the system and the thermostat increases. That means, at the value of the integral $K$ $K'=K'_e$ for fixed full value of the common energy of the system and the thermostat, the entropy of the system and the thermostat is not maximal which contradicts the fact that $K'_e$ — the equilibrium value of the integral $K$. The same way we will demonstrate that at the fixed energy of the system the pressure $P_1$, corresponding to the value of the integral $K'_1$ Is equal to the ambient pressure $P$. The pressure is defined by formula $P=(\frac{\partial E}{\partial \lambda})_S$. Let us assume, for definiteness, that $P_1>P$. As it was demonstrated above the temperature corresponding to the value of the integral $K'_1$ is equal to the ambient temperature $T$. Let us make the system “expand” isothermally, doing work against the external forces. Meanwhile the system pressure $P_1$ will be decreasing by virtue of the known thermodynamic inequation, fair in our case as well, that states that the module of the isothermal compressibility is positive. We will allow the system to expand quasistatically until its pressure is equal to the pressure of the thermostat. At that the system + thermostat will do a certain positive work, i.e. the full energy of the system + thermostat will decrease by this value. But as the process of the system expansion is quasistatic and therefore reversible, the full entropy of the system + thermostat will not change. Therefore we can conclude that at the value of the integral $K'_e$, and for fixed full entropy of the system and the thermostat, the energy minimum is not reached which means the value of the integral $K'_e$ is not equilibrium. So, we demonstrated that for fixed energy of the system $E$, The values of the pressure and temperature are constant on the whole interval $[\alpha(\lambda_0),\beta(\lambda_0)]$ of the values of the integral $K'$. Let us now adiabatically change the $\lambda$, $\lambda_0\mapsto\lambda'=\lambda_0+\delta\lambda$. At this change the entropy of the system will not change and the energy of the system will increase by the value $P\delta\lambda$, i.e. the value that does not depend on $K'$. In other words, at the transformation $K'\mapsto K'+\overline{\frac{\partial K}{\partial \lambda}}\delta\lambda$ the points of the plateau will pass again into points of the plateau but corresponding to the changed parameter $\lambda$. When making the inverse adiabatic change of the parameter $\lambda$ $\lambda'\mapsto\lambda_0=\lambda'-\delta\lambda$, we see that the formula $K'\mapsto K'+\overline{\frac{\partial K}{\partial \lambda}}\delta\lambda$ sets a one-to-one correspondence between the points of the plateau $[\alpha(\lambda_0),\beta(\lambda_0)]$ and $[\alpha(\lambda'),\beta(\lambda')]$. We remind that we proceed from the assumption that the system is in equilibrium with the environment but is not in equilibrium with itself, being in a stationary non-equilibrium state. Let us describe a possible way for the system to change from stationary non-equilibrium to equilibrium. Let us use the (\[Delta\]) for the distribution function. This expression uses $k$ independent first integrals in involution $K_1,...,K_k$. However that does not mean that there no other integrals in the system except $K_1,...,K_k$, being in involution with them. For example in the formula for the ordinary microcanonical distribution only the energy function takes part but, as it was discussed in the section 1, even the most realistic systems of the statistical mechanics are nonergodic at the thermodinamyc limit, i.e. there are other first integrals except the energy. We will call first integrals that are included in formula (\[Delta\] active. A possible way of changing to equilibrium is that the more and more first integrals from the list $K_1,...,K_k$ cease being active. Let us assume that in this system the active integrals were $K_1,...,K_k$ and the integral $K_k$ ceased being active. Let us use the symbol $(\delta S)_E$ to denote the change of the entropy at this process (at constant energy). Let us prove the following important proposition: **Proposition.** $(\delta S)_E$ does not change at quasistatic processes, that is to say $$\begin{aligned} (d(\delta S)_E)_{qs}=0.\end{aligned}$$ **Proof.** Let us assume that the active integrals are $K_1,...,K_k$. Then the generalized microcanonical distribution corresponding to these integrals will take the form: $$\begin{aligned} w_k(E,K'_1,...,K'_k,x)=c\Delta(H(x)-E) \prod \limits_{i=1}^k\Delta(K_i(x)-K'_i).\end{aligned}$$ Let us note that $w_k(E,K'_1,...,K'_k,x)$ is in proportion to the characteristic function of a certain set. If the integral $K_k$ becomes inactive then the generalized microcanonical distribution oi $w_{k-1}(E,K'_1,...,K'_{k-1},x)$ turns out to be composed of a certain number $M$ of microcanonical distributions $w_k(E,K'_1,...,K'_k,x)$, the entropies of which are the same. It is evident that $(\delta S)_E=\ln M$. Let us consider the adiabatic change of the parameter $\lambda\mapsto\lambda'=\lambda_0+\delta\lambda$. We can assume that $\forall i=1,...,k$ $$\begin{aligned} \overline{(\frac{\partial K_i}{\partial \lambda})}_{K'_1,...,K'_k}|_{\lambda=\lambda_0}=0.\end{aligned}$$ In this formula the low indices $K'_1,...,K'_k$ mean that the averaging corresponding to the vinculum is taken over the microcanonical distribution $w_k(E,K'_1,...,K'_k,x)$. This can be reached by replacing $\forall i=1,...,k$ $K_i\mapsto K_i-(\lambda-\lambda_0)f_i(K_1,...,K_k,H)$, where $f_i(K_1,...,K_k, H)$ is the suitable functions of the first integrals and Hamiltonian. We will suppose that at this adiabatic process the point in the phase space that represents the state of our system, constantly moves in such a way that the values $K_1,...,K_k$ are the motion integrals. Meanwhile the microcanonical distribution $w_k(E,K'_1,...,K'_k,x)$ transforms into the microcanonical distribution $w_k(E',K'_1,...,K'_k,x)$. But, because of the above, the value $\delta E:=E'-E$ will not depend on the value $K'_k$ and will be the same if we calculat it for the distribution $w_{k-1}(E,K'_1,...,K'_{k-1},x)$. In the same adiabatic process the microcanonical distribution $w_{k-1}(E,K'_1,...,K'_{k-1},x)$, corresponding to the value of the parameter $\lambda=\lambda_0$, will transform into the distribution $w_{k-1}(E',K'_1,...,K'_{k-1},x)$, and by virtue of the fact that $\delta E$ can be calculated over the distribution $w_{k-1}(E,K'_1,...,K'_{k-1},x)$, the entropies corresponding to the distribution $w_{k-1}(E,K'_1,...,K'_{k-1},x)$ with $\lambda=\lambda_0$ and the distribution $w_{k-1}(E',K'_1,...,K'_{k-1},x)$ with $\lambda=\lambda'$ will be the same. But at the adiabatic process (such as the one described in the beginning of the proof) the evolution of the distributions simply resolves into the canonical substitution of the arguments of these distributions. That is why with $\lambda=\lambda'$, the distribution $w_{k-1}(E',K'_1,...,K'_{k-1},x)$ will turn out to be composed from the same number of $M$ generalized microcanonical distributions $w_k(Ey,K'_1,...,K'_k,x)$. That is to say $M=\rm const \mit$. That is to say $((\delta S)_E)_{qs}=\ln M=\rm const \mit$, which was to be proved. In case if there are too many of first integrals and if the process of decrease of the number of active integrals can be considered continuous, it is reasonable to introduce the continuous parameter $s$, characterizing the number of active integrals the way that $s \in [0,1]$, the value $s=0$ corresponds to the case when no integrals are active and the decrease of the number of active integrals corresponds to the decrease of the parameter $s$. Therefore we have: $$\begin{aligned} \frac{dS}{ds}\leq 0.\end{aligned}$$ The formula $\frac{dS}{ds}\leq 0$ is correct when it is assumed that the system energy and volume are kept constant. Let us assume, however, that the system temperature and volume are kept constant. Then: $$\begin{aligned} 0 \geq(\frac{dS}{ds})_{E,V}=(\frac{dS}{ds})_{T,V}-(\frac{\partial S}{\partial E})_{V,s}(\frac{dE}{ds})_{V,T}=\nonumber\\ =(\frac{dS}{ds})_{T,V}-\frac{1}{T}(\frac{dE}{ds})_{V,T}=\nonumber\\ =\frac{1}{T}(\frac{d(ST-E)}{ds})_{T,V}.\end{aligned}$$ That is to say: $$\begin{aligned} (\frac{dF}{ds})_{T,V}\leq 0.\end{aligned}$$ The same way let us assume that the system temperature and pressure are kept constant. Then: $$\begin{aligned} 0\geq (\frac{d F}{d s})_{T,V}=(\frac{d F}{d s})_{T,P}-(\frac{\partial F}{\partial V})_{s,T}(\frac{d V}{d s})_{P,T}=\nonumber\\ =(\frac{d F}{d s})_{T,P}-P(\frac{d V}{d s})_{P,T}=\nonumber\\ =(\frac{d \Phi}{d s})_{P,T}.\end{aligned}$$ So, in case when the temperature and pressure are kept constant: $$\begin{aligned} (\frac{d \Phi}{d s})_{P,T}\leq 0.\end{aligned}$$ The statement, that the number of independent first integrals in involution for the system must be very large can be established, for example, by the following way. Let us decompose the system $\mathfrak{S}$ into disjoint union of enough large number $N$ of small but macroscopic subsystems $\mathfrak{S}_i,\;i=1,...,N$. Let $H_i,\;i=1,...,N$ be a Hamiltonian of subsystems $\mathfrak{S}_i$ and $H_{ij},\;i,j=1,...,N,\;i\neq j$ be the total interaction Hamiltonian of subsystems $\mathfrak{S}_i$ and $\mathfrak{S}_j$. Then $\sum \limits_{i=1}^N H_i$ is proportional to the volume of system $\mathfrak{S}$ and $\sum \limits_{1\leq i<j\leq N}^N H_{i,j}$ is proportional to the total area of boundaries between different subsystems $\mathfrak{S}_i$. But each subsystem $\mathfrak{S}_i$ is macroscopic. So we can neglect by the interaction Hamiltonian of different subsystems in total hamiltonian $H$ of system $\mathfrak{S}$. Therefore we can write that $$\begin{aligned} H=\sum\limits_{i=1}^N H_i \label{additiv}\end{aligned}$$ But each subsystem $\mathfrak{S}_i$ is macroscopic. So, according to the nonergodic theorem for each subsystem $\mathfrak{S}_i$ there exists at least one its first integral $K_i$. It is easy to see that for different $i=1,...,N$ integrals $K_i$ are independent and in involution. It follows from (\[additiv\]) that integrals $K_i$ are the first integrals of whole system $\mathfrak{S}$ at the same time. E.S. Bauer’s stable non-equilibrium principle and the resting state of G. Ling. ================================================================================ For the sake of further analysis we will consider the contribution that E.S. Bauer and G. Ling made to the issue at hand. According to Bauer \[4\] the substance of the living cell can be in one of two states: the stable-non-equilibrium (resting) and thermodynamic equilibrium (active). The work The work produced by living system a living system is done when the system passes from stable non-equilibrium to equilibrium. Bauer used this principle (the two basic states and the change between them) as the basis for his “theoretical biology” and deductively derived the basic properties of such biological processes as assimilation, growth, excitability, adaptability and reproduction. That is what Bauer himself says \[4, p. 143\] about this principle: “Nonliving systems are ever in equilibrium and due to their free energy they do work against equilibrium required by the laws of physics and chemistry in the existing ambient conditions.”Further Bauer expands on it: “We will call this principle as stable none-equilibrium principle of the living systems. This name clearly expresses the meaning of the principle and the unique feature of living systems from the point of view of thermodynamics. The same way as a system in stable equilibrium when disturbed returns back to it, a living system in stable non-equilibrium also keeps coming back to it. Our principle also characterises the unique feature of living systems as we do not know any nonliving systems that would be in stable non-equilibrium.” The ground rule of Ling’s Association-Induction Hypothesis \[1,2,3\] is that the cell as a quasi-solid body is a system with a non-maximal entropy in a resting state.We proceed from the approach based on the fact that the resting state of the living cell, according to both Ling and Bauer, is a stationary non-equilibrium state, the possibility of existence of which was proven by one of us for a large class of realistic systems in statistical mechanics \[7\]. Here we simply apply it to the living cell. The appropriateness of the proposed approach to the living will be verified by comparing the results of the theoretical analysis and the available experimental data. Based on the assumption that the state of the living cell is stationary non-equilibrium we will give a thermodynamic description of the following phenomena: 1\) When the cell is excited and when it dies heat is generated. 2\) When the cell is excited and when it dies the cells’ size changes (mostly decreases). 3\) When the cell is excited and when it dies , the cell’s key protein molecules fold. 4\) When the cell is excited and when it dies, efflux of potassium ions from the cell takes place. Let us begin with the explanation of the first phenomenon. Let us assume that the substance inside the cell is in stationary non-equilibrium corresponding to a number of active first integrals in involution and the process of excitation and death manifests through some of these first integrals becoming inactive. Let’s assume that just like in the section 3 that the number of the active first integrals is characterised by a continuous $s \in [0,1]$, Where $s = 0$ corresponds to the case, when neither of first integrals are active and the value of $s = 1$ to the case when the maximum number of the first integrals are active. The question why there must be so many nontrivial integrals that their number can be described by a continuous parameter, can be answered by the fact that the muscle fibres can be contracted to different degrees and is able to algebraically summate the nerve impulses that enter it. It will be clear when we consider the changes in the size of the Ling cell when it’s activated and when it dies. We can choose $s$ in such a way that $s$ is proportionate to the number of active first integrals. The system’s volume is supposed to be constant. Let us make an infinitesimal change in the parameter $s$: $$\begin{aligned} s\mapsto s'=s-\delta s,\end{aligned}$$ where $\delta >0$, $\delta s$ is are infinitely small. It has been established that the knowledge of $k$ of the first integrals in involution allows decreasing the number of degrees of freedom by $k$ units (system deflation according to Whit-taker). For each $s$ we have $\rm const\, \mit s$ of active first integrals and the entropy that corresponds to the parameter $s$ is the entropy of the reduced system, corresponding to these constant $s$ first integrals. If s $s\mapsto s'=s-\delta s$, then the number of degrees of freedom of reduced system increases by $\rm const \mit \delta s$ units, and we may consider that these new ’turned-on’ degrees of freedom corresponding to the new system $\delta \mathfrak{S}$, that is in thermodynamic equilibrium with the initial system $\mathfrak{S}$. We will change the parameter $s$ at a fixed temperature $T$. If $s\mapsto s'=s-\delta s$ the system’s energy changes: $$\begin{aligned} E\mapsto E'=E+(\delta E)_T,\end{aligned}$$ and entropy: $$\begin{aligned} S\mapsto S'=S+(\delta S)_T.\end{aligned}$$ We can equate $(\delta E)_T$ to the energy of the system $\delta \mathfrak{S}$, and $(\delta S)_T$ to the entropy of the same system. As the system $\delta \mathfrak{S}$ is in thermodynamic equilibrium with the system $\mathfrak{S}$, it is essential that the temperature of the system $\delta \mathfrak{S}$ should also be equal to $T$, in other words, it is necessary that: $$\begin{aligned} \frac{1}{T}=\frac{d (\delta S)_T}{d (\delta E)_T}.\label{RR}\end{aligned}$$ But this equality can also be derived directly. We have: $$\begin{aligned} \frac{dS}{dE}=\frac{1}{T},\nonumber\\ \frac{dS+d (\delta S)_T}{dE+d (\delta E)_T}=\frac{1}{T}.\end{aligned}$$ From these two equalities we derive (\[RR\]) by using the elementary transformation. But in order for $\delta \mathfrak{S}$ to be in equilibrium with $ \mathfrak{S}$ the equation (\[RR\] is not enough. It is necessary that $$\begin{aligned} \frac{d^2 (\delta S)_T}{(d (\delta E)_T)^2}<0. \label{TDuneq}\end{aligned}$$ This inequality results from the requirement for maximum entropy in the system $ \mathfrak{S}+\delta \mathfrak{S}$ under condition that the energy remains constant, if we take into account the fact that the number of degrees of freedom of the system $\delta \mathfrak{S}$ is far less than the number of degrees of freedom of the system $\mathfrak{S}$. The detailed discussion of this inequality is found in Appendix 1. This inequality leads to $$\begin{aligned} \frac{d}{d (\delta E)_T} (\frac{1}{T})<0,\nonumber\\ -\frac{1}{T^2} \frac{d T}{d (\delta E)_T}<0,\end{aligned}$$ or $$\begin{aligned} \delta c_v:=\frac{d (\delta E)_T}{dT}>0.\end{aligned}$$ We will make the following assumption about the value of $ (\delta E)_T$. There is a temperature $T_0$, that makes the temperature $T> T0$ incompatible with the existence of life in the cell. This means that if the temperature $T\rightarrow T_0$ the interval $\Delta$ from the formula (\[Delta\]) becomes more and more comparable to the range spaces of the integrals $K_1,...K_k$ values, and the generalized micro-canonical distribution degenerates into a usual micro-canonical distribution. Then, if $T > T_0$ $$\begin{aligned} (\delta S)_E=\rm const \mit, \label{Gran}\end{aligned}$$) i.e. if $T> T_0$ the dependence of $T$ on $E$ for $\mathfrak{S}$ will be the same as the one for $\mathfrak{S}+\delta \mathfrak{S}$, i.e. $(\delta E)_T=0$. This boundary condition is discussed in more detail in the Appendix 2. We have: $$\begin{aligned} (\delta E)_T=(\delta E)_{T_0}-\int \limits_{T}^{T_0} \frac{d (\delta E)_{T'}}{dT'} dT'=\nonumber\\ = (\delta E)_{T_0}-\int \limits_{T}^{T_0}\delta c_v(T') dT'=\nonumber\\ =-\int \limits_{T}^{T_0}\delta c_v(T') dT'<0.\end{aligned}$$ I.e. $(\delta E)_T<0$, and that’s what we set out to prove initially. Let us proceed now to the explanation of why the Ling cell changes size when it is excited and when it dies. Let us examine a resting cell in equilibrium with the environment at a constant temperature $T$ and pressure $P$, and show that when the cell is excited and when it dies its size decreases. So, we assume that as above $s\mapsto s'=s-\delta s$, $\delta s>0$, and that $(\delta S)_E$ means the change in the entropy of the cell if its energy is constant. In section 3 we demonstrated that $(\delta S)_E=\rm const \mit$ in an adiabatic process and that is why $(\delta S)_E$ is a function of the entropy only, that is to say: $$\begin{aligned} (\delta S)_E=\delta f(S),\end{aligned}$$ for some function $\delta f$ of one variable. Let us show now that $\delta f$ is a decreasing function. We have: $$\begin{aligned} (\delta S)_E=(\delta S)_T-(\delta E)_T \frac{\partial S}{\partial E},\nonumber\\ (\delta S)_E=(\delta S)_T-\frac{1}{T}(\delta E)_T.\end{aligned}$$ Because of the thermodynamic inequality $c_V>0$ (($c_V$ -the thermal capacity of the cell at constant volume) the energy, and as a consequence, the entropy are strictly increasing functions of the temperature. That is why it is enough to show that $\frac{\partial}{\partial T} ((\delta S)_E)<0$. But $$\begin{aligned} \frac{\partial}{\partial T} ((\delta S)_E)=\frac{\partial}{\partial T} ((\delta S)_T)-\frac{1}{T}\frac{\partial}{\partial T} ((\delta E)_T)+\frac{1}{T^2}(\delta E)_T=\nonumber\\ =\frac{1}{T^2}(\delta E)_T.\end{aligned}$$ And by virtue of this, as it was already proved, $(\delta E)_T<0$ we find: $\frac{\partial}{\partial T} ((\delta S)_E)<0$, i.e. $\delta f$ is a decreasing function of its argument. Now let us define the minor additions that appear in the thermodynamic potentials under $s\mapsto s'=s-\delta s$. We have: $$\begin{aligned} (\delta S)_E=(\delta S)_T-(\delta E)_T \frac{\partial S}{\partial E},\nonumber\\ (\delta S)_E=(\delta S)_T-\frac{1}{T}(\delta E)_T.\end{aligned}$$ I.e. under $s\mapsto s'=s-\delta s$, $$\begin{aligned} E(V,S)\mapsto E(V,S)-T\delta f(S).\end{aligned}$$ According to the minor additions theorem for free energy $F$ and the thermodynamic potential $\Phi$ we have: $$\begin{aligned} F(V,T)\mapsto F(V,T)-T\delta f(S),\nonumber\\ \Phi(P,T)\mapsto\Phi(P,T)-T\delta f(S).\end{aligned}$$ As $$\begin{aligned} V=(\frac{\partial \Phi(P,T)}{\partial P})_T,\end{aligned}$$ Then change in the cell size with $s\mapsto s'=s-\delta s$ takes the form: $$\begin{aligned} \delta V=-T(\delta f)'(S)(\frac{\partial V}{\partial P})_T(\frac{\partial S}{\partial V})_T.\end{aligned}$$ As $(\delta f)'<0$ by virtue of what was proven above and $(\frac{\partial V}{\partial P})_T<0$ (thermodynamic inequality, which is always true), the sign $\delta V$ is opposite to the sign $(\frac{\partial S}{\partial V})_T$. Let us set the sign $(\frac{\partial S}{\partial V})_T$. But because of a known thermodynamic relation $$\begin{aligned} (\frac{\partial S}{\partial V})_T=-(\frac{\partial V}{\partial T})_P(\frac{\partial P}{\partial V})_T.\end{aligned}$$ That is why the statement that $(\frac{\partial S}{\partial V})_T>0$ is therefore equivalent to the statement that the cell expands when heated (at constant pressure). This behavior is demonstrated by most solid bodies when heated. If we accept that this is true (the cell expands when heated at constant pressure) then we find $\delta V<0$, which was to be proved. We thus conclude that the cell’s size decrease. We consider our description to be a general thermodynamic prerequisite that explains the phenomenon of the Ling cell contraction, a prototype of muscle contraction. Before we consider the aggregation of key proteins in the Ling’s model, let us go back to the question of the transformation of the internal energy of the cell when it dies. We have demonstrated above that death occurs then $$\begin{aligned} (\delta E)_{T,V}<0.\end{aligned}$$ If we use the symbol $\delta Q$ to denote the amount of heat that entered the cell, we will find: $$\begin{aligned} \delta Q=(\delta E)_{T,V}+T(\frac{\partial S}{\partial V})_T\delta V.\end{aligned}$$ But we already know that $\delta V$ and have opposite signs, i.e. $$\begin{aligned} \delta Q<0.\end{aligned}$$ In other words, the death and excitation of the cell are exothermic reactions. Let us consider now the thermodynamic prerequisites of the aggregation of proteins that are present in a resting Ling’s cell in an expanded state. Our analysis draws on the fact that the energy of an unfolded protein molecule is greater than the energy of a folded protein molecule and is based on the protein model in which the configuration of the protein molecule can be characterized by a real parameter $x \in [0,1]$, describing the degree to which the molecule has unfolded. We assume that the energy of the protein molecule is a function only of $x$ and it is an increasing function. We assume also that $x = 0$ corresponds to a completely folded molecule whose energy is at a minimum; and $x = 1$ corresponds to a completely unfolded molecule whose energy is at its maximum. Let us use the following supporting procedure. We will identify a single protein molecule in the cell’s protoplasm and call it $\mathfrak{X}$, and we’ll use $x$ as a parameter characterizing the degree to which our single molecule is unfolded. If we call the system’s temperature $T$, then all the thermodynamic equations derived in the previous section can be applied to $\mathfrak{X}$, more particularly we have the equation: $$\begin{aligned} dE=TdS-hdx,\label{ttt}\end{aligned}$$ for a certain function $h(E, S)$ of the energy and entropy $\mathfrak{X}$. The value $h(E, S)$ is a generalised force that is thermodynamically conjugated to $x$. When the external forces that affect our protein molecule, are absent, $h(E, S) = 0$. The relation (\[ttt\]) is equivalent to the equation $$\begin{aligned} dE=TdS-PdV,\end{aligned}$$ And that is why we can more or less directly apply to case the conclusions for the cell interacting with the environment, changing $x\mapsto V$, $P\mapsto h$. So, Let’s assume again that $s\mapsto s'=s-\delta s$, then $$\begin{aligned} (\delta x)_{T,h}=-T (\delta f)'(S)(\frac{\partial S}{\partial h})_{T,x}.\end{aligned}$$ We see that the sign of $\delta x$ is the same as the sign of $(\frac{\partial S}{\partial h})_T$. But by virtue of the known thermodynamic relation: $$\begin{aligned} (\frac{\partial S}{\partial h})_T=-(\frac{\partial x}{\partial T})_h.\end{aligned}$$ So, the sign $(\frac{\partial S}{\partial h})_T=-(\frac{\partial x}{\partial T})_h$ will only be negative if the protein molecules unfold and the cell temperature increases. So we have proved that the protein molecules fold when the protoplasm is activated and when the cell dies only if they unfold when the cell is heated. It is clear that this statement is true only for the protein model used here. But the property of the protein molecule to unfold when heated appears to be physically evident. This assumption seems to be reasonable as the protein molecules are supposed to change to states with more energy when heated, and the energy is an increasing function of the parameter $x$. That is why the protein molecule folds under s $s\mapsto s'=s-\delta s$, which was to be proved. The question of how we apply this general method we used for the analysis of more realistic protein models needs further research. Let us consider now the question of why the Ling cell’s nonergodicity includes ion flows between the cell and the environment when the physiological state of the Long cell is changed. When analyzing this phenomenon we will restrict ourselves to potassium ions that play an important role in the physiological processes of the cell. We will try to answer the question of why this positive ions are released by excited or dead cells. The answer to this question will explain another phenomenon well-known from the classical cytology: why the damaged area in a cell always has the negative charge to the adjoining intact areas that are not affected by excitation or injury. It is clear that our approach must be improved in order for us to be able to consider a broader spectrum of ion characteristics, such as: ion chemical nature, their physical properties, their concentration in solution, etc. Let us assume that $N$ means the number of potassium ions in the cell $\mathfrak{C}$. We can regard $N$ as a parameter, and it will be now clear why. Our analysis is carried out in the thermodynamic limit and from this point of view the cell can be considered very big. Let us attach a very small vessel $\mathfrak{S}$, to the cell which will be separated from it a semipermeable membrane that is permeable only for potassium ions. We will assume that in $\mathfrak{S}$ we have only potassium ions and that $\mathfrak{S}$ is so small that the entropy of the ions inside it is equal to zero. As a model for $\mathfrak{S}$ we can take a very narrow potential well, the quantum energy levels of which are not degenerate and the distance between adjoining energy levels is much greater than the temperature $T$. Then potassium ions in $\mathfrak{S}$ will be mostly in a basic energy state and the thermodynamic probability $W=1$, and $S_{\mathfrak{S}}=\ln W=0$. Let us also introduce an external electric field the potential of which $\varphi$ everywhere except $\mathfrak{S}$ is equal to zero, and inside $\mathfrak{S}$ is constant and equal to $\varphi_0$. Adiabatically changing $\varphi_0$, we can pump the potassium ions from $\mathfrak{S}$ to the cell and back. Meanwhile the entropy of the whole system will be constant and equal to the entropy of the cell $S$. By virtue of the fact that $\varphi_0$ is a parameter in the Hamiltonian of the whole system, $((\delta S)_E)_{qs}=\rm const \mit$ during adiabatic change $\varphi_0$. Here $E$ means the energy of the entire system. Let us assume that $E_{\mathfrak{C}}$ means the energy of the cell. All the motion integrals that are related to the entire system $\mathfrak{S}+\mathfrak{C}$ are equally related to the cell $\mathfrak{C}$ because the potassium ions in $\mathfrak{S}$ are immovable, and their coordinates and momentums can be removed from the list of the dynamic variables. But $$\begin{aligned} (\delta S)_E=(\delta S)_{E_\mathfrak{C},N}+\frac{\partial S}{\partial E_{\mathfrak{C}}}\delta E_\mathfrak{C}+\frac{\partial S}{\partial N}\delta N.\end{aligned}$$ But $\delta E_\mathfrak{C}=-\varphi_0\delta N$, as the full energy of the cell and of the vessel is constant. Considering that in equilibrium: $$\begin{aligned} -\frac{\partial S}{\partial E_{\mathfrak{C}}}\varphi_0\delta N+\frac{\partial S}{\partial N}\delta N=0,\end{aligned}$$ we find: $$\begin{aligned} (\delta S)_E=(\delta S)_{E_\mathfrak{C},N}.\end{aligned}$$ That is why during an adiabatic change in the parameter $\varphi_0$ or, of the number of particles in the cell $N$ $$\begin{aligned} ((\delta S)_{E_\mathfrak{C},N})_{qs}=\rm const \mit,\end{aligned}$$ Which means that $N$ can be considered as a parameter. Let us proceed now to the question of the amount of potassium ions released by the cell during its activation and death. We assume that the cell is immersed in a potassium ion solution and the chemical potential of the potassium ions is constant and equal to $\mu$. The following thermodynamic relation will hold true: $$\begin{aligned} dE=TdS+\mu dN\end{aligned}$$ and all its corollaries. This equation similar to the thermodynamic equation $$\begin{aligned} dE=TdS-PdV,\end{aligned}$$ if we equate $N\leftrightarrow V$, $\mu\leftrightarrow -P$. Therefore, to prove that the potassium ion yield from the cell takes place when $s\mapsto s'=s-\delta s$ all we have to show is that $$\begin{aligned} (\frac{\partial S}{\partial N})_T>0.\end{aligned}$$ To find $(\frac{\partial S}{\partial N})_T$ we will note that the concentration of potassium ions in the cell is low and that is why we can use the strong electrolyte theory (see \[13\]). This theory suggests that in a low concentration of potassium ions the cell’s entropy (at fixed temperature) $S=S_0+S^{K^+}+N\varphi(P,T)$, where $S_0$ - the cell entropy at zero ion concentration, $S^{K^+}$ - the entropy of only the potassium ions, calculated for a perfect gas, and $\varphi(P,T)$ is a function of the cell temperature and pressure. That is why for the entropy of the potassium ions we can use the standard formula (see \[13\]): $$\begin{aligned} S^{K^+}=N \ln \frac{e V}{N}-Nf'(T), \label{entropy}\end{aligned}$$ where $$\begin{aligned} f(T)=-T \ln ((\frac{mT}{2 \pi})^{3/2}).\end{aligned}$$ Here $m$ is the mass of a potassium ion. It is clear that if the potassium ion solution in the cell is thin enough $(\frac{\partial S}{\partial N})_T>0$. So, $\delta N<0$, which was to be proved. For simplicity of the analysis we assumed above that there are only potassium ions in the cell. But in reality there are different kinds of ions and all the derived formulas should be true for each kind of ions. Particularly, during the activation and death of the cell all kinds of ions should come out of the cell if their concentration is low enough. But it is known that some ions, for example sodium ions, are on the contrary, absorbed by the cell during activation or death (according to Ling’s model). . However, the cell becomes negatively charged during activation, i.e. our explanation leads to the correct answer “on average”. This gives us hope that our explanation is genuinely true and can be improved in a proper way. To conclude of this section let us note that we considered the resting state of the cell to be stationary non-equilibrium and the state corresponding to the excitation and death to be equilibrium. We considered these states as given and never asked the question why there are transitions between these two states. One of the possible hypotheses is that such transitions appear due to changes in the conditions on the cell’s boundaries. That is why in section 3 we discussed in detail the role that the conditions on the system’s boundaries play in the establishment of thermodynamic equilibrium inside the cell. The question of how true this hypothesis is will be the subject of further research. Conclusion. =========== We used the results of \[7\] to analyse the Ling cell in this paper, assuming that this kind of cell is a nonergodic system. Using the property of nonergodicity as a starting point for our research, we first discussed the nonergodicity of statistical mechanics systems, we then developed the thermodynamics of stationary non-equilibrium states assuming the existence of several nontrivial first integrals that commutate between them. These thermodynamics are applicable to any system in statistical mechanics, including the Ling cell. The analysis of the Ling cell properly demonstrated that using the approach based on our thermodynamics of stationary non-equilibrium we managed to derive such realistic properties of the excited Ling’s cell as heat generation, decrease of the cell size, the folding of the proteins that are unfolded when the cell is resting and the releasing of potassium ions. The applicability of the proposed approach to the living cell model, such as the Ling cell, testifies to the applicability and adequacy of our analysis for studying the living state of matter. We are very grateful to P. Agutter, A.V. Koshelkin, Yu. E. Lozovik, A.V. Zayakin and E.N. Telshevskiy for valuable critical comments on this article and very useful discussions. Appendix 1. Discussion of the inequality (\[TDuneq\]). ====================================================== In this section we will try to justify the inequality (\[TDuneq\]). Let us see the Hamiltonian dynamical system with $n+k$ degrees of freedom ($k\ll n$) and $k$ independent first integrals in involution $K_1,...,K_k$. Locally we can always choose the canonical coordinates $(p_1,q_1)$,...,$(p_{n+k},q_{n+k})$ in such a way that $K_1,...,K_n$ will only be the functions of $(p_{n+1},q_{n+1})$,...,$(p_{n+k},q_{n+k})$. First, let us assume that the coordinates that measure up this property can be chosen globally. So, the phase space $M$ of our system represents as a direct product of the phase spaces $M=M_1\times M_2$, where $M_1$ corresponds to the coordinates $(p_1,q_1)$,...,$(p_{n},q_{n})$, and $M_2$ — to the coordinates $(p_{n+1},q_{n+1})$,...,$(p_{n+k},q_{n+k})$. Let us denote through $d \Gamma$ the element of the phase volume on $M$, and through $d \Gamma^1$ and $d \Gamma^2$ the elements of the phase volumes on $M_1$ and $M_2$, correspondingly. Let us assume that $H(x,y)$ — the Hamiltonian of our system, $x \in M_1$, $y \in M_2$. Let us assume that $S(E)$ is the entropy calculated for the microcanonical distribution and $S_1(E,K_1',...,K_k')$ is the entropy calculated for the generalized microcanonical distribution corresponding to the $K'_1,...,K'_k$ of the first integrals $K_1,...,K_k$. Let us recall that $$\begin{aligned} S_1(E,K_1',...,K_k')=\ln \int d\Gamma \delta(H-E)\prod \limits_{i=1}^k\delta(K_i-K'_i).\end{aligned}$$ We assume that $S_1(E,K_1',...,K_n')$ measures up the equivalence principle (see section 4). Locally on $M_2$ we can build the function $\varphi_1,...,\varphi_k$, and do it in such a way that the set of functions $(K_1,\varphi_1)$,...,$(K_k,\varphi_k)$ will represent the set of canonical coordinates, i.e. the Poisson bracket in these coordinates will have the standard form. In other words, we can build a partition $M_2$ in the area $V_i$, $i=1,2,...$ with a piecewise-smooth boundary and in each of them we can choose the functions $\varphi_1^i,...,\varphi_k^i$, that satisfy the requirement mentioned above. Let us assume that $\Sigma$ — a certain joint surface of the level of the integrals $K_1,...,K_k$, and $x \in \Sigma$. If (without losing generality) we assume that $\Sigma$ is connected, then the Hamiltonian phase flows generated by $K_1,...,K_k$ transitively acts on $\Sigma$. And as these phase flows keep the phase volume on $M$ and the Hamiltonian, then $$\begin{aligned} S_1(E,K'_1,...,K'_k)=\ln \int d\Gamma_x^1 \delta(H(x,y)-E)+c, \label{Rel}\end{aligned}$$ where $x \in M_1$,and $c$ is the constant equal to: $$\begin{aligned} c=\ln \sum \limits_{i=1}^{\infty} \int \limits_{\Sigma\cap V_i}d\varphi_1^i...d\varphi_k^i. \label{CC}\end{aligned}$$ The constant $c$ does not depend on $E$ and can be discarded. Let us now turn to the calculation of $(\delta S)_T$. If the system $M=M_1\times M_2$ is described by the Gibbs distribution then the probability to find the system $M$ at the point $(x,y),\;x\in M_1,\;y\in M_2$ takes the form: $$\begin{aligned} w_{12}(x,y)=\frac{1}{Z(T)}e^{-\frac{H(x,y)}{T}}, \label{Cond}\end{aligned}$$ where $$\begin{aligned} Z(T):=\int d \Gamma_x^1 d\Gamma_y^2 e^{-\frac{H(x,y)}{T}}.\end{aligned}$$ The probability to find the system $M_2$ at the point $y$ is given then by the expression: $$\begin{aligned} w_2(y)=\int d \Gamma_x^1 w_{12}(x,y).\end{aligned}$$ Let us introduce the conditional probability $w_{1|2}(x|y)$ to find a system $M_1$ at the point $x$ at condition that the system $M_2$ is at the point $y$ by the formula: $$\begin{aligned} w_{12}(x,y)=w_{1|2}(x|y)w_2(y). \label{Cond1}\end{aligned}$$ Von Neumann entropy of the system $M$ takes the form: $$\begin{aligned} S(E(T))=-\int d \Gamma_x^1 d \Gamma_y^2 w_{12}(x,y) \ln w_{12}(x,y).\end{aligned}$$ Using (\[Cond1\]) $S(E(T))$ we can transform to the form: $$\begin{aligned} S(E(T))=\langle S_1(E(T),y)\rangle_{M_2}+S_2(T),\end{aligned}$$ where $$\begin{aligned} S_1(E(T),y)=-\int d \Gamma_x^1 w_{1|2}(x|y) \ln w_{1|2}(x|y),\end{aligned}$$ $\langle\cdot\rangle_{M_2}$ means the averaging over $M_2$ by the measure $w_2(y) d \Gamma_y$, and $$\begin{aligned} S_2(T)=-\int d \Gamma_y^2 w_2(y) \ln w_2(y).\end{aligned}$$ Note that only averaged values can be measured at experiment. We find: $$\begin{aligned} (\delta S)_T=S_2(T).\end{aligned}$$ At that we meant the formula (\[Rel\]) and the fact that within the thermodynamic limit the descriptions by the microcanonical and canonical assemblies coincide. At the same time we also consider, of course, that the constant $c$ from the formula (\[CC\]) does not depend on $K'_1$,...,$K'_k$. When we will proceed to the discussion of the general case that will not require the presentability of $M$ in the form of the direct product $M_1$ and $M_2$ we will see that this assumption is justified. Let us try to describe the dynamics of the system $M_2$, using the fact that the number of degrees of freedom $M_2$ $n$ is much greater that the number of degrees of freedom $M_1$ $k$. The Hamiltonian canonical equations for $M_2$ take the form: $$\begin{aligned} \dot{p}_i=-\frac{\partial H(p_1,q_1,...,p_{n+k},q_{n+k})}{\partial q_i},\nonumber\\ \dot{q}_i=\frac{\partial H(p_1,q_1,...,p_{n+k},q_{n+k})}{\partial p_i},\nonumber\\ i=n+1,...,n+k.\end{aligned}$$ As the number of degrees of freedom $M_2$ is much less that the number of degrees of freedom $M_1$ we can average the right parts of the last equation by $w_{1|2}(x,y)$. As a result, skipping some quite trivial calculations we will find: $$\begin{aligned} \dot{p}_i=-\frac{\partial F(p_{n+1},q_{n+1},...,p_{n+k},q_{n+k}|T)}{\partial q_i},\nonumber\\ \dot{q}_i=\frac{\partial F(p_{n+1},q_{n+1},...,p_{n+k},q_{n+k}|T)}{\partial p_i},\nonumber\\ i=n+1,...,n+k,\end{aligned}$$ where $$\begin{aligned} F(p_{n+1},q_{n+1},...,p_{n+k},q_{n+k}|T)=-T\ln Z_1(y|T),\nonumber\\ Z_1(y|T)=\int d \Gamma_x^1 e^{-\frac{H(x,y)}{T}}, \nonumber\\ y=(p_{n+1},q_{n+1},...,p_{n+k},q_{n+k}).\end{aligned}$$ So, this way the dynamics of the system $M_2$ will be described by the canonical Hamiltonian equations with the Hamiltonian $F(y|T)$. Let us note that this way the defined Hamiltonian is not defined uniquely but accurate to the temperature arbitrary function $f(T)$. Let us define the new Hamiltonian of the system $M_2$ $$\begin{aligned} H_2(y|T)=F(y|T)+f(T),\end{aligned}$$ in such a way that makes true the following: $$\begin{aligned} (\delta E)_T=\langle H_2(y|T)\rangle_{M_2}.\end{aligned}$$ Let us show that in this case $$\begin{aligned} \langle\frac{d}{dT} H_2(y|T)\rangle_{M_2}=0. \label{aaa}\end{aligned}$$ For this we will proceed from the identical equation: $$\begin{aligned} \frac{\frac{d}{dT} (\delta S)_T}{\frac{d}{dT} (\delta E)_T}=\frac{1}{T}.\end{aligned}$$ But $$\begin{aligned} (\delta S)_T=S_2(E(T)) =-\int d \Gamma_y^2 w_2(y) \ln w_2(y).\nonumber\\ \frac{d}{dT}S_2(E(T))=\int d \Gamma_y^2 \frac{H_2(y|T)}{T} \frac{d}{dT} \frac{e^{-\frac{H_2(y|T)}{T}}}{Z_2(T)},\end{aligned}$$ where $$\begin{aligned} Z_2(T):=\int d \Gamma_y^2 e^{-\frac{H_2(y|T)}{T}}.\end{aligned}$$ We have: $$\begin{aligned} \frac{d}{dT}S_2(E(T))=-\langle\Delta(\frac{H_2(y|T)}{T})\Delta(\frac{d}{dT}\frac{H_2(y|T)}{T})\rangle_{M_2}, \label{174}\end{aligned}$$ where we assumed by definition: $$\begin{aligned} \Delta g(y):=g(y)-\langle f(y) \rangle_{M_2}.\end{aligned}$$ Making the similar calculations, we find: $$\begin{aligned} \frac{d}{dT}\langle H_2(y|T)\rangle=\langle \frac{d}{dT} H_2(y|T)\rangle-\langle\Delta({H_2(y|T)})\Delta(\frac{d}{dT}\frac{H_2(y|T)}{T})\rangle_{M_2}. \label{176}\end{aligned}$$ Comparing (\[174\]) and (\[176\]) we will find: $$\begin{aligned} \langle\frac{d}{dT} H_2(y|T)\rangle_{M_2}=0,\end{aligned}$$ which was to be proved. Let us describe now the effective Hamiltonian of the whole system $M_1\times M_2$. The dynamics of the system $M_1$ can be described by the Hamiltonian $H(x,y)$ where $y \in M_2$ and depends on time, and $x\in M_1$. But because of $K_1,...,K_k$ are the motion integrals, the value $H(x,y)$ at fixed $x$ is the same for all $y$ corresponding to the same value of the motion integrals. That is why the system $M_1$ is described by the Hamiltonian $H(x,y)$, where $y$ is constant with time and is chosen the same way as in the formula (\[Rel\]). Further the effective Hamiltonian of the system $M_2$ $H_2(y|T)$ depends on the temperature $T$ of the system $M_1$, but temperature is an intensive parameter and instead of $T$ we can use the energy of the system $M_1$, falling within one degree of freedom. Let us assume by definition: $$\begin{aligned} H'_2(y|\frac{E_1}{n})=H_2(y|T),\end{aligned}$$ where $E_1$ — the energy of the system $M_1$. As an effective Hamiltonian of the system $M=M_1\times M_2$ the following expression will be used: $$\begin{aligned} H_{eff}(y,z)=H(y,x)+H'_2(z|\frac{H(y,x)}{n}). \label{eff}\end{aligned}$$ Let us recall that according to this formula $x$ is chosen the same way as in (\[Rel\]). Let us show that within the limit $n\rightarrow \infty$ the Hamiltonian equation for $H_{eff}$ will coincide with the Hamiltonian equations for the initial Hamiltonian if the point $x$ is on the same joint surface of the level of the integrals $K_1,...,K_k$, that $z$. Let us assume that $(y,z) \in M=M_1\times M_2$. Let us analyze the derivative with time of the point $z$ by virtue of the Hamiltonian equations for the Hamiltonian $H_{eff}$. This will be exactly the derivative by virtue of the Hamiltonian motion equations corresponding to the Hamiltonian $H_2(z|T)$ and the fact that this derivative coincide with the derivative with time by virtue of the Hamiltonian motion equations built by $H(y,z)$ was shown by us above. Let us assume now that $R(y)$ is the dynamic variable on $M_1$. Let us calculate its derivative with time by virtue of $H_{eff}$. We have: $$\begin{aligned} \dot{R}(y)=(H(y,x),R(y))(1+\frac{1}{n}H^{''}_2(z,\frac{H(y,x)}{n})). \label{123}\end{aligned}$$ Here the symbol $H^{''}_2(\varepsilon,z)$ denotes the derivative $H^{'}_2(\varepsilon,z)$ by the first argument. Within the limit $n \rightarrow \infty$ the right part of (\[123\]) evidently transforms into $(H(y,x),R(y))$, i.e. within this limit $$\begin{aligned} \dot{R}(y)=(H(y,x),R(y)),\end{aligned}$$ which was to be proved. Let us note as well that $H_{eff}$ depends on the choice of the point $x \in M_2$. However, because of the equivalence principle, the entropic properties of the system $M=M_1\times M_2$ will not depend on this choice and these last ones are the only ones that are important for us. That is why from the point of view of the calculations of the entropic properties the Hamiltonian $H_{eff}(y,z)$ is as good as the initial $H(y,z)$ and we will further use only the first one. Let us note finally that the effective Hamiltonian with the form of (\[eff\]) is defined accurate to the arbitrary function $f(\frac{H(y,x)}{n})$. I.e. the effective Hamiltonians corresponding to the different choices of such normalization induce the same motion equations. That is why all the thermodynamic characteristics calculated by $H_{eff}$ corresponding different choices of the normalization must coincide. And this is really so. Actually at the old temperature the adding of $f(\frac{H(y,x)}{n})$ to $H_{eff}$ the arbitrary function can be considered by the corresponding change $H_1(y,x)$. But at the same time with $n \rightarrow \infty$ the average value of the energy $E_1$ of the system $M_1$ will change by the value $\sim 1$, and that is why the entropy of the system $M_2$ will change by the value $\sim \frac{1}{n}$, i.e. it will not change at all within the thermodynamic limit. Let us use $S(T)$ to denote the entropy of the system $M$, as the function of the temperature $T$, calculated using Gibbs distribution for the Hamiltonian $H_{eff}$. Let us use $E_1$ to denote the average $H_1(y,x)$, and $E_2$ to denote the average $H_2(y|T)$ by the measures on $M_1$ and $M_2$ respectively, induced by the same Gibbs distribution. Let us assume also $E=E_1+E_2$. The direct calculation shows that $\frac{d S}{dE}=\frac{1}{T}$. From the other point of view we normalized $H_{2}(y|T)$ in such a way that $\frac{d S_2(T(E_2))}{d E_2}=\frac{1}{T}$ would be true. From these two equalities we obtain $\frac{d S_1(E_1)}{d E_1}=\frac{1}{T}$. I.e. in the thermodynamic equilibrium state the temperatures of the systems $M_1$ and $M_2$ coincide, and this equality is true precisely and not only in the thermodynamic limit $k=\rm const \mit$, $n \rightarrow \infty$. If $k$ is very large (being at the same time much less than $n$), then according to the main principles of the statistical mechanics, instead of the description of $M_2$ by means of Gibbs canonical distribution we can use the description by means of the microcanonical distribution and the function of the distribution for the whole system $M$, at which the energies of the systems $M_1$ and $M_2$ will be equal to $E_1$ and $E_2$ respectively, will take the form: $$\begin{aligned} \delta(H(y,x)-E_1)\delta(H_2'(z|\frac{E_1}{n})-E_2),\end{aligned}$$ $y \in M_1,\; z\in M_2$. The entropies of the systems $M_1$ and $M_2$ will be equal to $$\begin{aligned} S_1(E_1)=\ln \int \delta(H(y,x)-E_1) d \Gamma_y^1,\nonumber\\ S_2(E_2,E_1)=\ln \int \delta(H_2'(z|\frac{E_1}{n})-E_2) d \Gamma_z^2.\end{aligned}$$ To obtain the inequality that we need let us make the following conceptual experiment. Let us make the system $M_2$ contact the thermostat $T$, which is at the temperature $T$. As for the system $M_1$, we will take a very large system $M_3$ that will be connected to the system $M_1$ via a heat pump $N$, working by Carnot cycle. We will consider the working body of this pump to be so small that its thermal capacity can be neglected. Let us also assume that during the thermal contact with $M_1$ or $M_3$ the temperature of the working body of the pump is equal to the temperature of $M_1$ or $M_3$ (depending on which one has thermal contact with the working body of the pump), i.e. the heat pumping is done without the entropy increase. The working body of the pump is connected to a shaft by something like a connecting rod gear, this shaft can rotate and on the axis of which there is a spring $P$, possessing the energy $E_P$. Let us note that because of our assumptions about the effective Hamiltonian $H_{eff}(y,z)$ the dynamics given by it on $M_1$ are exactly the dynamics given by $H(y,x)$ and that is why the system $M_1$ can be considered closed. Or more precisely as the system $M_1$ is related to $M_3$ by means of the heat pump $N$ the system that consists from $M_1$, $M_3$, $N$, $P$ can be considered closed. We will now assume that the elastic characteristics of the spring $P$ are picked the way that the work that must be done to the working body $N$ is exactly equal to the change of the spring energy. So this way we have a whole continuum of equilibrium states of the system consisting of $M_1$, $M_3$, $N$, $P$, which is parametrized by, for example, the angle of rotation of the pump shaft and at $S_1+S_3=\rm const \mit$, $E_1+E_3+E_P=\rm const \mit$. When our whole system obtains the equilibrium then the temperature of the thermostat is equal to the temperature of $M_1$ and the temperature of $M_3$. From the condition of maximal entropy of the whole system that consists of $M_1$, $M_2$, $M_3$, $T$, $N$, $P$ at fixed full energy considering the equalities: $S_1+S_3=\rm const \mit$, $E_1+E_3+E_P=\rm const \mit$ we will find that in the equilibrium the following value $$\begin{aligned} S_2(E_2,E_1)+S_T(E_T)\end{aligned}$$ should reach the maximum at the additional condition $$\begin{aligned} E_2+E_T=\rm const \mit.\end{aligned}$$ Here $E_T$ — the energy of the thermostat and $S_T(E_T)$ — the entropy of the thermostat. Instead of the dependence $S_2(E_2,E_1)$ on $E_1$ it appears more convenient to consider the dependence $S_2$ from the system temperature $M_1$ $\lambda(E_1)$. Let us introduce a new function $$\begin{aligned} S'(E,\lambda(E_1)):=S_2(E,E_1).\end{aligned}$$ The condition of the maximality of the entropy of the system $M_2$ and the thermostat lead to the equalities: $$\begin{aligned} \frac{\partial S'(E,\lambda)}{\partial\lambda}|_{\lambda=T,\;E=E(T)}=0,\nonumber\\ \frac{\partial S'(E,\lambda)}{\partial E}|_{\lambda=T,\;E=E(T)}=\frac{1}{T}. \label{Nor}\end{aligned}$$ But these equalities can be obtained directly. The second equality is obtain from the ordinary formulas of the statistical mechanics if we assume that the system $M_2$ is described by the Gibbs canonical distribution with Hamiltonian $H_2(y,\lambda)$. Let us set up the first equality. Let us adiabatically change the parameter $\lambda$, $\lambda\mapsto T+\delta\lambda$. At such change of the parameter $\lambda$ the entropy of the system $M_2$ will not change. But by virtue of (\[aaa\]) the energy $M_2$ will not change. I.e. the derivative of the entropy of the system $M_2$ by $\lambda$ at fixed energy is equal to zero, which is exactly the equality that we needed. But for the maximality of the entropy of the system $M_2$ and the thermostat it is also necessary that the matrix of the second derivatives $S'(E,\lambda)$ at the point $E=E(T),\lambda=T$ would be defined negative. To prove now the inequality (\[TDuneq\]), we should prove that $$\begin{aligned} \frac{d^2}{dE^2}S'(E,T(E))\leq 0.\end{aligned}$$ But, $$\begin{aligned} \frac{d^2}{dE^2}S'(E,T(E))=\frac{\partial^2 S'(E,T)}{\partial E^2}+2\frac{d T(E)}{dE}\frac{\partial^2 S'(E,T)}{\partial E\partial T}+\nonumber\\ +(\frac{d T(E)}{dE})^2\frac{\partial^2 S'(E,T)}{\partial ^2 T}\leq 0,\end{aligned}$$ And the last inequality is true by virtue of the negative determination of the matrix of the second derivatives $S'(E,\lambda)$ at the point $E=E(T),\lambda=T$. Which was to be proved. In a general case when $M$ is impossible to be conceived as a direct product $M=M_1\times M_2$ it can be shown that for some covering $(\tilde{M},\tilde{H})$ of our Hamiltonian system $(M,H)$ such factorization is possible. Deriving from the very beginning all the thermodynamic relations for $(\tilde{M},\tilde{H})$, as it was done in section 3, we come again to the inequality (\[TDuneq\]). Or more exactly we want to say the following. In all our thermodynamic analysis the Hamiltonian $H$ and the integrals $K_1,...,K_k$ depended on a certain parameter $\lambda$ ( volume), $H(\lambda)$, $K_1(\lambda),...,K_k(\lambda)$. Further the dynamic variable $X(\lambda)$, depending on the parameter $\lambda$ at $\lambda=0$ will be denoted simply as $X$. We can build a covering Hamiltonian system $(\tilde{M},\tilde{H})$ of the system $(M,H)$, so that $\tilde{M}$ would be conceived as a direct product $\tilde{M}=\tilde{M}_1\times\tilde{M}_2$ and so that will measure up one more additional condition. Let us assume that $\pi$ is a canonical projection of $\tilde{M}$ on $M$ and let us assume that $\tilde{H}(\lambda)$, $\tilde{K}_1(\lambda)$,...,$\tilde{K}_k(\lambda)$ are the lift $H(\lambda)$, $K_1(\lambda)$,...,$K_k(\lambda)$ on $\tilde{M}$. The additional condition mentioned above is that the canonical coordinates on $\tilde{M}_2$ are $\tilde{K}_1,...,\tilde{K}_k$ and the conjugated variables $\varphi_1,...,\varphi_k$, and $\forall i=1,...,k$ $\varphi_i$ runs all the real axis. We will now derive all the thermodynamic relations for $\tilde{M},\tilde{H}$. But here we get the difficulty related to the fact that $\tilde{M}_2$ is not compact. There is how we propose to overcome this difficulty. We propose to deal with unnormalized distributions $w$ for which the following is true $\int \limits_{\tilde{M}}w(x)d \tilde{\Gamma}_x=\infty$, where $d \tilde{\Gamma}_x$ is the element of the phase volume on $\tilde{M}$. Let us assume that $R_L=\{x \in \tilde{M}|\forall i=1,...,k\; |\varphi_i(x)|<L\}$. To calculate the average of the dynamic variable $\tilde{f}$ on $\tilde{M}$ by $w$ we propose to use the following formula: $$\begin{aligned} \langle\tilde{f}\rangle=\lim \limits_{L\rightarrow \infty} \frac{\int \limits_{R_L} \tilde{f}(x)w(x)d\tilde{\Gamma}_x}{\int \limits_{R_L} w(x)d\tilde{\Gamma}_x}. \label{Lim}\end{aligned}$$ The existence of the limit (\[Lim\]) — a nontrivial question but we are going to show that in all the cases we are interested the limit (\[Lim\]) exists. We will say that the function $\tilde{f}$ on $\tilde{M}$ is periodical if $\forall x,y \in \tilde{M}$ such that $\pi(x)=\pi(y)$ $\tilde{f}(x)=\tilde{f}(y)$. All the lifts of the functions given on $M$ on $\tilde{M}$ are periodical and all the periodical functions can be obtained this way. All the interesting for us dynamic variables on $\tilde{M}$ appear exactly as lifts of the functions given on $M$ and that is why are periodical. Further the generalized microcanonical distributions on $\tilde{M}$, built by $\tilde{H}(\lambda),\;\tilde{K}_1(\lambda),...,\tilde{K}_k(\lambda)$ are also periodical. We are going to demonstrate that if $\tilde{f}(x)$ and $w(x)$ are periodical functions then the limit(\[Lim\]) exists. Let us assume that $d\tilde{\Gamma}^1$ is the element of the phase volume on $\tilde{M}_1$. Let us assume that $\tilde{M}_{K'_1,...,K'_k}$ is the surface of the level on $\tilde{M}$, corresponding to the values $K'_1,...,K'_k$ of the integrals $\tilde{K}_1,...,\tilde{K}_k$. Let us assume that $d\tilde{\nu}=d\tilde{\Gamma}^1 d\varphi_1,...,d\varphi_k$ is the measure on $\tilde{M}_{K'_1,...,K'_k}$. The projection$\pi$ reflects $\tilde{M}_{K'_1,...,K'_k}$ on ${M}_{K'_1,...,K'_k}$ is the surface of the level on $M$ corresponding to the values $K'_1,...,K'_k$ of the integrals $K_1,...,K_k$. On ${M}_{K'_1,...,K'_k}$ we have a measure $d\nu$, That can be characterized as a quotient of the phase volume $d \Gamma$ by $dK_1...dK_k$: $d\Gamma=d\nu dK_1...dK_k$. It is evident that the measure $d\tilde{\nu}$ is a lift of the measure $d{\nu}$ from ${M}_{K'_1,...,K'_k}$ to $\tilde{M}_{K'_1,...,K'_k}$ by virtue of the canonical projection $\pi$. To prove the conclusion about the existence of the limit (\[Lim\]), it is evidently enough to prove that for any $K'_1,...,K'_k$ the following limit $$\begin{aligned} \lim \limits_{L \rightarrow \infty} \frac{1}{L^k} \int \limits_{R_L\cap \tilde{M}_{K'_1,...,K'_k}} \tilde{f}(x) d \tilde{\nu}(x)\end{aligned}$$ exists for any (good enough) periodical function $\tilde{f}(x)=f\circ\pi(x)$. And for this it is enough, evidently, to prove that for any differentiable function $\psi$ on $\tilde{M}_1$ with a compact support there is a limit: $$\begin{aligned} \lim \limits_{L \rightarrow \infty} \frac{1}{L^k} \int \limits_{\tilde{M}_1}\psi(x)d\Gamma^1_x \int \limits_{|\varphi_i|<L}\tilde{f}(x,\varphi_1,...,\varphi_k).\end{aligned}$$ But on second thought, the existence of the last limit succeeds directly from $k$- dimensional analogue of the von Neumann ergodic theorem \[14\], applied to the commutated flows on ${M}_{K'_1,...,K'_k}$, that are induced by the Hamiltonians $K_1,...,K_k$ keeping the measure $d\nu$ and to restriction of the function $f$ on ${M}_{K'_1,...,K'_k}$. Now in order to establish all the necessary to us thermodynamic relations, for example the relation $dE=TdS-PdV$, we derive them for $\tilde{M}$ cut of $R_L$, and then tend $L$ to infinity. The only difficulty that can occur is the divergence of the entropy at $L\rightarrow \infty$, but this divergence is removed by the diminution of the logarithmically divergent constant$k \ln L$ from the entropy and this constant is the same at any value of the parameter $\lambda$. The inequality (\[TDuneq\]) that we need is now derived as before but at $\lambda=0$. It is enough to obtain again all our conclusions concerning the Ling cell. Appendix 2. Discussion of the boundary condition \[Gran\]. ========================================================== The aim of this appendix is to make more or less obvious the existence of such temperature $T_0$, above which $(\delta S)_E=\rm const \mit$. Let us assume that at $s\mapsto s+\delta s$ the integral$K$ becomes inactive. The entropy of the system $S(E,K')$ at fixed energy $E$ depends on the value $K'$ of this integral and the values of the other integrals that we will not name here. When discussing the equivalency principle we said that $S(E,K')$ should not depend on $K'$, but it would be more precise to say that $S(E,K')$ reaches the maximum on some interval $[\alpha,\beta]$ of the values $K$, and decreases drastically outside it. It is clear that $(\delta S)_E =\ln (\beta-\alpha)$. But this kind of dependence $S(E,K')$ from $K'$ is not analytical and is proper to the systems with an infinite number of degrees of freedom, i.e. in the thermodynamic limit. As an example of non-analyticity within the thermodynamic limit we can see the phase transition liquid $\leftrightarrow$ vapor. Let us see the dependence of the system pressure from its volume at temperature lower that critical. When the volume is decreasing from infinity the system pressure first begins to increase and then on some interval of volume change will remain constant. This interval corresponds to the co-existence of the liquid and vapor phases. At further volume decrease the pressure will increase again. So this way this isotherm is not describe by the analytical function. When temperature is increasing the analyticity is restored. The temperature at which the analyticity is restored is called critical. Above this temperature the difference between vapor and liquid disappears. Another example of the non-analytical behavior of the thermodynamic values which is more close to the Ling cell is borrowed from the van der Waals theory of the second-order phase transitions \[13\]. The second-order phase transition is usually related to some symmetry breaking and the breaking of this symmetry is described by a certain parameter of the order $\eta$. We will analyze here the simplest case of the real-valued $\eta$ and $\mathbb{Z}_2$-symmetry that only changes the sign $\eta$; $\eta\mapsto-\eta$. In the neighborhood of the point of phase transition the Taylor approximation of the thermodynamic potential $\Phi(\eta,T)$ accurate to the inappreciable terms with higher orders, takes the form: $$\begin{aligned} \Phi(\eta,T)=c_0\eta^4+c_1(T-T_c)\eta^2,\end{aligned}$$ where $T_c$ — critical temperature and $c_0$, $c_1$ — certain positive constants. At $T\geq T_c$ $\Phi(\eta,T)$ has by $T$ one minimum in zero. At $T<T_c$ $\Phi(\eta,T)$ has one local maximum by $\eta$ in zero and two minimums in the points $$\begin{aligned} \eta_{1,2}=\pm \sqrt{\frac{c_1}{2c_0}(T_c-T)}. \label{1,2}\end{aligned}$$ We assume that the parameter of the order $\eta$ is additive and that is why is lower that critical temperature, the analyzed system is separated into domains and in each of them $\eta$ takes one of the two opposite values, corresponding to the minimum of the thermodynamic potential of the domain. The resulting value $\eta$ will be bounded between two values $\eta_1$ and $\eta_2$. From the other side, as the domains are big enough we can neglect the interaction between the different domains when calculating the thermodynamic potential and the thermodynamic potential will be equal to the minimal value of $\Phi(\eta,T)$ by $\eta$. So this way we come to conclusion that instead of $\Phi(\eta,T)$ we should use its upper envelope $\tilde{\Phi}(\eta,T)$ (by $\eta$) and at $T>T_c$ $\tilde{\Phi}(\eta,T)$ will have the plateau, i.e. $\tilde{\Phi}(\eta,T)$ will not be the analytic function anymore. The recovery of the analyticity at the temperature increase is explained in the following way. At the temperatures high enough the kinetic energy of every particle becomes much more than the potential energy of the particle interaction and the particles can be considered non-interacting. I.e. at high temperatures all the particles that compose the system can be considered free and for such system all the characteristics must be analytical. The same way we will assume that at the temperature high enough $S(E,K')$ will be the analytical function of the parameter $K'$. Then it is either constant or reaches the maximum at an isolated point. But $(\delta S)_E =\ln (\beta-\alpha)$, as it was shown above, decreases with the temperature increase, i.e. at temperature high enough $S(E,K')$ can not be constant on the whole interval of the change of $K'$. I.e. $S(E,K')$ reaches the maximum at an isolated point and there is no more difference between using the generalized microcanonical distribution or just the microcanonical distribution. [99999]{} \[1\] Ling, G.N., *A Physical Theory of the Living State: the Association-Induction Hypothesis*, (Blaisdell Publ., Waltham, Mass., 1962). \[2\] Ling, G.N., *Life at the Cell and Below-Cell Level. The Hidden History of a Fundamental Revolution in Biology*, (Pacific Press, New York, 2001). \[3\] Ling G.N., *Debunking the alleged resurrection of the sodium pump hypothesis*, Physiol Chem Phys Med NMR. 1997; 29(2):123-198. \[4\] Bauer, E.S. *Theoretical Biology*, (M.; L., 1935. S. 2.). \[5\] Prigogin, I., *Introduction to Thermodynamics of Irreversible Processes*, (John Wiley and Sons Inc., 3rd edition, 1968). \[6\] Schrodinger, E., *What is Life? Mind and Matter*, (Cambrige University Press, 1944). \[7\] D.V. Prokhorenko, *On Non Ergodic Property of Bose Gas with Weak Pair Interaction*, Journal of Physical Mathematics, vol 1, 2009, Article ID S090701 (math-ph/0904.2868). \[8\] Glimm, J., Jaffe, A., *Quantum Physics, A Functional Integral Point of View*, (Springer-Verlag, New York, Heidelberg, Berlin, 1981). \[9\] Penrose, R.,*Structure of Space-time*, (N.Y.-Amsterdam, 1968). \[10\] Arnold, V.I., *Mathematical Methods of Classical Mechanics*, (Graduate Texts in Math., Vol 60, Springer-Verlag, New York and Berlin, 1978). \[11\] von Neumann, J., *Mathematishe Grundlagen der Quantenmechanik*, (Verlag von Julius Springer, Berlin, 1932) \[12\] Dirac, P.A.M., *The Principles of Quantum Mechanics*, (fourth edition, Oxford, at the Claredon Press, 1958). \[13\] Bogoliubov, N.N., Bogoliubov N.N. (jr.), *Introduction to Quantum Statistical Mechanics*, (Nauka, Moscow, 1984) \[14\] Lifshitz, E.M. and Pitaevskii, L.P. *Theoretical Physics, vol. 9, Statistical Physics, part 2*, (Nauka, Moscow 1978). \[15\] Lifshitz, E.M. and Pitaevskii, L.P. *Theoretical Physics, vol. 9, Statistical Physics*, (Nauka, Moscow 1995). \[16\] Riesz, F., Sz.-Nagy B., *Leçons d’Analyse Fonctionnelle*, (Akadémiai Kiadó, Budapest, 1972). [^1]: Institute of Spectroscopy, Russian Academy of Sciences, 142190 Moskow Region, Troitsk, [email protected] [^2]: Institute of Cytology, Russian Academy of Sciences, Saint Petersburg, [email protected]
{ "pile_set_name": "ArXiv" }
--- abstract: 'We introduce the notion of fully Hilbertian fields, a strictly stronger notion than that of Hilbertian fields. We show that this class of fields exhibits the same good behavior as Hilbertian fields, but for fields of uncountable cardinality, is more natural than the notion of Hilbertian fields. In particular, we show it can be used to achieve stronger Galois theoretic results. Our proofs also provide a step toward the so-called Jarden-Lubotzky twinning principle.' address: - 'Einstein Institute of Mathematics Edmond J. Safra Campus, Givat Ram, The Hebrew University of Jerusalem, Jerusalem, 91904, Israel' - 'School of Mathematical Sciences, Tel Aviv University, Ramat Aviv, Tel Aviv, 69978, Israel' author: - 'Lior Bary-Soroker' - Elad Paran title: Fully Hilbertian Fields --- Introduction ============ A field $K$ is called Hilbertian if it satisfies the following property: for every irreducible polynomial $f(X,Y)\in K[X,Y]$ which is separable in $Y$, there exist infinitely many $a\in K$ for which $f(a,Y)$ is irreducible in $K[Y]$. The name Hilbertian is derived from Hilbert’s irreducibility theorem, which asserts that number fields are Hilbertian. Hilbert’s original motivation for his irreducibility theorem was the inverse Galois problem. Assume one wants to realize a finite group $G$ over a Hilbertian field $K$. Then given a regular Galois extension $F/K(x)$ with group $G$, one can use Hilbert’s irreducibility theorem to specialize $x$ to $a\in K$ and obtain a Galois extension $\Fgag$ of $K$ with group $G$. Moreover, there exist infinitely many such specialized fields, linearly disjoint over $K$, obtained by suitable choices of $a \in K$. Even nowadays Hilbert’s irreducibility theorem remains a central approach for the inverse Galois problem, see [@MalleMatzat1999; @Voelklein1996]. Moreover, this theorem has numerous other applications in number theory, see e.g.[@Serre1989]. It is yet unknown if the inverse Galois problem has a positive solution over a Hilbertian field $K$, or equivalently, if it has a positive solution over $K(x)$. If a field $K$ is ample (e.g. $K$ is algebraically closed, PAC, or Henselian) then a theorem of Pop asserts that the inverse Galois problem has a positive solution over $K(x)$. Thus the inverse Galois problem has a positive solution over ample Hilbertian fields. The cardinality of the set of solution fields $\Fgag$ to each instance of the inverse Galois problem affects the structure of the absolute Galois group of the field $K$. Surprisingly, if $K$ is uncountable, it might happen that a regular Galois extension $F/K(x)$ has only countably many specialized solutions $\Fgag$ (see Example \[exa:jarden\]). Moreover, in order to study the absolute Galois group, one needs to study not only the possible realizations of finite groups, but also how the obtained fields fit together. This is achieved using finite embedding problems, which give rise to non-regular extensions of $K(x)$. More precisely, let $f(X,Y)$ be an irreducible polynomial that is separable in $Y$. Let $F=K(X)[Y]/(f(X,Y))$ and let $L$ be the algebraic closure of $K$ in $F$. Each $a$ in the corresponding Hilbert set $H(f)=\{a\in K \mid f(a,Y) \mbox{ irreducible}\}$ provides a specialized field $\Fgag_a$ that contains $L$ with degree $[\Fgag_a:K]=[F:K]$. If $F=L(X)$, then $L=\Fgag_a$ for all $a$, i.e., $L$ is the unique specialized field. In the non-trivial case, Hilbertianity implies the existence of infinitely many specialized fields that are linearly disjoint over $L$. However, as mentioned above, even if $K$ is uncountable it is possible that there exist only countably many such fields (and even though the cardinality of $H(f)$ always equals that of $K$). In this work we study fields which have as much and as distinct as possible specialized fields $\Fgag_a$. That is to say, there exists a subset $A\subseteq H(f)$ of cardinality $|K|$ such that the specialized fields $\Fgag_a$ are linearly disjoint over $L$ (where $a$ runs over $A$). We call a field with this feature **fully Hilbertian**. It is important to note that a *countable* field $K$ is Hilbertian if and only if it is fully Hilbertian. In particular, number fields are fully Hilbertian. The objective of this work is to initiate the study of fully Hilbertian fields. Therefore we give several equivalent definitions of fully Hilbertian fields, construct many fundamental examples of fully Hilbertian fields, and study the behavior of this notion under extensions. We also show how one can apply this notion and obtain strong Galois theoretic results over fields of large cardinality. We note that as Hilbert’s irreducibility theorem found many other applications outside the scope of Galois theory, this notion may be useful for other problems over uncountable fields. Characterizations of fully Hilbertian fields -------------------------------------------- A recent characterization of Hilbertian fields by the first author [@Bary-Soroker2008IMRN] reduces the Hilbertianity property to absolutely irreducible polynomials. We show that a similar property holds for fully Hilbertian fields: The following are equivalent for a field $K$. 1. $K$ is fully Hilbertian. 2. For every absolutely irreducible $f(X,Y)\in K[X,Y]$ that is separable in $Y$ there exist $|K|$ many $a\in H(f)$ and $b_a$ a root of $f(a,Y)$ such that $K(b_a)$ are linearly disjoint over $K$. 3. For every finite *Galois* extension $F/K(x)$ with $L$ the algebraic closure of $K$ in $F$, there exist $|K|$ many $a\in K$ such that $[\Fgag_a:K]=[F:K]$ and all $\Fgag_a$ are linearly disjoint over $L$. Finitely generated field extensions ----------------------------------- The most fundamental family of Hilbertian fields is the family of number fields. As mentioned above, since countable Hilbertian fields are fully Hilbertian, number fields are fully Hilbertian. The second important family is that of function fields: \[IT:function\_field\] Let $F$ be a finitely generated transcendental extension of an arbitrary field $K$. Then $F$ is fully Hilbertian. Extensions of fully Hilbertian fields ------------------------------------- Hilbertian fields have interesting behavior under algebraic extensions which has been well studied; [@FriedJarden2005 Chapter 13] offers a good treatment of this subject. We show that fully Hilbertian fields exhibit the same behavior. Any algebraic extension $L/K$ factors into a tower of fields $K \subseteq E \subseteq L$ in which $E/K$ is purely inseparable and $L/E$ is separable. We study each case separately. First, full Hilbertianity is preserved under purely inseparable extensions: \[IT:purely\_iseparable\] Let $E$ be a purely inseparable extension of a fully Hilbertian field $K$. Then $E$ is fully Hilbertian. The case of separable extensions is much more interesting. First, it is clear that not every extension of a fully Hilbertian (or Hilbertian) field is Hilbertian. For example, a separably closed field is not Hilbertian, and hence not fully Hilbertian. The most general result for Hilbertian fields is Haran’s diamond theorem [@Haran1999Invent]. We prove an analog of the diamond theorem, and all other permanence criteria for fully Hilbertian fields. \[IT:separable\] Let $M$ be a separable extension of a fully Hilbertian field $K$. Then each of the following conditions suffices for $M$ to be fully Hilbertian. 1. $M/K$ is finite. 2. $M$ is an abelian extension of $K$. 3. \[case:weissauer\] $M$ is a proper finite extension of a Galois extension $N$ of $K$. 4. (The diamond theorem) there exist Galois extensions $M_{1}, M_{2}$ of $K$ such that $M\subseteq M_{1}M_{2}$, but $M\not\subseteq M_{i}$ for $i=1,2$. (Our full result appears in Section \[sec:sep-ext\].) In order to prove Theorem \[IT:separable\] we identify and exploit a connection between fully Hilbertian fields and so-called semi-free profinite groups. This is a refinement of the **twinning principle** [@JardenLubotzky1992] suggested by Jarden and Lubotzky. They note a connection between results about freeness of subgroups of free profinite groups and results about Hilbertianity of separable extensions of Hilbertian fields. Furthermore, Jarden and Lubotzky state that the proofs in both cases have analogies. In spite of that they add that *it is difficult to see a real analogy between the proofs of the group theoretic theorems and those of field theory*. In [@Haran1999Group; @Haran1999Invent] Haran provides more evidence to the twinning principle by proving his diamond theorem in both cases (see also [@FriedJarden2005 Theorems 13.8.3 and 25.4.3]). Haran’s main tool in both proofs is twisted wreath products. (Using twisted wreath product one can induce embedding problems and then, on the other direction, induce weak solutions via Shapiro’s map. Haran shows that under some conditions those weak solutions are in fact solutions, i.e., surjective. We refer to this method henceforth as the **Haran-Shapiro induction**.) In this work we prove all the permanence results for fully Hilbertian fields by reducing them to the group theoretic case. In fact we show that any group theoretic result obtained via the Haran-Shapiro induction transfers an analogous field theoretic result. First we explain why a better analogy is between *semi-free profinite* groups of rank $m$ and *fully Hilbertian* fields of cardinality $m$[^1]. A semi-free group is, in a sense, a free group without projectivity, i.e., a semi-free projective group is free. These groups recently appeared with connection to Galois theory, see [@Bary-SorokerHaranHarbater]. The idea of the proof is to show that all the constructions in the Haran-Shapiro induction are “field theoretic”. This is based on the theory of rational embedding problems and geometric solutions developed in the first author’s PhD dissertation, see [@Bary-SorokerPACext]. Then all the permanence results about semi-free profinite groups of [@Bary-SorokerHaranHarbater] carry over to fully Hilbertian fields, proving Theorem \[IT:separable\]. To the best of our knowledge, this is the first time a reduction from the field theoretic case to its group theoretic counterpart has been made. We hope this is a step towards a rigorous formulation of the twinning principle. Galois theory of complete local domains --------------------------------------- Complete local domains play an important role in number theory and algebraic geometry, and their algebraic properties have been described by Cohen’s structure theorem in 1946. However, the Galois theoretic properties of their quotient fields have only recently been understood. The first result in this direction has been made in [@HarbaterStevenson2005], where Harbater-Stevenson essentially prove that the absolute Galois group of $K=K_0((X,Y))$ is semi-free of rank $|K|$, for an arbitrary base field $K_0$ (see also [@Bary-SorokerHaranHarbater]). The general case was treated independently by Pop [@Pop2009] and the second author [@Paran2010]: \[IT:com-krull-dom\] Let $K$ be the quotient field of a complete local domain of dimension exceeding 1. Then ${\textnormal{Gal}}(K)$ is semi-free of rank $|K|$. The fields $K$ in the above theorem are ample [@Pop2009]. For such fields, we have the following result: If $K$ is a fully Hilbertian ample field, then ${\textnormal{Gal}}(K)$ is semi-free of rank $|K|$. We prove a stronger results about these, namely their maximal purely inseparable extension is fully Hilbertian. \[IT:com-krull-dom-hil\] Let $K$ be the quotient field of a complete local domain of dimension exceeding 1. Then the maximal purely inseparable extension $K_{ins}$ of $K$ is fully Hilbertian. Note that if $K$ is of characteristic $0$, then $K$ itself is fully Hilbertian. Theorem \[IT:com-krull-dom-hil\] provides a new proof of Theorem \[IT:com-krull-dom\] since ${\textnormal{Gal}}(K) = {\textnormal{Gal}}(K_{ins})$. However, we note that Theorem \[IT:com-krull-dom-hil\] is stronger than Theorem \[IT:com-krull-dom\], since there exist ample fields of any characteristic with semi-free absolute Galois group that are not fully Hilbertian (Remark \[rem:free-no-hil\]). We do not know in general whether $K$ is fully Hilbertian or not. Our key result states that there is $A\subseteq K$ of independent irreducible specializations provided that all the ramification points are separable over $K$. To prove Theorem \[IT:com-krull-dom-hil\] we use the Hilbertianity of $K$. Using the density theorem of Hilbertian sets, we find irreducible specializations with the following nice feature. The geometric inertia groups of the function fields are “specialized” to inertia groups of valuations of $K$. During the Oberfolwach workshop on the arithmetic of fields after reporting this work, we learned from Pop, that he uses a similar approach to prove Theorem \[IT:com-krull-dom\]. Acknowledgment {#acknowledgment .unnumbered} -------------- We wish to thank Arno Fehm, Dan Haran, Moshe Jarden, and Franz-Viktor Kuhlmann for useful discussions about this work. Embedding problems, Hilbertian fields, and fully Hilbertian fields ================================================================== Fully Hilbertian fields in terms of polynomials ----------------------------------------------- Recall that, for an irreducible polynomial $f(X,Y)\in K[X,Y]$ that is separable in $Y$, the Hilbert set $H(f)\subseteq K$ is the set of all $a\in K$ such that $f(a,Y)$ is irreducible. It will be convenient sometimes to restrict to the co-finite subset $H'(f)\subseteq H(f)$ of all $a\in H(f)$ for which $\deg f(a,Y) = \deg_Y f(X,Y)$ and $f(a,Y)$ separable. By definition $K$ is Hilbertian if and only if all Hilbert sets are non-empty, and hence infinite. In fact the cardinality of each Hilbert set must be as large as possible. \[prop:largeHilbertsets\] Let $K$ be a Hilbertian field and $H = H(f)$ a Hilbert set over $K$. Then $|H| = |K|$. If $K$ is countable the result is trivial. Assume $m = |K| > \aleph_0$. Let $\{t_i \mid i\leq m\}$ be a transcendence basis of $K$ over its prime field $K_0$. Define a valuation $v$ on $K_0(t_i \mid i\leq m)$ inductively such that $v(t_1) = 1$ and $$j< i \longrightarrow v(t_j) > v(t_i).$$ We have $v(t_i - t_j) = v(t_i) \leq v(t_1) = 1$, for all $j< i \leq m$. Extend $v$ to $K$ arbitrarily. For each $i<m$ let $B_i = \{ x\in K \mid v(x-t_i) > 1\}$ be an open ball. If $x\in B_i$, then $v(x) = v(x-t_i + t_i) = v(t_i)$. This implies that these balls are disjoint. The proof is now done, since $H$ is dense in the topology defined by $v$ [@GeyerJarden1975 Lemma 4.1]. Let us now focus on the fields generated by a root of $f(a,Y)$ where $a$ varies over $H(f)$. The last result asserts that $|H(f)|=|K|$, hence it is somewhat surprising that the cardinality of the set of fields generated by a root of $f(a,X)$, $a\in H(f)$ can be smaller than $|K|$. The following example gives a Hilbertian field $K$ such that $|K|>\aleph_0$, but the cardinality of the set of specialized fields is countable. \[exa:jarden\] Let $K$ be a pseudo algebraically closed (PAC) field with free absolute Galois group of rank $> \aleph_0$. Then $|K|>\aleph_0$. Now the free profinite group $G$ of countable rank is a subgroup of ${\textnormal{Gal}}(K)$; let $L$ be its fixed field. Then ${\textnormal{Gal}}(L) =G$. By [@FriedJarden2005 Corollary 11.2.5] $L$ is PAC. Clearly $|L|=|K|$. Furthermore, $L$ is Hilbertian, as an $\omega$-free PAC field [@FriedJarden2005 Corollary 27.3.3]. Every finite separable extension of $L$ corresponds to an open subgroup of $G$. But the rank of $G$ is countable; hence $G$ has only countably many open subgroups, and thus there are only countably many fields generated by roots of $f(a,Y)$, $a\in H(f)$. We will use the following notation notation: Let $K$ be a field and fix a separable closure $K_s$ of $K$. Consider an irreducible polynomial $f(X,Y)\in K[X,Y]$ that is separable in $Y$ (i.e. $\frac{\partial f}{\partial Y} \neq 0$). Let $E=K(X)[Y]/(f(X,Y))$. Then $E/K(X)$ is a finite separable extension. Let $L = E\cap K_s$ be the algebraic closure of $K$ in $E$. For every $a\in H'(f)$ there is a unique prime ${{\mathfrak p}}$ of $E/L$ lying above $(X-a)$. However the residue field $E_a$ of $E$ under ${{\mathfrak p}}$ is not unique. Since we already fixed $E$ and $K_s$, and hence $L$, $E_a$ is unique up to an element $\sigma\in {\textnormal{Gal}}(L)$. In any rate, $L\subseteq E_a$ independent of the choice of $E_a$. We define fully Hilbertian fields to have as much as possible specialized fields that are linearly disjoint over $L$. \[def:ful-hil-gen\] A field $K$ is called fully Hilbertian if for every $f(X,Y)$ the following property holds. Let $E=K(X)[Y]/(f(X,Y))$ and $L = E\cap K_s$. Then there exists a set $A\subseteq H(f)$ of cardinality $|A| = |K|$ and a set of specialized fields $\{E_a \mid a\in A\}$ that are linearly disjoint over $L$. If $E \cong L(X)$, then $E_a = L$, for all $a$. However, formally, $L$ is linearly disjoint of $L$ over $L$. So in this trivial case, there always exists a set $A$ as in the definition. Definition \[def:ful-hil-gen\] is the most naïve definition. The definition is not canonical, in the sense that there are many choices of the fields $E_a$. This will be solved below, when we give a characterization of fully Hilbertian fields in terms of Galois extensions $E/K(X)$, or more accurately, in terms of rational embedding problems. We give a weaker definition of a ‘regular fully Hilbertian’ field. Later we prove that in fact it is the same notion as fully Hilbertian, cf. [@Bary-Soroker2008IMRN] for the classical case. \[def:ful-hil-abs-irr\] A field $K$ is called **regular fully Hilbertian** if for any absolutely irreducible polynomial $f(X,Y)$ there exists a set $A\subseteq H(f)$ of cardinality $|A|=|K|$ and for each $a\in A$, a root $b_a$ of $f(a,X)$ such that the fields $K(b_a)$, $a\in A$ are linearly disjoint over $K$. It is natural to consider Galois fully Hilbertian fields – fields having the strong specialization property for every irreducible polynomial $f(X,Y)$ separable and Galois in $Y$. before doing so, we introduce some useful terminology. Embedding problems ------------------ The following notions of rational embedding problem and geometric solutions are essential in this work, cf. [@Bary-SorokerPACext]. Recall that an **embedding problem** for a field $K$ is a diagram $$\xymatrix{ &{\textnormal{Gal}}(K) \ar[d]^{\nu}\\ B\ar[r]^{\alpha} &A, }$$ where $B,A$ are profinite groups and $\alpha,\nu$ are (continuous) epimorphisms. The embedding problem is **finite** (resp.**split**) if $B$ is finite (resp. $\alpha$ splits). It is **non-trivial** if $\alpha$ is not isomorphism, or equivalently, $\ker \alpha\neq 1$. A **weak solution** of $(\nu,\alpha)$ is a homomorphism $\theta\colon {\textnormal{Gal}}(K) \to B$ such that $\theta\nu = \alpha$. If $\theta$ is also surjective, we say that $\theta $ is a **proper solution**, or in short **solution**. We can always replace $A$ with ${\textnormal{Gal}}(L/K)$, where ${\textnormal{Gal}}(L) = \ker \nu$ and $\nu$ with the restriction map to get an equivalent embedding problem. In what follows we show that $B$ can also be considered as a Galois group of some geometric object. In fact, the interesting object of study are the geometric embedding problems. Let $E$ be a finitely generated regular transcendental extension of $K$, let $F/E$ be a Galois extension, and let $L = F\cap K_s$. Then the restriction map $\alpha \colon {\textnormal{Gal}}(F/E)\to {\textnormal{Gal}}(L/K)$ is surjective, since $E\cap K_s = K$. Therefore $$\label{eq:geometricep} (\nu\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(L/K), \alpha\colon {\textnormal{Gal}}(F/E)\to {\textnormal{Gal}}(L/K))$$ is an embedding problem for $K$. We call such an embedding problem a **geometric embedding problem**. If $E=K(x_1,\ldots,x_d)$ is a field of rational functions over $K$ then we call a **rational embedding problem**. In the definition the transcendence degree $d$ of $E/K$ can be arbitrary. However, in the setting of this work, it suffices to consider only the case $d=1$. This observation is based on the Bertini-Noether lemma and the Matsusaka-Zariski theorem, see [@FriedJarden2005 Proposition 13.2.1] or [@Bary-Soroker2008IMRN Lemma 2.4]. In [@FriedJarden2005 Lemma 11.6.1] it is shown (in different terminology) that every finite embedding problem for a field $K$ is equivalent to a geometric embedding problem. We shall see below that if $K$ is infinite, then any rational embedding problem has a weak solution. Hence the existence of a weak solution is necessary for an embedding problem to be equivalent to a rational embedding problem. A central result in Galois theory of Pop [@Pop1996] asserts that if $K$ is ample, then every finite split embedding problem over $K$ is regularly solvable, and in our terminology, equivalent to a rational embedding problem. In fact the regularity of finite split embedding problems immediately implies that the above necessary condition, i.e., existence of a weak solution, also suffices for an embedding problem over an ample field to be regularly solvable. See Lemma \[lem:basic-rat-eps\]. The next natural step is to consider solutions to geometric embedding problems that come from field theory, i.e., are defined by places. Before giving the definition, we introduce additional notation. Let $E/K$ be an extension of fields and let ${\varphi}$ be a place of $E$. We say that ${\varphi}$ is a $K$-place if ${\varphi}|_{K} = \id_K$. We denote by $\Egag_{\varphi}$ (or simply $\Egag$ if there is no risk of confusion) the residue field of $E$ under ${\varphi}$. \[def:geometricsolution\] Consider a geometric embedding problem . Let ${\varphi}$ be a $K$-place of $E$ and let $\Phi$ be an extension of ${\varphi}$ to an $L$-place of $F$. Assume that 1. \[def:geo-sol-a\] $\Egag = K$ and 2. \[def:geo-sol-b\] $\Phi/{\varphi}$ is unramified. Then the decomposition group $D = D_{\Phi/{\varphi}}$ is isomorphic to ${\textnormal{Gal}}(\bar F/K)$ (for a proof, apply [@FriedJarden2005 Lemma 6.1.4] to the valutation rings of $\Phi/{\varphi}$). Thus we have a canonical map $\Phi^*\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(F/E)$ whose image is $D$. Clearly $\Phi^* \nu = \alpha$, since $\Phi^*$ respects the restriction maps (recall that $\nu$ is a restriction map). In other words $\Phi^*$ is a weak solution of . Let $\theta\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(F/E)$ be a weak solution of . We say that $\theta$ is **geometric**, if there exists $\Phi/{\varphi}$ satisfying and for which $\theta = \Phi^*$. The following lemma shows that if we take ‘larger’ or ‘smaller’ embedding problems from a given rational embedding problem, the new ones are also rational. \[lem:basic-rat-eps\] Let $(\nu \colon {\textnormal{Gal}}(K)\to {\textnormal{Gal}}(L/K), \alpha\colon {\textnormal{Gal}}(F/K(x)) \to {\textnormal{Gal}}(L/K))$ be a rational embedding problem for a field $K$. 1. If $\alpha=\beta\gamma$, for some epimorphisms $\gamma\colon {\textnormal{Gal}}(F/K(x)) \to B$, $\beta\colon B\to {\textnormal{Gal}}(L/K)$, then $(\nu,\beta)$ is rational. 2. Let $\Phi^*$ be a geometric weak solution of $(\nu,\alpha)$. Then $\gamma \Phi^*$ is a geometric weak solution of $(\nu,\beta)$. 3. \[case:fib-pro-ep-is-rat\] Let $N$ be a Galois extension of $K$ that contains $L$, let $\mu\colon {\textnormal{Gal}}(K)\to {\textnormal{Gal}}(N/K)$ be the restriction map, let $C = {\textnormal{Gal}}(F/K(x))\times_{{\textnormal{Gal}}(L/K)} {\textnormal{Gal}}(N/K)$ be the fiber product, and let $\delta\colon C \to {\textnormal{Gal}}(N/K)$ and $\epsilon\colon C\to {\textnormal{Gal}}(F/K(x)$ be the quotient maps. Then there exists a natural isomorphism $C\cong {\textnormal{Gal}}(FN/K(x))$ under which $\delta,\epsilon$ are the restriction maps . In particular, $(\mu,\delta)$ is rational. 4. \[case:fib-pro-geo-sol\] Let $\Phi$ be an $N$-place of $FN$ which is unramified over $K(x)$. If $\theta=\Phi^*$ is a weak geometric solution of $(\mu,\beta)$, then $\epsilon \theta =(\Phi|_{F})^*$. In particular, $\epsilon \theta$ is a weak geometric solution of $(\nu,\alpha)$. 5. \[case:fin-pro-geo-sol-split\] If $F/K(x)$ is finite and $K$ infinite, then there exists a finite separable extension $N_0/K$ such that if $N_0\subseteq N$, then the embedding problem $(\mu,\beta)$ splits. Let $F_0$ be the fixed field of $\ker \gamma$. Then $\gamma$ induces an isomorphism $\gammagag\colon {\textnormal{Gal}}(F_0/K(x))\to B$. Thus, the following commutative diagram (recall that $\alpha$ is the restriction map) implies that $(\nu,\beta)$ is rational. $$\xymatrix{ {\textnormal{Gal}}(F/K(x))\ar[r]^-{\res}\ar[dr]^-{\gamma} &{\textnormal{Gal}}(F_0/K(x))\ar[d]^{\gammagag}\ar[dr]^{\res}\\ &B\ar[r]^{\beta} &{\textnormal{Gal}}(L/K) }$$ Now under these identifications, $\gamma \Phi^* = (\Phi|_{F_0})^*$, and hence is a geometric weak solution. To see , recall the natural isomorphism $${\textnormal{Gal}}(FN/K(x)) \cong {\textnormal{Gal}}(F/K(x)) \times_{{\textnormal{Gal}}(L/K)} {\textnormal{Gal}}(N/K) = C$$ is given by $\sigma \mapsto (\sigma|_{F}, \sigma|_{N})$. Part  is immediate since the correlation $\Phi \mapsto \Phi^*$ respects restrictions, and since $\epsilon$ is the restriction map. To prove , choose $a\in K$ such that $(x-a)$ is unramified in $F$. Extend the specialization $x\mapsto a$ to a place ${\varphi}$ of $F/L$. Let $N_0$ be the residue field of $F$ under ${\varphi}$. Then ${\varphi}^*$ is a geometric weak solution of $(\nu,\alpha)$ that factors via ${\textnormal{Gal}}(N_0/K)$. Choose $N$ as in with $N_0\subseteq N$. We need to show that $\beta\colon C\to {\textnormal{Gal}}(N/K)$ splits. Indeed, let $\beta'\colon {\textnormal{Gal}}(N/K) \to C$ be defined as follows. For each $\sigma\in {\textnormal{Gal}}(N/K)$ we write ${\varphi}^* (\sigma) = {\varphi}^*(\sigma')$, where $\sigma'\in {\textnormal{Gal}}(K)$ is an extension of $\sigma$ to $K_s$. Since ${\varphi}^*$ factors via ${\textnormal{Gal}}(N_0/K)$ it also factors via ${\textnormal{Gal}}(N/K)$, and hence ${\varphi}^*(\sigma)$ is well defined. Now we set $\beta'(\sigma) = ({\varphi}^*(\sigma), \sigma)$. Clearly $\beta'$ is a section of $\beta$, as needed. Characterization of fully Hilbertian fields ------------------------------------------- We start by formulating the Hilbertianity property in terms of geometric solutions of rational embedding problems. \[prop:hilbertian-ep\] A field $K$ is Hilbertian if and only if every rational embedding problem is geometrically solvable. We can assume that $K$ is infinite. Assume that every rational embedding problem is geometrically solvable. Let $f(X,Y)\in K[X,Y]$ be an irreducible polynomial that is separable in $Y$ and with a leading coefficient $a(X)\neq 0$. Choose some finite set $A\subseteq K$ with more elements than the numbers of roots of $a$. Let $x$ be a variable, take $F$ to be the splitting field of $\prod_{\alpha\in A} f(x+\alpha ,Y) f(\frac{1}{x}+\alpha,Y)$ over $K(x)$, and let $L=F\cap K_s$. Let $\Phi^*$ be a geometric solution of the rational embedding problem $$(\nu\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(L/K), \alpha\colon {\textnormal{Gal}}(F/K(x))\to {\textnormal{Gal}}(L/K)).$$ Assume without loss of generality that $\beta = \Phi(x) \in K$ (otherwise we repeat the following argument with $\beta= \Phi(\frac{1}{x})$). Since $|A|$ is bigger than the number of roots of $a(x)$, there exists $\alpha\in A$ for which $a(\beta+\alpha)\neq 0$. Let $y\in F$ be a root of $f(x+\alpha,Y)$ and let $\gamma = \Phi(y)$. Then $\gamma$ is finite. But since $\Phi^*$ is surjective, $[F:K(x)] = [\Fgag:K]$. In particular $$\deg_Y f(x+\alpha,Y) = [K(x,y):K(x)] = [K(\gamma),K],$$ and thus $f(\beta+\alpha, Y)$ is irreducible (recall that $f(\beta+\alpha,\gamma)=0$). Next assume that $K$ is Hilbertian and consider a rational finite embedding problem $$(\nu\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(L/K), \alpha\colon {\textnormal{Gal}}(F/K(x))\to {\textnormal{Gal}}(L/K)).$$ Let $f(x,Y)\in K[x,Y]$ be an irreducible polynomial that is separable in $Y$, a root of which generates $F/K(x)$. Let $0\neq g(x) \in K[x]$ be the product of the discriminant of $f$ over $K(x)$ and its leading coefficient. Choose an irreducible specialization $x\mapsto \alpha\in K$ of $f(x, Y)$ such that $g(\alpha)$ does not vanish. Extend this specialization to an $L$-place of $F$. Then $\Phi$ is unramified over $K(x)$. Moreover, since $f(\alpha,Y)$ is irreducible, we have that $[\Fgag:K] = [F:K(x)]$. Thus the decomposition group equals ${\textnormal{Gal}}(F/K(x))$, i.e., $\Phi^*$ is surjective. \[rem:free-no-hil\] In the above result, it is necessary that the solutions are geometric. Indeed, there exists a non-Hilbetian field $K$ whose absolute Galois group is $\omega$-free (i.e., every finite embedding problem is solvable): Let $K_0 = {\mathbb{C}}(x)$, then ${\textnormal{Gal}}(K_0)$ is an infinitely generated free group. The restriction map $\alpha\colon {\textnormal{Gal}}(K_0((x)) ) \to {\textnormal{Gal}}(K_0)$ splits ([@GeyerJarden1988 Corollary 4.1(e)]). Let $K$ be the fixed field of the image of a section of $\alpha$. Then ${\textnormal{Gal}}(K) \cong {\textnormal{Gal}}(K_0)$ is free of infinite rank. But $K$ is Henselian as an algebraic extension of the Henselian field $K_0((x))$. By [@FriedJarden2005 Lemma 15.5.4] $K$ is not Hilbertian. We wish to characterize fully Hilbertian fields in terms of geometric solutions of rational embedding problems. In the meanwhile we shall define *Galois fully Hilbertian* fields to be fields with the proper amount of geometric solutions to a rational embedding problem. Moreover we shall require that the solutions are independent in the following sense. Then we shall prove that Galois fully Hilbertian fields, regular fully Hilbertian fields and fully Hilbertian fields are the same. We call a family of solutions $\{\theta_i\mid i\in I\}$ of a non-trivial embedding problem **independent** (or **linearly disjoint**) if the solution fields, i.e. the fixed fields of $\ker \theta_i$, are linearly disjoint over $L$. The name derives from the group theoretic analog, cf.[@Bary-SorokerHaranHarbater Definition 2.3]. Before continuing, we strengthen Lemma \[lem:basic-rat-eps\] and show that independency of solutions is preserved under mild changes of the embedding problems. \[lem:basic-rat-eps\_2\] In the notation of Lemma \[lem:basic-rat-eps\] the following extra assertions hold: 1. Let $\{\Phi_i^* \mid i \in I\}$ be a family of independent geometric solutions of $(\nu,\alpha)$. Then $\{\gamma \theta_i \mid i\in I\}$ is a family of independent geometric solutions of $(\nu,\beta)$. 2. If $\{\theta_i \mid i\in I\}$ is a set of independent geometric solutions of $(\mu,\beta)$, then $\{\epsilon \theta_i \mid i\in I\}$ is a set of independent geometric solutions of $(\nu,\alpha)$. We start with the first case. Let $\Fgag_i$ be the residue field of $\Phi_i$. By assumptions, $\{\Fgag_i \mid i\in I\}$ is a linearly disjoint set over $L$. Let $\Egag_i$ be the residue field of $F_0$ under $\Phi_i|_{F_0}$. (Recall that $B\cong {\textnormal{Gal}}(F_0/K(x))$.) Then $\Egag_i\subseteq \Fgag_i$, and hence $\{\Egag_i\mid i\in I\}$ is a linearly disjoint set over $L$. Thus, by definition, $\gamma\Phi^* = (\Phi_i|_{F_0})^*$ are independent solutions of $(\nu,\beta)$. Next we prove the second case. Denote by $G,B,A$ the groups ${\textnormal{Gal}}(F/K(x))$, ${\textnormal{Gal}}(N/K)$, ${\textnormal{Gal}}(L/K)$, respectively. By [@Bary-SorokerHaranHarbater Lemma 2.5], for any $i_1,\ldots, i_n\in I$, the image of $\theta_{i_1}\times\cdots\times \theta_{i_n}$ is the fiber product $C^n_B$ of $n$ copies of $C$ over $B$. Hence by [@Bary-SorokerHaranHarbater Lemma 2.4] the image of the corresponding $\epsilon\theta_{i_1}\times\cdots\times \epsilon\theta_{i_n}$ equals the fiber product $G^n_A$ of $n$ copies of $G$ over $A$, and thus $\{\epsilon\theta_i \mid i\in I\}$ is an independent set of solutions, by [@Bary-SorokerHaranHarbater Lemma 2.5]. By Lemma \[lem:basic-rat-eps\] these solutions are geometric. A field $K$ is called **Galois fully Hilbertian** if every non-trivial rational finite embedding problem for $K$ has $|K|$ independent geometric solutions. Note that the definition of Galois fully Hilbertian is equivalent to saying that the property of fully Hilbertian fields holds for all irreducible $f(X,Y)\in K[X,Y]$ that are Galois over $K(X)$. Finite fields are not Galois fully Hilbertian, since the absolute Galois group of a finite field is pro-cyclic. The requirement on the solutions in the definition of Galois fully Hilbertian fields can be weakened, as we show in the next lemma. \[lem:splitep\] A field $K$ is Galois fully Hilbertian if and only if every non-trivial rational finite split embedding problem for $K$ has $|K|$ pairwise-independent geometric solutions. We may assume that $K$ is infinite. Consider a rational embedding problem $$\xymatrix{ &{\textnormal{Gal}}(K)\ar[d]^{\nu}\\ {\textnormal{Gal}}(F/K(x))\ar[r]^{\alpha} &{\textnormal{Gal}}(L/K). }$$ Let $N_0$ be as in Lemma \[lem:basic-rat-eps\]. Choose some finite Galois extension $N/K$ that contains $LN_0$. Then Lemma \[lem:basic-rat-eps\_2\] implies that it suffices to find a set of $m$ independent geometric solutions of the rational finite split embedding problem $(\mu,\beta)$, where $\mu\colon {\textnormal{Gal}}(K)\to {\textnormal{Gal}}(N/K)$ is the restriction map, and $$\beta\colon {\textnormal{Gal}}(F/K(x))\times_{{\textnormal{Gal}}(L/K)} {\textnormal{Gal}}(N/K) \to {\textnormal{Gal}}(N/K)$$ is the quotient map of the fiber product. By assumption there exists a set of pairwise-independent geometric solutions $\{\theta_i\mid i\in I\}$ of $(\mu,\beta)$ with $|I|=|K|$. By [@Bary-SorokerHaranHarbater Proposition 3.6] there exists a subset $I'\subseteq I$ of the same cardinality $|I'|=|K|$ for which $\{\theta_i\mid i\in I'\}$ is an independent set of geometric solutions, as needed. Standard set theoretic considerations show that it suffices to solve infinite embedding problems of cardinality $< |K|$. \[lem:&lt;m\_ep\] For a field $K$ to be Galois fully Hilbertian it suffices that any rational embedding problem in which $\rank{\textnormal{Gal}}(F/K(x)) < |K|$ and $\ker \alpha$ is finite is geometrically solvable. Consider a rational finite embedding problem $$(\nu\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(L/K), \alpha\colon {\textnormal{Gal}}(F/K(x))\to {\textnormal{Gal}}(L/K))$$ for $K$ and write $B = {\textnormal{Gal}}(F/K(x))$ and $A = {\textnormal{Gal}}(L/K)$. In particular $F$ is regular over $L$. For each $i<m$ we inductively construct a geometric solution $\theta_i$ such that the corresponding solution field $N_i$ is linearly disjoint of the compositum $\prod_{j<i} N_j$ over $L$. This would imply that $\{\theta_i \mid i< m \}$ is a family of independent geometric solutions, and hence the proof. Assume we have constructed geometric solutions $\theta_j$ for each $j < i$ that are independent. Let $N_j$ be the solution field of $\theta_j$, let $B_j= {\textnormal{Gal}}(N_j/K)$. Then $B \cong B_j$. Write $N = \prod_{j < i} N_j$ and let $\nu_i \colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(N/K)$ be the restriction map. Since the $\{\theta_j\mid j<i\}$ are independent it follows that ${\textnormal{Gal}}(N/K)$ is canonically isomorphic to the fiber product $\prod_A B_j$ of $B_j$ over $A$, where $j<i$. The isomorphism is given by $\sigma \mapsto ( (\sigma|_{N_j}))_{j<i}$ [@Bary-SorokerHaranHarbater Lemma 2.5]. Moreover, since $F$ is regular over $L$, it is linearly disjoint of $N$, and hence of $N(x)$ over $L(x)$. Hence $${\textnormal{Gal}}(FN/K(x)) \cong {\textnormal{Gal}}(F/K(x))\times_{{\textnormal{Gal}}(L/K)} {\textnormal{Gal}}(N/K) \cong B\times_A (\prod_A B_j),$$ so $\rank {\textnormal{Gal}}(FN/K(x)) \leq i+1 <m$. Let $\alpha_i \colon {\textnormal{Gal}}(FN/K(x)) \to {\textnormal{Gal}}(N/K)$ be the restriction map. The assumption gives a geometric solution $\Phi^* \colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(FN/K(x))$ of the rational embedding problem $$(\nu_i \colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(N/K), \alpha_i \colon {\textnormal{Gal}}(FN/K(x)) \to {\textnormal{Gal}}(N/K)).$$ Let $M$ and $N_i$ be the residue fields of $FN$ and $F$ under $\Phi$, respectively, and let ${\varphi}_i = \Phi|_{F}$. Then $\theta_i = {\varphi}_i^*\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(F/K(x))$ is a geometric solution. Since the residue field of $N(x)$ is $N$ and $\Phi^*$ is an isomorphism, we have $[M:N] = [FN : N(x)] = |B|/|A|$. But $M = NN_i$; so $[NN_i : N] = |B|/|A| = [N_i:L]$. Therefore $N_i$ is linearly disjoint of $N$ over $L$, as required. We are now ready to prove that the notions of Galois fully Hilbertian fields, regular fully Hilbertian fields, and fully Hilbertian fields are equivalent. \[thm:equiv-def\] Let $K$ be a field. The following conditions are equivalent. 1. $K$ is fully Hilbertian. 2. $K$ is regular fully Hilbertian. 3. $K$ is Galois fully Hilbertian. It is clear that full Hilbertianity implies regular full Hilbertianity. Assume $K$ is regular fully Hilbertian. We prove that $K$ is Galois fully Hilbertian. Let $$(\nu\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(L/K), \alpha \colon {\textnormal{Gal}}(F/K(x)) \to {\textnormal{Gal}}(L/K))$$ be a rational non-trivial finite split embedding problem. By Lemma \[lem:splitep\] it suffices to find $|K|$ independent geometric solutions of $(n_0,\alpha_0)$. By Zorn’s lemma there exists a maximal set of geometric independent solutions $\{ \theta_i \mid i\in I\}$ of $(\nu, \alpha)$. We claim that $|I| = |K|$. Assume otherwise, i.e., $|I|<|K|$. Let $N$ be the compositum of the fixed fields of $\ker \theta_i$, $i\in I$. Since $|I|<|K|$ the rank of ${\textnormal{Gal}}(N/K)$ is less than $|K|$. Thus $$\label{eq:rank<|K|} \mbox{ there exist at most $|I|$ finite subextensions of $N/K$. }$$ By assumption $\alpha$ splits, let $\alpha'\colon {\textnormal{Gal}}(L/K)\to {\textnormal{Gal}}(F/K(x)$ be a section of $\alpha$ and let $E$ be the fixed field of the image of $\alpha'$. Then $E L = F$ and $E\cap L = K$. In particular $E/K$ is regular. Choose a primitive element $y\in E$ of $E/K(x)$ that is integral over $K[x]$. Then $E=K(x,y)$ and the irreducible polynomial $f(x,Y)\in K[x,Y]$ of $y$ over $K(x)$ is monic and absolutely irreducible. Since $K$ is regular fully Hilbertian there exists a set $A\subseteq H(f)$ of cardinality $|A|=|K|$ and a choice of roots $\ygag_a$ of $f(a,Y)$ such that the fields $E_a=K(\ygag_a)$ are linearly disjoint over $K$. In particular $E_a\cap E_b = K$ for all $a\neq b$. We claim that there exist infinitely many $a\in A$ such that $E_a\cap N = K$, and hence $E_a$ and $N$ are linearly disjoint over $K$ (recall that $N/K$ is Galois). Otherwise, we would have that $E'_a = E_a \cap N\neq K$ (where $a$ runs on a co-finite subset of $A$) are $|K|$ finite subextensions of $N/K$. Since $$K\subseteq E'_a\cap E'_b \subseteq E_a\cap E_b \subseteq K,$$ it follows that $E'_a\cap E'_b = K$, hence $E'_a$, $E'_b$ are distinct. This contradicts . Choose $a\in A$ such that $E_a$ and $N$ are linearly disjoint and if ${\varphi}$ is a place of $F$ that lies over $(x-a)$, then $\Fgag = \overline{EL} = \Egag L = E_a L$. Since $E_a$ and $L$ are also linearly disjoint, we have $$[\Fgag:K] = [E_a L:K]=[E_a:K][L:K] = [E:K(x)] [L(x):K(x)] = [F:K(x)],$$ hence ${\varphi}^*\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(F/K(x))$ is surjective. Now $E_a,N$ are linearly disjoint over $K$, hence $\Fgag$, $N$ are linearly disjoint over $L$. But this means that $\{{\varphi}^*, \theta_i\}$ is a set of geometric independent solutions, a contradiction to the maximality of $\{\theta_i\mid i\in I\}$. It remains to prove that a Galois fully Hilbertian field $K$ is fully Hilbertian. Let $f(X,Y)$ be an irreducible polynomial that is separable in $Y$. Let $x$ be a variable, and let $y$ be a root of $f(x,Y)$ in some fixed algebraic closure of $K(x)$. Let $E= K(x,y)$. Then $E/K(x)$ is a finite separable extension. Let $L_0=E\cap K_s$. Choose a Galois extension $F/K(x)$ such that $E\subset F$ and let $L=F\cap K_s$. Then $EL\subseteq F$ and, since $E$ is regular over $L_0$, $$\label{eq:EL:L=E:L_0} [EL:L] = [E:L_0].$$ Since $K$ is Galois fully Hilbertian, it follows that there exists a family $\{{\varphi}_i^*\mid i\in I\}$ of independent geometric solutions of the non-trivial finite rational embedding problem $(\mu \colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(L/K), \alpha\colon {\textnormal{Gal}}(F/K(x)) \to {\textnormal{Gal}}(L/K))$, and $|I|=|K|$. The set of $i\in I$ for which one of the following conditions is not satisfied is finite, hence we can assume that every $i\in I$ satisfies these conditions. 1. $a_i = {\varphi}_i(x)$ and $b_i = {\varphi}(y_i)$ are finite, for all $i\in I$. 2. $f(a_i,b_i) = 0$. 3. The residue field of $E$ under ${\varphi}_i$ is $\Egag_i=K(b_i)$. 4. The residue field of $EL$ under ${\varphi}_i$ is $\overline{EL}_i = \Egag_i L = L(b_i)$ Since ${\varphi}_i^*$ is surjective for all $i\in I$, we have $[\Fgag_i :K] = [F:K(x)]$, and hence $[\Rgag_i : \Sgag_i]$ for any tower $K(x)\subseteq S\subseteq R\subseteq F$. Hence to finish the proof, it suffices to prove that the $\Egag_i$ are linearly disjoint over $L_0$, for any finite subset $J$ of $I$. Let $J$ be a finite subset of $I$. Since $\{{\varphi}^*_i\mid i\in J\}$ are independent, the fields $\Fgag_i$, $i\in J$ are linearly disjoint over $L$. In particular $\Egag_i L$ are linearly disjoint over $L$. By we have $$\begin{aligned} [E:L_0]^{|J|} &=& \prod_{i\in J} [E:L_0] = \prod_{i\in J} [\Egag_i:L_0] \geq [\prod_{i\in J} \Egag_i :L_0] \geq [\prod_{i\in J} (\Egag_i L) : L] \\ &=& \prod_{i\in J} [\Egag_i L : L] = \prod_{i\in J} [E L : L] = [EL:L]^{|J|} = [E:L_0]^{|J|}.\end{aligned}$$ We thus get that all inequalities are equalities, and thus $\Egag_i$ are linearly independent over $L_0$. For countable fields, Hilbertianty is the same as full Hilbertianity. This is an immediate consequence of Proposition \[prop:hilbertian-ep\] and Lemma \[lem:&lt;m\_ep\]. \[cor:countableHilbertian\] Let $K$ be a countable field. Then $K$ is Hilbertian if and only if $K$ is fully Hilbertian. The absolute Galois group of fully Hilbertian fields ---------------------------------------------------- In this section we consider the group theoretic properties of the absolute Galois group of fully Hilbertian fields. Recall the definition of semi-free groups [@Bary-SorokerHaranHarbater]. A profinite group $\Gamma$ of infinite rank $m$ is called **semi-free** if every finite split embedding problem for $\Gamma$ has a set of $m$ independent solutions. The following conjecture is the analog of Conjecture “Split EP/$_{K hilb.}$" of [@DebesDeschamps1997]. \[conj:semi-free\] The absolute Galois group of a fully Hilbertian field $K$ is semi-free of rank $|K|$. In [@DebesDeschamps1997] Dèbes and Deschamps show that “Split EP/$_{K hilb.}$" follows from the conjecture they call “Split EP/$_{k(t)}$." We show that a slightly stronger property implies Conjecture \[conj:semi-free\]. \[prop:Cond-semi-free\] Let $K$ be a fully Hilbertian field and assume that any finite split embedding problem for $K$ is equivalent to a rational embedding problem. Then ${\textnormal{Gal}}(K)$ is semi-free of rank $|K|$. Any split embedding problem over $K$ is equivalent to a rational embedding problem. Therefore, by definition there exist $|K|$ independent geometric solutions. Recall that a field $K$ is called **ample** (or **large**) if every curve defined over $K$ with a simple $K$-rational point has infinitely many $K$-rational points. In [@Pop1996], Pop proves that the assumption of Proposition \[prop:Cond-semi-free\] holds for an ample field, see also [@HaranJarden1998]. Therefore we get: \[cor:f-Hil+ample=semi-free\] Let $K$ be a fully Hilbertian ample field. Then ${\textnormal{Gal}}(K)$ is semi-free of rank $|K|$. A stronger property than being ample is being PAC: A field $K$ is PAC if every curve defined over $K$ has infinitely many $K$-rational points. The next result extends Roquette’s Theorem [@FriedJarden2005 Corollary 27.3.3]. Let $K$ be a PAC field. Then $K$ is fully Hilbertian if and only if ${\textnormal{Gal}}(K)$ is free of rank $|K|$. Assume $K$ is fully Hilbertian. Then Corollary \[cor:f-Hil+ample=semi-free\] implies that ${\textnormal{Gal}}(K)$ is semi-free of rank $|K|$. But [@FriedJarden2005 Theorem 11.6.2] implies that ${\textnormal{Gal}}(K)$ is projective, and thus ${\textnormal{Gal}}(K)$ is free of rank $|K|$ [@Bary-SorokerHaranHarbater Theorem 3.5]. Assume that ${\textnormal{Gal}}(K)$ is free of rank $|K|$. Then any finite embedding problem has $m$ independent solutions [@FriedJarden2005 Propsition 25.1.6]. Since $K$ is PAC, any solution of a geometric embedding problem is geometric [@Bary-SorokerPACext Corollary 3.4]. Therefore any rational finite embedding problem for $K$ has $m$ independent geometric solutions, i.e., $K$ is fully Hilbertian. We now consider arbitrary fields. We give a result about embedding problems with abelian kernel. \[prop:abelian-ker\] Let $K$ be a fully Hilbertian field of cardinality $m$, $A$ an abelian group, $L/K$ a finite Galois extension with Galois group $G={\textnormal{Gal}}(L/K)$. Assume that $G$ acts on $A$. Then the infinite group $A^m \rtimes G$ occurs as a Galois group over $K$. In particular $A^m$ occurs as a Galois group over $K$. Let $\mu\colon {\textnormal{Gal}}(K) \to G$ be the restriction map, and $\alpha\colon A\rtimes G \to G$ the quotient map. By Ikeda’s theorem [@FriedJarden2005 Proposition 16.4.4], the embedding problem $(\mu,\alpha)$ is equivalent to a rational embedding problem. Since $K$ is fully Hilbertian, there exists a set $S$ of $m$ independent solutions. The fiber product of $m$ copies of $A\rtimes G$ over $G$ is isomorphic to $A^m\rtimes G$ [@Bary-SorokerHaranHarbater Lemma 4.3] and the corresponding solution $$\prod_{\theta\in S} \theta \colon {\textnormal{Gal}}(K) \to A^m\rtimes G$$ is surjective [@Bary-SorokerHaranHarbater Lemma 2.5]. For any field $K$, the rank of ${\textnormal{Gal}}(K)$ is bounded by $|K|$. Taking $A={\mathbb{Z}}/2{\mathbb{Z}}$ in Proposition \[prop:abelian-ker\], we get the following result: If $K$ is fully Hilbertian, then the rank of ${\textnormal{Gal}}(K)$ equals $|K|$. Rational Function Fields ======================== In this section we show that the basic families of Hilbertian fields — number fields and function fields — are fully Hilbertian. Since for a countable field full Hilbertianity is equivalent to Hilbertianity, we are left with function fields over an infinite field. Our proof is based on an explicit description of elements in Hilbert sets and on the fact that full Hilbertianity is preserved by finite extensions. The latter assertion will be proved below, see Corollary \[cor:finite-ext\]. Let $K_0$ be an infinite field and $K=K_0(t)$ a rational function field over $K_0$. Then each Hilbert set contains a set $$\{ a + bt \mid g(a,b)\neq 0\},$$ for some nonzero polynomial $g(A,B)\in K_0[A,B]$ [@FriedJarden2005 Proposition 13.2.1]. In this section we use this description in order to show that $K$ is fully Hilbertian. \[thm:ratff\] Let $K_0$ be a field and let $K=K_0(t)$ be a rational function field. Then $K$ is fully Hilbertian. If $K_0$ is countable, then Hilbertianity is the same as fully Hilbertianity. Hence $K$ is fully Hilbertian. Assume that $m = |K_0|$ is infinite. By Lemma \[lem:&lt;m\_ep\] it suffices to geometrically solve a rational embedding problem $$(\nu\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(L/K), \alpha\colon {\textnormal{Gal}}(F/K(x)) \to{\textnormal{Gal}}(L/K))$$ under the assumption that $\rank {\textnormal{Gal}}(F/K(x)) = n < m$. Since the number of finite subextensions of $F/K(x)$ is bounded by $n$ (and actually equals $n$ if $n$ is infinite), we can order the finite subextensions $E\neq K(x)$ of $F/K(x)$, say $\{E_i \mid i\leq n\}$. For each $E_i$ there exists a Hilbert set $H_i$ such that for each $u\in H_i$ and for each $K$-place ${\varphi}$ of $E_i$ that lies over $(x-u)$ the degree of the residue field extension equals to $[E_i:K(x)]$. In particular ${\varphi}$ is unramified in $E_i$. Let $g_i(A,B)\in K_0[A,B]$ be the nonzero polynomial such that $\{a +bt\mid a,b\in K_0\mbox{ and }g(a,b)\neq 0\} \subseteq H_i$. Write $C_i(K_0) = \{ (a,b)\in K_0^2 \mid g(a,b) = 0\}$ for the corresponding curve. Now since $n<m$ $$K_0^2 \neq \bigcup_{i\leq n} C_i(K_0).$$ Let $(a,b)\in K_0^2 \smallsetminus \bigcup_{i\leq n} C_i(K_0)$. Extend the specialization $x\mapsto a + b t\in K$ to a $K$-place ${\varphi}$ of $F$. For each $i\leq n$, $g_i(a,b)\neq 0$, hence ${\varphi}|_{E_i}$ is unramified over $K(x)$. Therefore ${\varphi}$ is unramified over $K$, and thus the geometric weak solution ${\varphi}^*\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(F/K(x))$ of $(\nu,\alpha)$ is well defined. We claim that ${\varphi}^*\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(F/K(x))$ is a solution, i.e., surjective. To show that it suffices to show that $D_{\varphi}={\textnormal{Gal}}(F/K(x))$, where $D_{\varphi}$ is the corresponding decomposition group. Let $\Delta$ be the decomposition field (i.e. the fixed field of $D_{\varphi}$ in $F$). We need to show that $\Delta = K(x)$. If $\Delta \neq K(x)$, then there exists a finite extension $E_i$ of $K(x)$ such that $E_i \subseteq \Delta$. Hence the residue field $\bar E_i$ of $E_i$ is $K$. But since $g_i(a,b) \neq 0$ we have $[\bar E_i : K] = [E_i:K(x)] > 1$. This contradiction implies that $\Delta = K(x)$, and hence ${\varphi}^*$ is an epimorphism, which concludes the proof. Let $F$ be a finitely generated transcendental extension of an arbitrary field $K$. Then $F$ is fully Hilbertian. Let $(t_1,\ldots, t_r)$ be a transcendence basis of $F/K$. Then $F$ is a finite extension of $K_0(t_1, \ldots, t_r)$. By Theorem \[thm:ratff\] the field $K(t_1,\ldots, t_r) = K(t_1,\ldots,t_{r-1})(t_r)$ is fully Hilbertian. By Corollary \[cor:finite-ext\] below, $F/K(t_1,\ldots, t_r)$ is fully Hilbertian. The twinning principle {#sec:twinning} ====================== We mentioned in the introduction the *twinning principle* formulated by Lubotzky-Jarden. This principle suggests there should be a connection between results about free subgroups of a free profinite group and results about Hilbertian separable extensions of a Hilbertian field. We also remarked in the introduction that the *Haran-Shapiro induction* gives a strong evidence to this principle, since it proves the diamond theorem in both settings, and the proofs are analogous. In fact results about Hilbertian fields transfer to results about free profinite groups of countable rank, using Roquette’s theorem, see [@JardenLubotzky1992 Weak twinning principle]. In this section we further study the twinning principle. We first claim that the connection should be between semi-free profinite groups of rank $m$ and fully Hilbertian fields of cardinality $m$. 1. The connection is evident from the characterization of fully Hilbertian fields as Galois fully Hilbertian fields (Theorem \[thm:equiv-def\]). A semi-free group $\Gamma$ of rank $m$ is a profinite group for which every non-trivial finite split embedding problem has $m$ independent solutions, while a Galois fully Hilbertian field of cardinality $|K|=m$ is a field for which every non-trivial *rational* embedding problem has $m$ independent *geometric* solutions. 2. The absolute Galois group of a fully Hilbertian field is not projective in general, and free groups are projective. In the meanwhile semi-free groups are, in a sense, free group without projectivity ([@Bary-SorokerHaranHarbater Theorem 3.5]). 3. A fully Hilbertian ample field has a semi-free group rank $|K|$ (Corollary \[cor:f-Hil+ample=semi-free\]), and conditionally this is always the case (Proposition \[prop:Cond-semi-free\]). Maybe the most convincing argument for this connection is the following 1. \[case:twinning-carry-over\] The proofs about sufficient conditions for a subgroup of a semi-free group to be semi-free carry over to the field theoretic setting. If one wants to consider the more classical setting of Hilbertian fields, then our method gives the connection with the property that each finite split embedding problem has a solution. One can even consider only semi-free groups of countable rank. Let us go into more details. First note that there is a certain trade-off here – in the group theoretic setting we need to solve all finite split embedding problems, while in field theory only rational embedding problems come into the picture. On the other hand, in the field theoretic setting we need the solution to have an extra property, namely only geometric solutions are considered. The way we transfer group theoretic proofs into field theory is via the Haran-Shapiro induction. This method uses fiber products and twisted wreath products in order to induce solutions of embedding problems from the group to its subgroup. So what we need to do is to prove, or essentially to observe, that all constructions of the Haran-Shapiro induction are field theoretic constructions. That is to say, (1) if one starts from a rational embedding problem, then the constructed embedding problems are also rational (2) the solutions induced by the constructions preserve the property of being geometric. As an immediate consequence of this, all the permanence properties of semi-free groups carry over to fully Hilbertian fields. In particular we get all cases of Theorem \[IT:separable\]. Fiber products and twisted wreath products ------------------------------------------ As indicated before in results about semi-free subgroups of semi-free groups there are two group theoretic constructions involved, fiber products and twisted wreath products. Let us start with the more elementary one, fiber products. Write $\Gamma = {\textnormal{Gal}}(K)$ and consider an embedding problem $(\mu \colon \Gamma \to G, \alpha\colon H\to G)$ for $K$. Let $\Delta$ be a normal subgroup of $\Gamma$ contained in $\ker \mu$. Then $\mu$ factors as $\mu=\mugag \muhat$, where $\muhat\colon \Gamma \to \Ghat:=\Gamma/\Delta$ is the quotient map and $\mugag\colon \Ghat \to G$ is canonically defined. This data defines a corresponding **fiber product embedding problem**: $$\xymatrix{ &\Gamma\ar[d]_{\muhat}\ar@/^/[dd]^{\mu}\\ H\times_{G} \Ghat \ar[r]^{\alphahat}\ar[d]^{\beta} &\Ghat\ar[d]_{\mugag}\\ H\ar[r]^{\alpha} &G. }$$ Here $H\times_{G} \Ghat = \{ (h,\ghat)\in H\times \Ghat \mid \alpha(h) = \mugag(\ghat)\}$ and $\alphahat, \beta$ the canonical projections. A (weak) solution $\thetahat\colon \Gamma\to H\times_G \Ghat$ of the fiber product embedding problem induces the (weak) solution $\theta=\beta\thetahat$ of $(\mu,\alpha)$. The following result asserts that this construction preserves field theoretic aspects. \[lem:fiber\_split\] In the above notation, if $(\mu,\alpha)$ is geometric, then so is $(\muhat,\alphahat)$. Moreover, if $\thetahat$ is geometric, so does $\theta$. More precisely, assume $H = {\textnormal{Gal}}(F/K(x))$, $G={\textnormal{Gal}}(L/K)$ for $L=F\cap K_S$, and $\alpha,\mu$ are the restriction maps. Let $N$ be the fixed field of $\Delta$. Then $\Ghat = {\textnormal{Gal}}(N/K)$, $H\times_{G} \Ghat = {\textnormal{Gal}}(FN/K(x))$, and $\alphahat,\beta$ are the corresponding restriction maps. Moreover, if $\thetahat = \hat {{\mathfrak P}}^*$ for some unramified prime extension $\hat {{\mathfrak P}}/(x-a)$ of $FN/K(x)$, then $\theta = (\hat{{\mathfrak P}}\cap F)^*$. See [@Bary-SorokerPACext Lemma 2.6]. Now we move to the more sophisticated construction of twisted wreath products. For the reader’s convenience we give a short survey on the Haran-Shapiro induction, starting from the definitions. See [@Bary-SorokerHaranHarbater] for full details and proofs. Let $A$, $G_0 \leq G$ be finite groups. Assume that $G_0$ acts on $A$ (from the right). Let $${{\rm Ind}}_{G_0}^G(A) = \{ f\colon G\to A\mid f(\sigma\rho) = f(\sigma)^\rho,\ \forall \sigma\in G, \rho\in G_0\} \cong A^{(G:G_0)}.$$ Then $G$ acts on ${{\rm Ind}}_{G_0}^G(A)$ by $f^\sigma(\tau) = f(\sigma \tau)$. We define the **twisted wreath product** to be the semidirect product $$A{\wre}_{G_0} G = {{\rm Ind}}_{G_0}^G(A) \rtimes G,$$ i.e., an element in $A{\wre}_{G_0} G$ can be written uniquely as $f\sigma$, where $f\in {{\rm Ind}}_{G_0}^G(A)$ and $\sigma\in G$. The multiplication is then given by $(f\sigma) (g\tau) = f g^{\sigma^{-1}} \sigma\tau$. The twisted wreath product is equipped with the quotient map $\alpha\colon A{\wre}_{G_0} G\to G$ defined by $\alpha(f\sigma) = \sigma$. The map $\pi \colon {{\rm Ind}}_{G_0}^G(A) \rtimes G_0 \to A\rtimes G_0$ defined by $\pi (f \sigma) = f(1) \sigma$ is called the **Shapiro map**. It is an epimorphism, since, for $f\in {{\rm Ind}}_{G_0}^G(A)$ and $\sigma\in G_0$, we have $$\pi(f^\sigma)=f(\sigma)=f(1)^\sigma = \pi(f)^\sigma.$$ A tower of fields $$K\subseteq L_0 \subseteq L\subseteq F\subseteq \Fhat$$ is said to **realize** the twisted wreath product $A{\wre}_{G_0} G$ if $\Fhat/K$ is a Galois extension with Galois group isomorphic to $A{\wre}_{G_0} G$ and the tower of fields corresponds to the subgroups (via taking fixed fields in $\Fhat$) $$A{\wre}_{G_0} G \geq {{\rm Ind}}_{G_0}^G(A)\rtimes G_0\geq {{\rm Ind}}_{G_0}^G(A) \geq \ker \pi \geq 1.$$ We note that the extension $F/L_0$ is Galois with group $A\rtimes G_0$, and the restriction map ${\textnormal{Gal}}(\Fhat/L_0)\to {\textnormal{Gal}}(F/L_0)$ is the corresponding Shapiro map of $A{\wre}_{G_0}G$. For more details see [@FriedJarden2005 Remark 13.7.6] or [@Haran1999Invent Remark 1.2]. Let $\Lambda \leq \Gamma$ be profinite groups. Consider a finite split embedding problem $(\mu_1\colon \Lambda \to G_1, \beta_1\colon A\rtimes G_1\to G_1)$ for $\Lambda$. To use the Haran-Shapiro induction, first we choose an open normal subgroup $\Delta$ of $\Gamma$ such that $\Delta\cap \Gamma\leq \ker \mu_1$. Let $\mu\colon \Gamma\to G:=\Gamma/\Delta$ be the quotient map and $\mu_0=\mu|_{\Lambda}\colon \Lambda\to G_0:=\Lambda/\Lambda\cap\Delta$. Then $\mu_1$ factors as $\mu_1 = \mugag_1\mu_0$. Moreover we have the following embedding problems $$\label{eq:EPinHSI} \xymatrix@C=40pt{ &\Lambda\ar[d]_{\mu_0} \ar@/^10pt/[ddd]^{\mu_1}\\ {{\rm Ind}}_{G_0}^G(A)\rtimes G_0 \ar[r]^{\beta} \ar[d]^{\pi} &G_0\ar@{=}[d]\\ A\rtimes G_0 \ar[r]^{\beta_0}\ar[d]^{\rho} &G_0\ar[d]_{\mugag_1}\\ A\rtimes G_1 \ar[r]^{\beta_1} &G_1. } \qquad \xymatrix@C=40pt{ &\Gamma\ar[d]^{\mu}\\ A{\wre}_{G_0} G \ar[r]^{\alpha} &G. }$$ Here $\beta$ is the restriction of $\alpha$ to ${{\rm Ind}}_{G_0}^G(A)\rtimes G_0$, $\pi$ is the Shapiro map, $A\rtimes G_0 \cong (A\rtimes G_1)\times_{G_1} G_0$, and $\rho,\beta_0$ are the projection maps. Note the the lower square in the left diagram is the fiber product embedding problem. Now, if $\theta\colon \Gamma\to A{\wre}_{G_0} G$ is a weak solution, then $\theta_0 = \rho\pi \theta|_{\Lambda}$ is a weak solution of the original embedding problem $(\mu_1,\beta_1)$. The problem here is that if $\theta$ is a solution, i.e., surjective, it does not imply that $\theta_0$ is a solution. So we are lead to Haran’s contribution to the method: Under some conditions on the subgroup $\Lambda$ (e.g., open, $\Gamma'\leq \Lambda$, or most generally, contained in a diamond) the Haran-Shapiro induction gives that, taking $\Delta$ small enough, or equivalently, $G$ large enough, we have $$\theta \mbox { surjective} \Longrightarrow \theta_0 \mbox { surjective}.$$ Moreover, if $\{\theta_i \mid i\in I\}$ is a family of pairwise-independent solutions of $(\mu,\alpha)$, then the induced solutions $\{\theta_{i0} \mid i\in I\}$ of $(\mu_1, \beta_1)$ are also pairwise-independent. Now we show that these constructions carry over to the field theoretic setting. For this assume that $M$ is a separable extension of an *infinite* field $K$, $\Gamma = {\textnormal{Gal}}(K)$, and $\Lambda = {\textnormal{Gal}}(M)$. Also denote by $L$ the fixed field of $\Delta$, i.e., $G = {\textnormal{Gal}}(L/K)$. \[lem:wre\_prod\] If $(\mu_1,\beta_1)$ is a rational embedding problem for $M$, then there exists a finite separable extension $K'/K$ such that if $K'\subseteq L$, then all the embedding problems in are rational, and all maps are restriction maps. Moreover, if a weak solution $\theta$ of $(\mu,\alpha)$ is geometric, then the induced weak solution $\theta_0 = \rho\pi\theta|_{{\textnormal{Gal}}(M)}$ of $(\mu_1,\beta_1)$ is also geometric. More precisely, assume that $A\rtimes G_1 = {\textnormal{Gal}}(F_1/M(x))$, $G_1 = {\textnormal{Gal}}(N_1/M)$ for $N_1 = F_1\cap M_s$, and $\beta_1,\mu_1$ are the restriction maps. Then there exists a finite separable extension $K'/K$ such that for every finite Galois extension $L$ of $K$ that contains $K'$ we have: 1. \[lem:wre\_prod\_a\] $N_1 \subseteq ML$. Let $L_0=L\cap M$ and $G_0 = {\textnormal{Gal}}(ML/L) \cong {\textnormal{Gal}}(L/L_0)$. 1. \[lem:wre\_prod\_b\] There exist $x_0\in L_0(x)$ such that $L_0(x_0) = L_0(x)$ and a tower of fields $$K(x_0) \subseteq L_0(x) \subseteq L(x) \subseteq F \subseteq \Fhat$$ that realizes $A{\wre}_{G_0} G$. In particular, we identify ${\textnormal{Gal}}(\Fhat/K(x_0)) = A{\wre}_{G_0} G$. 2. \[lem:wre\_prod\_c\] $\Fhat$ is regular over $L$. 3. \[lem:wre\_prod\_d\] Let $\Fhat_0 = \Fhat M$, $F_0 = FM$, and $N_0 = LM$. Then $N_0/L$, $\Fhat_0/M(x)$, and $F_0/M(x)$ are Galois, we have $$\begin{aligned} {\textnormal{Gal}}(N_0/M) &\cong G_0,\\ {\textnormal{Gal}}(\Fhat_0/ M(x)) &\cong {{\rm Ind}}_{G_0}^G (A) \rtimes G_0,\\ {\textnormal{Gal}}(F_0/M(x))& \cong A\rtimes G_0,\end{aligned}$$ and under these identification the maps in are the restriction maps. 4. \[lem:wre\_prod\_e\] Let $\theta\colon {\textnormal{Gal}}(K) \to A{\wre}_{G_0} G$ be a geometric weak solution, i.e., $\theta = \phihat^*$ for some place $\phihat$ of $\Fhat$ that is unramified over $K(x)$, and extends a specialization $x\mapsto a\in K$. Extend $\phihat$ to an $M$-place $\psihat$ of $\Fhat_0$. Then $\theta_0 := \rho\pi\theta|_{{\textnormal{Gal}}(M)} = (\psihat|_{F_1})^*$. Since the restriction map ${\textnormal{Gal}}(F_1/M(x)) \to {\textnormal{Gal}}(N_1/M)$ splits, we can lift $G_1 = {\textnormal{Gal}}(N_1/M)$ to a subgroup of ${\textnormal{Gal}}(F_1/M(x))$. Let $E$ be its fixed field. Then $E\cap N_1 = M$ and $EN_1 = F_1$. Let $y\in E$ be a primitive element of $E/M(x)$, i.e., $E = M(x,y)$. Then $y$ is also a primitive element of $F_1/N_1(x)$. Then the irreducible polynomial $f(x,Y)\in M[x,Y]$ of $y$ over $M(x)$ is Galois over $N_1(x)$. Let $K'/K$ be a finite extension such that $N_1 \subseteq MK'$, the coefficients of $f(x,Y)$ are in $K'\cap M$, and $f(x,Y)$ is Galois over $K'(x)$. Let $L$ be a finite Galois extension of $K$ that contains $K'$. Then, in particular, holds, $f(x,Y)\in L_0[x,Y]$ for $L_0 = L\cap M$, and $f(x,Y)$ is Galois over $L(x)$. Choose a basis $c_1, \ldots, c_n$ of $L_0/K$ and let $x_0, x_1, \ldots, x_n$ be an $(n+1)$-tuple of variables. Now by [@Haran1999Invent Lemma 3.1] there exist fields $F', \Fhat'$ such that (denoting ${\mathbf x}= (x_1,\ldots, x_n)$) $$K({\mathbf x}) \subseteq L_0({\mathbf x}) \subseteq L({\mathbf x}) \subseteq F'\subseteq \Fhat'$$ realizes $A{\wre}_{G_0} G$ and $\Fhat'$ is regular over $L$. Furthermore, $F'$ is generated by a root of $f(\sum_{i=1}^n c_i x_i, Y)$ over $L({\mathbf x})$. A usual combination of the Bertini-Neother Lemma and the Matsusaka-Zariski Theorem (see, e.g., [@Bary-Soroker2008IMRN Lemma 2.4]) implies that we can substitute $x_i$ with $\alpha_i+\beta_i x_0$ for some $\alpha_i,\beta_i\in K^\times$ and extend this substitution to a $K$-place of $\Fhat'$ with residue field $\Fhat$ with the same properties. That is, if we denote by $F$ the residue field of $F'$, then $$K(x_0) \subseteq L_0(x_0) \subseteq L(x_0) \subseteq F \subseteq \Fhat$$ realizes $A{\wre}_{G_0} G$, $\Fhat$ is regular over $L$, and $F$ is generated by a root of $f(\sum_{i=1}^n c_i (\alpha_i + \beta_i x_0),Y)\in L_0[x_0,Y]$ over $L(x_0)$. Since $c_1,\ldots,c_n$ are linearly independent over $K$, we have that $\sum_{i=1}^n \beta_ic_i\neq 0$. Therefore $u=\sum_{i=1}^n c_i (\alpha_i + \beta_i x_0)$ is a transcendental element and $L_0(x_0) = L_0(u)$. We may apply an isomorphism that sends $u$ to $x$ to assume that $$K(x_0) \subseteq L_0(x) \subseteq L(x) \subseteq F \subseteq \Fhat$$ realizes $A{\wre}_{G_0} G$, $\Fhat$ regular over $L$, and $F$ is generated by the root $y$ of $f(x,Y)$. This finishes the proof of and . Let $N_0=ML$, $F_0 = F M $, $\Fhat_0 = \Fhat L$, as in . Note that $F_0 = FM =K(x,y) LM= N_0(x,y) $. Since $\Fhat$ is regular over $L$, we have $L_0 = \Fhat\cap M = L\cap M$, and thus ${\textnormal{Gal}}(N_0/M) \cong {\textnormal{Gal}}(L/L_0)=G_0$. Similarly we have $${\textnormal{Gal}}(\Fhat_0/M(x)) \cong {\textnormal{Gal}}(\Fhat/L_0(x)) \cong {{\rm Ind}}_{G_0}^G(A)\rtimes G_0$$ and ${\textnormal{Gal}}(F_0/M(x)) \cong {\textnormal{Gal}}(F/L_0(x)) \cong A\rtimes G_0 $. Since all isomorphisms are defined canonically, the maps in are the restriction maps. It remains to prove . Let $\phihat^*\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(\Fhat/K(x))$ be a weak geometric solution. Extend $\phihat$, $M$-linearly, to a place $\psihat$ of $\Fhat_0$. Then $\psihat$ is unramified over $M(x)$, and $\phihat^*|_{{\textnormal{Gal}}(M)} = \psihat^*$. Moreover, since all maps in are the restriction maps, we get that $\rho\pi (\phihat^*|_{{\textnormal{Gal}}(M)}) = \rho\pi (\psihat^*) = (\psihat|_{F_1})^*$. Separable extensions of fully Hilbertian fields {#sec:sep-ext} ----------------------------------------------- The following result is more general than Theorem \[IT:separable\], cf. Main Theorem of [@Bary-SorokerHaranHarbater]. \[thm:sep-ext\] Let $M$ be a separable extension of a fully Hilbertian field $K$. Then each of the following conditions suffices for $M$ to be fully Hilbertian. 1. \[case:sep-ext-finite\] $[M:K]<\infty$. 2. \[case:sep-ext-fin-gen\] The Galois group $\Gamma$ of the Galois closure of $M/K$ is finitely generated. 3. The cardinality of the set of all finite subextensions of $M/K$ is less than $|M|$. 4. $M$ is a proper finite extension of a Galois extension $N/K$. 5. $M$ is an abelian extension of $K$. 6. \[case:sep-ext-diamond\] There exist Galois extensions $M_{1}, M_{2}$ of $K$ such that $M\subseteq M_{1}M_{2}$, but $M\not\subseteq M_{i}$ for $i=1,2$. 7. $M$ is contained in a Galois extension $N/K$ with pronilpotent group and $[M:K]$ is divisible by at least two primes. 8. $M/K$ is sparse. 9. $[M:K] = \prod p^{\alpha(p)}$, where $\alpha(p)<\infty$ for all $p$. Let $m=|K|$ and $(\mu_1 \colon {\textnormal{Gal}}(M) \to {\textnormal{Gal}}(N_1/M), \alpha \colon {\textnormal{Gal}}(F_1/M(x)) \to {\textnormal{Gal}}(N_1/M))$ be a rational split embedding problem for $M$. Then if we write $G_1 = {\textnormal{Gal}}(N_1/M)$, then ${\textnormal{Gal}}(F_1/M(x)) \cong A\rtimes G_1$ and $\alpha$ the quotient map. It suffices to find a set of $m$ pairwise-independent geometric solutions of $(\mu_1,\beta_1)$. Let $K'/K$ be the finite separable extension given in Lemma \[lem:wre\_prod\]. Choose a ‘large’ finite Galois extension $L/K$ such that $K'\subseteq L$ and let $\mu\colon {\textnormal{Gal}}(K) \to G:={\textnormal{Gal}}(L/K)$ be the restriction map. In the notation of Lemma \[lem:wre\_prod\] the embedding problem $$(\mu\colon {\textnormal{Gal}}(K) \to G, \alpha\colon A{\wre}_{G_0} G \to G)$$ is rational, and hence has a set of $m$ independent geometric solutions $\{\theta_i\mid i\in I\}$ of $(\mu,\alpha)$. Cases  and : Note that is a special case of . In [@Bary-SorokerHaranHarbater Section 5.1.2] it is proved that there exists an open normal subgroup $\Lambda$ of ${\textnormal{Gal}}(K)$ (in the notation of loc. cit. $L=\Lambda$) such that if ${\textnormal{Gal}}(L) \leq \Lambda$, then $S=\{\rho\pi \theta_i \mid i\in I\}$ is a set of pairwise-independent solutions of $(\mu_1,\beta_1)$. Choose $L$ sufficiently large, so that it contains both $K'$ and the fixed field of $\Lambda$. Then on the one hand $S$ is a set of pairwise-independent solutions and on the other hand the solutions in $S$ are geometric (Lemma \[lem:wre\_prod\]). Case : As in the previous case, in [@Bary-SorokerHaranHarbater Section 5.1.3] it is proved, basing on the proof of [@FriedJarden2005 Theorem 25.4.3], that there exists a normal subgroup $\Lambda$ of ${\textnormal{Gal}}(K)$ such that if ${\textnormal{Gal}}(L)\leq \Lambda$, then $S$ (as above) is a set of pairwise-independent solutions of $(\mu_1,\beta_1)$. Hence the same argument as above finishes the proof. The proof of all the other cases other follows from the above cases, as shown in [@Bary-SorokerHaranHarbater] or proved in a similar manner. In the proof above, the only new property we exploited is that the constructions of fiber products and twisted wreath product, and the induced solutions, are field theoretic. Purely inseparable extensions {#sec:p.i.} ============================= In the previous section we dealt with separable extensions of a fully Hilbertian field, using the twinning principle. Here we show that full Hilbertianity, as Hilbertianity, is preserved under purely inseparable extensions. This will be useful later, for example in order to prove that full Hilbertianity is preserved under all finite extensions. \[prop:p.i-ful-hil\] Let $K$ be a fully Hilbertian field and $E/K$ a purely inseparable extension. Then $E$ is fully Hilbertian. Let $(\nu\colon {\textnormal{Gal}}(E) \to {\textnormal{Gal}}(M/E), \alpha \colon {\textnormal{Gal}}(P/E(x)) \to {\textnormal{Gal}}(M/E))$ be a rational embedding problem for $E$. Since $E/K$ is purely inseparable and $F/E(x)$, $M/E$ are Galois, the extensions $F/K(x)$, $M/K$ are normal. Let $F$ (resp. $L$) be the maximal separable extension of $K(x)$ (resp. $K$) in $P$ (resp. $M$). Then $P=FE$ and $M=LE$. $$\xymatrix{ E(x)\ar@{-}[r] &M(x)\ar@{-}[r] &P\\ K(x)\ar@{-}[r]\ar@{-}[u] &L(x)\ar@{-}[r]\ar@{-}[u] &F\ar@{-}[u]. }$$ Then $P/F$ and $M/L$ are purely inseparable and ${\textnormal{Gal}}(F/K(x)) \cong {\textnormal{Gal}}(P/E(x))$, ${\textnormal{Gal}}(L/K) \cong {\textnormal{Gal}}(M/E)$ via the restriction maps. Note that $P=FE$ is regular over $E$, i.e., linearly disjoint of the algebraic closure ${{\tilde E}}$ of $E$ over $E$. As $F/K$ is separable and $E/K$ purely inseparable, $F$ and $E$ are linearly disjoint over $K$. Thus $F$ and ${{\tilde E}}( = {{\tilde K}})$ are linearly disjoint. Thus $F$ is regular over $K$. Let $\mu\colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(L/K)$ and $\beta\colon {\textnormal{Gal}}(F/K(x)) \to {\textnormal{Gal}}(L/K)$ be the restriction maps. As $K$ is fully Hilbertian, there exists a set $\{{\varphi}^*_i \mid i\in I\}$ of independent solutions of $(\mu,\beta)$ with $|I|=|K|$. Extend $E$-linearly each ${\varphi}_i$ to a place $\psi_i$ of $P$. Since $\mu = \nu \res_{E_s,K_s}$ and $\beta = \alpha \res_{P,F}$, and since the restriction maps $\res_{E_s,K_s} \colon {\textnormal{Gal}}(E) \to {\textnormal{Gal}}(K)$, $\res_{P,F}\colon {\textnormal{Gal}}(P/E(x)) \to {\textnormal{Gal}}(F/K(x))$ are isomorphisms, we get that $\{\psi^*_i\mid i\in I\}$ is a set of independent solutions of $(\nu,\alpha)$ of cardinality $|I|=|K|=|E|$. \[cor:finite-ext\] A finite extension of a fully Hilbertian field is fully Hilbertian. Let $L/K$ be a finite extension. Let $E$ the maximal purely inseparable extension of $K$ inside $L$. Then $L/E$ is separable. By Proposition \[prop:p.i-ful-hil\] $E$ is fully Hilbertian and by Theorem \[thm:sep-ext\] $L$ is fully Hilbertian. Dominating Embedding Problems ============================= This section provides some auxiliary results that reduce the necessary embedding problems needed for full Hilbertianity. This reduction is needed in the sequel in order to prove Theorem \[IT:com-krull-dom\]. Hurwitz’s Formula ----------------- Let $F/E$ be a finite separable extension of function fields over a base field $L$. Denote by $g_F,g_E$ the genera of $F,E$, respectively, and let $n=[F:E]$. Hurwitz’s formula asserts that $$\label{hurwitz} 2g_F - 2 = n (2g_E - 2) + \deg R,$$ where $R$ is the ramification divisor. If the characteristic of $L$ is $p>0$, then the primes of $F$ divide into two sets: the tamely ramified primes $P_{tr}=\{ {{\mathfrak P}}\mid p\nmid e_{{\mathfrak P}}\}$ and the wildly ramified primes $P_{wr} = \{ {{\mathfrak P}}\mid p \mid e_{{\mathfrak P}}\}$. If the characteristic of $L$ is $0$, then all the primes are, by definition, tamely ramified. We have $$\label{degR} \deg R = \sum_{{{\mathfrak P}}\in P_{tr}} (e_{{{\mathfrak P}}}-1) + \sum_{{{\mathfrak P}}\in P_{wr}} l_{{\mathfrak P}},$$ and $l_{{\mathfrak P}}> e_{{\mathfrak P}}-1$, see [@Hartshorne1977 Corollary IV.2.4]. If ${{\mathfrak p}}$ is a prime of $L(x)/L$, then either ${{\mathfrak p}}$ corresponds to some irreducible polynomial in $L[x]$, in this case we say that ${{\mathfrak p}}$ is finite (w.r.t. $x$), or ${{\mathfrak p}}$ corresponds to $\infty$, and then we write ${{\mathfrak p}}= {{\mathfrak p}}_\infty$. The following result is an immediate consequence of Hurwitz’s formula. \[cor:non-void\_ram\] Let $F/L(x)$ be a separable extension of degree $n\geq 2$. Assume that all primes of $F$ lying above ${{\mathfrak p}}_\infty$ are tamely ramified in $F$. Then there exists a prime ${{\mathfrak P}}$ of $F$ lying over a finite prime ${{\mathfrak p}}$ of $L(x)$ such that ${{\mathfrak P}}/{{\mathfrak p}}$ is ramified. By we have $$2g_F - 2 = -2n +\deg R \Rightarrow \deg R = 2g_F + 2(n-1) \geq 2(n-1) >n-1.$$ (Recall that $n\geq 2$.) Using the formula $\sum_{{{\mathfrak P}}/{{\mathfrak p}}} e_{{\mathfrak P}}f_{{\mathfrak P}}= n$ with ${{\mathfrak p}}={{\mathfrak p}}_\infty$ we get $$\sum_{{{\mathfrak P}}/{{\mathfrak p}}_{\infty}} (e_{{\mathfrak P}}- 1) \leq n-1 < \deg R.$$ Thus there must be a prime ${{\mathfrak P}}$ of $F$ not lying above ${{\mathfrak p}}_\infty$ that is ramified. \[lem:gen-by-ine\] Let $L$ be a field and let $F/L(x)$ be a finite Galois extension that is not wildly ramified at infinity. Suppose $F/L(x)$ is regular, and that ${{\rm Ram}}(F/L(x)) = \{c_1,c_2,\ldots,c_k\} \subseteq L$. For each $1 \leq i \leq k$ let ${{\mathfrak P}}_{i,1},\ldots,{{\mathfrak P}}_{i,g_i}$ be the primes of $F$ lying over $(x-c_i)$. Denote the inertia group of $F/L(x)$ at ${{\mathfrak P}}_{i,j}$ by $I_{i,j}$. Then $G$ is generated by all the $I_{i,j}$. Let $H\leq G$ be the subgroup generated by all $I_{j,j}$. Let $E = F^H$ the corresponding field. Then, by definition, $E$ is ramified at most at infinity, and this ramification, if exists, is tame. Thus by Corollary \[cor:non-void\_ram\], $E=L(t)$, i.e., $H=G$. Before we continue we recall some notation. Fix a field $L$. Let $a\in {{\tilde L}}$, $f(X)$ its irreducible polynomial over $L$, and ${{\mathfrak p}}$ the prime of $L(x)$ that is defined by $f$. We say that $a$ is a **ramification point** of $F/L(x)$ if ${{\mathfrak p}}$ is ramified in $F$. We denote by ${{\rm Ram}}(F/L(x))\subseteq {{\tilde L}}$ the set of all ramification points of $F/L(x)$ in ${{\tilde L}}$. We emphasize that ${{\rm Ram}}(F/L(x))$ depends on $x$. Let ${{\mathfrak P}}/{{\mathfrak p}}$ be an extension of primes of $F/L(x)$ with ramification index $e$. If the characteristic of $L$ is positive, say $p>0$, we write $e=e' q$, where $\gcd(e',p)=1$ and $q$ is a power of $p$. We will call $e'$ the **tame ramification index** of ${{\mathfrak P}}/{{\mathfrak p}}$ and $q$ the **wild ramification index**. If $L$ is of characteristic $0$, then the tame ramification index equals the ramification index, and the wild ramification index is $1$. If $F/L(x)$ is Galois, then the tame and wild ramification indices are independent of the choice of the prime ${{\mathfrak P}}$ that lies over ${{\mathfrak p}}$, so we abbreviate and say tame (resp., wild) ramification index of ${{\mathfrak p}}$ in $F$. \[lem:con\_F\] Let $K$ be an infinite field, let $F_0/K(x_0)$ be a finite Galois extension, and let $L_0 ={{\tilde K}}\cap F_0$. Assume that ${{\rm Ram}}(F_0/L_0(x_0)) \subseteq K_s$. Then there exist $x \in K(x_0)$ and a finite Galois extension $F/K(x)$ such that, for $L = K_s\cap F$, the following conditions hold: 1. $K(x) = K(x_0)$. 2. The prime ${{\mathfrak p}}_\infty$ of $L(x)$ is tamely ramified in $F$. 3. $F_0\subseteq F$, 4. ${{\rm Ram}}(F/L(x))\subseteq L$. 5. Let $c\in {{\rm Ram}}(F/L(x))$, ${{\mathfrak P}}_c$ a prime of $F$ lying above ${{\mathfrak p}}_c = (x-c)$, and $e'$ (resp., $q$) be the tame (resp., wild) ramification index of ${{\mathfrak P}}_c/{{\mathfrak p}}_c$. Then 1. $(x-c)^{1/e'}\in F$, 2. the primitive $e'$-th root of unity $\zeta_{e'}$ is contained in $L$, and 3. the residue field of $F$ at ${{\mathfrak P}}_c$ is a purely inseparable extension of $L$. Since $K$ is infinite, there is a Möbieus transformation that sends $x_0$ to $x=\frac{u + v x_0}{w + z x_0}$, for some $\left(\begin{matrix}u & v \\w & z\end{matrix}\right) \in PSL_2(K)$ under which $\infty$ is mapped to an unramified point. Clearly $K(x) = K(x_0)$ and ${{\rm Ram}}(F_0/L_0(x)) \subseteq K_s$. Let $n = |{{\rm Ram}}(F_0/L_0(x))|$, write ${{\rm Ram}}(F_0/L_0(x)) =\{c_1,\ldots, c_n\}$, and for each $i$, let $e_i'$, (resp., $q_i$) be the tame (resp., wild) ramification index of $(x-c_i)$ in $F_0$. Then $e_i = e_i' q$, $i=1,\ldots, n$ are the corresponding ramification indices. Let $L$ be a finite Galois extension of $K$ that contains $L_0$, $c_1, \ldots, c_n$. Write $F_1=F_0 L$. Then ${{\rm Ram}}(F_1/L(x)) = \{c_1,\ldots, c_n\}$, and each $c_i$ has the same ramification indices $e_1,\ldots, e_n$ (since $L(x)/L_0(x)$ is unramified). Let $F= F_1((x-c_i)^{1/e'_i}\mid i=1,\ldots, n)$. By [@FriedJarden2005 Example 2.3.8] it follows that $F/F_1$ is unramified over the primes of $F_1$ that lie above $(x-c_i)$, $i=1,\ldots, n$. By the multiplicity of ramification indices, we get that each $(x-c_i)$ has ramification $e_i$ in $F$. On the other hand, let ${{\mathfrak p}}\neq (x-c_i)$, $i=1,\ldots,n$ be a finite prime of $L(x)$. Then ${{\mathfrak p}}$ is unramified both in $F_1/L(x)$ and in $K((x-c_j)^{1/e'_i}\mid i=1,\ldots, n)$. Thus it is unramified in their compositum $F$. Replace $L$ with a larger extension, if necessary, to assume that the residue field of $F$ w.r.t. ${{\mathfrak P}}_{c_i}$ is a purely inseparable extension of $L$. Finally note that the ramification of ${{\mathfrak p}}_\infty$ in $F$ divides $e_1' \cdots e_n'$, and hence is co-prime to the characteristic of $K$, if not $0$. Let $K$ be an infinite field and $$(\mu_0 \colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(L_0/K(x_0)), \alpha_0 \colon {\textnormal{Gal}}(F_0/K(x_0) )\to {\textnormal{Gal}}(L_0/K(x_0))$$ a non-trivial rational finite embedding problem. Then there exist $x\in K(x_0)$ and a non-trivial rational finite embedding problem $$(\mu \colon {\textnormal{Gal}}(K) \to {\textnormal{Gal}}(L/K(x)), \alpha \colon {\textnormal{Gal}}(F/K(x)) \to {\textnormal{Gal}}(L/K(x))$$ satisfying Properties (a)-(e) of Lemma \[lem:con\_F\] such that an independent set $\{\theta_i \mid i \in I\}$ of proper geometric solutions of $(\mu,\alpha)$ induces an independent set $\{\res_{F,F_0} \theta_i\mid i\in I\}$ of proper geometric solutions of $(\mu_0,\alpha_0)$. Let $F$ be the field given by Lemma \[lem:con\_F\]. Consider the commutative diagram $$\xymatrix{ &&{{\textnormal{Gal}}(K)}\ar[d]^{\mu}\ar@/^30pt/[dd]^{\mu_0}\\ {\textnormal{Gal}}(F/K(x))\ar[r]_-{\gamma}\ar@/^10pt/@{.>}[rr]^{\alpha} &{\textnormal{Gal}}(F_0L/K(x))\ar[r]_-{\beta}\ar[d]^\pi &{\textnormal{Gal}}(L/K)\ar[d]^{\mu'}\\ &{\textnormal{Gal}}(F_0/K(x))\ar[r]^-{\alpha_0} &{\textnormal{Gal}}(L_0/K) }$$ in which all the maps are the restriction maps. We are given an independent set $\{\theta_i\mid i\in I\}$ of geometric proper solutions of $(\mu, \alpha)$. By Lemma \[lem:basic-rat-eps\_2\] $\{\gamma\theta_i\mid i\in I\}$ is an independent set of geometric proper solutions of $(\mu,\beta)$. Then Lemma \[lem:basic-rat-eps\_2\] implies that $\{\pi\gamma\theta_i\mid i\in I\}$ is an independent set of geometric proper solutions of $(\mu_0,\alpha_0)$. The following is an immediate conclusion. \[cor:suf\] For a field $K$ to be fully Hilbertian it suffices that every non-trivial rational embedding problem $$(\mu \colon {\textnormal{Gal}}(K)\to {\textnormal{Gal}}(L/K), \alpha\colon {\textnormal{Gal}}(F/K(x))\to {\textnormal{Gal}}(L/K))$$ satisfying Properties (a)-(e) of Lemma \[lem:con\_F\] has $|K|$ independent geometric solutions. Wild ramification ----------------- This part deals with wild ramifications. The reader interested only with fields of characteristic zero can skip this part. For the rest of this section fix a field $E$ with positive characteristic $p>0$ which is equipped with a discrete valuation $v$. \[Artin-Schreier\] Let $F/E$ be a Galois extension of degree $p$. There exists $a_0\in E$ such that for every $a\in a_0 + E^p - E$ and every $x$ satisfying $x^p-x=a$ we have $F = E(x)$. Let $\wp(x) = x^p - x$. Then the short exact sequence $$\xymatrix@1{ 0\ar[r]& {\mathbb{F}}_p \ar[r] & E_s \ar[r]^-{\wp} & E_s\ar[r] &0 }$$ of additive ${\textnormal{Gal}}(E)$-modules gives rise to the following exact sequence of cohomology $$\xymatrix@1{ E \ar[r]^-{\wp} & E\ar[r] & H^{1} ({\mathbb{F}}_q) \ar[r] & H^{1} (E_s). }$$ By Hilbert’s 90th, $H^1(E_s) = 0$. Since ${\textnormal{Gal}}(E)$ acts trivially on ${\mathbb{F}}_p \cong {\mathbb{Z}}/p{\mathbb{Z}}$ we have $H^1({\mathbb{F}}_q) = \Hom({\textnormal{Gal}}(E) , {\mathbb{Z}}/p {\mathbb{Z}})$. Therefore $\Hom ({\textnormal{Gal}}(E), {\mathbb{Z}}/p{\mathbb{Z}}) \cong E/\wp(E)$. Applying this to the restriction map ${\textnormal{Gal}}(E) \to {\textnormal{Gal}}(F/E)$ proves the assertion. We give a criterion for $v$ to be totally ramified in a Galois $p$ extension. \[prop:crit-tot-ram\] Let $F/E$ be a Galois extension of degree $p$. Then $v$ is totally ramified in $F$ if and only if there exists $a\in E$ and $x\in F$ such that $F=E(x)$, $x^p-x=a$, and $p\nmid v(a)<0$. First assume there exist $x\in F$ and $a\in E$ as in the proposition. Let $w$ be a normalized discrete valuation of $F$ lying above $v$. Then $w(x) < 0$, since otherwise $v(a) \geq 0$. We thus get that $w(x^p) < w(x)$, hence $$e(w/v) v(a) = w(a) = w(x^p -x) = w(x^p) = p w(x).$$ Since $\gcd(v(a),p)=1$, we get that $p$ divides $e(w/v)$, hence $p=e(w/v)$. Thus $v$ is totally ramified in each $F$. Conversely, assume that $v$ is totally ramified in $F$ and let $w$ be its normalized extension. Fix $1 \leq i \leq m$. We show how to choose suitable $a,x$. Choose a uniformizer $\pi \in E$ for $v$, i.e., $v(\pi)=1$. Then $w(\pi)=p$. Let $K$ be the residue field of $E$. Since $w/v$ is totally ramified, $K$ is the residue field of $F$ as well. Let ${\varphi}$ be the corresponding place of $F$. Apply Lemma \[Artin-Schreier\] to get an element $a_0 \in E$ such that for every $a \in a_0 + E^p - E$ and $x \in F$ such that $x^p - x = a$ we have $F = E(x)$. Denote $A = a_0 + E^p - E$. By [@FriedJarden2005 Example 2.3.9] $v(A)$ is a set of negative integer numbers. Let $l\in {\mathbb{N}}$ be minimal such that $-lp\in v(A)$. Choose some $a\in A$ satisfying $v(a) = -lp$. It suffices to show that $-lp\neq \max v(A)$. Let $x\in F$ be a root of $X^p-X=a$. Then $F = E(x)$. We have $w(x^p - x) = w(a) =pv(a) < 0$. Hence $w(x)<0$, so $w(x^p)<w(x)$ which implies $pw(x) = w(x^p) = w(x^p-x) = p v(a)$. Therefore $w(x) = v(a) = -lp$. Write $a = \pi^{-lp} \alpha$ and $x = \pi^{-l} \xi$. Note that $v(\alpha) = v(a) + lp=0$ and $w(\xi) = w(x) + l w(\pi) = -lp+lp=0$. Thus $\alpha,\xi$ are $w$-invertible, or equivalently, $\bar \alpha, \bar\xi \neq 0,\infty$, where the notation $\bar \cdot$ indicates the image after applying ${\varphi}$. We apply ${\varphi}$ to the equation $$\xi^p - x \pi^{lp} = x^p \pi^{lp} - x \pi^{lp} = a \pi^{lp} = \alpha$$ to get the equality $\bar \xi^p = \bar \alpha$. Since $\Egag = K$, there exists $b\in E$ with $\bgag = \bar \xi$. In particular $v(b) = 0$. Also, as $\bgag^p - \bar \alpha =0$, we have $v(b^p-\alpha)>0$. Let $a' = a - \pi^{-lp} b^p + \pi^{-l} b \in A$. Then $d:=v(a-\pi^{-lp} b^p) = v(-\pi^{-lp}(\alpha-b^p)) =-lp + v(\alpha - b^p) > -lp$. Thus $v(a') \geq \min\{d, -m\} > -lp = v(a)$. This implies $v(a)\neq \max v(A)$, as needed. Now we are ready to state and prove the main result of this part which sets a necessary condition for a totally ramified Galois extension of order a power of $p$ to remain totally ramified when moving to the residue extension w.r.t. some other place. \[cor:wildly-ram\] Let $N/E$ be a Galois extension of degree a power of $p$. Let $v$ be a discrete valuation of $E$ that is totally ramified in $N$. Let ${\varphi}$ be a place of $N$, let $\bar N,\bar E$ be the respective residue fields of $N,E$ under ${\varphi}$, and assume the characteristic of $\bar E$ is $p$ and $\bar N/\bar E$ is a Galois extension of degree $[\bar N:\bar E] = [N:E]$. Then there exists $a_1,\ldots,a_m \in E$ with $v(a_i)<0$ and $\gcd(v(a_i),p)=1$ for each $1 \leq i \leq m$, such that for every discrete valuation $\vgag$ of $\Egag$, if $\agag_i:={\varphi}(a_i)\neq \infty$, $\vgag(\agag_i)< 0$ and $\gcd(\vgag(\agag_i),p)=1$ for each $1 \leq i \leq m$, then $\vgag$ is totally ramified in $\Ngag$. Let $F_1, \ldots, F_m$ be all the minimal extension of $E$ inside $N$. Then $F_i/E$ is a Galois extension of degree $p$. Proposition \[prop:crit-tot-ram\] gives $a_1,\ldots,a_m\in E, x_1\in F_1,\ldots,x_m \in F_m$ such that $F_i = E(x_i)$, $x_i^p - x_i = a_i$, $v(a_i)<0$ and $\gcd(v(a_i),p)=1$, for each $1 \leq i \leq m$. Let $\Fgag$ be the residue field of $F$ under ${\varphi}$. By assumption $\Ngag/\Egag$ is a Galois extension of the same degree as $N/E$, hence ${\textnormal{Gal}}(\Ngag/\Egag) \cong {\textnormal{Gal}}(N/E) = G$. Thus ${\textnormal{Gal}}(\Ngag/\Fgag) = \Phi$. Let $H$ be the inertia group of $\vgag$ (w.r.t. some extension of it to $\Ngag$) in $G$, and let $\Egag'$ be its fixed field. Then, by definition $\vgag$ is unramified in $\Egag'$. If $\Egag'\neq \Egag$, then $\Egag'$ contains some $\Fgag_i$. Since $\agag_i$ is finite, so is $\xgag_i$, $\Fgag_i=\Egag_i(\xgag_i)$, and $\xgag_i^p-\xgag_i =\agag_i$. Proposition \[prop:crit-tot-ram\] implies that $\vgag$ is totally ramified in $\Fgag_i$, hence ramified in $\Egag'$. This contradiction implies that $\Egag'=\Egag$, i.e., $\vgag$ is totally ramified in $\Ngag$. Complete domains {#sec:gett-many-dist} ================ We now prove Theorems \[IT:com-krull-dom\] and \[IT:com-krull-dom-hil\] about quotient fields of complete local domains. To get independent solutions of a rational embedding problem we show that if $a$ in a Hilbert set approximates all the ‘ramification points’ of the embedding problem w.r.t. some valuations of our base field $K$, then the geometric inertia groups become inertia groups w.r.t. those valuations of $K$. \[lem:unram-prime-val\] Let $L$ be a field, let $E/L(t)$ be a finite separable extension, and let $\beta_1,\ldots,\beta_l \in E$. Let ${{\mathfrak P}}/(t-c)$ be a prime extension of $E/L(t)$ such that $e({{\mathfrak P}}/(t-c)) =f({{\mathfrak P}}/(t-c)) =1$. Then there exists a finite set $\Lambda \subseteq L$, such that for any discrete valuation $v$ of $L$, satisfying $v(\lambda) = 0$ for each $\lambda \in \Lambda$, there exists $M \in {\mathbb{R}}$ such that for every $\alp \in L$ satisfying $v(\alp - c) > M$ the following properties hold: 1. \[cond:unramified\] There exists a prime ${{\mathfrak P}}_\alp$ of $E$ lying over $(t - \alp)$ such that $v$ is unramified in the residue field $\bar{E}_\alpha$. 2. \[cond:rel-prime\] Let $v_s$ be an extension of $v$ to $L_s$. Let $m_i$ be the order of $\beta_i$ at ${{\mathfrak P}}$ and let $\nu\in {\mathbb{Z}}$ be co-prime to $m_i$, for each $1 \leq i \leq l$. Suppose $m_i < 0$ for each $1 \leq i \leq l$. Then if $v(\alpha - c)$ is co-prime to $\nu$, then $v_s(\bar \beta_i) \in {\mathbb{Z}}$, $v_s(\bar \beta_i) < 0$ and $\gcd(v_s(\bar \beta_i), \nu) = 1$, for each $1 \leq i \leq l$. (Here $\betagag_i$ is the image of $\beta_i$ modulo ${{\mathfrak P}}_\alpha$.) Let $\hat{E} =\Ehat_{{\mathfrak P}}$ be the completion of $E$ at ${{\mathfrak P}}$, and consider the completion $L((t-c))$ of $L(t-c)$ w.r.t. $(t - c)$. Then $\hat{E}/L((t-c))$ is unramified and of residue field degree $1$. Hence $\hat{E} = L((t-c))$. Let $\beta_0 \in E$ be a primitive element of $E/L(t)$. Then $\beta_0, \beta_1, \ldots, \beta_l \in L((t-c))$, thus we have $$\beta_i = \sum_{n=m_i}^{\infty} \lambda_{i,n} (t-c)^n, \qquad m_i\in {\mathbb{Z}}, \lambda_{i,n}\in L,$$ for each $0 \leq i \leq l$. Without loss of generality we can assume that $m_0 > 0$ (by multiplying $\beta_0$ with a sufficiently large power of $(t-c)$). Put $\Lambda = \{ \lambda_{1,m_1},\ldots,\lambda_{l,m_l}\}$ Suppose $v$ is a discrete valuation of $L$, satisfying $v(\lambda)= 0$ for each $\lambda \in \Lambda$, and $v_s$ an extension of $v$ to $L_s$. Let $\hat{L}$ be the completion of $L$ at $v$, and extend $v$ to the algebraic closure of $\hat{L}$ compatibly with $v_s$. Since $\beta_i \in L((t-c)) \subseteq \hat{L}(((t-c)))$ is algebraic over $L(t-c) \subseteq \hat{L}(t-c)$ for each $0 \leq i \leq l$, we get by [@Artin1968 Theorem 2.14] that $\beta_i$ converges at some $b_i \in \hat{L}$. That is, $v(\lambda_{i,n} b_i^n) = v(\lambda_{i,n}) + n v(b_i)\to \infty$. Choose $b\in L$ with sufficiently large value $M := v(b)$. That is, we assume that $M \geq \max\{v(b_0),v(b_1),\ldots,v(b_l)\}$, and hence $\beta_0,\beta_1,\ldots,\beta_l$ converge at $b$, and for every $r>M, 1 \leq i \leq l$ and $n>m_i$ the following inequality holds. $$\label{eq:r-suff-large} v(\lambda_{i,m_i}) + m_i r < v(\lambda_{i,n}) + nr.$$ Denote $x = \frac{t - c}{b}$. Let $R = \hat{L}\{x\}$ be the ring of convergent power series in $x$ over $\hat{L}$. That is, all power series in $\hat{L}[[x]]$ whose series of coefficients converges to $0$ with respect to $v$. In particular, $\beta_i = \sum_{n = 1}^\infty {\lambda_{i,n} b^n} x^n \in R$, for all $i=0,1,\ldots,l$. Now, suppose $\alp \in L$ satisfies $v(\alp - c) > M$. Then the element ${\varepsilon}= {\alp - c \over b} \in \hat{L}$ satisfies $v({\varepsilon}) > 0$. Since $\hat{L}$ is complete with respect to $v$, we have a substitution homomorphism ${\varphi}_{\alp} \colon R \to \hat{L}$, given by ${\varphi}_{\alp} (\sum \mu_n x^n) = \sum \mu_n {\varepsilon}^n$. Extend ${\varphi}_{\alp}$ to a place of $Q = {{\rm Quot}}(R)$. We also have ${\varphi}_\alpha(\beta_i) = \sum_{n=m_i}^{\infty} \lambda_{i,n} (\alpha-c)^n$. Applying to $r = v(\alpha-c)$ we get that $v(\lambda_{i,m_i} (\alpha-c)^{m_i})< v(\lambda_{i,n} (\alpha-c)^n)$, for every $n>m_i$. Therefore, if we further assume that $v(\alp - c)$ is co-prime to $\nu$, then, for all $i\geq 1$, $$v({\varphi}_\alpha(\beta_i)) = v(\sum_{n=m_i}^{\infty} \lambda_{i,n} (\alpha-c)^n) = v(\lambda_{i,m_i} (\alpha-c)^{m_i}) = m_i v(\alpha - c)$$ is a negative integer (since $m_i$ is negative), co-prime to $\nu$. This concludes the proof of . Let us continue and prove . Since $\beta_0 \in R$, the field $E$ is contained in $Q$. The restriction of ${\varphi}_{\alp}$ to $E$ fixes $L$ and maps $t$ to $\alp$ (by the definition of ${\varphi}_{\alp}$). Let ${{\mathfrak P}}_\alpha$ be the corresponding prime of $E$. Then ${{\mathfrak P}}_\alpha$ lies over the prime $(t-\alp)$ of $L(t)$. By the Weierstrass division theorem for convergent power series, each element of $R$ can be written in the form $g \cdot (x - {\varepsilon}) + r$, with $g \in R, r \in \hat{L}$. It follows that the kernel of ${\varphi}_{\alp}$ (as a homomorphism of $R$) is a principal ideal, generated by $x - {\varepsilon}$. Moreover, by the Weierstrass preparation theorem for convergent power series, $R$ is a unique factorization domain (whose primes are the primes of $\hat{L}[x]$). Hence each element $h \in Q$ can be written in the form $(x - {\varepsilon})^n {f \over g}$, with $f,g \in R$ not divisible by $x - {\varepsilon}$, and $n \in {\mathbb{Z}}$. Then ${\varphi}_{\alp}( {f \over g}) = {{\varphi}_{\alp}(f) \over {\varphi}_{\alp}(g)} \in \hat{L}^\times$. Thus, if $n \geq 0$, then ${\varphi}_{\alp}(h) \in \hat{L}$, and if $n < 0$, then ${\varphi}_{\alp}(h) = \infty$. So the residue field of $Q$ at ${\varphi}_{\alp}$ is $\hat{L}$. Thus $\bar{E}_\alpha$ is contained in $\hat{L}$, hence $\bar{E}_\alpha/L$ is unramified at $v$. \[lem:exact-cond\] Let $K$ be a field, let $F/K(x)$ be a finite Galois extension with $L=F\cap K_s$, let $c\in {{\rm Ram}}(F/K(x)) \cap L$, let ${{\mathfrak P}}$ a prime of $F$ lying above ${{\mathfrak p}}=(x-c)$, let $I$ be the corresponding inertia group, let $e$ be the ramification index, and let $E = F^I$ the inertia field. Assume conditions (e1)-(e2) of Lemma \[lem:con\_F\] are satisfied and that the residue field of $F$ at ${{\mathfrak P}}$ is $L$, in addition to (e3). Then there exists a finite set $\Lambda \subseteq L$, such that if $v_s$ is a valuation of $K_s$ that restricts to a discrete normalized valuation $v$ of $K$, and satisfies $v(\lambda) = 0$ for each $\lambda \in \Lambda$, then there exists $M\in {\mathbb{R}}$ such that if $\alpha\in K$ satisfies 1. $v(\alpha-c)> M$, 2. \[cond:val-alp-c=rho\] $v(\alpha-c)$ is co-prime to $e$, 3. $(x-\alpha)$ is totally inert in $F$, then ${\textnormal{Gal}}(\Fgag/ \Egag)$ is an inertia group of $v$ in $\Fgag$, where $\Fgag, \Egag$ are the respective residue fields of $F, E$ at the unique extension ${{\mathfrak P}}_\alpha$ of $(x-\alpha)$ to $F$. For any separable extension $N$ of $K$ we write $v_{N}$ for the normalized valuation that lies above $v$ and below $v_s$. Similarly for every subfield $F'$ of $F$ containing $L(x)$, we let ${{\mathfrak P}}_{F'}$ be the restriction of ${{\mathfrak P}}$ to $F'$. We recall that $e = e' q$, where $e'$ is the tame ramification index and $q$ is the wild ramification index. If the characteristic of $K$ is $0$, then $e'=e$ and $q=1$, and if the characteristic is $p>0$, then $\gcd(e',p)=1$ and $q=p^r$, for some $r\geq 0$. <span style="font-variant:small-caps;">Construction in the wild ramification case:</span> The following construction is needed only in the the case where $q>1$. Since $I$ is an inertia group, it has a unique $p$-sylow subgroup $J$. Then $J\normal I$, $|J| = q$, $(I:J) = e'$. Let $E'$ be the fixed field of $J$ in $F$. Then $[F:E']=q$, $[E':E]=e'$. Furthermore both ${{\mathfrak P}}_{E'}/{{\mathfrak P}}_{E}$ and ${{\mathfrak P}}_{F}/{{\mathfrak P}}_{E'}$ are totally ramified, while the former ramification is tame and the latter is wild. Let $a_1,\ldots,a_l \in E'$ be the elements that Corollary \[cor:wildly-ram\] gives when applying it to $F/E'$ (instead of $N/E$ in the notation of the corollary) with the normalized valuation $v_{{{\mathfrak P}}_{E'}}$ (instead of $v$) arising from the prime ${{\mathfrak P}}_{E'}$ and with the place ${\varphi}$ arising from the prime ${{\mathfrak P}}_\alpha$. Let $m_i = v_{{{\mathfrak P}}_{E'}} (a_i)$ for each $1 \leq i \leq l$. Then $m_i<0$ is co-prime to $p$, for each $1 \leq i \leq l$. Let $\Lambda$ be the finite set given by Lemma \[lem:unram-prime-val\] (if $q = 1$, $\Lambda$ may be chosen to be empty). Suppose $v$ satisfies the conditions of the proposition (we fix such $v$ now, even if $q = 1$ and the construction we are making in this part is redundant), and let $M $ be the bound given by Lemma \[lem:unram-prime-val\]. Note that $\agag_i := {\varphi}(a_i)$ is finite for all $\alpha$ but a finite set, hence, without loss of generality, we can assume it is finite (by enlarging $M$). Then if $$\label{eq:v-Egag'-cond} p\nmid v_{\Egag'}(\agag) < 0,$$ then $v_{\Egag'}$ is totally ramified in $F$. We now prove that holds. Let $E_0 = L((x-c)^{1/e'})$. We have $$e' q =e = e({{\mathfrak P}}_{F}/{{\mathfrak P}}_{E_0}) e({{\mathfrak P}}_{E_0}/ {{\mathfrak p}}) = e({{\mathfrak P}}_{F}/{{\mathfrak P}}_{E_0}) e',$$ hence $e({{\mathfrak P}}_{F}/{{\mathfrak P}}_{E_0})=q$. $$\xymatrix{ E\ar@{-}[r]^{e=e'} &E'\ar@{-}[r]\ar@{.}@/^10pt/[rr]^{e=q} &E'E_0\ar@{-}[r] &F \\ L(x)\ar@{-}[u]^{e=1}\ar@{-}[rr]_{e=e'} & &E_0\ar@{-}[u]\ar@{.}[ur]_{e=q} }$$ Since $e({{\mathfrak P}}_{E' E_0}/{{\mathfrak P}}_{E_0})$ divides $e({{\mathfrak P}}_{E'}/{{\mathfrak P}}_{E})=e'$, we get that ${{\mathfrak P}}_{E' E_0}/{{\mathfrak P}}_{E_0}$ is unramified. Similarly ${{\mathfrak P}}_{E' E_0}/{{\mathfrak P}}_{E'}$ is unramified. Hence $m_i$ is also the order of $a_i$ at ${{\mathfrak P}}_{E' E_0}$ for each $1 \leq i \leq l$. Since $v(\alpha - c)$ is co-prime to $e$, it is co-prime to $p$, hence so is $v((\alpha - c)^{1 \over e'})$. Lemma \[lem:unram-prime-val\], applied to $\beta_i=a_i, t=(x-c)^{1/e'}, \nu = p$ and with $c$ there replaced by $0$ here, we get that $v_{\Egag' \Egag_0}(\agag_i)<0$ is co-prime to $p$ for each $1 \leq i \leq l$. Since $[\Egag' \Egag_0 : \Egag']$ divides $e'$ which is relatively prime to $p$, follows. For conclusion, Corollary \[cor:wildly-ram\] implies that $v_{\Egag'}$ is totally ramified in $\Fgag$. <span style="font-variant:small-caps;">Conclusion of the proof:</span> It remains to show that $v_{\Egag}/ v$ is unramified and that $v_{\Egag'}/v_{\Egag}$ is totally ramified. The former follows from Lemma \[lem:unram-prime-val\] (applied to the prime extension ${{\mathfrak P}}_{E}/{{\mathfrak p}}$ of $E/L(x-c)$). To prove that $v_{\Egag'}/v_{\Egag}$ is totally ramified, let $\tgag = (\alpha-c)^{1/e'}\in \Fgag$. Since $v(\alpha-c)$ is co-prime to $e'$, it is also co-prime to $e'$. Since $${\mathbb{Z}}\ni v_{\Fgag}(\tgag) = \frac{1}{e'} v_{\Fgag}(\alpha - c) = \frac{e(v_{\Fgag}/v)}{e'} v(\alpha-c).$$ we get that $e' \mid e(v_{\Fgag}/v)$. Note that the condition that $(x-\alpha)$ is totally inert in $F$ means that the degrees of the residue field extensions are preserved. Since $[\Fgag:\Egag']=q$ is relatively prime to $e'$ and since $v_{\Egag}/v$ is unramified, we get that $e' \mid e(v_{\Egag'}/v_{\Egag})$. On the other hand, $e(v_{\Egag'}/v_{\Egag}) \leq [\Egag':\Egag] = e'$, so $e' = e(v_{\Egag'}/v_{\Egag})$, i.e., $v_{\Egag'}/v_{\Egag}$ is totally ramified. \[cor:vel-geo-sol\] Let $K$ be a Hilbertian field. Consider a rational finite embedding problem $(\mu\colon {\textnormal{Gal}}(K)\to {\textnormal{Gal}}(L/K), \alpha\colon {\textnormal{Gal}}(F/K(x)) \to {\textnormal{Gal}}(L/K))$ that satisfies (a)-(e) of Lemma \[lem:con\_F\]. For each $c_i\in {{\rm Ram}}(F/L(x))$ let ${{\mathfrak P}}_{i,1},\ldots,{{\mathfrak P}}_{i,g_i}$ be the primes of $F$ lying above ${{\mathfrak p}}_i = (x-c_i)$. Assume that the residue field of $F$ at ${{\mathfrak P}}_c$ is $L$. Then there exists a finite set $\Lambda \subseteq L$, such that if $\{v_{i,j} \st 1 \leq i \leq k, 1 \leq j \leq g_i\}$ is a family of non-equivalent discrete valuations of $K$, totally split in $L/K$, and trivial on $\Lambda$, then there exists a geometric solution ${{\mathfrak P}}^*$ of $(\mu,\alpha)$ with solution field $L'$, and extensions $v'_{i,j}$ of the $v_{i,j}$ to $L'$, such that ${\textnormal{Gal}}(L'/L)$ is generated by the inertia groups of the $v'_{i,j}$ at $L'/L$. Denote the inertia group of $F/L(x)$ at ${{\mathfrak P}}_{i,j}$ by $I_{i,j}$. By Lemma \[lem:gen-by-ine\], ${\textnormal{Gal}}(F/L(x))$ is generated by the $I_{i,j}$. Let $E_{i,j} = F^{I_{i,j}}$. Then ${{\mathfrak p}}_{i,j}={{\mathfrak P}}_{i,j}\cap E_{i,j}$ is unramified over $L(x)$ and of degree $1$. Let $M_{i,j}$ be the bound given by Lemma \[lem:exact-cond\] applied to ${{\mathfrak p}}_{i,j}$, $E_{i,j}$, and $v_{i,j}$. Let $n_{i,j} \in {\mathbb{Z}}$ be co-prime to $e_i$ and such that $n_{i,j} > M_{i,j}$. Since $v_{i,j}$ totally splits in $L/K$, $K$ is $v_{i,j}$-dense in $L$. Choose $c_{i,j} \in K$ such that $v_{i,j}(c_{i,j} - c_i) > n_{i,j}$. Since $K$ is Hilbertian, we may choose $\alp \in K$ such that $(x-\alpha)$ is totally inert in $F$, or equivalently, the $K$-specialization $x \mapsto \alp$ extends to a place ${{\mathfrak P}}$ of $F$ which defines a geometric solution ${{\mathfrak P}}^*$ of $(\mu,\alpha)$. Moreover, since Hilbert sets are dense with respect to any finite family of valuations [@Jarden1991 Proposition 19.8], we may assume that $v_{i,j}(\alp - c_{i,j}) = n_{i,j}$. Hence, $v_{i,j}(\alp - c_i) = v_{i,j}((\alp - c_{i,j}) + (c_{i,j} - c_i)) = n_{i,j}$. By Lemma \[lem:exact-cond\], we get that ${\textnormal{Gal}}(L'/\Egag_{i,j})\cong I_{i,j}$ is the inertia group of $v'_{i,j}/v_{i,j}$ in $L'$. Thus, since the $I_{i,j}$ generate ${\textnormal{Gal}}(F/L(x))$, we get that ${\textnormal{Gal}}(L'/L)$ is generated by the ${\textnormal{Gal}}(L'/\Egag_{i,j})$, as asserted. \[prop:fully-Hilbertian\] Let $K$ be a Hilbertian field, equipped with a family ${{{\mathcal F}}}$ of discrete valuations, satisfying: 1. \[cond:fin-many-non-zero\] For each $a \in K^\times$, $v(a) = 0$ for almost all $v \in {{{\mathcal F}}}$. 2. \[cond:many-tot-split\] For each finite separable extension $L/K$, there exist $|K|$ valuations in ${{{\mathcal F}}}$ that totally split in $L/K$. 3. \[cond:trv-prime-field\] Each $v \in {{{\mathcal F}}}$ is trivial on the prime field of $K$. Then the maximal purely inseparable extension $K_{ins}$ of $K$ is fully Hilbertian. Let $\kappa = |K|$. Then any algebraic extension of $K$ has cardinality $\kappa$. Let $$(\mu'\colon {\textnormal{Gal}}(K_{ins})\to {\textnormal{Gal}}(L'/K_{ins}) , \alpha' \colon {\textnormal{Gal}}(F'/K_{ins}(x)) \to {\textnormal{Gal}}(L'/K_{ins}))$$ be a finite rational embedding problem for $K$. We need to produce $\kappa$ pair-wise independent geometric solutions of ${{\mathcal E}}_{ins}$. Let $K_0$ be a finite purely inseparable extension of $K$. Since $K_{ins}/K_0$ is purely inseparable, as in the proof of Proposition \[prop:p.i-ful-hil\], the maximal separable extension $F$ (resp., $L$) of $K_0(x)$ (resp., $K_0$) in $F'$ (resp., $L'$) is a Galois extension, with same Galois group as ${\textnormal{Gal}}(F'/K(x))$ (resp., ${\textnormal{Gal}}(L'/K)$). Moreover, that proof shows that a set of $\kappa$ pair-wise independent geometric solutions of $$(\mu \colon {\textnormal{Gal}}(K_0) \to \colon {\textnormal{Gal}}(L/K_0), \alpha\colon {\textnormal{Gal}}(F/K_0(x)) \to {\textnormal{Gal}}(L/K_0))$$ gives a set of $\kappa$ pair-wise independent geometric solutions of $(\mu',\alpha')$. Therefore it suffices to find $\kappa$ pair-wise independent geometric solutions of $(\mu,\alpha)$. By Corollary \[cor:suf\] we may assume conditions (a)-(e) of Lemma \[lem:con\_F\] hold. Note that condition (e4) gives that the residue field of ${{\mathfrak P}}_c$, $c\in {{\rm Ram}}(F/L(x))$ is a purely inseparable extension of $L$. Thus if we replace $K_0$ with a sufficiently large finite extension we may assume that this residue field is $L$. Since $K_0/K$ is purely inseparable, each $v\in {{\mathcal F}}$ can be extended uniquely to a discrete valuation $v_0$ of $K_0$. Let ${{\mathcal F}}_0 = \{ v_0 \mid v\in {{\mathcal F}}\}$. Clearly and hold for ${{\mathcal F}}_0$. The last assumption also holds, since if $N/K$ is separable of degree $n$ and $v\in {{\mathcal F}}$ is totally split in $N$, i.e., there exist $w_1,\ldots, w_n$ lying above $v$, then the corresponding $v_0\in {{\mathcal F}}_0$ is totally split in $NK_0$, since $[NK_0:K_0]=n$, and the extensions of $w_1,\ldots, w_n$ to $NK_0$ lie above $v_0$. Thus, without loss of generality, we may assume that $K=K_0$. Let ${{{\mathcal F}}}_L$ be the family of valuations in ${{{\mathcal F}}}$ which are totally split in $L/K$. Then $|{{{\mathcal F}}}_L| = \kappa$. Let $\{\theta_\del \mid \del \in \Del\}$ be a maximal family of pairwise-independent geometric solutions of $(\mu,\alpha)$ with corresponding solution fields $L_i$. We wish to show that $|\Delta| = \kappa$. Suppose $|\Del| < \kappa$. Each $L_\del$ is defined by a polynomial over $K$. Adjoining all the coefficients of all these polynomials to the prime field of $K$, we get that the solutions are defined over a smaller subfield. That is, there exists a field $M \subset K$ with $|M| < \kappa$, such that each $L_\del$ is of the form $M_\del K$, for some finite Galois extension $M_\del$ of $M$. Let ${{{\mathcal F}}}' = \{v \in {{{\mathcal F}}}_L \st v(a) = 0 \hbox{ for all } a \in M^\times\}$. Then ${{{\mathcal F}}}_L{\smallsetminus}{{{\mathcal F}}}' = \bigcup _{a \in M^\times} \{v \in {{{\mathcal F}}}_L \st v(a) \neq 0\}$ is of cardinality $|M|$ at most. In particular ${{{\mathcal F}}}'$ is infinite. Let $\Lambda$ be the finite set given in Corollary \[cor:vel-geo-sol\]. By discarding finitely many valuations, we can assume that each $v \in {{{\mathcal F}}}'$ is trivial on $\Lambda$. Choose distinct $v_{i,j} \in {{{\mathcal F}}}', 1 \leq i \leq k, 1 \leq j \leq g_i$. Then all the $v_{i,j}$ are trivial on $M$, hence on all $M_i$. By [@FriedJarden2005 Lemma 2.3.6], all of the $v_{i,j}$ are unramified in $L_\del/K$ for each $\del \in \Del$. By Corollary \[cor:vel-geo-sol\] there exists a geometric solution ${{\mathfrak P}}^*$ with residue field $L'$, and extensions $v'_{i,j}$-s of the $v_{i,j}$-s to $L'$, such that ${\textnormal{Gal}}(L'/L)$ is generated by the inertia groups of the $v'_{i,j}$ at $L'/L$. Denote these groups by $I_{i,j}$. It suffices to show that $L'$ is linearly disjoint from of each $L_\del$ over $L$, $\del \in \Del$, since then $\{\theta_\del \mid \del\in \Del\}$ is not maximal. Indeed, let $\del \in \Delta$, and let $1 \leq i \leq k, 1 \leq j \leq g_i$. The restriction ${\textnormal{Gal}}(L'/L) \to {\textnormal{Gal}}(L'\cap L_\del/L)$ maps $I_{i,j}$ into the inertia group of $v'_{i,j}$ at $L' \cap L_\del /L$. Since $v_{i,j}$ is unramified at $L_\del/L$, this inertia group is trivial. Hence each element of $I_{i,j}$ fixes $L' \cap L_\del$. So $I_{i,j} \subseteq {\textnormal{Gal}}(L'/L' \cap L_\del)$. Since the $I_{i,j}$ generate $G$, ${\textnormal{Gal}}(L'/L' \cap L_\del) = G$, hence $L' \cap L_\del = L$. Thus $L'$ is linearly disjoint from $L_\del$ over $L$, for each $\del \in \Delta$, as claimed. Let $K$ be the quotient field of a complete local domain $R$ of dimension exceeding 1. Then its maximal purely inseparable extension $K_{ins}$ is fully Hilbertian. By Cohen’s structure theorem for complete local domains, $R$ is a finite extension of a domain $A$ of the form $K_0[[X_1,\ldots,X_n]]$ for some field $K_0$, or of the $A = {\mathbb{Z}}_p[[X_1,\ldots,X_n]]$ for some prime integer $p$. Since $\dim R > 1$, we have $n > 1$ in the first case and $n \geq 1$ in the second case. In both cases, [@Paran2010 §5] proves the existence of a family of valuations satisfying conditions (a)-(c) of Proposition \[prop:fully-Hilbertian\]. Thus the maximal purely inseparable extension of ${{\rm Quot}}(A)$ is fully Hilbertian. Since $K_{ins}$ is a finite extension of it, $K_{ins}$ is also fully Hilbertian, by Corollary \[cor:finite-ext\]. Finally, we get Theorem \[IT:com-krull-dom\]: Let $K$ be the quotient field of a complete local domain $R$ of dimension exceeding 1. Then ${\textnormal{Gal}}(K)$ is semi-free of rank $|K|$. Since $K$ is the quotient field of a complete domain, every finite split embedding problem over $K$ is regularly solvable, by [@Paran2008] or [@Pop2009]. Hence every finite split embedding problem over $K_{ins}$ is regularly solvable. By the preceding theorem $K_{ins}$ is fully Hilbertian, hence Proposition \[prop:Cond-semi-free\], ${\textnormal{Gal}}(K_{ins})$ is semi-free of rank $|K|$. Finally ${\textnormal{Gal}}(K)\cong {\textnormal{Gal}}(K_{ins})$ via the restriction map, hence ${\textnormal{Gal}}(K)$ is semi-free. [10]{} Emil Artin, *Algebraic numbers and algebraic functions*, Nelson, 1968. Lior Bary-Soroker, *On pseudo algebraically closed extensions of fields*, 2008. Lior Bary-Soroker, *On the characterization of [H]{}ilbertian fields*, Int. Math. Res. Not. IMRN (2008), Art. ID rnn 089, 10. Lior Bary-Soroker, Dan Haran, and David Harbater, *Permanence criteria for semi-free profinite groups*, 2008. Pierre D[è]{}bes and Bruno Deschamps, *The regular inverse [G]{}alois problem over large fields*, Geometric [G]{}alois actions, 2, London Math. Soc. Lecture Note Ser., vol. 243, Cambridge Univ. Press, Cambridge, 1997, pp. 119–138. Michael D. Fried and Moshe Jarden, *Field arithmetic*, second ed., Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics \[Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics\], vol. 11, Springer-Verlag, Berlin, 2005. Wulf Dieter Geyer and Moshe Jarden, *Fields with the density property*, J. Algebra **35** (1975), 178–189. Wulf-Dieter Geyer and Moshe Jarden, *On the normalizer of finitely generated subgroups of absolute [G]{}alois groups of uncountable [H]{}ilbertian fields of characteristic [$0$]{}*, Israel J. Math. **63** (1988), no. 3, 323–334. Dan Haran, *Free subgroups of free profinite groups*, J. Group Theory **2** (1999), no. 3, 307–317. Dan Haran, *Hilbertian fields under separable algebraic extensions*, Invent. Math. **137** (1999), no. 1, 113–126. Dan Haran and Moshe Jarden, *Regular split embedding problems over complete valued fields*, Forum Math. **10** (1998), no. 3, 329–351. David Harbater and Katherine F. Stevenson, *Local [G]{}alois theory in dimension two*, Adv. Math. **198** (2005), no. 2, 623–653. Robin Hartshorne, *Algebraic geometry*, Graduate Texts in Mathematics, No. 52. Springer-Verlag, New York-Heidelberg, 1977. Moshe Jarden, *Intersections of local algebraic extensions of a [H]{}ilbertian*, Generators and relations in groups and geometries (Lucca, 1990), 343–405, NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci., **333**, Kluwer Acad. Publ., Dordrecht, 1991. Moshe Jarden and Alexander Lubotzky, *Hilbertian fields and free profinite groups*, J. London Math. Soc. (2) **46** (1992), no. 2, 205–227. Gunter Malle and B. Heinrich Matzat, *Inverse [G]{}alois theory*, Springer Monographs in Mathematics, Springer-Verlag, Berlin, 1999. Elad Paran, *Split embedding problems over complete domains*, Ann. of Math. (2), to appear. Elad Paran, *Galois theory over complete local domains*, manuscript. Florian Pop, *Embedding problems over large fields*, Ann. of Math. (2) **144** (1996), no. 1, 1–34. Florian Pop, *Henselian implies large*, manuscript. Jean-Pierre Serre, *Lectures on the [M]{}ordell-[W]{}eil theorem*, Aspects of Mathematics, E15, Friedr. Vieweg & Sohn, Braunschweig, 1989, Translated from the French and edited by Martin Brown from notes by Michel Waldschmidt. Helmut V[ö]{}lklein, *Groups as [G]{}alois groups*, Cambridge Studies in Advanced Mathematics, vol. 53, Cambridge University Press, Cambridge, 1996, An introduction. [^1]: If one wants to consider only Hilbertian fields, then our work shows the connection between Hilbertian fields and profinite groups having the property that any finite split embedding problem is solvable.
{ "pile_set_name": "ArXiv" }
--- abstract: | This paper considers the completion problem for a tensor (also referred to as a multidimensional array) from limited sampling. Our greedy method is based on extending the low-rank approximation pursuit (LRAP) method for matrix completions to tensor completions. The method performs a tensor factorization using the tensor singular value decomposition (t-SVD) which extends the standard matrix SVD to tensors. The t-SVD leads to a notion of rank, called tubal-rank here. We want to recreate the data in tensors from low resolution samples as best we can here. To complete a low resolution tensor successfully we assume that the given tensor data has low tubal-rank. For tensors of low tubal-rank, we establish convergence results for our method that are based on the tensor restricted isometry property (TRIP). Our result with the TRIP condition for tensors is similar to low-rank matrix completions under the RIP condition. The TRIP condition uses the t-SVD for low tubal-rank tensors, while RIP uses the SVD for matrices. We show that a subgaussian measurement map satisfies the TRIP condition with high probability and gives an almost optimal bound on the number of required measurements. We compare the numerical performance of the proposed algorithm with those for state-of-the-art approaches on video recovery and color image recovery.\ author: - 'An-Bao Xu [^1][^2]' bibliography: - 'ref.bib' title: 'Tensor Completion via a Low-Rank Approximation Pursuit' --- Tensor completion, low rank, tensor singular value decomposition, restricted isometry property, rank minimization, approximation pursuit Introduction ============ Tensors generalize vectors and matrices [@KB2009; @H2012]. Tensor completions recently have drawn much attention [@LMWY2013; @GRY2011; @SDLS2013; @TKK2010; @PC2013; @YZ2014]. We try to partially reconstruct the original tensor from a given low rank tensor formed by partial observations and do so sufficiently well. Such problems arise in a variety of applications, such as in signal processing [@LMWY2013; @ZEAHK2014], in multi-class learning [@OTJ2010], in data mining [@KS2008; @SPLCLQ2009], and in dimension reduction [@LLR1995]. Typically, tensor completion is formulated as an optimization problem that involves sums of nuclear norms of the unfolding matrices inside the unknown tensor and uses the notion of tensor rank defined in terms of the higher singular value decomposition (HOSVD) [@LMWY2013; @GRY2011; @SDLS2013; @TKK2010]. However, algorithms that use sums of matrix nuclear norms generally need to solve several subproblems via a full singular value decomposition at each iteration step. This limits their speed, especially for large tensor sizes due to the high cost of full SVDs. Moreover sums of nuclear norms do not allow exploiting the tensor structure and such processes lead to suboptimal procedures, see [@YZ2014] e. g.. Furthermore, if the tensor rank is determined by the CANDECOM/PARAFAC decomposition (CP) for example, a best multirank-k approximation may only exist under further assumptions. Computing the multirank-k approximation is highly nontrivial even when it exits, see [@ZSKA2018]. For tensor decompositions, our algebraic t-SVD framework differs from the classic multilinear algebraic framework [@H2012] and the tensor tubal rank (using t-SVD), differs from the CP rank (using CP) and the Tucker rank (using HOSVD). Therefore, bounds and conditions of tensor completion using low Tucker rank or low CP rank are not directly comparable to those in this paper. The t-SVD has recently been used in [@LFCLLY2019; @ZEAHK2014; @ZA2017] for tensor recoveries applied to computer vision. These papers define tensor incoherence conditions and obtain theoretical performance bounds for the corresponding algorithms. Our method differs from these as we employ the TRIP condition to obtain theoretical performance bounds. We were inspired by [@D2016], [@RSS2017] and their main tools, namely $\epsilon$-nets and covering numbers. With $\epsilon$-nets and covering numbers, we show that subgaussian measurement maps satisfy the TRIP condition with high probability and for a certain almost optimal bound on the number of measurements. Thus we can provide convergence results that hold with high probability when the TRIP conditions are satisfied. In this paper, the Y tensor and the X tensors all have the same original large tensor sizes in three dimensions. our basic tensor completion model involves a large data tensor $\bm{\mathcal{Y}}_{n_1,n_2,n_3} $, a set $\Omega$ comprised of horizontal, lateral and frontal slice indices $\{(O_h,O_l,O_f)\}$ with $1 \leq O_h \leq n_1$ and $1 \leq O_l \leq n_2$ and $1 \leq O_f \leq n_3$ and $|\Omega| \ll n_1 \cdot n_2 \cdot n_3$ that indicate the horizontal, lateral and frontal slice indices of a entry subset of $\bm{\mathcal{Y}}$. In this framework we search for a low rank completion tensor $\bm{\mathcal{X}}_{n_1,n_2,n_3}$ which has the same entries as $\bm{\mathcal{Y}}$ in the positions of $\Omega$ and has arbitrary entries in its complementary positions. The selection is such that $\bm{\mathcal{X}}$ has the minimal possible tensor rank. This tensor completion problem can be formalized as follows. $$\label{1.1} \text{Find } \ \mathop{\min}\limits_{\bm{\mathcal{X}}\in\mathbb{R}^{n_1\times n_2 \times n_3}} {\rm rank_t}(\bm{\mathcal{X}}) \quad \text{such that} \quad P_\Omega(\bm{\mathcal{X}})=P_\Omega(\bm{\mathcal{Y}}),$$ where $\rm rank_t$ is the tensor tubal rank of Definition \[definition 2.6\], see below, $\Omega$ is the set of all index pairs in $\bm{\mathcal{Y}}$ that $\bm{\mathcal{X}}$ shares with $\bm{\mathcal{Y}}$. Here $P_\Omega$ denotes the orthogonal projector onto the span of tensors with zeros at the positions not in $\Omega$. The aim of our tensor completion is to create a low rank tensor $\bm{\mathcal{Y}}$ from the partially observed tensor $P_\Omega(\bm{\mathcal{Y}})$ of $\bm{\mathcal{Y}}$. This tensor completion is based on that you are given a imprecise tensor and want to recreate a same sized tensor that mimicks the nonzero part of the original tensor $\bm{\mathcal{Y}}$ as best as you can. The nonzero part of the original tensor is $P_\Omega(\bm{\mathcal{Y}})$ which is fewer than the $n_1 \cdot n_2 \cdot n_3$ measurements in $\bm{\mathcal{Y}}$. This benefits many applications. Following Section 5 of [@WLLFDY2015], we extend the basic model (\[1.1\]) to the following tensor sensing problem $$\label{1.2} \text{Find } \ \mathop{{\rm min}}\limits_{\bm{\mathcal{X}} \in \mathbb{R}^{n_1 \times n_2 \times n_3}} {\rm rank_t} (\bm{\mathcal{X}}) \quad s.t. \ \Phi(\bm{\mathcal{X}})=\Phi(\bm{\mathcal{Y}}),$$ where $\bm{\mathcal{Y}}$ is a target low rank tensor and $\Phi =\phi \cdot {\rm vec}$ is a linear operator. Its inverse $\Phi^{-1} $ is the linear operator with $\Phi\Phi^{-1}(\mathbf b) = \mathbf b$ for any vector $\mathbf b$. Note that $\Phi^{-1}\Phi$ is not an identity operator. This paper proposes a simple and efficient algorithm to solve the more general problem (\[1.2\]). In every iteration, rank-one basis tensors are constructed by the truncated t-SVD of the currently known residual tensor. In our standard version we fully update the weights (or the coefficients) for all rank-one tensors in the current basis set in each iteration. The most time-consuming process in this version is the truncated t-SVD computation. Therefore we recommend readers to adopt the t-SVD algorithm of [@LFCLLY2019] or the rt-SVD method of [@ZSKA2018] here. We also adopt an economic weight updating rule from [@XX2017] to decrease the time and storage complexity further. Interestingly, both algorithms converge linearly.\ The notion of the TRIP condition here and our results are new and not directly implied by the results from matrix completions using the standard matrix RIP conditions and the matrix LRAP.\ The main contributions of our paper are: - We propose a computationally more efficient greedy algorithm for tensor completions, which extends the LRAP method for matrix completions to tensor completions. - This article consists in an analysis of the TRIP related to the tensor formats t-SVD for random measurement maps. We show that subgaussian measurement maps satisfy the TRIP with high probability under a certain almost optimal bound on the number of measurements. - Using this result of the TRIP condition, We show that the proposed algorithm achieve linear convergence. - We illustrate the efficiency of the proposed algorithm for tensor completions via numerical comparison with those for state-of-the-art approaches on video recovery and color image recovery.\ The next section introduces notations and preliminaries. In Section III, we construct our standard algorithm and a more economic version. Section IV defines the RIP condition and proves that subgaussian measurement ensembles satisfy the RIP condition with high probability. We prove linear convergence of our algorithms in Section V. Empirical numerical test evaluations and comparisons with other methods are presented in Section VI. They verify the efficiency of the proposed algorithms. Notations and Preliminaries =========================== Notations --------- Here we denote matrices by boldface capital letters and vectors by boldface lowercase letters. Tensors are represented in Euler bold script letters. For example, a third-order tensor is represented as $\bm{\mathcal{A}}$, and its $(i,j,k)$th entry is represented as $\bm{\mathcal{A}}_{ijk}$ or $a_{ijk}$. The Matlab notation $\bm{\mathcal{A}}(i,:,:),\bm{\mathcal{A}}(:,i,:)$ and $\bm{\mathcal{A}}(:,:,i)$ are used to denote respectively the $i$-th horizontal, lateral and frontal slices. Used frequently, $A^{(i)} $ denotes compactly the frontal slice $\bm{\mathcal{A}}(:,:,i)$ and $\bm{\mathcal{A}}(i,j,:)$ denotes the tube of $i,j$ in the third tensor dimension. The Frobenius inner product of two compatible tensors is $$\langle \bm{\mathcal{A}}, \bm{\mathcal{B}} \rangle=\sum\nolimits_{i_1, \cdots, i_t} {\bm{\mathcal{A}}_{i_1, \cdots, i_t}\bm{\mathcal{B}}_{i_1, \cdots, i_t}}.$$ Then $||\bm{\mathcal{A}}||_F = \sqrt{\langle \bm{\mathcal{A}}, \bm{\mathcal{A}} \rangle}$ is the corresponding Frobenius tensor norm and $d_F(\bm{\mathcal{A}}, \bm{\mathcal{B}}) = ||\bm{\mathcal{A}} - \bm{\mathcal{B}}||_F$ the induced tensor metric. For any $\bm{\mathcal{A}} \in \mathbb{C}^{n_1 \times n_2 \times n_3}$, the complex conjugate of $\bm{\mathcal{A}}$ is denoted as ${\rm conj}(\bm{\mathcal{A}})$, whose entries are the complex conjugates of the respective entry in $\bm{\mathcal{A}}$. The vector ${\rm vec}(\bm{\mathcal{Y}})$ contains the entries of $\bm{\mathcal{Y}}$ reshaped by concatenating all tensor entries. The vector $\mathop {\textbf{y}}\limits^{\textbf{.}}={\rm vec}_\Omega(\bm{\mathcal{Y}})=\{(y_{w_1}, \cdots, y_{w_{|\Omega|}})^T \ \forall w_i \in\Omega\}$ denotes the vector generated by concatenating all elements of $\bm{\mathcal{Y}}$ in the index set $\Omega$. Discrete Fourier Transformation (DFT) ------------------------------------- The DFT on $\mathbf v \in \mathbb{R}^n$, denoted as $\hat{\textbf v} $, is obtained by $$\hat{\mathbf v} = \mathbf F_n \mathbf v \in \mathbb{C}^n,$$ where $\mathbf F_n$ is the DFT matrix $$\mathbf F_n = \begin{bmatrix} 1 & 1 & 1& \cdots & 1 \\ 1 & \omega & \omega^2 & \cdots & \omega^{n-1} \\ \vdots & \vdots & \vdots & \ddots & \vdots \\ 1 & \omega^{n-1} & \omega^{2(n-1)} & \cdots & \omega^{(n-1)(n-1)} \\ \end{bmatrix} \in \mathbb{C}^{n \times n},$$ and $\omega = e^{-{2 \pi i \over n}}$ is a primitive $n$-th root of unity with $i = \sqrt{-1}$ [@LFCLLY2019]. Note that $\mathbf F_n / \sqrt{n}$ is an orthogonal matrix, i.e., $$\label{7} \mathbf F^*_n \mathbf F_n = \mathbf F_n \mathbf F^*_n = n \mathbf I_n.$$ Hence $\mathbf F^{-1}_n = \mathbf F^*_n /n$. For any tensor $\bm{\mathcal{A}} \in \mathbb{R}^{n_1 \times n_2 \times n_3}$, ${\bm{\hat\mathcal{A}}} \in \mathbb{C}^{n_1 \times n_2 \times n_3}$ denotes the result of the DFT on $\bm{\mathcal{A}}$ along the 3-rd dimension, i.e., using the DFT on all $i,j$ tubes in $\bm{\mathcal{A}}$. By using the Matlab command ${\rm fft}$, we obtain $$\hat{\bm{\mathcal{A}}} = {\rm fft}(\bm{\mathcal{A}},[],3).$$ Analogously $\bm{\mathcal{A}}$ is computed from $\hat{\bm{\mathcal{A}}}$ by performing the inverse FFT, i.e. , $$\bm{\mathcal{A}} = {\rm ifft}(\hat{\bm{\mathcal{A}}},[],3).$$ $\overline{\mathbf A} \in \mathbb{C}^{n_1 n_3 \times n_2 n_3}$ is the block diagonal matrix whose $i$-th diagonal block is the $i$-th frontal slice $\hat{\mathbf A}^{(i)} \text{ of } \hat{\bm{\mathcal{A}}}$, i.e., $$\overline{\mathbf A} = {\rm bdiag}(\hat{\bm{\mathcal{A}}}) = \begin{bmatrix} \hat{\mathbf A}^{(1)} & 0 & \cdots & 0 \\ 0 & \hat{\mathbf A}^{(2)} & 0 & \cdots \\ \vdots & \ddots & \ddots & \ddots \\ 0 & \cdots & 0 & \hat{\mathbf A}^{(n)} \\ \end{bmatrix}.$$ Here ${\rm bdiag}$ denotes the operator that maps the tensor $\hat{\bm{\mathcal{A}}}$ to the block diagonal matrix $\overline{\mathbf A}$. The block circulant matrix ${\rm bcirc}(\bm{\mathcal{A}}) \in \mathbb{R}^{n_1 n_3 \times n_2 n_3}$ of $\bm{\mathcal{A}}$ is given by $${\rm bcirc}(\bm{\mathcal{A}}) = \begin{bmatrix} \mathbf A^{(1)} & \mathbf A^{(n_3 )} & \cdots & \mathbf A^{(2)} \\ \mathbf A^{(2)} & \mathbf A^{(1)} & \cdots & \mathbf A^{(3)} \\ \vdots & \vdots & \ddots & \vdots \\ \mathbf A^{(n_3 )} & \mathbf A^{(n_3 -1)} & \cdots & \mathbf A^{(1)} \\ \end{bmatrix}.$$ The block circulant matrix can be block diagonalized, i.e. , $$\label{10} (\mathbf F_{n_3} \otimes \mathbf I_{n_1}) \cdot {\rm bcirc}(\bm{\mathcal{A}}) \cdot (\mathbf F^{-1}_{n_3} \otimes \mathbf I_{n_2}) = \overline{\mathbf A}$$ where $\otimes$ denotes the Kronecker product and $(\mathbf F_{n_3} \otimes \mathbf I_{n_1}) / \sqrt{n_3}$ is orthogonal [@LFCLLY2019]. Based on (\[7\]), we have $$\label{12} ||\bm{\mathcal{A}}||_F = {1 \over \sqrt{n_3}} ||\overline{\mathbf A}||_F ,$$ $$\label{13} \langle \bm{\mathcal{A}}, \bm{\mathcal{B}} \rangle = {1 \over n_3} \langle \overline{\mathbf A}, \overline{\mathbf B} \rangle \ .$$ T-product and T-SVD ------------------- For $\bm{\mathcal{A}} \in \mathbb{R}^{n_1 \times n_2 \times n_3}$, the ${\rm unfold}$ operator map of $\bm{\mathcal{A}}$ is the matrix of size $n_1 n_3 \times n_2$ $${\rm unfold}(\bm{\mathcal{A}}) = \begin{bmatrix} \mathbf A^{(1)} \\ \mathbf A^{(2)} \\ \vdots \\ \mathbf A^{(n_3)} \end{bmatrix}.$$ Its inverse operator ${\rm fold}$ operates so that $${\rm fold}({\rm unfold}(\bm{\mathcal{A}})) = \bm{\mathcal{A}}.$$ \[definition 2.1\] [@KM2011] **(T-product)** Let $\bm{\mathcal{A}} \in \mathbb{R}^{n_1 \times n_2 \times n_3}$ and $\bm{\mathcal{B}} \in \mathbb{R}^{n_2 \times l \times n_3}$. Then the t-product $\bm{\mathcal{A}} * \bm{\mathcal{B}} $ is the following tensor of size $n_1 \times l \times n_3$ : $$\label{14} \bm{\mathcal{A}} * \bm{\mathcal{B}} = {\rm fold}({\rm bcirc}(\bm{\mathcal{A}}) \cdot {\rm unfold}(\bm{\mathcal{B}}))\ .$$ The t-product is analogous to matrix multiplication. The only difference between them is that the circular convolution substitutes multiplication between elements. Therefore the t-product of tensors is matrix multiplication in the Fourier domain; namely, $\bm{\mathcal{C}} = \bm{\mathcal{A}} * \bm{\mathcal{B}} $ is equivalent to $\overline{\mathbf C} = \overline{\mathbf A} \cdot \overline{\mathbf B}$ due to (\[10\]) [@LFCLLY2019]. This property provides an efficient way (based on the FFT) to calculate t-product instead of performing (\[14\]). See Algorithm 1 in [@LFCLLY2019]. The t-product has many similar properties as the matrix-matrix product. \[definition 2.2\][@LFCLLY2019] **(Conjugate transpose)** The conjugate transpose of a tensor $\bm{\mathcal{A}} \in \mathbb{C}^{n_1 \times n_2 \times n_3}$ is the tensor $\bm{\mathcal{A}}^* \in \mathbb{C}^{n_1 \times n_2 \times n_3}$ obtained by conjugate transposing each of the frontal slices and then reversing the order of transposed frontal slices 2 through $n_3$. \[definition 2.3\][@KM2011] **(Identity tensor)** The identity tensor $\bm{\mathcal{I}} \in \mathbb{R}^{n \times n \times n_3}$ is the tensor with the $n \times n$ identity matrix $\mathbf I_n$ as its first frontal slice and all other frontal slices being $\mathbf O_n$. It is obviously that $\bm{\mathcal{A}} * \bm{\mathcal{I}} = \bm{\mathcal{A}}$ and $\bm{\mathcal{I}} * \bm{\mathcal{A}} = \bm{\mathcal{A}}$ for appropriate dimensions. For the tensor $\hat{\bm{\mathcal{I}}} = {\rm fft}(\bm{\mathcal{I}}, [], 3)$ for example, the frontal slice is the identity matrix. \[definition 2.4\][@KM2011] **(Orthogonal tensor)** A tensor $\bm{\mathcal{Q}} \in \mathbb{R}^{n \times n \times n_3}$ is orthogonal if it satisfies $\bm{\mathcal{Q}}^* * \bm{\mathcal{Q}} = \bm{\mathcal{Q}} * \bm{\mathcal{Q}}^* = \bm{\mathcal{I}}.$ The notion of partial orthogonality can be defined and is analogous to that a tall, thin matrix has orthogonal columns [@KM2011]. In this case if $\bm{\mathcal{Q}}$ is $n \times q \times n_3$ and **partially orthogonal**, then this implies that $\bm{\mathcal{Q}}^* * \bm{\mathcal{Q}}$ is well defined and equivalent to the $q \times q \times n_3$ identity tensor. \[definition 2.5\][@KM2011] **(F-diagonal tensor)** A tensor is called f-diagonal if each of its frontal slices is a diagonal matrix. \[Theorem 2.2\][@LFCLLY2019] **(T-SVD)** Let $\bm{\mathcal{A}} \in \mathbb{R}^{n_1 \times n_2 \times n_3}$. Then it can be factored as $$\label{16} \bm{\mathcal{A}} = \bm{\mathcal{U}} * \bm{\mathcal{S}} * \bm{\mathcal{V}}^*$$ where $\bm{\mathcal{U}} \in \mathbb{R}^{n_1 \times n_2 \times n_3}$ and $\bm{\mathcal{V}} \in \mathbb{R}^{n_1 \times n_2 \times n_3}$ are orthogonal, and $\bm{\mathcal{S}} \in \mathbb{R}^{n_1 \times n_2 \times n_3}$ is an f-diagonal tensor. Theorem \[Theorem 2.2\] implies that any 3 order tensor can be factorized into 3 factors, two of which are orthogonal tensors and the central factor is an f-diagonal tensor, as depicted in Figure \[Figure 1\]. ![The construction of the t-SVD is similar to the matrix substitute the equivalent matrix operations. Like the matrix SVD, the t-SVD can also be formalized as the sum of outer tensor products[]{data-label="Figure 1"}](t_SVD2.eps) \[definition 2.6\] [@LFCLLY2019] **(Tensor tubal rank)** For $\bm{\mathcal{A}} \in \mathbb{R}^{n_1 \times n_2 \times n_3}$ the tensor tubal rank, denoted as ${\rm rank_t} (\bm{\mathcal{A}})$, is the number of nonzero singular tubes of $\bm{\mathcal{S}}$, where $\bm{\mathcal{S}}$ is the central factor of the t-SVD of $ \bm{\mathcal{A}} = \bm{\mathcal{U}} * \bm{\mathcal{S}} * \bm{\mathcal{V}}^* $. \[definition 2.9\][@ZSKA2018] **(Truncated t-SVD)** Given a tensor $\bm{\mathcal{A}} \in \mathbb{R}^{n_1 \times n_2 \times n_3}$, the truncated t-SVD of $\bm{\mathcal{A}}$ is $$\bm{\mathcal{A}}_k = \sum^k_{i=1} \bm{\mathcal{U}}(:,i,:) * \bm{\mathcal{S}}(i,i,:) * \bm{\mathcal{V}}(:,i,:)^T,$$ where $k \leq {\rm min}(n_1, n_2)$ is a target truncation term. Here $\bm{\mathcal{A}}_k = \bm{\mathcal{U}}_k * \bm{\mathcal{S}}_k * \bm{\mathcal{V}}_k^T,$ where $\bm{\mathcal{U}}_k \in \mathbb{R}^{n_1 \times k \times n_3}$ and $\bm{\mathcal{V}}_k \in \mathbb{R}^{n_2 \times k \times n_3}$ are partially orthogonal tensors and $\bm{\mathcal{S}}_k \in \mathbb{R}^{s \times k \times n_3}$ is an f-diagonal tensor. A nice feature of the truncated t-SVD is that it gives us the best “multirank-k” approximation of tensors [@ZSKA2018]. If a tensor $\bm{\mathcal{A}}$ has tubal rank $r$, we have $\bm{\mathcal{A}}$ = $\bm{\mathcal{A}}_r$ and we can use $\bm{\mathcal{A}}_r$ as a reduced version of the t-SVD [@ZA2017] . \[Lemma 2.2\] Let $\bm{\mathcal{M}}_{1}= \bm{\mathcal{U}} (:,1,:) * {\bm{\mathcal{S}} (1,1,:) \over ||\bm{\mathcal{S}} (1,1,:)||_F}* \bm{\mathcal{V}} (:,1,:)^T$ be the rank-one tensor from the t-SVD of $\bm{\mathcal{A}}$ .\ Then $\langle {\bm{\mathcal{M}}_1, \bm{\mathcal{A}}} \rangle \ge {\| \bm{\mathcal{A}} \|_F \over \sqrt{{\rm min}(m,n)}}$ for all $k \ge 1$. The optimum $\bm{\mathcal{M}}_1$ in our algorithm satisfies $$\langle \bm{\mathcal{M}}_1, \bm{\mathcal{A}}\rangle^2 = ||\bm{\mathcal{S}} (1,1,:)||^2_F \cdot \langle \bm{\mathcal{M}}_1, \bm{\mathcal{M}}_1 \rangle^2 = ||\bm{\mathcal{S}} (1,1,:)||^2_F$$ $$= || ({\mathbf F_{n_3} \over \sqrt{n_3}} \otimes I) \ {\rm unfold}(\bm{\mathcal{S}} (1,1,:)) ||^2_F$$ $$= {1 \over n_3} ||\hat{\mathbf S}^{(1)}(1,1)||^2_F+ \cdots + {1 \over n_3} ||\hat{\mathbf S}^{(n_3)}(1,1)||^2_F$$ $$\geq {1 \over n_3} {\sum_i [\hat{\mathbf S}^{(1)}(i,i)]^2 \over {\rm rank}(\hat{\mathbf S}^{(1)} ) }+ \cdots + {1 \over n_3} {\sum_i [\hat{\mathbf S}^{(n_3)}(i,i)]^2 \over {\rm rank}(\hat{\mathbf S}^{(n_3)} ) }$$ $$\geq {1 \over n_3} {\sum_i [\hat{\mathbf S}^{(1)}(i,i)]^2 \over {\rm min}(n_1,n_2) }+ \cdots + {1 \over n_3} {\sum_i [\hat{\mathbf S}^{(n_3)}(i,i)]^2 \over {\rm min}(n_1,n_2) }$$ $$= {1 \over n_3} { ||\overline{\mathbf S}||^2_F \over {\rm min}(n_1,n_2) }= {1 \over n_3} { ||\overline{\mathbf A}||^2_F \over {\rm min}(n_1,n_2) } = {1 \over n_3} { n_3 ||\bm{\mathcal{A}}||^2_F \over {\rm min}(n_1,n_2) }.$$ This completes the proof. $\Box$ This result will be used to prove Theorem \[Theorem 3.1\]. Low-Rank Approximation Pursuit for Tensor Sensing (LRAP4TS) =========================================================== Based on Theorem \[Theorem 2.2\] and Figure \[Figure 1\], any tensor $\bm{\mathcal{X}}\in \mathbb{R}^{n_1\times n_2 \times n_3}$ can be written as a linear combination of rank-one tensors, namely $$\bm{\mathcal{X}}= \bm{\mathcal{U}} * \bm{\mathcal{S}} * \bm{\mathcal{V}}^T=\mathop{\sum}\limits_{i=I} \bm{\mathcal{U}}(:,i,:) * \bm{\mathcal{S}}(i,i,:) * \bm{\mathcal{V}}(:,i,:)^T$$ $$=\mathop{\sum}\limits_{i=I} ||\bm{\mathcal{S}}(i,i,:)||_F \cdot \bm{\mathcal{U}}(:,i,:) * {\bm{\mathcal{S}}(i,i,:) \over ||\bm{\mathcal{S}}(i,i,:)||_F}* \bm{\mathcal{V}}(:,i,:)^T$$ $$=\mathop{\sum}\limits_{i=I}\theta_i\bm{\mathcal{M}}_i=\bm{\mathcal{M}} (\bm{\theta})$$ where $\{\bm{\mathcal{M}}_i = {\bm{\mathcal{S}}(i,i,:) \over ||\bm{\mathcal{S}}(i,i,:)||_F} : i\in I \}$ is the set of all $n_1\times n_2\times n_3$ rank-one tensors with unit Frobenius norm and $\theta_i = ||\bm{\mathcal{S}}(i,i,:)||_F $ is the norm of the $i$-th singular value tube. Hence the original low rank tensor sensing problem (\[1.2\]) can be rewritten as $$\label{2.2} \mathop{\min}\limits_{\bm{\theta}} \| {\bm{\theta}} \|_0 \quad \text{s.t.} \quad \Phi(\bm{\mathcal{M}}(\bm{\theta}))=\Phi(\bm{\mathcal{Y}}),$$ where $\| \bm{\theta}\|_0$ denotes the number of nonzero elements of vector $ \bm{\theta}$. An alternative formula of Problem (\[2.2\]) is $$\mathop{\min}\limits_{\bm{\theta}} \| \Phi(\bm{\mathcal{M}}(\bm{\theta}))-\Phi(\bm{\mathcal{Y}}) \|_F^2 \quad \text{s.t.} \quad \ \| {\bm{\theta}} \|_0 \leqslant r.$$ This problem can be solved by our algorithm, which is an LRAP [@XX2017] type algorithm using rank-one tensors as the basis. Below we show the main steps of our LRAP4TS and its economic version ELRAP4TS in Algorithm \[alg:Framwork\]. Let the orthogonal projector $P_{\Omega}$ be the linear operator $\Phi$, then Algorithm \[alg:Framwork\] is suitable for tensor completion and we will be refer to it as LRAP4TC or ELRAP4TC. Now we give details of the iteration procedure details for both versions of the tensor sensing algorithm. Both greedy algorithms add $s$ new rank-one tensors to the basis set in each iteration. Both algorithms alternate between three iteration steps: (1) pursuit s rank-one basis tensors; (2) update the weights of the tensors; and (3) renew the residual tensor.    $\bm{\mathcal{R}}_0 = \Phi^{-1} \Phi (\bm{\mathcal{Y}})$, the tubal rank $r$ of the estimated tensor $\bm{\mathcal{Y}}$ and the number $s$ of candidates searched in each iteration.\ **Initialize:** Set $\bm{\mathcal{X}}_0=0$, $\bm{\mathcal{R}}_1=\bm{\mathcal{R}}_0$, $\bm{\theta}^0=0$, $\widehat{\bm{\mathcal{Y}}}_{0}=0$ and $k=1$.\     $k \leqslant \ceil{r/s}$  **do :**\ **Step 1:** [(This step yields the best multirank-s approximation of $\bm{\mathcal{R}}_k$)\ Search for the $s$ leading principle left and right singular vectors of tubes $\{ (\bm{\mathcal{U}}_k (:,j,:)$, $\bm{\mathcal{S}}_k (j,j,:), \bm{\mathcal{V}}_k (:,j,:)), j=1,\cdots, s \}$ of $\bm{\mathcal{R}}_k$, and set up the $s$ rank-one basis tensors $\{ \bm{\mathcal{M}}_{k,j}=\bm{\mathcal{U}}_k (:,j,:) * {\bm{\mathcal{S}}_k (j,j,:) \over ||\bm{\mathcal{S}}_k (j,j,:)||_F}* \bm{\mathcal{V}}_k (:,j,:)^T, j=1,\cdots, s \}$.\ ]{} \ **Step 2:** [Solve the following least squares problem:\ \[-4mm\] $$\label{2.6} 1) \mathop{\min}\limits_{\bm{\theta}=(\theta_{1,1},\cdots,\theta_{1,s}, \cdots, \theta_{k,1},\cdots \theta_{k,s})^T \in \mathbb{R}^{sk}}\| \sum_{i=1}^{k} {[\theta_{i,1} \Phi^{-1} \Phi (\bm{\mathcal{M}}_{i,1}) + \cdots + \theta_{i,s}\Phi^{-1} \Phi (\bm{\mathcal{M}}_{i,s})] - \bm{\mathcal{R}}_0} \|^2. \ \ ({\rm LRAP4TS})$$\ \[-4mm\] Or\ \[-4mm\] $$\label{2.9} 2) \mathop{\min}\limits_{\bm{\alpha}=(\alpha_0, \alpha_1,\cdots, \alpha_s )^T \in \mathbb{R}^{s}}\| \alpha_0 \bm{\mathcal{X}}_{k-1} + \alpha_1 \Phi^{-1} \Phi (\bm{\mathcal{M}}_{k,1} +\cdots+\alpha_s \Phi^{-1} \Phi (\bm{\mathcal{M}}_{k,s} )-\bm{\mathcal{R}}_0 \|^2. \ \ ({\rm ELRAP4TS})$$\ ]{} \ **Step 3:** [1) Set $\bm{\mathcal{X}}_{k}=\sum_{i=1}^{k} [\theta_{i,1}^{k} \Phi^{-1} \Phi(\bm{\mathcal{M}}_{i,1})+ \cdots + \theta_{i,s}^{k} \Phi^{-1} \Phi(\bm{\mathcal{M}}_{i,s})]$ and $\bm{\mathcal{R}}_{k+1}= \Phi^{-1} \Phi(\bm{\mathcal{Y}}) - \bm{\mathcal{X}}_{k}$; $k \leftarrow k+1$\ Or\ 2) Set $\bm{\mathcal{X}}_{k}=\alpha_0^k \bm{\mathcal{X}}_{k-1} + \alpha_1^k \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1} )+\cdots+\alpha_s^k \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,s} )$,\ $\widehat{\bm{\mathcal{Y}}}_{k}=\alpha_0^k \widehat{\bm{\mathcal{Y}}}_{k-1} + \alpha_1^k \bm{\mathcal{M}}_{k,1} +\cdots+\alpha_s^k \bm{\mathcal{M}}_{k,s}$ and $\bm{\mathcal{R}}_{k+1}= \Phi^{-1} \Phi(\bm{\mathcal{Y}}) - \bm{\mathcal{X}}_{k}$; $k \leftarrow k+1$.\ ]{} \ \    For 1) LRAP4TS: Construct tensor $\widehat{\bm{\mathcal{Y}}}=\sum_{i=1}^{k}[\theta_{i,1}^{k} (\bm{\mathcal{M}}_{i,1})+ \cdots + \theta_{i,s}^{k} (\bm{\mathcal{M}}_{i,s})]$.\ For 2) ELRAP4TS: $\widehat{\bm{\mathcal{Y}}}=\widehat{\bm{\mathcal{Y}}}_{k}$.\ \(1) In this step, we find a set of $s$ rank-one basis tensors {$\bm{\mathcal{M}}_{k,1},\cdots, \bm{\mathcal{M}}_{k,s}$} with unit Frobenius norm, which are related to the currently known residual tensor $\bm{\mathcal{R}}_k$. The tensors $\{\bm{\mathcal{M}}_{k,1},\cdots, \bm{\mathcal{M}}_{k,s}\}$ are constructed based on the $s$ leading principle left and right singular vectors of the tubes $\{ (\bm{\mathcal{U}}_k (:,j,:), \bm{\mathcal{S}}_k (j,j,:), \bm{\mathcal{V}}_k (:,j,:)), j=1,\cdots, s \}$ from the t-SVD of $\bm{\mathcal{R}}_k$. By construction, the $s$ rank-one basis tensors $\{ \bm{\mathcal{M}}_{k,j}=\bm{\mathcal{U}}_k (:,i,:) * {\bm{\mathcal{S}}_k (i,i,:) \over ||\bm{\mathcal{S}}_k (i,i,:)||_F}* \bm{\mathcal{V}}_k (:,i,:)^T, j=1,\cdots, s \}$ are orthogonal to each other and have Frobenius norm one. This step locates the best multirank-s approximation to $\bm{\mathcal{R}}_k$ by computing a truncating t-SVD. The t-SVD can be obtained by the t-SVD algorithm or the rt-SVD method. \(2) In the standard version of this step, the weights $\bm{\theta}^k$ are estimated for all current basis tensors $\{\bm{\mathcal{M}}_{i,1},\cdots, \bm{\mathcal{M}}_{i,s}, i=1,\cdots, k\}$ by solving the least squares problem (\[2.6\]). Using the orthogonal projector $P_{\Omega}$ as the linear operator $\Phi$ in formula (\[1.2\]), Problem (\[2.6\]) can be rewritten as $$\mathop{\min}\limits_{\bm{\theta} \in \mathbb{R}^{sk}} \| \sum_{i=1}^{k} \sum_{j=1}^{s} {\theta_{i,j} P_{\Omega}(\bm{\mathcal{M}}_{i,j}) - P_{\Omega}(\bm{\mathcal{Y}})} \|^2.$$ By reshaping the tensors $P_{\Omega}(\bm{\mathcal{Y}})$ and $P_{\Omega}(\bm{\mathcal{M}}_{i,j})$ into column vector form $\mathop {\textbf{y}}\limits^{\textbf{.}}$ and ${\mathop {\textbf{m}}\limits^{\textbf{.}}}_{i,j}$, the above overdetermined system can be reformulated as $\mathop{\min}\limits_{\bm{\theta} \in \mathbb{R}^{sk}} \| \overline {\mathbf M}_k \bm{\theta} - \mathop {\textbf{y}}\limits^{\textbf{.}} \|^2$. Here $\overline {\mathbf M}_k = [ {\mathop {\textbf{m}}\limits^{\textbf{.}}}_{1,1},\cdots,{\mathop {\textbf{m}}\limits^{\textbf{.}}}_{1,s}, \cdots, {\mathop {\textbf{m}}\limits^{\textbf{.}}}_{k,1},\cdots, {\mathop {\textbf{m}}\limits^{\textbf{.}}}_{k,s} ]$ is the matrix formed by all reshaped basis tensors. The row size of $\overline {\mathbf M}_k$ is equal to the total number of observed entries $p=|\Omega|$. In case of using ELRAP4TS, the orthogonal projection step consists of only tracking the estimated tensor $\bm{\mathcal{X}}_{k-1}$ and the $s$ rank-one basis tensors $\bm{\mathcal{M}}_{k,j}$ for $j=1,\cdots,s$. This step of ELRAP4TS updates the weights for $s+1$ tensors based on the solution of the least squares problem (\[2.9\]). \(3) Here we update the residual tensor $\bm{\mathcal{R}}_{k+1}=\mathcal{P}_\Omega(\bm{\mathcal{Y}})-\bm{\mathcal{X}}_{k}$ as follows: In the standard LRAP4TS algorithm we use $$\bm{\mathcal{X}}_{k}=\Phi(\bm{\mathcal{M}}(\bm{\theta}^{k}))=\sum_{i=1}^{k}[\theta_{i,1}^{k} \Phi(\bm{\mathcal{M}}_{i,1})+ \cdots + \theta_{i,s}^{k} \Phi(\bm{\mathcal{M}}_{i,s})]=$$ $$=\sum_{i=1}^{k}\sum_{j=1}^{s}\theta_{i,j}^{k} \Phi(\bm{\mathcal{M}}_{i,j}).$$ In the ELRAP4TS version we use $$\bm{\mathcal{X}}_{k}=\alpha_0^k X_{k-1} + \alpha_1^k \Phi(\bm{\mathcal{M}}_{k,1} )+\cdots+\alpha_s^k \Phi(\bm{\mathcal{M}}_{k,s} ).$$ These three steps are performed iteratively until our stopping criterion is reached. The Flow charts in Figures \[Figure 2\] and \[Figure 3\] illustrate the process for $r=4$ and $s=2$. Both of our methods find a rank-4 tensor as the approximate solution. For Problem (\[1.2\]) we need just 2 iterations in LRAP4TS and in ELRAP4TS. Figure \[Figure 2\] shows the process of LRAP4TS: the red arrows represent the first iteration step and the blue arrows illustrate the second step. After two iterations LRAP4TS has computed 4 basis tensors and 4 coefficients (or weights), as well as the rank-4 tensor $\widehat{\bm{\mathcal{Y}}}=\theta_{1,1}^2 \bm{\mathcal{M}}_1 + \theta_{1,2}^2 \bm{\mathcal{M}}_2 + \theta_{1,3}^2 \bm{\mathcal{M}}_3 +\theta_{1,4}^2 \bm{\mathcal{M}}_4 $ (when $r=4$). If we were to continue iterating then in the $h$-th iterations LRAP4TS algorithm would have to handle 2$h$ basis tensors and $2h$ weights. This data growth induces expanded storage requirements for these tensors and it increases the computational complexity. Figure \[Figure 3\] illustrates that ELRAP4TS builds only 3 basis tensors and computes only 3 coefficients in each iteration which is a slight improvement over LRAP4TS. ; (R1) –(M1); (R1) –(M2); (M1) –(LS1); (M2) –(LS1); (LS1) –(theta111); (LS1) –(theta121); (LS1) –(X1); (X1) –(R2); (R2) –(M3); (R2) –(M4); (M1) –(LS2); (M2) –(LS2); (M3) –(LS2); (M4) –(LS2); (LS2) –(theta112); (LS2) –(theta122); (LS2) –(theta132); (LS2) –(theta142); (LS2) –(X2); (X2) –(R3); (R3) –(Space1); (R3) –(Space2); ; (R1) –(M1); (R1) –(M2); (X0) –(LS1); (M1) –(LS1); (M2) –(LS1); (LS1) –(theta101); (LS1) –(theta111); (LS1) –(theta121); (LS1) –(X1); (X1) –(R2); (R2) –(M3); (R2) –(M4); (X1) –(LS2); (M3) –(LS2); (M4) –(LS2); (LS2) –(theta102); (LS2) –(theta112); (LS2) –(theta122); (LS2) –(X2); (X2) –(R3); (R3) –(Space1); (R3) –(Space2); TRIP ==== \[definition 1\][@ZWHWW2019] **(TRIP).** Let the linear operator $\Phi : \mathbb{R}^{n_1 \times n_2 \times n_3} \rightarrow \mathbb{R}^m$ be a linear map on the linear space of tensors of size $n_1 \times n_2 \times n_3$ with $n_1 \leq n_2$. Then for the t-SVD decomposition and every integer $r$ with $1 \leq r \leq n_1$, the TRIP constant $\delta_r$ of $\Phi$ is the smallest quantity such that $$(1-\delta_r) ||\bm{\mathcal{X}}||^2_F \leq ||\Phi(\bm{\mathcal{X}})||^2_2 \leq (1+\delta_r) ||\bm{\mathcal{X}}||^2_F$$ for all tensors $\bm{\mathcal{X}} \in \mathbb{R}^{n_1 \times n_2 \times n_3}$ of rank at most $r$. \[Theorem 2\] For $\delta, \epsilon \in (0,1)$, a random draw of an L-subgaussian measurement ensemble $\Phi : \mathbb{R}^{n_1 \times n_2 \times n_3} \rightarrow \mathbb{R}^m$ satisfies $\delta_r \leq \delta$ with probability at least $1-\epsilon$, provided that $$m \geq C \alpha^2 \delta^{-2} \max \{(r \cdot r \cdot n_3+ n_1 \cdot r \cdot n_3 + n_2 \cdot r \cdot n_3) \log(2), \log( \eta^{-1}) \},$$ where $r$ is tensor tubal rank. The constants $C > 0$ only depend on the subgaussian parameter $L$. To prepare for the proof of Theorem \[Theorem 2\], we state several useful Lemmas. The proof of Theorem \[Theorem 2\] also uses $\epsilon$-nets and covering numbers, see [@V2012] for background on these topics. \[Definition 2\][@V2012] **(Nets, covering numbers)** A set $\mathscr{N}^{\mathscr{X}}_{\epsilon} \subset \mathscr{X}$ with $\mathscr{X}$ a subset of a normed space is called an $\epsilon$-net of $\mathscr{X}$ with respect to the norm $||\cdot ||$ if for each $\mathbf v \in \mathscr{X}$ there exists $\mathbf v_0 \in \mathscr{N}^{\mathscr{X}}_{\epsilon}$ with $||\mathbf v_0 - \mathbf v || \leq \epsilon$. The minimal cardinality of an $\epsilon$-net of $\mathscr{X}$ with respect to the norm $||\cdot ||$ is denoted by $\mathscr{N} (\mathscr{X}, ||\cdot ||, \epsilon) $ and called the covering number of $\mathscr{X}$ (at scale $\epsilon$). Equivalently, $\mathscr{N} (\mathscr{X}, ||\cdot ||, \epsilon) $ is the minimal number of balls with radii $\epsilon$ and with centers in $\mathscr{X}$ needed to cover $\mathscr{X}$. The following is well-known [@V2012; @RSS2017] and will be used frequently in what follows. \[Lemma 1\][@V2012] **(Covering numbers of the sphere)** Let $\mathscr{X}$ be a subset of a vector space of real dimension $k$ with norm $||\cdot ||$, and let $0 < \epsilon < 1$. Then there exists an $\epsilon$-net $\mathscr{N}^{\mathscr{X}}_{\epsilon} \subset \mathscr{X}$ with $$|\mathscr{N}^{\mathscr{X}}_{\epsilon}| \leq {Vol(\mathscr{X} + {\epsilon \over 2} \mathscr{B}) \over Vol({\epsilon \over 2} \mathscr{B})},$$ where ${\epsilon \over 2} \mathscr{B}$ is an $\epsilon / 2$ ball with respect to the norm $||\cdot ||$ and $$\mathscr{X} + {\epsilon \over 2} \mathscr{B} = \{\mathbf x+\mathbf y : \mathbf x \in \mathscr{X}, \mathbf y \in {\epsilon \over 2} \mathscr{B}\}.$$ Specifically if $\mathscr{X}$ is a subset of the $||\cdot ||$-unit ball then $\mathscr{X} + {\epsilon \over 2} \mathscr{B}$ is contained in the $(1+{\epsilon \over 2})$-ball and thus $$|\mathscr{N}^{\mathscr{X}}_{\epsilon}| \leq {(1 + \epsilon / 2)^k \over (\epsilon / 2)^k} = (1 + {2 \over \epsilon})^k < (3 / \epsilon)^k.$$ The covering number Lemmas are critical for the proof of Theorem \[Theorem 2\] where we need to calculate the covering number for the set of rank $r$ tensors with unit Frobenius norm. \[Lemma 2\] (**Covering numbers related to the t-SVD**). The covering numbers of $$\mathscr{S}_r = \{\bm{\mathcal{X}} \in \mathbb{R}^{n_1 \times n_2 \times n_3} : {\rm rank_t} (\bm{\mathcal{X}}) \leq r, ||\bm{\mathcal{X}}||_F = 1 \}$$ with respect to the Frobenius norm are bounded by $$\label{47} \mathscr{N} (\mathscr{S}_r, ||\cdot ||_F, \epsilon) \leq (9 / \epsilon)^{r \cdot r \cdot n_3+ n_1 \cdot r \cdot n_3 + n_2 \cdot r \cdot n_3}.$$ The proof follows a same strategy as the Lemma 3.1 of [@CP2011] and Lemma 3 of [@RSS2017]. The (truncating) t-SVD decomposition $\bm{\mathcal{X}} = \bm{\mathcal{U}} * \bm{\mathcal{S}} * \bm{\mathcal{V}}^T$ of any $\bm{\mathcal{X}} \in \mathscr{S}_r$ obeys $|| \bm{\mathcal{S}}||_F = 1$ ($ \bm{\mathcal{S}}$ is a f-diagonal tensor with singular tubes on the diagonal, and $ \bm{\mathcal{U}}$ and $\bm{\mathcal{V}}$ are partially orthogonal tensors of left- and right-singular vectors of tubes). We constructs an $\epsilon$-net for $ \mathscr{S}_r$ by covering the sets of tensors $ \bm{\mathcal{U}}$, $ \bm{\mathcal{V}}$ with partially orthogonal lateral slices and the set of unit Frobenius tensors $ \bm{\mathcal{S}}$. Let $\mathscr{D}$ be the set of f-diagonal tensors $\bm{\mathcal{X}} \in \mathbb{R}^{r \times r \times n_3}$ with unit Frobenius norm, which is included in $\mathscr{F} = \{ \bm{\mathcal{X}} \in \mathbb{R}^{r \times r \times n_3} : ||\bm{\mathcal{X}} ||_F = 1\}$. Hence Lemma \[Lemma 1\] gives an $\epsilon / 3$-net $\mathscr{N}^{\mathscr{F}}_{\epsilon / (d + 1)}$ in respect of the Frobenius norm of cardinality $$|\mathscr{N}^{\mathscr{F}}_{\epsilon / (d + 1)} | \leq (9 /\epsilon)^{r \cdot r \cdot n_3}.$$ For covering $\mathscr{O}_{n_1,r,n_3} = \{\bm{\mathcal{U}} \in \mathbb{R}^{n_1 \times r \times n_3}: \bm{\mathcal{U}}^T * \bm{\mathcal{U}} = \bm{\mathcal{I}} \}$ and $\mathscr{O}_{n_2,r,n_3} = \{\bm{\mathcal{V}} \in \mathbb{R}^{n_2 \times r \times n_3}: \bm{\mathcal{V}}^T * \bm{\mathcal{V}} = \bm{\mathcal{I}} \}$, it is crucial to use the norm $|| \cdot ||_{1,F}$, which is defined as $$|| \bm{\mathcal{Y}} ||_{1,F} = \max_{i,j} || \hat{\bm{\mathcal{Y}}} (:,j,k) ||_F,$$ where $\hat{\bm{\mathcal{Y}}} (:,j,k)$ denotes the $j,k$-th tube fiber in $\hat{\bm{\mathcal{Y}}}$ from $\bm{\mathcal{Y}}$ using the DFT. Obviously, we have that $ || \bm{\mathcal{Y}} ||_{1,F}=|| \overline{\mathbf Y} ||_{1,F}.$ Because the elements of $\mathscr{O}_{n_1,r,n_3}$ have normed lateral slices, it holds $\mathscr{O}_{n_1,r,n_3} \subset \mathscr{Q}_{n_1,r,n_3} = \{ \bm{\mathcal{Y}} \in \mathbb{R}^{n_1 \times r \times n_3}: || \bm{\mathcal{Y}} ||_{1,F} \leq 1 \}$. Lemma \[Lemma 1\] provides $$\mathscr{N} (\mathscr{O}_{n_1,r,n_3}, ||\cdot ||_{1,F}, \epsilon / 3) \leq 9 / \epsilon)^{n_1 \cdot r \cdot n_3 },$$ i.e. there exists an $ \epsilon / 3$-net $\mathscr{N}^{\mathscr{O}_{n_1,r,n_3}}_{\epsilon / 3}$ of this cardinality. Then the set $$\mathscr{N}^{\mathscr{S}_r}_{\epsilon} := \{ \tilde{\bm{\mathcal{U}}} * \tilde{\bm{\mathcal{S}}} * \tilde{\bm{\mathcal{V}}}^T : \tilde{\bm{\mathcal{S}}} \in \mathscr{N}^{\mathscr{D}}_{\epsilon / 3}, \ \tilde{\bm{\mathcal{U}}} \in \mathscr{N}^{\mathscr{O}_{n_1,r,n_3}}_{\epsilon / 3},$$ $$\ and \ \ \tilde{\bm{\mathcal{V}}} \in \mathscr{N}^{\mathscr{O}_{n_2,r,n_3}}_{\epsilon / 3} \}$$ obeys $$|\mathscr{N}^{\mathscr{S}_r}_{\epsilon} | \leq \mathscr{N} (\mathscr{D}, ||\cdot ||_{F}, \epsilon / 3) \cdot \mathscr{N} (\mathscr{O}_{n_1,r,n_3}, ||\cdot ||_{1,F}, \epsilon / 3) \cdot$$ $$\cdot \mathscr{N} (\mathscr{O}_{n_2,r,n_3}, ||\cdot ||_{1,F}, \epsilon / 3) \leq (9 / \epsilon)^{r \cdot r \cdot n_3+ n_1 \cdot r \cdot n_3 + n_2 \cdot r \cdot n_3}.$$ It remains to show that $\mathscr{N}^{\mathscr{S}_r}_{\epsilon}$ is an $\epsilon$-net for $\mathscr{S}_r$, i.e. that for all $\bm{\mathcal{X}} \in \mathscr{S}_r$ there exists $\tilde{\bm{\mathcal{X}}} \in\mathscr{N}^{\mathscr{S}_r}_{\epsilon}$ with $||\bm{\mathcal{X}} - \tilde{\bm{\mathcal{X}}}||_F \leq \epsilon$. At last, we fix $\bm{\mathcal{X}} \in \mathscr{S}_r$ and decompose $\bm{\mathcal{X}}$ as $\bm{\mathcal{X}} = \bm{\mathcal{U}} * \bm{\mathcal{S}} * \bm{\mathcal{V}}^T$. Then there exists $\tilde{\bm{\mathcal{X}}} = \tilde{\bm{\mathcal{U}}} * \tilde{\bm{\mathcal{S}}} * \tilde{\bm{\mathcal{V}}}^T \in \mathscr{N}^{\mathscr{S}_r}_{\epsilon}$ with $\tilde{\bm{\mathcal{U}}} \in \mathscr{N}^{\mathscr{O}_{n_1,r,n_3}}_{\epsilon / 3}$, $\tilde{\bm{\mathcal{V}}} \in \mathscr{N}^{\mathscr{O}_{n_2,r,n_3}}_{\epsilon / 3}$ and $\tilde{\bm{\mathcal{S}}} \in \mathscr{N}^{\mathscr{D}}_{\epsilon / 3}$ obeying $||\bm{\mathcal{U}} - \tilde{\bm{\mathcal{U}}}||_{1,F} \leq \epsilon / 3$, $||\bm{\mathcal{V}} - \tilde{\bm{\mathcal{V}}}||_{1,F} \leq \epsilon / 3$ and $||\bm{\mathcal{S}} - \tilde{\bm{\mathcal{S}}}||_F \leq \epsilon / 3$. This gives $$\label{48} \begin{array}{lll} ||\bm{\mathcal{X}} - \tilde{\bm{\mathcal{X}}}||_F = || \bm{\mathcal{U}} * \bm{\mathcal{S}} * \bm{\mathcal{V}}^T- \tilde{\bm{\mathcal{U}}} * \tilde{\bm{\mathcal{S}}} * \tilde{\bm{\mathcal{V}}}^T||_F \\ \ \ \ \ \leq || (\bm{\mathcal{U}} - \tilde{\bm{\mathcal{U}}}) * \bm{\mathcal{S}} * \bm{\mathcal{V}}^T||_F + || \tilde{\bm{\mathcal{U}}} *( \bm{\mathcal{S}} - \tilde{\bm{\mathcal{S}}}) * \bm{\mathcal{V}}^T||_F \\ \ \ \ \ \ \ \ + ||\tilde{\bm{\mathcal{U}}} * \tilde{\bm{\mathcal{S}}} * (\bm{\mathcal{V}} - \tilde{\bm{\mathcal{V}}})^T||_F \end{array}$$ For the first term, since $\bm{\mathcal{V}}$ is an partially orthogonal tensor and have unitary invariance [@KM2011], $ || (\bm{\mathcal{U}} - \tilde{\bm{\mathcal{U}}}) * \bm{\mathcal{S}} * \bm{\mathcal{V}}^T||_F = || (\bm{\mathcal{U}} - \tilde{\bm{\mathcal{U}}}) * \bm{\mathcal{S}}||_F$, and $$\begin{array}{lll} || (\bm{\mathcal{U}} - \tilde{\bm{\mathcal{U}}}) * \bm{\mathcal{S}}||^2_F =|| \bm{\mathcal{C}} * \bm{\mathcal{S}}||^2_F \\ \qquad = {1 \over n_3}|| \overline{\mathbf C} * \overline{\mathbf S}||^2_F \leq {1 \over n_3}||\overline{\mathbf S}||^2_F \cdot || \overline{\mathbf C} ||_{1,F} \\ \qquad = ||\bm{\mathcal{S}}||^2_F \cdot || \overline{\mathbf C} ||_{1,F} = ||\bm{\mathcal{S}}||^2_F \cdot || \bm{\mathcal{C}} ||_{1,F} \\ \qquad = ||\bm{\mathcal{S}}||^2_F \cdot || (\bm{\mathcal{U}} - \tilde{\bm{\mathcal{U}}}) ||^2_{1,F} \\ \qquad \leq (\epsilon / 3)^2, \end{array}$$ where $\bm{\mathcal{C}} = (\bm{\mathcal{U}} - \tilde{\bm{\mathcal{U}}})$. Hence, $|| (\bm{\mathcal{U}} - \tilde{\bm{\mathcal{U}}}) * \bm{\mathcal{S}} * \bm{\mathcal{V}}^T||_F \leq \epsilon / 3$. The same argument gives $||\tilde{\bm{\mathcal{U}}} * \tilde{\bm{\mathcal{S}}} * (\bm{\mathcal{V}} - \tilde{\bm{\mathcal{V}}})^T||_F \leq \epsilon / 3$. To bound the middle term, observe that $|| \tilde{\bm{\mathcal{U}}} *( \bm{\mathcal{S}} - \tilde{\bm{\mathcal{S}}}) * \bm{\mathcal{V}}^T||_F = || \bm{\mathcal{S}} - \tilde{\bm{\mathcal{S}}}||_F \leq \epsilon / 3$. Therefore, we have the desired result. The proof of Theorem \[Theorem 2\] also requires the following result from Corollary 5.4 of [@D2016]. \[Corollary 5.4\] [@D2016] Let $\mathscr{S}_{(1)}, \cdots , \mathscr{S}_{(k)}$ be subsets of a Hilbert space $H$ and let $\mathscr{S} = \cup^k_{i=1} \mathscr{S}_{(i)}$. Set $$\mathscr{S}_{(i),nv} = \{\mathbf x / ||\mathbf x ||_2 : \mathbf x \in \mathscr{S}_{(i)} \}.$$ suppose that $\mathscr{S}_{(i),nv}$ has covering dimension $K_i$ with parameter $c_i$ and base covering $\mathscr{N}_{0,(i)}$ with respect to $d_H$. Set $K = \max_i K_i, c = \max_i c_i$ and $\mathscr{N}_0 = \max_i \mathscr{N}_{0,(i)}$. Let $\Phi : \Omega \times H \rightarrow \mathbb{R}^m$ be a subgaussian map on $\mathscr{S}_{nv}$. Then, for any $0 < \delta, \eta < 1$, we have $\mathbb{P}(\delta_{\mathscr{S},\Phi} \leq \eta)$, provided that $$m \geq C \alpha^2 \delta^{-2} max \{\log k + \log \mathscr{N}_0 + K \log(c), \log( \eta^{-1)} \}.$$ Proof of Theorem \[Theorem 2\]. The proof follows the same strategy as that of Example 5.8 of [@D2016]. We consider the Frobenius inner product $ \langle \bm{\mathcal{X}}, \bm{\mathcal{Y}} \rangle$, the corresponding norm $||\bm{\mathcal{X}}||_F = \langle \bm{\mathcal{X}}, \bm{\mathcal{Y}} \rangle^2$ and the induced metric $d_F(\bm{\mathcal{X}}, \bm{\mathcal{Y}}) = ||\bm{\mathcal{X}} - \bm{\mathcal{Y}}||_F$. As said earlier, $$\mathscr{S}_r = \{\bm{\mathcal{X}} \in \mathbb{R}^{n_1 \times n_2 \times n_3} : {\rm rank_t} (\bm{\mathcal{X}}) \leq r, ||\bm{\mathcal{X}}||_F = 1 \},$$ then $\delta_r = \delta_{\mathscr{S}_r, \Phi} $. It is shown in Lemma [\[Lemma 2\]]{} that for any $0 < \epsilon \leq 1$, the covering number is $$\mathscr{N} (\mathscr{S}_r, ||\cdot ||_F, \epsilon) \leq (3(2+1) / \epsilon)^{r \cdot r \cdot n_3+ n_1 \cdot r \cdot n_3 + n_2 \cdot r \cdot n_3}.$$ In other words, $\mathscr{S}_r $ has covering dimension $K = r \cdot r \cdot n_3+ n_1 \cdot r \cdot n_3 + n_2 \cdot r \cdot n_3$ with parameter $c = 9$. Corollary \[Corollary 5.4\] means that for any subgaussian map $\Phi$ and $0< \delta, \eta < 1$, we have $\mathbb{P}(\delta_r \geq \delta) \leq \eta$, provided that $$m \geq C \alpha^2 \delta^{-2} \max \{(r \cdot r \cdot n_3+ n_1 \cdot r \cdot n_3 + n_2 \cdot r \cdot n_3) \log(2), \log( \eta^{-1}) \}.$$ This completes the proof. $\Box$ Convergence Analysis ==================== We will proof that Algorithm \[alg:Framwork\] converges linearly in this section. This is shown in Theorem \[Theorem 3.1\]. \[Theorem 3.1\] Algorithm \[alg:Framwork\] in the standard LRAP4TS and economic LRAP4TS versions both have the linear convergence rate of $$\left\|\bm{\mathcal{R}}_k\right\| \leqslant \left( {\sqrt{1-\frac{1}{\min(m,n)}}} \right)^{k-1} \cdot \left\|\Phi^{-1}(\mathbf{b})\right\| \text{ for all } k\geqslant 1,$$ where $\mathbf{b}=\Phi(\bm{\mathcal{Y}})=\phi \cdot vec(\bm{\mathcal{Y}})$. This holds for all tensors $\bm{\mathcal{Y}}$ of rank at most $r$. First we give that both LRAP4TS and ELRAP4TS satisfy the following inequality: $$\|\bm{\mathcal{R}}_{k+1}\|^2 \leqslant \|\bm{\mathcal{R}}_{k}\|^2-\langle {\bm{\mathcal{M}}_{k,1}, \bm{\mathcal{R}}_k} \rangle^2.$$ For the LRAP4TC algorithm we have $$\begin{array}{lll} \|\bm{\mathcal{R}}_{k+1} \|^2 &=& \mathop{\min}\limits_{\bm{\theta} \in \mathbb{R}^{sk}}\| \Phi^{-1} \Phi(\bm{\mathcal{Y}})-\sum_{i=1}^{k}\sum_{j=1}^{s}\theta_{i,j} \Phi^{-1} \Phi(\bm{\mathcal{M}}_{i,j}) \|^2 \\ &\leqslant& \mathop{\min}\limits_{\theta_{k,1}\in \mathbb{R}}\| \Phi^{-1} \Phi(\bm{\mathcal{Y}})- \bm{\mathcal{X}}_{k-1}- \theta_{k,1} \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1}) \|^2 \\ &=& \mathop{\min}\limits_{\theta_{k,1}\in \mathbb{R}}\| \bm{\mathcal{R}}_k- \theta_{k,1} \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1}) \|^2. \\ \end{array}$$ And for the ELRAP4TS algorithm we have $$\begin{array}{lll} \|\bm{\mathcal{R}}_{k+1} \|^2 = \\ = \mathop{\min}\limits_{\bm{\alpha}\in \mathbb{R}^{s+1}}\| \Phi(\bm{\mathcal{Y}})- \alpha_0 \bm{\mathcal{X}}_{k-1} - \alpha_1 \Phi(\bm{\mathcal{M}}_{k,1}) -\cdots-\alpha_s \Phi(\bm{\mathcal{M}}_{k,s}) \|_\Omega^2 \\ \leqslant \mathop{\min}\limits_{\alpha_1\in \mathbb{R}}\| \Phi^{-1} \Phi(\bm{\mathcal{Y}})- \bm{\mathcal{X}}_{k-1} - \alpha_1 \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1}) \|^2 \\ = \mathop{\min}\limits_{\alpha_1\in \mathbb{R}}\|\bm{\mathcal{R}}_{k} - \alpha_1 \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1}) \|^2 . \\ \end{array}$$ In both cases we get closed form solutions for $\theta_{k,1}^*=\frac{\langle \bm{\mathcal{R}}_{k},\Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1})\rangle}{\langle \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1}), \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1})\rangle}=\frac{\langle \bm{\mathcal{R}}_{k}, \bm{\mathcal{M}}_{k,1}\rangle}{\langle \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1}), \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1})\rangle}$ and $\alpha_1^*=\frac{\langle \bm{\mathcal{R}}_{k}, \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1})\rangle}{\langle \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1}), \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1})\rangle}=\frac{\langle \bm{\mathcal{R}}_{k}, \bm{\mathcal{M}}_{k,1}\rangle}{\langle \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1}), \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1})\rangle}$. Plugging the optima $\theta_{k,1}^*$ and $\alpha_1^*$ back into the above formulas, we get $$\label{2.7} \begin{array}{lll} \|\bm{\mathcal{R}}_{k+1} \|^2 &\leqslant&\| \bm{\mathcal{R}}_k-\frac{\langle \bm{\mathcal{R}}_{k}, \bm{\mathcal{M}}_{k,1}\rangle}{\langle \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1}), \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1})\rangle} \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1}) \|^2 \\ &=& \| \bm{\mathcal{R}}_k\|^2-\frac{\langle \bm{\mathcal{R}}_{k}, \bm{\mathcal{M}}_{k,1}\rangle^2}{\langle \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1}), \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1})\rangle} \\ &=& \| \bm{\mathcal{R}}_k\|^2-\langle \bm{\mathcal{R}}_{k}, \bm{\mathcal{M}}_{k,1}\rangle^2,\\ \end{array}$$ since $\langle \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1}), \Phi^{-1} \Phi(\bm{\mathcal{M}}_{k,1})\rangle \leqslant 1$ from $\Phi \Phi^{-1}$ is an identity operator and $\Phi = \phi \cdot vec(\cdot)$. Using inequality (\[2.7\]) and Lemma \[Lemma 2.2\], we obtain that $$\|\bm{\mathcal{R}}_{k+1}\|^2\leqslant \|\bm{\mathcal{R}}_{k}\|^2-\langle \bm{\mathcal{R}}_{k}, \bm{\mathcal{M}}_{k,1}\rangle^2 \leqslant \left(1-\frac{1}{min(m,n)}\right)\|\bm{\mathcal{R}}_{k}\|^2.$$ This completes the proof. $\Box$ Based on Theorem \[Theorem 2\] some linear operators $\Phi$ satisfies the TRIP condition with high probability. By assuming that the TRIP condition holds, we prove the following approximation result. \[Theorem B1\] Let $\bm{\mathcal{Y}}$ be a tensor of tensor tubal rank $r$. Suppose the measurement mapping $\Phi(\bm{\mathcal{X}})$ satisfies TRIP for rank-$r_0$ with $\delta_{r_0}=\delta_{r_0}(\Phi)<1$ with $r_0\geqslant 2r$. The output tensor $\bm{\mathcal{M}}(\bm{\theta}^k)$ approximates the exact tensor $\bm{\mathcal{Y}}$ in the following sense: there is a positive constant $\tau$ such that $$\|\bm{\mathcal{M}}(\bm{\theta}^k)-Y\|_F\leqslant \frac{C}{\sqrt{1-\delta_{r_0}}} \tau^k$$ for all $k=1,\cdots,\lfloor{(r_0-r)/s}\rfloor$, where $\lfloor{(r_0-r)/s}\rfloor$ denotes the largest inter less than or equal to $r/s$ and $C>0$ is a constant depending on $\Phi$. Based on the definition of $\delta_{r_0}$, for $s \red{\cdot} k+r\leqslant r_0$, we have $$\begin{array}{lll} (1-\delta_{r_0})\|\bm{\mathcal{M}}(\bm{\theta}^k)-\bm{\mathcal{Y}}\|^2_F &\leqslant& \|\Phi(\bm{\mathcal{M}}(\bm{\theta}^k))-\Phi(\bm{\mathcal{Y}})\|^2_2 \\ &=& \|\Phi(\bf{\mathcal{R}}_k)\|^2_2=\|\phi \cdot vec(\bm{\mathcal{R}}_k)\|^2_2\\ &\leqslant&\|\phi\|^2_2\|vec(\bm{\mathcal{R}}_k)\|^2_2=\|\phi\|^2_2\|\bm{\mathcal{R}}_k\|^2_F\\ &\leqslant& \|\phi \|^2_2\tau^{2k}\|\Phi^{-1}(\mathbf{b})\|^2_F \end{array}$$ where $\tau=\sqrt{1-\frac{1}{\min(m,n)}}$ from Theorem \[Theorem 3.1\]. So we have $$\|\bm{\mathcal{M}}(\bm{\theta}_k)-\bm{\mathcal{Y}}\|^2_F \leqslant\frac{ \|\phi \|^2_2\tau^{2k}}{1-\delta_{r_0}}\|\Phi^{-1}(\mathbf{b})\|^2_F.$$ Therefore, we have the desired result. $\Box$ The above convergence result require $s \cdot k+r\leqslant r_0$, which guarantees the TRIP condition for all estimated tensors during the iteration process. Numerical Tests and Applications ================================= All our methods have been implemented for tensor completion in MATLAB. All test experiments were run on a Macbook Pro with OSX High Sierra, an Intel Core i5 2.6 GHz processor and with 8G RAM. We have noted that the speed of the RAM is very important for Matlab speeds. Our experimental results are compared with respect to the root-mean-square error (RMSE), defined as $RMSE = \sqrt{|| \bm{\mathcal{X}} - \bm{\mathcal{Y}} ||^2_F / (n_1 n_2 n_3) }$. The proposed ELRAP4TC method reduces the CPU time significantly from that of LRAP4TC with a small decrease in quality. The test results with ELRAP4TC are only included for clarity. The t-SVD algorithm to find the truncated t-SVD in the ELRAP4TC algorithm. For further speed up significantly, readers might adopt the rt-SVD method instead. Video Recovery --------------- A video clip, shot with a 1/50 sec or faster shutter speed every 1/25 sec contains 25 digital data frames for every second of real time. In our video recovery example below, we use a $144 \times 256 \times 40$ black and white gray-scale “basketball” video. Our video clip depicts 1.6 sec of a game and contains 40 digital images in avi format. Each frame holds 144 by 256 pixels of raw sensor data. To visualize, consider these data frames vertically, 40 in total, one behind the other along the time line, capturing 1.6 second of time. This visualization helps us to interpret the data as a 3-dimensional tensor. Such a video generated tensor will have low tubal rank because if one were to look perpendicularly to the video image data planes in the direction of time over a small area of the images themselves. Then one would (locally) see almost identical images with little changes in any one small tube. Looking at the same small cut out area for all 40 consecutive video frames, there will generally be only small changes because the angle of view does not change all that quickly, fixed objects do not move, people do not run that fast, not in a low resolution video and not over one second or two. Hence if some of the 40 frames are missing or damaged – as we assume – we would start from a low tubal rank tensor and try to reconstruct the video on that premise. We have set rank $r=100$ in our first experiment, i.e., we retain $50\%$ of the given pixels as known entries in $\Omega$. The recovery of the damaged video has thus become a tensor completion problem that we can solve with our ELRAP4TC algorithm. Since ELRAP4TC is always applied with $s \leqslant r$ we work with $s=1, 2, 3$ here. In Figure \[Figure 4\] the convergence characteristics are shown for the ELRAP4TC algorithm. The running times decrease in the ELRAP4TC method for increasing values of $s$. ![Linear convergence of ELRAP4TC with $s=1, 2, 3$ (on the left) and run times in seconds (on the right)[]{data-label="Figure 4"}](RMSE_ELRAP.eps "fig:") ![Linear convergence of ELRAP4TC with $s=1, 2, 3$ (on the left) and run times in seconds (on the right)[]{data-label="Figure 4"}](Time_ELRAP.eps "fig:") To evaluate our LRAP4TC we compare it with two state-of-the art algorithms such as HoMP [@YMS2015] and ADMM with the t-SVD as subroutine (ADMM-t-SVD) [@ZA2017]. The two previous algorithm own the theoretical recovery guarantee. For ADMM-t-SVD, the parameter $\lambda$ is set to $\lambda = 1/ \sqrt{3 \max{(n_1,n_2)}}$. For HoMP, we empirically set $r =100$ which is the same with our algorithm. From the numerical experiments in Section V of [@YMS2015], the HoMP algorithm is the fastest state-of-the-art algorithms four years ago for the tensor completion problem. The comparison result is shown in Figure \[Figure 5\]. ![The 30th frame of completion result for a basketball video.[]{data-label="Figure 5"}](Compares.eps) We measure the running time and RMSE for all three methods. Running time and RMSE are listed in Table \[Table 2\]. From these results, we show the following observations. First, these experiments show obviously that the ELRAP4TC algorithm is overall the fastest methods that offers satisfactory results. Second, ELRAP4TC outperforms HoMP in terms of their RMSE in Table \[Table 2\]. Third, ADMM-t-SVD demands the highest cost in this experiment to the best RMSE result. These not only demonstrates the superiority of our ELRAP4TC, but also validate our recovery guarantee in Theorem \[Theorem 3.1\] on video data. Completion Approach Running Time RMSE --------------------- -------------- ------------- ADMM-t-SVD 124.76 **11.7117** HoMP 38.40 29.6814 ELRAP4TC (s=1) 17.55 20.2427 ELRAP4TC (s=3) **9.61** 20.8546 : *Runing time and RMSE of tensor completion result on the basketball video*[]{data-label="Table 2"} In the second case, we will explore the performance of our algorithm with a variety of specified missing ratio from $30\%$ to $90\%$. For each missing ratio, we test all algorithms 50 times and get the mean of their running time as vertical axis in Figure \[Figure 6\] or their RMSE as vertical axis in Figure \[Figure 7\]. The experiment setup is the same with the first experiment. From Figure \[Figure 6\] and Figure \[Figure 7\], we can observe that the performance of all algorithms, for each specified missing ratio, are the almost same with the first experiment: ELRAP4TC is the fast method and outperforms HoMP to obtain the better accuracy solution. ![The trends of running times with different missing ratio.[]{data-label="Figure 6"}](Ratio_Time.eps) ![The trends of the RMSE values with different missing ratio.[]{data-label="Figure 7"}](Ratio_RMSE.eps) In the third experiment, we give recovery performance comparison of all algorithm with a variety of frontal slice (video frame) number from $2$ to $40$. The experiment setup is the same with the first experiment. We test all algorithms 50 times, for each frontal slice number, and obtain the mean of their running time as vertical axis in Figure \[Figure 10\] or their RMSE as vertical axis in Figure \[Figure 11\]. Figures \[Figure 10\] and \[Figure 11\] show that the performances of all algorithms and each specified frontal slice generally are nearly the same as for our first experiment: namely ELRAP4TC is the fastest overall and it outperforms HoMP with better solutions for frontal slice numbers above 8. ![The trends of running times following different number of frontal slice.[]{data-label="Figure 10"}](Frontal_Time.eps) ![The trends of the RMSE values following different number of frontal slice.[]{data-label="Figure 11"}](Frontal_RMSE.eps) Color Image Recovery --------------------- In this subsection, we consider tensor completion based on the color image to test the stability of our algorithms. We format a $n_1 \times n_2$ sized color image as a tensor of size $n_1 \times n_2 \times 3$. Here the Matlab function, ${\rm imnoise}$, is used to generate blurring noise to the image and also add Gaussian noise with a mean zero and a standard deviation $\sigma=5e-3$. We will present that the recovery proformance of ELRAPTC is still satisfactory. 80 color images are used for the test from the Berkeley Segmentation Dataset [@MFTM2001]. The sizes of images are $481 \times 321$ or $321 \times 481$. For each image, we remain $50 \%$ of the given pixels as known entries in $\Omega$ and set the desired minimal rank $r=100$. See Figure \[Figure 9\] (b) for some sample images with noises. We compare our ELRAP4TC with ADMM-t-SVD and HoMP. Figure \[Figure 8\] gives the comparison of running time and RMSE on all 80 images. Some examples with the recovered images are represented in Figure \[Figure 9\]. Based on these results, we have the observation that our ELRAP4TC is overall the fast methods to obtain reasonable solution in the presence of noise. ![image](Runtime_RMSE.eps) ![image](Compares_Images1.eps) [|c|c|c|c|c|]{} Index & ADMM-t-SVD & HoMP & ---------- ELRAP4TC (s=1) ---------- : *Comparison of running time on the 5 images in Figure \[Figure 9\]*[]{data-label="Table 3"} & ---------- ELRAP4TC (s=3) ---------- : *Comparison of running time on the 5 images in Figure \[Figure 9\]*[]{data-label="Table 3"} \ 1 & 109.80& 22.25& 7.76& **4.89**\ 2 & 94.21& 17.93& 7.24& **4.50**\ 3 & 96.22& 18.09& 7.31& **4.03**\ 4 & 97.36& 18.28& 7.23& **4.35**\ 5 & 129.56& 17.93& 7.17& **4.41**\ conclusion ========== In this context, an efficient and scalable algorithms are proposed for tensor completion and tensor sensing. In order to obtain the convergence of them, we define a new TRIP condition which is based on t-SVD. We show that subgaussian measurement ensemble satisfy the TRIP condition with high probability under the optimal bound on the number of measurements. Using this result, we present that both algorithms perform linear convergence rate. Numerical experiments on real datas are contained that show the accuracy and efficiency of our algorithms. [^1]: This work was supported by the National Natural Science Foundation of China under Grant 11801418. [^2]: The author is with the College of Mathematics and Physics, Wenzhou University, Zhejiang 325035, China (e-mail: [email protected]).
{ "pile_set_name": "ArXiv" }
--- abstract: | By exploiting the database of early-type galaxies (ETGs) members of the $WINGS$ survey of nearby clusters, we address here the long debated question of the origin and shape of the Fundamental Plane (FP). Our data suggest that different physical mechanisms concur in shaping and [*’tilting’*]{} the FP with respect to the virial plane (VP) expectation. In particular, an “hybrid solution” in which the structure of galaxies and their stellar population are the main contributors to the FP [*tilt*]{} seems to be favoured. We find that the bulk of the [*tilt*]{} should be attributed to structural non-homology, while stellar population effects play an important but less crucial role. In addition, our data indicate that the differential FP [*tilt*]{} between the $V$- and $K$-band is due to a sort of entanglement between structural and stellar population effects, for which the inward steepening of color profiles ($V-K$) tends to increase at increasing the stellar mass of ETGs. The same kind of analysis applied to the $ATLAS3D$ and $SDSS$ data in common with $WINGS$ ($WSDSS$ throughout the paper) confirms our results, the only remarkable difference being the less important role that our data attribute to the stellar mass-to-light-ratio (stellar populations) in determining the FP [*tilt*]{}. The $ATLAS3D$ data also suggest that the FP [*tilt*]{} depends as well on the dark matter (DM) fraction and on the rotational contribution to the kinetic energy ($V_{rot}/\sigma$), thus again pointing towards the above mentioned “hybrid solution”. We show that the global properties of the FP,  its [*tilt*]{} and tightness, can be understood in terms of the underlying correlation among mass, structure and stellar population of ETGs, for which, at increasing the stellar mass, ETGs become (on average) ’older’ and more centrally concentrated. Finally, we show that a Malmquist-like selection effect may mimic a differential evolution of the mass-to-light ratio for galaxies of different masses. This should be taken into account in the studies investigating the amount of the so called “downsizing” phenomenon. author: - | M. D’Onofrio$^{1}$[^1], G. Fasano$^{2}$ , A. Moretti$^{1}$, P. Marziani$^{2}$, D. Bindoni$^{1}$, J. Fritz$^{3}$, J. Varela$^{7}$, D. Bettoni$^{2}$, A. Cava$^{4}$, B. Poggianti$^{2}$, M. Gullieuszik$^{2}$, P. Kj[æ]{}rgaard$^{6}$, M. Moles$^{7}$, B. Vulcani$^{8}$, A. Omizzolo$^{9,2}$, W.J. Couch$^{10}$, A. Dressler$^{11}$\ $^1$Astronomy Department, Vicolo Osservatorio 3, I-35122 Padova, Italy\ $^2$INAF/Astronomical Observatory of Padova, Vicolo Osservatorio 5, I-35122 Padova, Italy\ $^3$Sterrenkundig Observatorium, University of Gent Krijgslaan 281 S9, B-9000 Gent, Belgium\ $^4$Astrophysics Department of the University Complutense of Madrid, 28040 Madrid, Spain\ $^6$The Niels Bohr Institute for Astronomy Physics and Geophysics, Juliane Maries Vej 30, 2100, Copenhagen, Denmark\ $^7$Centro de Estudios de Fisica del Cosmos de Aragón, Plaza de San Juan 1, 44001 Teruel, Spain\ $^8$Kavli Institute for the Physics and Mathematics of the Universe, University of Tokyo, Kashiwa, 277-8583, Japan\ $^9$Vatican Observatory Research Group, University of Arizona, Tucson, AZ 85721, USA\ $^{10}$Center for Astrophysics and Supercomputing, Swinburne University of Technology, PO Box 218, Hawthorn Victoria 3122, Australia\ $^{11}$Observatories of the Carnegie Institution of Washington, Pasadena, CA 91101, USA title: The hybrid solution for the Fundamental Plane --- \[firstpage\] Galaxies: early-types – Galaxies: structures and dynamics – Galaxies: photometry (Visible and Infrared) – Fundamental Plane. Introduction {#Intro} ============ Since its discovery, due to @DjorgDavis and @Dressetal, the Fundamental Plane (FP),  the relation linking the effective radius (), the central velocity dispersion ($\sigma$) and the average effective surface brightness () of early-type galaxies (ETGs): $$\label{fp} \fp1$$ has been considered a key tool for investigating the physical mechanisms driving their formation and evolution. In fact, the observed [*tilt*]{} of the FP with respect to the virial plane (VP) and its tightness imply a peculiar connection between the structure of the galaxies, their history of star formation and their dark matter (DM) content, offering useful constraints for the theoretical models. The [*tilt*]{} problem arose from the observation that, under the assumption of homology and constant $M/L$, the FP coefficients ($a$, $b$) deviate significantly from the virial expectation ($a=2$, $b=0.4$). The typical observed values are in fact $a\sim1.2$ and $b\sim0.3$ in the $V$ band, with small variations mainly depending on the adopted waveband [see @Scodeggio; @LaBarbera]. Since ETGs were considered for a long time “homologous” stellar systems, the FP [*tilt*]{} was originally attributed to stellar population effects, on the basis of the argument that its very existence, under virial equilibrium conditions (and assuming homology) implies a correlation between the dynamical mass-to-light ratio and the galaxy mass: $M/L \sim M^{\alpha}$. @Faber87 found $\alpha\sim0.25$ and later, independent analyses found similar values of $\alpha$ [see @Pahre; @Gerhard; @Borriello; @Treu05]. Among the factors potentially causing the variation of the mass-to-light ratio along the FP, the stellar metallicity, age and initial mass function (IMF), as well as the DM fraction were first considered. The trend in the mean metallicity seemed a viable option [see @Gerhard], but its effect was estimated to produce only a small fraction of the [*tilt*]{} [see @DjorgDavis; @DjorgSant]. Stellar population synthesis models failed to reproduce the [*tilt*]{} [@Renzini1995], but [@Chiosi] and [@ChiosiCarr], in the context of a “monolithic” scenario of galaxy formation, were able to explain it using a varying IMF and a different SFH for galaxies of different masses. The existence of a variable IMF is now supported by several papers [see  for a review @Kroupa2; @Cappellari2]. However, a drastic IMF variation seems required to produce the observed [*tilt*]{} [see @Renzini], while a SFH that smoothly varies with mass is more difficult to reconcile with the widely accepted “hierarchical” merging paradigm of the $\Lambda$CDM cosmology. A significant contribution of the DM to the FP [*tilt*]{} was excluded by @Ciotti [hereafter, C96] on the basis of a fine-tuning argument, but [@Tortora], estimating the total $M/L$ ratio from simple Jeans dynamical models, found that the DM fraction within the effective radius  is roughly constant for galaxies fainter than $M_B \sim -20.5$ and turns out to increase for brighter galaxies, thus implying a systematic variation of the dark-to-bright matter ratio along the FP. [@Padmanabhan] and [@Hyde] also found evidence that the DM fraction ($M_{tot}/M^*$) increases with mass. The finding by [@Pahre] that the [*tilt*]{} is still substantial in the $K$-band, where the luminosity maps the bulk of stellar mass, prompted the search for new explanations of its origin, not directly related to stellar population effects. Many works proposed the alternative scenario in which the “broken structural and dynamical homology” of ETGs is at the origin of the [*tilt*]{} [@Hjorth; @PrugSimien; @Busarello; @GrahamColless; @Pahre; @Bertin; @Trujillo; @Nipoti; @LaBarbera]. This interpretation was supported by the observation that ETGs are “non-homologous” stellar systems in both their structure and dynamics [@Capaccioli87; @deCarvalho88; @Capaccioli89; @Burkert93; @Michard85; @Schombert86; @Caon; @YoungCurrie; @PrugSimien]. However, C96 claimed that again a strong fine–tuning between stellar mass-to-light ratio and structure (Sersic index $n$) is required to explain with just structural non-homology both the [ *tilt*]{} of the FP and the small scatter around it. The role of non-homology was also excluded by [@Cappellari] and [@Cappellari3] using integral models of the ETGs mass distribution based on 2D kinematic maps. Along the same vein, [@Bolton], using the galaxies masses estimated from the gravitational lensing, claimed that structural non-homology does not have a significant role in [ *tilting*]{} the FP. The tightness of the FP relation is particularly important because it provides the strongest constraints on the SFH of galaxies. The origin of the FP scatter was investigated by [@Forbes] and [@Terlevich], who found a correlation between the residuals of the FP and the age of the galaxies (ETGs with higher/lower surface brightness have younger/older ages). [@Gargiulo] found that the FP residuals anti-correlate with the mean stellar Age, while a strong correlation exists with $\alpha/Fe$. In this case, the distribution of galaxies around the FP is tightly related to enrichment, and hence to the timescale of star-formation. [@Gravesetal] found that the stellar population variations contribute at most 50% of the total thickness and that correlated variations in the IMF or in the central DM fraction make up the rest. Recently, [@Magoulas], using a sample of $10^4$ ETGs extracted from the 6dF Galaxy Survey, found that the residuals about the FP show significant trends with environment, morphology and stellar population, the strongest trend being with age. This short review of the FP problem makes it clear that a general consensus about the origin of its properties is still lacking. In particular the role played by non-homology is far from being fully understood. In @Donofrio1 [Paper-I], we studied the FP of a sample of 1550 ETGs, obtained cross-matching the $V$-band, surface photometry dataset of ETGs from $WINGS$ [@Fasano; @Varela Wide-field Imaging of Nearby Galaxy-clusters Survey] with velocity dispersions from literature data [@Smith; @Bernardi]. Our main conclusions were the following: 1) the FP coefficients depend on the luminosity range of the ETGs sample under analysis, as well as on the fitting strategy; 2) the FP coefficients do depend on the local density, while they do not depend on the global cluster properties (such as  X-ray emission); 3) the stellar mass-to-light ratio ($M^*/L$) does not correlate with the $V$-band luminosity, so that a possible role of non-homology in causing the FP [*tilt*]{} should be considered. In this paper we exploit the spectroscopic and photometric, $K$-band database of ETGs in the $WINGS$ survey [@Fritz1; @Valentinuzzi] to complete the analysis of the FP problem. The paper is organized as follows: in Sec. \[sec1\] we present the main equations defining the FP problem. In Sec. \[sec2\] we describe the $WINGS$ data samples used in this work. In Sec. \[sec3\] we derive the FP coefficients for the $V$- and $K$-band. In Sec. \[sec4\] we discuss the origin of the bulk of the FP [*tilt*]{} and compare our results with those obtained using the $ATLAS3D$ and $WSDSS$ databases. In Sec. \[sec6\] we investigate the origin of the differential [*tilt*]{} observed between the $V$ and $K$ wavebands. In Sec. \[sec7\] we address the problem of the thickness of the FP and its connection with the non-homology of ETGs through the structure–stellar population conspiracy. In Sec. \[sec8\] we discuss the variation of the $M/L-M$ relation with redshift and the occurrence of selection effects. Our conclusions are summarized in Sec. \[sec9\], where we also try to probe our findings against the present theoretical models of galaxy formation and evolution. In this paper we use $H_0=70$ km sec$^{-1}$ Mpc$^{-1}$, $\Omega_{\Lambda}=0.7$ and $\Omega_m=0.3$. The FP problem in a nutshell {#sec1} ============================ We assume that ETGs are gravitationally bound stellar systems which satisfy the virial theorem equation: $$\label{eqvir} \langle V^2 \rangle \propto \frac{GM_{tot}}{\langle R \rangle},$$ where $M_{tot}$ is the total galaxy mass, $\langle R \rangle$ is a proxy for the gravitational radius, and $\langle V^2 \rangle$ the mean kinetic energy per unit mass. Ideally, all virialized systems should be placed onto the VP in the space defined by the variables $M_{tot}$, $\langle R \rangle$ and $\langle V^2 \rangle$. Unfortunately, these are not observable quantities. Therefore, in the case of ETGs, the virial equation  (\[eqvir\]) is usually written as follows: $$\label{eqMtot} M_{tot} \propto \frac{K_V\sigma^2\re}{G},$$ where $\sigma$ is the central velocity dispersion within a fixed aperture,  is the equivalent radius of the isophote enclosing half the total galaxy luminosity and $K_V=1/(k_vk_r)$ takes into account projection effects, density distribution and stellar orbits distribution. The term $K_V$ parametrizes our ignorance about orientation, 3D structure and dynamics of ETGs. The formal expression of $K_V$ assumes $\langle V^2 \rangle = k_v \sigma^2$ and $\langle R \rangle = k_r\re$. Introducing the mean effective surface brightness $\langle I \rangle_e = L/2\pi \re^2$, one gets: $$\label{eqvir5} \re \propto \frac{K_V}{2\pi G}\ (\frac{M_{tot}}{L})^{-1}\ \langle I\rangle_e^{-1}\ \sigma^2,$$ or, in log units and after some algebra: $$\label{eqvirlog} \log(\re) = 2\log(\sigma)+0.4\muem +C_\lambda +$$ $$\ \ \ \ \ \ \ \ \ \ \ \ +\log(K^*_V)-\log(\frac{M^*}{L}),$$ where $C_\lambda = - 0.4[M_\odot(\lambda) + 21.572] -\log(2\pi G)$ and $M_\odot(\lambda)$ is the absolute magnitude of the Sun in the given band. In the equation (\[eqvirlog\]), we have replaced the total mass $M_{tot}$ with $M^*$ (stellar mass) and $K_V$ with $K^*_V=K_V/k_m=1/(k_vk_rk_m)$, where $k_m=M_{tot}/M^*$ parametrizes our ignorance about the DM content. This formulation of the Virial theorem is directly comparable with the FP equation (\[fp\]) empirically derived from observations. It formally illustrates the problem of the FP [*tilt*]{}, given that, in all photometric bands, the observed coefficients of $\log(\sigma)$ and $\muem$ turn out to be remarkably different from the virial expectation. Since we assume that ETGs are in virial equilibrium, the reason for the observed deviation of the FP coefficients from the virial expectation must reside in the term: $$\label{eqKV1} \Delta_{FP}=\log(K^*_V)-\log(\frac{M^*}{L}).$$ In fact, any systematic dependence of this expression on the position along the FP would produce a displacement ([*tilt*]{}) of the FP from the VP. According to eq. (\[eqKV1\]), besides the stellar populations ($M^*/L$), the structural and dynamical non-homology ($k_r$ and $k_v$, respectively), as well as the dark matter fraction ($k_m$) might contribute to the FP [*tilt*]{}. The relative importance of each factor can be in principle estimated by determining how strongly it correlates with the position along the FP. In order to parametrize such position, we note that, since each galaxy must simultaneously lie onto the FP and on its proper VP, the right terms of equations (\[fp\]) and (\[eqvirlog\]) can be equated, thus giving: $$\label{eqvirobs} (a-2)\log(\sigma)+(b-0.4)\muem+c-C_\lambda =$$ $$\ \ \ \ \ \ \ \ \ \ \ \ =\log(K^*_V)-\log(M^*/L).$$ The left side of this equation can be computed for each galaxy and represents, for the proper values of $\log(\sigma)$ and $\muem$, the difference between the observed FP and a fixed, reference VP \[that for which $\log(K^*_V)-\log(M^*/L)=0$\]. In Section \[sec4\] we use such a difference: $$\label{eqvirobs1} \Delta_{FP}=(a-2)\log(\sigma)+(b-0.4)\muem+c-C_\lambda$$ to parametrize the position of galaxies onto the FP and we examine how $\Delta_{FP}$ correlates with the observed proxies of the physical factors to which the FP [*tilt*]{} could be ascribed. The WINGS dataset {#sec2} ================= The present work is based on two data samples extracted from the $WINGS$ database. The first one ([*Sample I*]{}) cross-matches the sample of 1550 ETGs used in Paper-I to study the FP in the V-band with the galaxies in a subsample of 26 $WINGS$ clusters for which we obtained $K$-band surface photometry [@Valentinuzzi; @Donofrio]. In total we got 620 ETGs. The second dataset ([*Sample II*]{}) contains 214 ETGs and is obtained cross-matching the [*Sample I*]{} with the catalogues of galaxy masses provided by @Fritz. Likewise the $V$-band, the $K$-band surface photometry was obtained using the purposely devised automatic software GASPHOT [@Pignatelli], which measured the total luminosity, the effective radii, the mean effective surface brightness and the Sersic index of each galaxy in the $WINGS$ clusters. These quantities were derived from a simultaneous best-fit of the major and minor axes growth curves with a Sersic law ( ; @Sersic) convolved with the local $PSF$. The quality of the GASPHOT surface photometry is discussed in [@Donofrio], but several tests of its robustness can be found in the papers already published by the $WINGS$ team (see @Valentinuzzi1 [@Varela; @Vulcani]). The average uncertainties in the surface photometry parameters are $\sim 10\%$, $\sim 20\%$ and $\sim 20\%$ for the luminosity, the effective radius and the Sersic index, respectively. It is worth mentioning that the previously defined samples do not include the galaxies for which the uncertainties of the best-fit parameters exceeded three times the upper quartiles of the corresponding distributions ($\sim 25\%$, $\sim 35\%$ and $\sim 35\%$ for the luminosity, the effective radius and the Sersic index, respectively). The procedure used to determine the stellar masses of galaxies from the $WINGS$ spectra database [@Cava09] has been exhaustively discussed by [@Fritz1; @Fritz] and [@Vulcani]. Here we recall that it is based on a spectrophotometric model that reproduces the main features of observed spectra by summing the theoretical spectra of simple stellar populations of different ages. Besides the stellar masses, the tool is able to derive star formation histories, average age and dust attenuation of galaxies. The models rely onto the Padova evolutionary tracks [@Bertelli] and use the standard Salpeter [@Salpeter] IMF, with masses in the range 0.15-120 $M_{\odot}$. Besides the stars which still are in the nuclear-burning phase, the stellar mass includes remnants, such as white dwarves, neutron stars and stellar black holes (for details see @Fritz1). The stellar mass values relative to the whole galaxy bodies are computed by rescaling to the total V-band magnitudes the masses obtained by fitting the optical spectra, which are calibrated on the V-band fiber magnitudes. Since this procedure implicitly assumes there are no color gradients, we correct the stellar masses following the prescriptions of @BDJ and using the $\Delta(B-V)$ colors measured within the fiber aperture and at a fixed aperture of 5 kpc. In @Fritz it is shown that the (total) stellar masses computed in this way are in fairly good agreement with other independent estimates, leading to an estimated accuracy of $\sim 0.2$ dex. The galaxies were classified as early-types (Ellipticals or S0s) using the automatic tool MORPHOT [@Fasano1], purposely built for the $WINGS$ project. For details about MORPHOT and about the accuracy achieved for our morphological classification we refer to [@Fasano1], where an average $r.m.s.$ of $\sim$1.7 is reported for the difference between automatic and visual estimates of the [ *Revised Hubble Type*]{} (@deVauc). ![ The FPs for the ETGs of [*Sample I*]{} in the $V$- and $K$-bands are shown with black and grey dots respectively. The vertical shift of 0.5 is artificial. The derived coefficients are shown in the y-axis label.[]{data-label="Fig1_FP"}](Fig1_FP.ps){width="90mm"} The velocity dispersions have been taken from the literature [@Smith; @Bernardi]. They have been corrected for aperture effects to $\re/8$ according to @Jorg96 and have an average error of $\sim 7-10$ . The FP in the $V$ and $K$ bands {#sec3} =============================== In Paper-I we studied the $V$-band FP of the ETGs members of the $WINGS$ clusters. Here we exploit the $K$-band extension of the $WINGS$ database to derive a $K$-band solution for the FP coefficients. We remind that, according to @Donofrio1, the FP coefficients do depend on the wavelength of observations (see also @Scodeggio [@LaBarbera]), on the selection criteria of the galaxy sample and on the environment [see also @Desroches; @LaBarbera]. In comparing the $WINGS$ $V$- and $K$-band FPs, we are allowed to ignore the selection and environmental effects, since the same galaxies are used in both wavebands. The fitting tool adopted in this work is the same used in Paper-I,  the software MIST, kindly provided by @LaBarb. We refer to their paper for any detail about the fitting strategy adopted by MIST. We obtain the following best-fit FPs for our [*Sample I*]{} in the $V$- and $K$-bands: $$\label{ourFP} \log(R^V_e) = 1.10(\pm 0.03)\log(\sigma)+0.30(\pm 0.005)\langle\mu\rangle^V_e$$ $$-8.04(\pm 0.13),$$ $$\label{ourFPK} \log(R^K_e) = 1.29(\pm 0.03)\log(\sigma)+0.29(\pm 0.005)\langle\mu\rangle^K_e \nonumber$$ $$-7.34(\pm 0.12).$$ In our formulation of the FP,  is defined as the equivalent (circularized) effective radius in $kpc$, $\sigma$ is the velocity dispersion at $\re/8$ in  and $\langle\mu\rangle_e$ is the mean effective surface brightness in magarcsec$^{-2}$. Note that, in spite of the different samples used, the coefficients obtained here for the $V$-band FP are quite close to the values found in Paper-I ($a=1.15\pm02$, $b=0.32\pm0.004$, $c=-8.56\pm0.09$; see that paper for a discussion about the FP coefficients obtained using different data samples). Note also that the FP slopes we find are a bit shallower than those generally found in the literature. This systematic difference, already discussed in Paper-I, is likely due to the fitting procedure and to the magnitude limit of the data sample. The FPs in the $V$- and $K$-band are shown in Figure \[Fig1\_FP\] (see eq. \[ourFP\] and \[ourFPK\]). The scatter around the FP looks very similar in the two bands ($r.m.s.\sim$ 0.092 and 0.088 for the $V$- and $K$-band, respectively). Taking into account the individual uncertainties of the three observed quantities (  and $\sigma$), together with the covariance term involving the  and  uncertainties, we estimated the intrinsic scatter around the $V$- and $K$-band FP to be 0.078 and 0.062, respectively,  nearly half of the total observed scatter. &gt;From equations  (\[ourFP\]) and (\[ourFPK\]) the FP [*tilt*]{} increases going from the $K$- to the $V$-band. In particular, the gap between the $\log\sigma$ coefficients in the $V$- and $K$-band turns out to be $\sim0.2$, larger than the estimated fitting uncertainty and in agreement with the value found by [@Pahre]. It is worth noticing that, if we assume $\sigma$ as independent variable, we obtain the following best-fit in the two bands: $$\label{ourFP1} \log(\sigma) = 0.72(\pm 0.02)\log(R^V_e)-0.21(\pm 0.006)\langle\mu\rangle^V_e$$ $$+6.18(\pm 0.12),$$ $$\label{ourFP1K} \log(\sigma) = 0.65(\pm 0.02)\log(R^K_e)-0.19(\pm 0.004)\langle\mu\rangle^K_e \nonumber$$ $$+5.17(\pm 0.07).$$ Again, the gap between the coefficients (relative to $\log R_e$ in this case) in the $V$- and $K$-band turns out to be larger than the estimated fitting uncertainty, thus implying that the difference has a physical meaning. Origin of the FP tilt {#sec4} ===================== &gt;From now on, unless warning of the contrary, the analysis of the $WINGS$ data will be based on [*Sample II*]{}. Following the logical thread outlined in Section \[sec1\], we adopt the difference $\Delta_{FP}$ \[see eq. (\[eqvirobs1\])\] between the FP and the reference VP defined therein, to parametrize the position of the galaxy in the FP. Then, we analyze the systematic variation along the FP of the different physical factors that might originate the FP [*tilt*]{}. Structural non-homology and stellar population effects {#sec41} ------------------------------------------------------ [ccccccl]{}\ \ $X$ & $\alpha$ & r.m.s. & Pearson CC & Spearman CC & Significance & Sample\ $\log(M^*/M_{\odot})$ & -0.335 & 0.088 & -0.800 & -0.800 & e-58 & ALL\ $\log(M^*/L)$ & -0.196 & 0.138 & -0.225 & -0.241 & e-03 & ALL\ $\log(n)$ & -0.585 & 0.122 & -0.541 & -0.530 & e-18 & ALL\ $\varepsilon$ & 0.239 & 0.134 & 0.330 & 0.325 & e-06 & ALL\ $\log(M^*/M_{\odot})$ & -0.325 & 0.072 & -0.886 & -0.880 & e-38 & Es\ $\log(M^*/L)$ & -0.365 & 0.141 & -0.400 & -0.409 & e-04 & Es\ $\log(n)$ & -0.946 & 0.134 & -0.579 & -0.563 & e-09 & Es\ $\varepsilon$ & 0.477 & 0.153 & 0.214 & 0.140 & 0.100 & Es\ $\log(M^*/M_{\odot})$ & -0.307 & 0.085 & -0.731 & -0.758 & e-27 & S0s\ $\log(M^*/L)$ & -0.074 & 0.119 & -0.105 & -0.136 & 0.180 & S0s\ $\log(n)$ & -0.375 & 0.108 & -0.454 & -0.416 & e-07 & S0s\ $\varepsilon$ & 0.113 & 0.118 & 0.152 & 0.107 & 0.160 & S0s\ \[Tab0\] In this subsection, by exploiting our homogeneous sample of ETGs, we examine how $\Delta_{FP}$ correlates with $\log(M^*/L)$ (a proxy of the stellar population) and with Sersic index and ellipticity ($n$ and $\varepsilon=1-a_{min}/a_{maj}$). While the Sersic index is commonly considered a robust proxy of the structural (non)homology, the role of the ellipticity in this sense is more controversial. We think that, at least in a statistical sense, the ellipticity should be considered, to all intents and purposes, a crucial ingredient of the galaxy structure, thus being also deeply linked to the structural (non)homology issue. In fact, is well known that different (intrinsic) flattenings usually correspond in ETGs to different shapes (oblateness vs. triaxiality) and luminosity profiles, even leaving aside the obvious prevalence of the composite (bulge+disk) profiles at increasing the ellipticity, due to the increasing fraction of S0 galaxies. In addition, since more flattened ETGs are preferentially fast rotators [@binney; @capp07; @mccarthy; @Emsell], again in a statistical sense, $\varepsilon$ should be linked to the amount of rotational contribution to the total kinetic energy ($V_{rot}/\sigma$). On the other hand, it is hard to deny that a different amount of rotation, by itself, is likely indicative of a different dynamical structure. Therefore, roughly speaking, we might also consider $\varepsilon$ as a sort of indirect proxy of dynamical (non)homology. Figure \[Fig2\_FP\] illustrates how $\Delta_{FP}$, computed from eq. (\[eqvirobs1\]) in the $V$-band, correlates with stellar mass and mass-to-light ratio ($M^*/L$), Sersic index ($n$) and ellipticity ($\varepsilon$). Ellipticals and S0 galaxies are represented in the figure by red and green dots, respectively. The linear best-fits reported in the plots have been computed through standard least square regression analysis taking into account the individual uncertainties on both variables (the crosses located aside in the plots outline the median uncertainties). Table \[Tab0\] lists the parameters of both the linear fits and the correlations in Figure \[Fig2\_FP\] for the whole [*Sample II*]{} and for Es and S0 galaxies separately. Apart from the expected, strong correlation with the stellar mass, the plots in this figure (see also Table \[Tab0\]) show that, among the three quantities we tested for the FP [*tilt*]{}, the Sersic index is the most strongly correlated with $\Delta_{FP}$ (i.e. with the position along the FP), the correlations involving the ellipticity and the stellar mass to light ratio (stellar populations) turning out to be significantly weaker. This holds in both the $V$- and the $K$-band (not reported here), at least for the ETGs sample as a whole. Besides the slopes and the correlation coefficients (both Pearson’s and Spearman’s $CC$), what enforces the above conclusion are the significances associated to the $CCs$ (see Table \[Tab0\]). These are computed according to @press92 and express the probability that the null hypothesis of zero correlation is happening (small values indicate significant correlations). Looking at the $CCs$ and significances reported in Table \[Tab0\], we note that, for the Es alone, the $\Delta_{FP}$–$M^*/L$ correlation is much more significant than for the S0 galaxies alone. We also note in Table \[Tab0\] that, likely because of the reduced base-line, the correlation between $\Delta_{FP}$ and $\varepsilon$ looks rather weak for Es and S0s separately. ![The correlations of the galaxy stellar mass with $M^*/L_V$ (upper panel) and the Sersic index $n$ (lower panel) for the galaxies in our $Sample-I$ (red dots) and $Sample-II$ (green dots).[]{data-label="Fig3_FP"}](Fig3_FP.ps){width="90mm"} Since many studies have found stellar population properties (e.g. age and metallicity) to increase with galaxy stellar mass, it could look surprising that $\Delta_{FP}$ turns out to be strongly correlated with $M^*$, while the correlation between $\Delta_{FP}$ and $M^*/L$ is weak. Does this mean that in our galaxy sample $M^*$ and $M^*/L$ do not correlate? Furthermore, is there any internal correlation among the four tested parameters ($M^*$, $M^*/L_V$, $n_V$ and $\varepsilon_V$)? Which is the relevant correlation in Figure \[Fig2\_FP\] ? Figure \[Fig3\_FP\] shows that, for our galaxy sample too, the correlation between $M^*$ and $M^*/L$ is actually in place, being even slightly more significant than that between $M^*$ and $n_V$ (0.66 vs. 0.55 of $CC$). In Figure \[Fig4\_FP\] it is shown that such a correlation does not translate into a strong correlation between $M^*/L$ and $\Delta_{FP}$. This figure shows face-on views of the V-band FP for our [*Sample II*]{} (left panel) and for the subsamples of ellipticals and S0 galaxies (central and right panels, respectively). In each plot, the horizontal straight line marked with VP$\cap$FP represents the intersection between the FP (coefficients: $a$ and $b$) and an arbitrary, reference VP (coefficients: 2 and 0.4). Thus, the components of the direction vector of such stright line turn out to be: $b$-0.4, 2-$a$, $a$+2$b$. The red arrows in the figure lie onto the FP and, for each tested quantity ($M^*$, $L_V$, $M^*/L$, $n$ and b/a=1-$\varepsilon$), identify the direction along which it turns to show the strongest correlation. More precisely, spanning the whole round corner (0 to 2$\pi$), we rotate the coordinate axis along the FP and, for each angle, we compute the correlation coefficients of the various quantities, recording the directions which maximize them. The length of each arrow is proportional to the corresponding maximum value of $CC$ and the dotted curves correspond to $CC=1$ and $CC=0.5$. In this schematic picture, the arrow relative to $\Delta_{FP}$ must be perpendicular to the straight line VP$\cap$FP and must have size one by definition. In Figure \[Fig4\_FP\] both the stellar mass and the luminosity appear to be strongly correlated with the position along the FP, the direction of maximum correlation being in both cases almost coincident, although with opposite direction, with that relative to $\Delta_{FP}$ (upwards arrow with $CC$=1 by definition). Also the Sersic index ($n$) appears quite well correlated with the position along the FP (especially for Es) and substancially aligned with $M^*$. The axial ratio (b/a) is generally correlated as well with the position along the FP, but its behaviour depends on the morphological type. In fact, while for Es there is a fair alignment with $M^*$ and a weak correlation, for S0s there is a good correlation, but along a direction almost orthogonal to $M^*$ (this is the reason why the $CCs$ in Table \[Tab0\] are small in both cases). In general, the stellar mass-to-light-ratio ($M^*/L$) turns out to be less strongly correlated with the position along the FP and, again, it shows a morphology dependent behaviour. In this case, however, the correlation and the alignment are coupled, both being stronger for Es than for S0s. $ \begin{array}{ccc} \includegraphics[width=57mm]{Fig4_FP.ps} & \includegraphics[width=57mm]{Fig4E_FP.ps} & \includegraphics[width=57mm]{Fig4L_FP.ps} \end{array}$ The Figure \[Fig4\_FP\] should also help to clarify the question of which are the truly important physical factors originating the FP [ *tilt*]{}. The stellar mass is clearly the main driver of the FP [ *tilt*]{} and drags the luminosity together. However, this is in some sense an obvious thing, since we would like actually identify the ’[ *mass-driven*]{}’ physical quantities which mainly contribute to shape the FP through their influence on the observed parameters ($R_e$,  and $\sigma$). To this concern, we checked that the $\Delta_{FP}$ residuals relative to the $M^*-\Delta_{FP}$ relation do correlate with both the Sersic index ($CC=0.31$) and the ellipticity ($CC=0.30$), the significance being very high in both cases ($\sim 10^{-6}$). However, since all physical quantities strongly depend on the stellar mass, the above mentioned correlations could be due to the possible (spurious) correlation between $M^*-\Delta_{FP}$ residuals and $M^*$ itself. Thus, in order to properly check the existence of additional (mass-free) dependences of $\Delta_{FP}$, we must look at the correlations between $M^*-\Delta_{FP}$ residuals and the residuals of the $M^*-Log(n)$ and $M^*-\varepsilon$ correlations. The lower panels of Figure \[Fig5\_FP\] illustrate these correlations, while in the upper panel of the same figure the correlation between $M^*-\Delta_{FP}$ residuals and $M^*$ it is shown. From the correlation coefficients and significances reported in the figure, it is clear that $M^*-\Delta_{FP}$ residuals do not correlate with $M^*$, while the correlations with both the Sersic index and the ellipticity residuals are even stronger than previously found for the corresponding, mass-dependent quantities. This led us to conclude that both the Sersic index and the ellipticity are physical factors actively driving the FP [*tilt*]{}. ![ [**Upper panel**]{}: residuals of the correlation $M^*-\Delta_{FP}$ (see upper-left panel of Fig. \[Fig2\_FP\]) as a funtion of $M^*$. [**Lower panels**]{}: $M^*-\Delta_{FP}$ residuals as a function of the residuals of the $M^*-Log(n)$ and $M^*-\varepsilon$ correlations.[]{data-label="Fig5_FP"}](Fig5_FP.ps){width="90mm"} $X_i$ $c_i$ $r.m.s.$ $t$ value P($>\vert t\vert$) ----------------- -------- ---------- ----------- -------------------- $\log(M^*/L_V)$ -0.115 0.041 -2.793 0.0057 $\log(n)$ -0.430 0.052 -8.294 e-14 $b/a$ -0.108 0.042 -2.540 0.0118 : Multi-variate regression analysis coefficients of the relation: $\Delta_{FP}$=$\sum_{i=1}^3{c_i X_i}$+$const.$ for the $WINGS$ [*Sample II*]{}. \[MVRA\] Further support to this conclusion comes from both the Multi-Variate Regression and the Principal Component Analyses (MVRA and PCA, respectively). Table \[MVRA\] reports the estimated MVRA coefficients of the relation: $\Delta_{FP}$=$c_1\log(M^*/L_V)$+$c_2\log(n)$+$c_3(b/a)$+$const.$, together with the proper standard errors, the relative Student’s $t$ statistics and the probabilities that the $t$ values are exceeded. Table \[MVRA\] indicates that the coefficients of all three variables are significant within the errors. A similar indication comes from the PCA, which is not able to reduce the number of significant variables of the previous relation (including of course $\Delta_{FP}$), all the four eigenvectors turning out to provide a significant contribution to the total energy content of the relation (49.4%, 23.3%, 17.0% and 10.3%, in descending order). It is worth mentioning that all the results illustrated in this Section remain unchanged when computing $\Delta_{FP}$ from the coefficients of eq. \[ourFP1\] rather than those of eq. \[ourFP\], when using the velocity dispersion (instead of the effective radius) as independent variable in the fit of the FP. The only noticeable difference with respect to the previous finding turns out to be the almost absent correlation between $\Delta_{FP}$ and $M^*/L$, with $CC=0.1, 0.25$ and $-0.06$ \[P($>\vert t\vert$)$\sim 0.19, 0.02$ and $0.5$\] for the global, elliptical and S0 samples, respectively. This might indicate that the role of stellar populations in determining the FP [*tilt*]{} is marginal, as the strength of the correlation between $M^*/L$ and $\Delta_{FP}$ seems to depend on the fitting procedure. We believe that the above findings strongly enough support the conclusion that various physical quantities concur in shaping the FP. Still, in Section \[sec7\] we show evidences suggesting that the FP properties ([*tilt*]{} and thickness) altogether could be interpreted as due to the existence of a connection between stellar mass, structure and stellar populations in ETGs [nMML relation; @Donofrio2]. In this framework, it would be misleading to attribute the FP properties to the mere summing of the influences of the individual physical quantities, since they are actually entangled through the nMML relation. Comparison with ATLAS3D and SDSS data {#sec42} ------------------------------------- In order to check the robustness of the results obtained in the previous section from the analysis of the $WINGS$ data sample, we perform here a similar analysis using the $ATLAS3D$ Project data [@Cappellari2011 C11 hereafter] and a subsample of the $SDSS$ and follow-up data [@Abaza09]. This will allow us to compare with our results and to speculate about the influence of dynamical non-homology and dark matter fraction on the FP [ *tilt*]{}. Both the $ATLAS3D$ and the $SDSS$ magnitudes are corrected for galactic extinction and are given in the $r$-band, AB system. Among the 260 galaxies of the $ATLAS3D$ sample, we selected 232 galaxies classified as early-type (Es and S0s) by C11. For these galaxies, we took from @Krajn13 the Sersic index ($n$), derived fitting the $r$-band luminosity profiles with a single Sersic law, and from @Cappellari4 the dynamical mass-to-light-ratio ($M_{Tot}/L$), the FP parameters (, $L/L_{\odot}$ and $\sigma_e$) and the ellipticity ($\varepsilon_e$). All these quantities are derived using the JAM dynamical model [@capp08] and the MGE surface photometry tool [@Emsell94]. To improve the robustness of the analysis, we decided to remove from the sample those galaxies for which the effective radius and the Sersic index uncertainties exceed 2.5$\times rms$ of the relative distributions, as well as the galaxies with extremely faint surface brightness or very small velocity dispersion. We also remove from the sample those galaxies for which the quality flag $q$ of the JAM fitting given by @Cappellari4 is zero (bad fitting). In particular, we kept in the final $ATLAS3D$ sample just those galaxies obeying the following conditions: $n\le$8, $\log(\sigma_e)\ge$1.6, $\muem\le$24, $\delta n\le$1.5, $\delta n/n\le$0.4, $\delta \re/\re\le$0.3 and $q>$0, where $\delta n$ and $\delta \re$ are the uncertainties of $n$ and , respectively. For the remaining sample of 137 $ATLAS3D$ galaxies, we took from @Cappellari3 the JAM model stellar mass-to-light-ratio ($M^*/L$) and from @Emsell the ratio $V_{rot}/\sigma_e$ between the rotation speed and the velocity dispersion. The distances of the $ATLAS3D$ galaxies are taken from C11 and are always less than 45 Mpc. The sample of 817 $SDSS$ galaxies we use here ($WSDSS$ hereafter) is obtained cross-matching the whole $SDSS-DR7$ galaxy sample with the galaxies classified as early-type (Es and S0s) in the $WINGS$ database. The surface photometry parameters in the $r$(AB)-band, including the Sersic index $n$, $\varepsilon$,  and  are taken from @Blant03 and are derived fitting the azimuthally averaged radial luminosity profiles of galaxies with a single component Sersic law. The central velocity dispersions ($\sigma_c$) and the ellipticities ($\varepsilon$) come from the $SDSS$ database [@Abaza09]. The stellar masses $M^*$ are taken from the MPA-JHU DR7 release of spectrum measurements ($http://www.mpa-garching.mpg.de/SDSS/DR7/$) and are obtained by fitting the population synthesis model of @Bruz03 to the $SDSS$ broad bands $u,g,r,i,z$ magnitudes adopting the universal initial mass function (IMF) as parametrized by @Kroupa. The total luminosities in th V-band have been obtained from the $r$- and $g$-band total magnitudes from @Blant03 using the recipe of @Lupt. After conversion to the Vega-mag system, they have been used to compute the mass-to-light ratios $M^*/L_V$. Finally, the distance moduli and the morphological types are taken from the $WINGS$ database. For both the $ATLAS3D$ and the $WSDSS$ samples, we have obtained the best fitting of the FP in the $r$-band using the same fitting algorithm used for the $WINGS$ data (MIST). We obtained the following equations: $$\label{ATLASFP} \log(R_e) = 1.31(\pm 0.05)\log(\sigma)+0.33(\pm 0.009)\langle\mu\rangle_e$$ $$-8.58(\pm 0.22),$$ $$\label{SDSSFP} \log(R_e) = 1.17(\pm 0.05)\log(\sigma)+0.32(\pm 0.009)\langle\mu\rangle_e$$ $$-8.12(\pm 0.37),$$ for the $ATLAS3D$ and the $WSDSS$ samples, respectively. While the FP coefficients relative to the $WINGS$ sample fairly agree with $WSDSS$, they turn out to be significantly different from those relative to the $ATLAS3D$ sample. Apart from the different wavebands (V for $WINGS$ and $r$ for $ATLAS3D$) and luminosity functions of the two galaxy samples, this difference might originate from the different methods used to measure the quantities involved in the FP (,  and $\sigma$). Since it is well known [see for instance @Kelson] that the correlated variation of and from different fitting methods has negligible effects on the FP coefficients, we guess the differences we found in this case are caused by differences in the methods used to measure velocity dispersions. Actually, while the $\sigma$ of the $WINGS$ galaxies are normalized according to the canonical rule proposed by @Jorg96 ($R_e/8$), those given in $ATLAS3D$ are averaged within the effective radius. ![ $\Delta_{FP}$ as a function of $\log(V_{rot}/\sigma_e)$ and $\log(M^*/M_{Tot})$ for 137 galaxies in the ATLAS3D sample. Symbols are as in the previous figures.[]{data-label="Fig6_FP"}](Fig6_FP.ps){width="100mm"} As in the case of the $WINGS$ data, through the eq. (\[eqvirobs1\]), we computed the quantities $\Delta_{FP}$ for the galaxies of the two samples and correlated them with the various physical parameters supposed to be responsible of the FP [*tilt*]{}. The results are illustrated in the Figures \[Fig6\_FP\] and \[Fig7\_FP\] and in the Tables \[Tabatlas\] and \[Tabsdss\]. Note that, in order to compare with our results about the stellar mass-to-light-ratio, we converted the $SDSS$ stellar masses from Kroupa to Salpeter IMF adding 0.13 dex and reported in Fig \[Fig7\_FP\] the $M/L$ in the $V$-band. For the conversion from the $r$(AB)- to the $V$-band we used the equations given in @Blant07 and the $g$ and $r$ magnitudes from $SDSS$. For the 41 $ATLAS3D$ galaxies not included in the $SDSS$ database, we used the (V-R) colors from NED. $ \begin{array}{cc} \includegraphics[width=90mm]{Fig7a_FP.ps} & \includegraphics[width=90mm]{Fig7b_FP.ps} \end{array}$ [ccccccl]{}\ \ $X$ & $\alpha$ & r.m.s. & Pearson CC & Spearman CC & Significance & Sample\ $\log(M^*/M_{\odot})$ & -0.406 & 0.114 & -0.813 & -0.804 & e-39 & ALL\ $\log(M^*/L)$ & -0.799 & 0.171 & -0.629 & -0.622 & e-09 & ALL\ $\log(n)$ & -0.930 & 0.147 & -0.664 & -0.689 & e-22 & ALL\ $\varepsilon$ & 0.616 & 0.171 & 0.444 & 0.433 & e-08 & ALL\ $-\log(V_{rot}/\sigma_e)$ & -0.362 & 0.157 & -0.546 & -0.429 & e-10 & ALL\ $-\log(M^*/M_{Tot})$ & 0.008 & 0.184 & 0.170 & 0.178 & 0.040 & ALL\ $\log(M^*/M_{\odot})$ & -0.406 & 0.100 & -0.891 & -0.889 & e-19 & Es\ $\log(M^*/L)$ & -1.390 & 0.209 & -0.512 & -0.535 & e-04 & Es\ $\log(n)$ & -1.660 & 0.218 & -0.565 & -0.566 & e-06 & Es\ $\varepsilon$ & -0.364 & 0.229 & 0.161 & 0.159 & 0.310 & Es\ $-\log(V_{rot}/\sigma_e)$ & -0.376 & 0.176 & -0.585 & -0.594 & e-05 & Es\ $-\log(M^*/M_{Tot})$ & -0.358 & 0.216 & -0.401 & -0.417 & 0.006 & Es\ $\log(M^*/M_{\odot})$ & -0.386 & 0.116 & -0.694 & -0.716 & e-18 & S0s\ $\log(M^*/L)$ & -0.418 & 0.138 & -0.384 & -0.399 & e-04 & S0s\ $\log(n)$ & -0.702 & 0.116 & -0.655 & -0.665 & e-14 & S0s\ $\varepsilon$ & 0.399 & 0.135 & 0.426 & 0.394 & e-05 & S0s\ $-\log(V_{rot}/\sigma_e)$ & -0.271 & 0.145 & -0.285 & -0.237 & 0.011 & S0s\ $-\log(M^*/M_{Tot})$ & 0.137 & 0.148 & 0.140 & 0.141 & 0.167 & S0s\ \[Tabatlas\] [ccccccl]{}\ \ $X$ & $\alpha$ & r.m.s. & Pearson CC & Spearman CC & Significance & Sample\ $\log(M^*/M_{\odot})$ & -0.367 & 0.080 & -0.848 & -0.833 & 0.000 & ALL\ $\log(M^*/L)$ & -1.230 & 0.118 & -0.603 & -0.664 & e-90 & ALL\ $\log(n)$ & -0.880 & 0.144 & -0.441 & -0.468 & e-45 & ALL\ $\varepsilon$ & 0.249 & 0.141 & 0.286 & 0.287 & e-16 & ALL\ $\log(M^*/M_{\odot})$ & -0.363 & 0.077 & -0.888 & -0.873 & 0.000 & Es\ $\log(M^*/L)$ & -1.950 & 0.124 & -0.681 & -0.658 & e-47 & Es\ $\log(n)$ & -1.450 & 0.190 & -0.409 & -0.486 & e-18 & Es\ $\varepsilon$ & 0.561 & 0.173 & 0.104 & 0.095 & 0.077 & Es\ $\log(M^*/M_{\odot})$ & -0.359 & 0.076 & -0.828 & -0.834 & 0.000 & S0s\ $\log(M^*/L)$ & -0.914 & 0.107 & -0.575 & -0.668 & e-60 & S0s\ $\log(n)$ & -0.675 & 0.124 & -0.432 & -0.426 & e-24 & S0s\ $\varepsilon$ & 0.257 & 0.124 & 0.320 & 0.308 & e-12 & S0s\ \[Tabsdss\] Figure \[Fig7\_FP\] and Tables \[Tabatlas\] and \[Tabsdss\] allow a direct comparison with the results obtained for the $WINGS$ sample (see Fig. \[Fig2\_FP\] and Table \[Tab0\]). From this comparison, we draw the following general conclusions: [*(i)*]{} for the stellar mass and the structural parameters $n$ and $\varepsilon$, both relative to the whole samples (Es+S0s) and to the Es and S0s samples separately, the $ATLAS3D$ and $WSDSS$ data produce correlations very similar to the corresponding ones we found for the $WINGS$ galaxy sample. Actually, for both $n$ and $\varepsilon$, the $CCs$ from $WINGS$ are intermediate between those obtained from the two comparison samples; [*(ii)*]{} the correlation between $\Delta_{FP}$ and $M^*/L_V$ turns out to be much stronger for the $ATLAS3D$ and $WSDSS$ samples than for the $WINGS$ sample, although for the $WINGS$ Es alone the correlation is quite significant as well. One could speculate that such discrepancy is because the origin of the tilt and scatter of the FP might be not the same for galaxies residing in different environments (i.e. $WINGS$ vs. $WSDSS$ and $ATLAS3D$ samples). However, even though the dependence of the FP coefficients on the environment still is a controversial issue [@Bernardi; @LaBarbera], many works have shown the FP tilt and scatter to be nearly the same in different environments [@Jorg96; @Pahrea; @Pahre; @Kocha; @delarosa; @Magoulas]. Since it is hard to believe that different physical mechanisms operating in different environments produce the same FP shape, we are inclined to believe that the origin of the above discrepancy lies in the different recipes adopted to calculate the total stellar mass of the galaxies. We do not enter here in the ample literature debate concerning the theoretical models which better reproduce the properties of the stellar population. We only note that: [*a)*]{} the range of $M^*/L_V$ values found with the $WSDSS$ data is smaller than those provided by $WINGS$ and $ATLAS3D$ (see Fig. \[Fig7\_FP\]); [ *b)*]{} the $ATLAS3D$ $M^*/L$ ratios are measured within 1, so they cannot be directly compared with those of $WINGS$ and $WSDSS$, in particular if one recognizes that strong color gradients exist in ETGs of high mass (see below Fig. \[Fig13\_FP\]). Taking into account that the $WINGS$ masses are obtained from the analysis of the whole SED of our galaxies and are corrected for color gradients, at variance with those derived by $ATLAS3D$ that are based on the line strength index of $H_{\beta}$ [@Cappellari4], we are confident that our masses and $M^*/L_V$ measurements are quite robust; [*c)*]{} the observed discrepancy will remain an open problem until new more reliable stellar mass measurements will be available for galaxies. &gt;From Figure \[Fig6\_FP\] and Table \[Tabatlas\] (just $ATLAS3D$ data), we conclude that: [*(iii)*]{} $M^*/M_{Tot}$ significantly correlates with $\Delta_{FP}$ just for the Es sample; [*(iv)*]{} the correlation between $\Delta_{FP}$ and $V_{rot}/\sigma_e$ turns out to be very significant for Es and, to a lesser extent, for S0 galaxies. Concerning the last point, in the first paragraph of Sec. \[sec41\] we briefly discussed the role played by both $\varepsilon$ and $V_{rot}/\sigma$ in the structural and dynamical (non)homology issue. We argued that, at least in a statistical sense, the ellipticity is likely involved in both the luminous and dynamical structure of ETGs. We also argued that, even if specifically indicating the rotational contribution to the kinetic energy, the quantity $V_{rot}/\sigma$ could also be thought as a rough diagnostic (proxy) of the dynamical structure of ETGs. Here, the similarity of the correlations linking $\Delta_{FP}$ with $\varepsilon$ and $V_{rot}/\sigma$ for the $ATLAS3D$ sample (see left panels of Figure \[Fig7\_FP\] and upper panel of Figure \[Fig6\_FP\]), suggests to conclude that, in some sense, $\varepsilon$ and $V_{rot}/\sigma$ could be though as tokens of the same physical phenomenon. Indeed, in Figure \[Fig8\_FP\] it is shown that a strong correlation exists between $\varepsilon$ and $V_{rot}/\sigma$. A question could be raised: which of the two quantities is driving the correlation with $\Delta_{FP}$? We tried to answer this question through the analysis of the residuals, but our attempts were unsuccessful, since the CCs and the significances of the residual correlations turned out to be always inconclusive, although for the Es the residuals of the $V_{rot}/\sigma$-$\Delta_{FP}$ regression turn out to be marginally correlated with $\varepsilon$ (significance $\sim 0.07$). In the last part of this sub-section it is shown that the combined MVRA and PCA results seem to favour $V_{rot}/\sigma$ as driver quantity of the correlation with $\Delta_{FP}$. ![ The correlation between the ellipticity $\varepsilon_e$ and $V_{rot}/\sigma_e$ for the $ATLAS3D$ galaxy sample. Symbols are as in the previous figures.[]{data-label="Fig8_FP"}](Fig8_FP.ps){width="90mm"} Figure \[Fig9\_FP\] illustrates, for the samples $ATLAS3D$ and $WSDSS$ too, the results of the analysis already performed for the $WINGS$ sample about the directions of maximum correlation onto the FP for the various physical parameters possibly involved in its [*tilt*]{} (see Fig. \[Fig4\_FP\]). Comparing Fig \[Fig9\_FP\] with Fig. \[Fig4\_FP\], we note that: [*(v)*]{} the arrows relative to $M^*/L_V$ and $n$, for all three samples ($WINGS$, $ATLAS3D$ and $WSDSS$), are substantially in agreement with each other as far as both the angle and the orientation onto the FP are concerned. The alignment with $\Delta_{FP}$ is pretty good (although with opposite orientation) and the correlation is in general stronger for Es than for S0s; [*(vi)*]{} for the axial ratio (b/a) the agreement among the three samples is less good, but still reasonable if one takes into account the different surface photometry techniques used by the three surveys. In this case the strength of the correlation is greater for S0 than for E galaxies and the alignment with $\Delta_{FP}$ is usually poor. Finally, from the upper panels of Figure \[Fig9\_FP\] (just $ATLAS3D$ sample) we note that: [*(vi)*]{} the arrows relative to both $M^*/M_{Tot}$ and $V_{rot}/\sigma_e$ turn out to be very well aligned with $\Delta_{FP}$ for Es, while there is a substantial lack of alignment for the S0 sample. $ \begin{array}{ccc} \includegraphics[width=57mm]{Fig9a_FP.ps} & \includegraphics[width=57mm]{Fig9aE_FP.ps} & \includegraphics[width=57mm]{Fig9aL_FP.ps} \end{array}$ $\begin{array}{ccc} \includegraphics[width=57mm]{Fig9b_FP.ps} & \includegraphics[width=57mm]{Fig9bE_FP.ps} & \includegraphics[width=57mm]{Fig9bL_FP.ps} \end{array}$ &gt;From the previous points it is clear that the results illustrated in Section \[sec41\] for the $WINGS$ galaxy sample are substantially confirmed by the $ATLAS3D$ and $WSDSS$ galaxies, the main difference being the weaker correlation we found between $\Delta_{FP}$ and $M^*/L$ for the $WINGS$ galaxy sample with respect to the two comparison samples. In general, however, all three data samples support the conclusion that both the stellar populations ($M^*/L$) and the structural non-homology ($n$) are responsible for the FP [*tilt*]{}. As in the case of the $WINGS$ sample, also for the $ATLAS3D$ and $WSDSS$ data, this conclusion is supported by the MVRA (see the first two rows of each sample in Table \[MVRA1\]). The conclusions are not so straightforward for the other tested quantities (b/a, $V_{rot}/\sigma_e$ and $M^*/M_{Tot}$). In fact, for the $ATLAS3D$ and $WSDSS$ samples, the PCA provides four (marginally five: 46.0%, 22.3%, 12.9%, 9.2%, 6.1% and 3.5%) and three (marginally four: 50.8%, 26.8%, 14.6% and 7.8%) significant eigen-vectors, respectively. The combined inspection of the upper plots in Fig. \[Fig9\_FP\] and of Table \[MVRA1\], seems to indicate that, of the two ’twin’ quantities $b/a$ and $V_{rot}/\sigma_e$ (see Fig. \[Fig8\_FP\]), the first one might be the most significant (  that driving the correlation with $\Delta_{FP}$) for the S0 galaxies, while the last one might be the most significant for elliptical galaxies. Finally, the the dark matter fraction too could be a not negligible ingredient of the FP [*tilt*]{}, at least for Es (see again the upper plots in Fig. \[Fig9\_FP\]). [ccccc]{}\ $X_i$ & $c_i$ & $r.m.s.$ & $t$ value & P($>\vert t\vert$)\ \ $\log(M^*/L_V)$ & -0.439 & 0.064 & -6.878 & e-10\ $\log(n)$ & -0.456 & 0.059 & -7.692 & e-12\ $b/a$ & -0.026 & 0.063 & -0.409 & 0.6830\ $-\log(V_{rot}/\sigma_e)$ & -0.119 & 0.037 & -3.228 & 0.0016\ $-\log(M^*/M_{tot})$ & -0.788 & 0.181 & -4.354 & 0.0001\ \ $\log(M^*/L_V)$ & -0.921 & 0.042 & -21.86 & e-16\ $\log(n)$ & -0.341 & 0.028 & -11.86 & e-16\ $b/a$ & -0.064 & 0.020 & -3.241 & 0.0012\ \[MVRA1\] The amount of non-homology {#sec5} -------------------------- The first attempt to estimate the non-homology term $K_V$ defined in eq. \[eqMtot\] for a spherical galaxy can be traced back to [@Poveda], who derived values around $\sim3$. More recently, @Cappellari compared the virial ($M/L \propto \re\sigma^2/L$) and Schwarzschild estimates of $M/L$ obtaining $K_V=5.0\pm0.1$, while [@Gallazzi] found values between 5 and $\sim 7$. In the previous sub-sections we have found that the dark matter fraction ($1-M^*/M_{Tot}$) and the structural and dynamical non-homology ($n$, $\varepsilon$, $V_{rot}/\sigma$) play a role at least as important as the stellar population effects ($M^*/L$) in determining the bulk of the FP [*tilt*]{}. These results imply that the term $K_V$ in the expression (\[eqKV1\]) cannot be constant. This is at variance with the finding of @Cappellari and @Cappellari3, who claimed that the mass-to-light-ratio has the whole responsibility of the FP [*tilt*]{}, finding an almost constant value of $K_V$ and ruling out the non-homology as a possible driver of the [*tilt*]{} itself. Trying to clarify this point, we used the $ATLAS3D$ data to compute $K^*_V$ from eq. (\[eqKV1\]) and $K_V$ from the ralation $K_V=K^*_V*k_m$ (see Section \[sec1\]), where $k_m=M_{Tot}/M^*$. We also computed $K^*_V$ (again from eq. \[eqKV1\]) for both the $WINGS$ and $WSDSS$ data. In the left panel of Figure \[Fig10\_FP\] we show the distribution of $K_V$ for the $ATLAS3D$ sample. The $K_V$ distribution of the $ATLAS3D$ sample looks at variance with the very finding claimed by @Cappellari and @Cappellari3 about the small scatter of $K_V$ ($\sim 0.1$). We guess that, rather than to the scatter, such a small value actually corresponds to the estimated uncertainty found by @Cappellari for the best-fit slope of the relation between the virial and Schwarzschild $M/L$ estimates (see Fig.13 therein). In the right panel of Figure \[Fig10\_FP\] we compare the distributions of the $K^*_V$ obtained for all three samples ($WINGS$, $ATLAS3D$ and $WSDSS$). We remind that, according to its very definition (see Sec. \[sec1\]), besides the structural and dynamical non-homology, the $K^*_V$ term also parametrizes the dark matter contribution (unknown, in the case of the $WINGS$ and $WSDSS$ samples). The $K^*_V$ distribution for the $ATLAS3D$ sample, where the dark matter contribution has been directly estimated, turns out to be in fair agreement with (although narrower then) that of the $WINGS$ sample and slightly shifted upwards with respect to that of the $WSDSS$ sample. Actually, the three distributions are not so different and, in any case, the statement about the constancy of $K_V$ and $K^*_V$ does not seem to be supported by the observations. This is also illustrated in Figure \[Fig11\_FP\], where $K^*_V$ is shown to depend on the Sersic index in all three samples. ![ The correlation between the $K^*_V$ and $\log(n)$ for the $WINGS$, $ATLAS3D$ and $WSDSS$ samples. Symbols are as in the previous figures.[]{data-label="Fig11_FP"}](Fig11_FP.ps){width="90mm"} The hybrid solution ------------------- In this Section we have strongly promoted the so called ‘[*hybrid solution*]{}’ of the FP problem. As already pointed out in the introduction, the idea is actually not new at all. Several previous works have also found that different physical mechanisms might concur in shaping the FP [[*tilt*]{} and thickness; see e.g. @Ciotti; @PrugSimien; @Pahre among others]. In their pioneering paper, C96 first highlighted, from a theoretical point of view, that a [*fine-tuning*]{} is required to explain the FP [*tilt*]{} with single physical effects, such as  structural non-homology or DM distribution. They concluded that: “...it remains the possibility of a [*hybrid*]{} origin for the [*tilt*]{}, with more than one effect contributing to tilting the FP, for example a small progression of anisotropy, DM concentration, and shape $n$, coupled with a modest increase of $M^*/L$.” [@PrugSimien] suggested that, adding together the contributions of stellar population effects, rotational support, and non-homology, one can fully account for the [*tilt*]{} of the FP. [@Pahre] presented a comprehensive model in which stellar population gradients and systematic deviations of the internal dynamical structure of ETGs from the homology, simultaneously explain the differential [*tilt*]{} of the FP among different bandpasses, the slope of the near-infrared FP and the slope of the $Mg_2-\sigma$ relation. Our approach is in the groove of the [*hybrid*]{} solution, but differs from the previous ones since we do not try to attribute the FP properties to the mere addition of several contributors. Instead, supported by our data, we believe that the various physical mechanisms shaping the FP are mutually entangled. In our picture, according to the $n - M^* - M^*/L$ relation found by [@Donofrio2 see Section \[sec7\]], at increasing the stellar mass, ETGs become (on average) ’older’ and more centrally concentrated. Such twofold behaviour is due to the fact that the old stellar populations tend to be more centrally concentrated than the young ones (Fig. \[Fig13\_FP\]). In the following sections, we show that, besides to easily explain the differential FP tilt between the V- and K-band, this natural [*fine-tuning*]{} mechanism might also drive the main FP properties ([*tilt*]{} and thickness). To this concern, we note that it would be misleading to believe that, in the hyperspace defined by , , $\sigma$, $M/L$, $n$, $\varepsilon$ (and possibly $V_{rot}/\sigma$ and DM fraction), the hybrid solution implies the existence of an hyperplane around which galaxies crowd with a scatter significantly smaller than that around the FP. First of all, if we assume that the scalar Virial Theorem is governing the galaxy structure and dynamics, the only hyperplane around which all galaxies are distributed with a scatter just due to measurement errors is that expressed by the eq. \[eqvir5\]. In this equation, besides the observables ,  and $\sigma$, we have to deal with the unknown quantities $M/L$ and $K_V$. Since we adopt $n$, $\varepsilon$, $V_{rot}/\sigma$, $M^*/L$ and the dark matter fraction as proxies of $M/L$ and $K_V$, the fact that $\Delta_{FP}$ depends on these quantities just tells us that each galaxy contributes to the FP [*tilt*]{} through its own $M/L$ and $K_V$. Moreover, since the above proxies are only approximately correlated with $M/L$ and $K_V$, if we use them, together with , and $\sigma$, to form an hyperspace, we introduce the additional (intrinsic) scatter of the correlations between $K_V$ and the quantities $n$, $\varepsilon$ and $V_{rot}/\sigma$, as well as that relative to the correlation between $M/L$ and $M^*/L$. Thus, we expect that around the hyperplane the scatter is not smaller than that found around the FP. This unfailingly happens. In fact, the scatter of galaxies in our [*Sample-II*]{} around the hyperplane best fitting the quantities , , $\sigma$, $M/L$, $n$ and $\varepsilon$ turns out to be $\sim 0.09$, quite similar to that found in Sec. \[sec3\] around the FP in both the V- and K-band. Origin of the differential tilt {#sec6} =============================== Following the same logical thread of Sec. \[sec4\] and using our $WINGS$ [*Sample II*]{}, we investigate in this section the origin of the differential FP [*tilt*]{} observed between the $V$- and $K$-band. In this case we parametrize the position of galaxies on the FPs through the difference between the planes themselves \[eq. (\[ourFP\]) and (\[ourFPK\])\], averaged over the values of $\muem$ in the two bands (the velocity dispersion being the same in both equations). In practice, we assume the analogous of $\Delta_{FP}$ (see eq. \[eqvirobs1\]) to be, in this case, the quantity: $\Delta FP_{VK} = (a_V-a_K)\log(\sigma)+(b_V-b_K)(\langle\mu\rangle^V_e+\langle\mu\rangle^K_e)/2$, where $a_V$,$b_V$,$\muem^V$ and $a_K$,$b_K$,$\muem^K$ are the FP coefficients and average surface brightness in the V- and K-band, respectively, while the term $(\muem^V+\muem^K)/2$ parametrizes the band-averaged position of each galaxy along the  axis. Then, we look whether such quantity correlates with some proxies of the stellar population and luminosity structure. ![ [**Upper panels**]{}: the difference between the FPs in the $V$- and $K$-band (see text for details) as a function of The $(V-K)$ color (left panel) and the band-averaged logarithm of the Sersic index (right panel) for the galaxies in the $WINGS$ sample; [**Lower panels**]{}: the Sersic index (top panel), the effective radius (middle panel) and effective surface brightness (bottom panel) differences between the $V$- and $K$-band as a function of the stellar mass for the galaxies in the $WINGS$ sample. Symbols as in Fig. \[Fig2\_FP\].[]{data-label="Fig12_FP"}](Fig12_FP.ps){width="95mm"} In the two topmost panels of Figure \[Fig12\_FP\] the above correlations are shown for the galaxies in our sample. In particular, the color $V-K$ \[$\simeq \Delta\log(M^*/L)_{V-K}$\] and the band-averaged logarithm of the Sersic index are used as proxies of the stellar population and structure, respectively. In both cases the correlations are clear ($C.C.\sim 0.36$ and $0.38$, respectively, with high significances in both cases), thus indicating that both the stellar population and the luminosity structure contribute to the FP [*tilt*]{} difference between the $V$- and $K$-band. How these factors operate in producing such a result is clarified by the Figure \[Fig13\_FP\] and by the three bottom panels in Figure \[Fig12\_FP\]. Figure \[Fig13\_FP\] shows the average color profiles of ETGs in our sample for different stellar mass intervals. The color profiles are derived from the GASPHOT growth curves, corrected for galactic extinction and K-correction and binned as a function of $R/R_e$. Besides confirming the well known inward reddening of ETGs, the Figure \[Fig13\_FP\] shows that their average color gradients are quite shallow for galaxies in the low/intermediate mass regime \[$\log(M^*/M_{\odot})<11$\], becoming relevant for more massive objects. We do not investigate here the origin of such different gradients (which of course can be related to metallicity effects and to the different star formation histories of ETGs). We just note here that our result is in good agreement with the finding of [@LaBarb2], which is based on a much larger sample [but see @Tortora2 for a different result]. It suggests that the differential FP [*tilt*]{} between the $V$- and $K$-band is originated by the different behaviour of the average luminosity profiles for different ranges of stellar mass in the two wavebands. In fact, Figure \[Fig13\_FP\] shows that, at increasing the stellar mass, the old (red) stellar population tends to be more centrally concentrated, most of the difference concerning massive galaxies \[$\log(M^*)>11$\]. This color behaviour produces the trends observed in the bottom panels of Figure \[Fig12\_FP\]. In fact, it implies that, at increasing the stellar mass, the luminosity profiles of ETGs appear progressively steeper (inward) in the $K$- than in the $V$-band \[decreasing $\Delta\log n$ and increasing $\Delta\log R_e$ and $\Delta\muem$\]. Since the position of galaxies on the FP also corresponds to a mass sequence (see upper-left panel of Fig. \[Fig2\_FP\]; see also Fig. \[Fig4\_FP\]), the trends shown in the bottom panels of Figure \[Fig12\_FP\] produce, in turn, the differential [*tilt*]{} between the $V$- and $K$-band. In some sense, saying that the differential FP [*tilt*]{} between the V- and K-band is due to the mass-dependent spatial distribution of stellar populations inside ETGs ( stellar population effects) is equivalent to say that it is due to the different behaviour of the luminosity profiles in the two bands as a function of stellar mass ( structural effects). This means that a sort of entanglement between structural and stellar population effects is at the origin of the differential [*tilt*]{} between the V- and K-band. We note in passing that, if we try to figure out the above reasoning in the viewpoint of the usual $M/L$ vs. non-homology diatribe, the differential FP [*tilt*]{} between the $V$- and $K$-band must be ascribed to both the luminosity structure (different luminosity profiles in the two bands: non-homology) and stellar population ($M/L$) effects, the two contributions being mutually entangled. ![ [**Top panel:**]{} the mean $V-K$ color profiles of our ETGs corrected for galactic extinction and K-corrections for different bin of mass: filled circles ($9.0<\log(M^*)<9.5$, 92 objects), filled squares ($9.5<\log(M^*)<10.0$, 199 objects), open circles ($10.0<\log(M^*)<10.5$, 250 objects), open squares ($10.5<\log(M^*)<11.0$, 200 objects). Each bin is the average of several growth curves measured by GASPHOT. [**Bottom panel:**]{} the same for very massive galaxies: stars ($11.0<\log(M^*)<11.5$, 82 objects), open pentagons ($11.5<\log(M^*)<12.0$, 18 objects), open triangles ($12.0<\log(M^*)<12.5$, 5 objects). []{data-label="Fig13_FP"}](Fig13_FP.ps){width="90mm"} As already noted at the end of Sec. \[sec41\] for the results illustrated therein, the findings in this Section and the related analyses remain unchanged when computing $\Delta_{FP}$ from the coefficients of eqs. \[ourFP1\] and \[ourFP1K\] rather than those of eqs. \[ourFP\] and \[ourFPK\], when using the velocity dispersion (instead of the effective radius) as independent variable in the fit of the FP. [**However, similarly to what noted for $M^*/L$ in Sec. \[sec41\], the color $V-K$ turns out to be less strongly correlated with $\Delta_{FP}$ than in the usual FP formulation ($CC\sim 0.28$ vs. $-0.36$), again suggesting that the influence of the stellar populations on the FP [*tilt*]{} is rather uncertain.**]{} The thickness of the FP {#sec7} ======================= In Section \[sec3\] we have shown that the thickness of the FP cannot be entirely due to observational errors. These can contribute to nearly half of the observed scatter of $\sim0.09$. The additional (intrinsic) scatter of the FP implies that ETGs do not deviate from the plane more than $\sim 15\%$. As mentioned in the introduction, this scatter has been mainly attributed to stellar population effects through the age of galaxies. ![ Residuals of the best-fit FP in the $V$-band from $WINGS$ data (eq. \[ourFP\]) vs. mass-to-light ratio (upper-left panel), stellar mass (upper-right panel), $V$-band luminosity (bottom-left panel) and luminosity-weighted age (bottom-right panel) of galaxies in our sample.[]{data-label="Fig14_FP"}](Fig14_FP.ps){width="90mm"} Figure \[Fig14\_FP\] shows the residuals of the best-fit FP in the $V$-band (eq. \[ourFP\]) for the galaxies in our [*Sample II*]{} as a function of some interesting quantities. No significant trends are found, in particular with the luminosity weighted Age. The weak correlations with $(M^*/M_\odot)$ and $(L/L_\odot)_V$ is likely indicating that the main driver of the FP residuals is the selection effect connected with the data sample used ( with the range of stellar masses and total luminosities). This would confirm the suggestion given in Paper-I that the FP is likely a bent surface and that a careful choice of the data sample is mandatory before drawing any conclusion about the FP properties (including thickness; see also Section \[sec8\]). In lack of any clear indication coming from the analysis of the residuals, we may wonder about whether the observed scatter around the FP is consistent with our finding that the FP [*tilt*]{} is originated by the simultaneous influence of different physical factors (non-homology, mass-to-light-ratio, dark matter fraction). According to C96, even if the range of Sersic indices spanned by ETGs is in principle able to explain by itself the observed FP [*tilt*]{}, the small scatter around the plane would require in this case an [ *’ad-hoc’*]{} fine-tuning. Using dynamical, isotropic models of spherical galaxies with Sersic profiles and without DM, C96 found that the observed FP tightness would require a tight correlation between the Sersic index ($n$) and the $k_1$ parameter of the $k$-space defined by [@BBF] \[$k_1=(2\log\sigma+\log\re)/\sqrt{2}$\]. In particular, if the FP [*tilt*]{} is entirely due to Sersic index variation, the maximum allowed range of $n$ at any given $k_1$ should not exceed $\sim 1.0$ (Figure 5 therein). Instead, using the small dataset of Virgo galaxies studied by [@Caon], C96 claim that the observed scatter of the $n$-$k_1$ relation turns out to be nearly 3 times larger (Figure 6 therein). Here, using the ETGs in our sample, we show that the intrinsic scatter of the $n-k_1$ relation may be consistent with the required theoretical strip defined by C96. The upper panel of Figure \[Fig15\_FP\] shows the relationship between $n$ and $k_1$ for the $V$-band. Once the uncertainties on both coordinates ($n_V$ and $k_1$) are taken into account, the intrinsic scatter of the relation (gray strip in the figure) turns out to be: $2\times r.m.s.\sim$1.58. This means that the intrinsic variation of $n_V$ at each $k_1$ is just a factor $\sim 1.5$ larger than the theoretical strip. Taking into account that such strip is obtained assuming spherical, isotropic models without DM and that the structural non-homology is likely not the only contributor to the FP [*tilt*]{} (as assumed by C96), we can guess that for our galaxy sample the [*intrinsic*]{} range of variation of $n_V$ at each $k_1$ might be comparable to the theoretical one ($\sim 1$). Therefore, the [*’ad hoc’*]{} fine-tuning required by models seems to be actually in place and we are lead to conclude that even the structural non-homology alone might simultaneously explain the [*tilt*]{} and the tightness of the FP. Trying to identify the origin of such fine-tuning, we recall that @Donofrio2 discovered a bi-variate correlation linking $n$, $M^*$ and $M^*/L$ (nMML relation, hereafter). The best-fit of this relation for our galaxy [*Sample II*]{} in the V-band turns out to be: $$\label{eqfp} \log(n^*_V)=0.285*log(M^*)-0.617*\log(M^*/L_V)+const.$$ with $r.m.s. = 0.14$. The existence of a relationship between structure, mass and stellar populations in ETGs has been claimed in various ways in the literature. [@Valentinuzzi2] [see also @Poggianti] found that both in clusters and in the field, at a given galaxy size, more massive objects have older luminosity weighted ages, while at a given mass, larger galaxies have younger luminosity weighted ages,  there is a correlation between compactness and age of ETGs. More recently, [@Wuyts] confirmed that objects with different Sersic index populate in a different way the SFR-Mass relation, built for the whole Hubble sequence (their Figure 1). ![ The effective radius  as a function of $n_V$ in log units. The full line shows the linear best-fit and symbols are as in the previous figures.[]{data-label="Fig16_FP"}](Fig16_FP.ps){width="80mm"} It is worth stressing that the direction of the nMML relation is in agreement with the findings of @Valentinuzzi2 and [@Poggianti]. In fact, given the well known direct correlation between the Sersic index and the effective radius [see Fig. \[Fig16\_FP\]; see also @Caon Fig.5 therein], the equation (\[eqfp\]) confirms the above mentioned link between age and compactness. Therefore, the nMML relation coexists with the FP relation and could actually represent the conspiracy between stellar population and structure (fine-tuning) which, according to C96, is required to explain the FP tightness if non-homology is responsible of the FP [*tilt*]{}. As an exercise, we plot in the lower panel of Figure \[Fig15\_FP\] the same relation shown in the upper panel, but replacing the observed value of the Sersic index ($n_V$) with the one ($n^*_V$) obtained through eq. (\[eqfp\]). The uncertainties in this figure are computed by error propagation. It is interesting to note that the intrinsic scatter of the new relation $n^*_V$-$k_1$ ($r.m.s. \sim 0.50$) turns out to be fully consistent with the above mentioned, ’theoretical’ one. It is also worth stressing that this very tight relation involves four largely independent quantities,   and $\sigma$ for the abscissa, $M^*$ and $L_V$ for the ordinate. Finally, we note that the nMML relation is someway implicit in the above eq. (\[eqKV1\]). In fact, using the correlation $$\log(K^*_V) = -1.35\log(n)+0.634$$ shown in the uppermost panel of Figure \[Fig11\_FP\] between $n$ and $K^*_V$ (the last one computed from eq. \[eqKV1\]) and the correlation between $\log(M^*)$ and $\Delta_{FP}$ shown in the upper-left panel of Figure \[Fig2\_FP\] (see the corresponding linear coefficients in Table \[Tab0\]), the equation (\[eqKV1\]) takes the form: $$\label{eqfp1} \log(n^*_V)=0.274*log(M^*)-0.740*\log(M^*/L_V)+const.$$ which, given the large fitting uncertainties, turns out to be remarkably similar to eq. (\[eqfp\]). Thus, the nMML relation seems to simultaneously explain the [*tilt*]{} and the tightness of the FP. In some sense, the very existence of the FP [*tilt*]{} seems to imply the FP tightness,  the two things appear to be entangled through the nMML relation. The mass-to-light ratio of high redshift galaxies {#sec8} ================================================= Since structure and stellar population of ETGs seem to be linked each other, the common practice of deriving the total mass-to-light ratio ($M_{tot}/L$) of ETGs by means of the FP assuming the perfect homology ($K_V=constant$) might be not correct, in particular for high redshift galaxies. It is well known that the FP coefficients are believed to vary with redshift, both in clusters [see @vanDokkum; @Sperello; @Jorgensen] and in the field [@Treu05]. If one uses the FP relation to derive the dynamical $M_{tot}/L$ assuming the homology, the above variation implies a progressive change of the mass-to-light ratio with redshift: $d\log(M_{tot}/L_B)/dz \sim -0.7$ [see e.g., @Treu05]. This behavior has been commonly explained invoking the “downsizing” phenomenon (see @Sperello [@Jorgensen]),  the idea that the mass-to-light ratio evolves at different rates for low and high mass galaxies. Actually, while for massive galaxies the FP appears nearly unchanged since $z\sim1$, for low-mass galaxies a progressive displacement from the local FP is observed at increasing redshift [@vanderWel; @Treu05]. In this section we indicate two warnings about this kind of reasoning. The first one is that the existence of the nMML relation implies that, besides the stellar populations, the galaxy structure too varies with redshift [see e.g., @Chevance]. The second one is that a Malmquist bias might mimic the “downsizing” phenomenon. This should advise people to carefully take into account both effects before drawing any conclusion about the evolution of the mass-to-light ratio. In order to illustrate the above mentioned second warning, we first derived the FP coefficients using the high redshift data of [@Sperello], obtained from the K20 survey at redshift $z\sim1$ [@Cimatti]. We choose these data for our analysis because all the values of the Sersic index $n$, the dynamical masses, and the total luminosities in the $B$ band rest frame of the galaxies are given for this sample, allowing a direct comparison with the $WINGS$ data. The fit of the FP with our MIST procedure provides the following coefficients for the high redshift dataset: $$\log(\re)=0.74\log(\sigma)-0.81\log(\langle I\rangle_e)+3.85$$ with errors in the coefficients $\pm0.13$, $\pm0.04$ and $\pm0.34$, respectively. In this formulation  is given in $pc$ and $\langle I\rangle_e$ in $L_{\odot}/pc^2$. These values of the FP coefficients are very close to those obtained by @Sperello. Comparing these coefficients with those observed in the $B$ band rest-frame for Coma by [@Jorg96], the authors concluded that the strong [*tilt*]{} difference between low and high redshift FPs implies the existence of a “downsizing” mechanism,  a different time-scale variation of the dynamical $M/L$ ratio for low and high mass galaxies. In addition, they showed that the $M/L - M$ relation derived from the FP at high redshift is quite different from that derived using a low redshift dataset of ETGs. In Paper-I, based on a much larger galaxy sample, we found that the FP is likely a bent surface (see also @Desroches) and that this produces a systematic variation of the FP coefficients at varying the bright-end cut-off of the galaxy sample. Trying to check if this effect could result in a sort of [ *Malmquist bias*]{} when moving at high redshift, we have obtained, for each galaxy in our sample, a rough estimate of the dynamical mass $M_{tot}$ from the virial eq. (\[eqMtot\]) using the expression of $K_V$ (as a function of $n$) proposed by [@Bertin] for one-component, spherical, non-rotating, isotropic  stellar systems: $$\label{eq3} K_V(n)=\frac{73.32}{10.465+(n-0.94)^2}+0.954.$$ Figure \[Fig17\_FP\] reports the slopes of the best-fit regressions $M_{tot}/L-M_{tot}$ obtained for different cut-off magnitudes from our galaxy sample in the $V$- and the $K$-bands (black and grey dots, respectively). ![The slopes of the $M_{tot}/L - M_{tot}$ relation, obtained for different cuts in absolute magnitudes in the $V$ (black dots) and $K$ (grey dots) bands. The number of galaxies used for each fit are reported close to the dots. The horizontal line corresponds to the slope of the high redshift $M_{tot}/L_B-M_{tot}$ relation derived from the data of @Sperello[]{data-label="Fig17_FP"}](Fig17_FP.ps){width="90mm"} This figure shows that the slope of the $M_{tot}/L-M_{tot}$ relation in the $V$ and $K$ bands increases when the cut-off luminosity increases (as it occurs for high redshift data sample that progressively loose faint objects). The same result is obtained when the mass-to-light ratio is derived from the combination of the FP and VP equations (eq. \[fp\] and \[eqvir5\], respectively), taking into account the variation of the FP coefficients with the cut-off luminosity. Using the high redshift data provided by @Sperello [Table 2 therein], we derived a slope of 0.51 for the $M_{tot}/L_B - M_{tot}$ relation at high redshift (see the horizontal line in Figure \[Fig17\_FP\]). Note that Figure 8 in [@Sperello] gives a cut-off magnitude for their high redshift sample at $M_B\sim-20.1$, which roughly corresponds to $M_V\sim-21$ and $M_K\sim-24.2$ (assuming the typical colors of ETGs: $B-V\sim1$ and $B-K\sim4.2$). It follows that the trends shown in Figure \[Fig17\_FP\] could be (at least in part) due to the observed change in the FP coefficients and in the slopes of the $M/L - M$ relation that occur when the sample is progressively cut by simulating high redshift observations. We are aware that many works have shown in different ways that M/L evolves differently as a function of mass (see above references). Here, we just wish to warn that for high redshift samples the selection effects must be carefully taken into account before attributing entirely to the “downsizing” phenomenon the changes of the $M/L -M$ relation with redshift. Discussion and conclusions {#sec9} ========================== Our analysis strongly suggests that the FP properties can be explained only within an “hybrid” framework, in which several physical mechanisms are at work together. They are: [*(i)*]{} the non homologous structure and dynamics of ETGs; [*(ii)*]{} the stellar population variation along the sequence of galaxy masses; [*(iii)*]{} the increasing DM fraction within  at increasing the galaxy mass. In this work we have shown that the conspiracy between stellar population and galaxy structure, which has been invoked by C96 to explain the FP properties with just non-homology, does actually exist and takes the form of the nMML relation. According to this relation, for a given stellar mass, galaxies with high (low) Sersic index values, have high (low) values of $M^*/L$. The coupling of structure and the stellar population variations along the FP seems to be an important physical mechanism behind its properties, since it turns out to simultaneously explain the [*tilt*]{} and the scatter of the FP. The data also suggest that the above mentioned conspiracy between structure and stellar population is not the only mechanism at work. In fact, the DM fraction seems also to play a role in shaping the FP, although both the sign and the strength of the correlation with $\Delta_{FP}$ turn out to be different for Es and S0 galaxies (see Fig. \[Fig9\_FP\] and Tab. \[Tabatlas\]). This again support the “hybrid” scenario to explain the FP properties. Although the recent huge progresses of numerical simulations have produced massive ETGs resembling the real ones and obeying the most important scaling relations, at present we are not aware of theoretical models of galaxy formation and evolution able to simultaneously explain the FP, the nMML and the DM variation with stellar mass. Some interesting numerical simulations have shown that a series of minor merging events (either dry or partially wet), could modify the size of the galaxies and form extended stellar envelopes, while the stellar density and the velocity dispersion decreases with time [see @Nipotietal2009; @Bezanson; @Naab; @Hopkinsetal2009; @Oser]. This is the so called inside-out mechanism of galaxy formation, predicting that the oldest and most massive galaxies have very small sizes and are more compact than present day objects. Several observations seem to confirm such hypothesis [@Daddietal2005; @Trujilloetal2006; @vanDokkumetal2008; @Cimattietal2008; @vanderWeletal2008; @Damjanovetal2009; @Chevance see ], even if other works found that the size evolution of individual galaxies is modest when the age-size-mass relation is taken into account [@Valentinuzzi1; @Poggianti; @Saracco]. Unfortunately, the minor merging scenario largely fails to reproduce the numerical evolution of ETGs, given that the number density of ETGs turns out to be $\sim$25 times larger today than at z$\sim$2.5, thus preventing the size evolution of individual galaxies to be responsible of the claimed mass-size evolution. The semi-analytical models of hierarchical formation [see , @DeLuciaetal2006; @Almeidaetal2007; @Gonzalezetal2009; @Parryetal2009; @DeLuciaetal2011] have also successfully explained several observed features of ETGs, but fail in others. In particular they overestimate the number of faint objects, under predict the sizes of bright ETGs, and fail in reproducing the $\alpha$-enhancement that increases with the mass of the galaxies. Moreover, it is not clear how such models could be connected with the observed nMML relation. A primordial activity of merging at $z \ge 2$ is not excluded by [@MerlinChiosi2006; @MerlinChiosi2007], who suggested that a series of merging of small stellar subunits could lead to the formation of massive objects resembling the real ones in their morphology, density profiles, and metallicity. In their scheme, a first generation of star clumps inside primordial haloes of DM enriched the medium in metals and made up the building blocks of larger systems with masses up to $10^{12} M_{\odot}$. The hydro-dynamical simulations of @Merlinetal2012 include radiative cooling, star formation, stellar energy feedback, re-ionization, and chemical enrichment. They reproduce many observed features of ETGs, such as the mass-density profiles, the mass-radius relation, the mass-metallicity relation, etc. In their scheme objects with the same initial mass might have experienced different star formation histories depending on the initial halo over-density: the deeper the perturbation, the more peaked, early and intense the star forming activity. This kind of behavior could provide a viable explanation for the existence of the nMML relation. According to [@Nipoti], the dissipationless collapse of initially cold stellar distributions in preexisting DM haloes had an important role in determining the observed weak homology of ETGs. The end-products of their N-body simulations of collapses inside DM halos have significant structural non-homology, with the Sersic index spanning the interval $1.9 \leq n \leq 12$. Remarkably, their parameter $n$ correlates with the DM fraction present within , being smaller for larger dark-to-visible mass ratios. Unfortunately, such models do not follow the star formation activity, so we do not know if they are consistent with the nMML relation. @Hopkins found that also the dissipational collapse following the major merging of gas rich spirals could be usefully employed to interpret the FP properties, the size-mass and velocity dispersion-mass correlations. Their results, however, do not assign to non-homology a key role in [*tilting*]{} the FP, contrary to our finding. It is also not clear whether the major merging hypothesis, coupled with the dissipational collapse, might be in agreement with the nMML relation. We actually believe that still much theoretical work should be done before the problem of the formation of ETGs will find a robust and definitive solution within the “hybrid” scenario strongly supported by our analysis, whose main conclusions are the following: - the [*tilt*]{} of the FP in the $K$-band is significantly smaller than in the $V$-band, but still substantial; - other than to the change of the stellar populations with galaxy mass, the bulk of the [*tilt*]{} in both wavebands can be attributed to a systematic variation of the structure of ETGs as a function of mass. This is proved by the observed dependence of the quantity $\Delta_{FP}$ (difference between the FP and a reference VP) on the Sersic index $n$ and on the axis ratios $b/a$ of the galaxies. This “hybrid” interpretation of the FP [*tilt*]{}, based on our WINGS galaxy sample, is also confirmed by the analysis of the $ATLAS3D$ and $WSDSS$ samples, for which the influence of the stellar mass-to-light-ratio on the FP [*tilt*]{} turns out to be greater than for the $WINGS$ sample; - using the data from the $ATLAS3D$ project and at variance with their claims, we find that again the “hybrid” interpretation of the FP should be preferred, since dynamical non-homology and DM effects seems also to play a significant role in producing the [ *tilt*]{}; - both the Principal Component and the Multi-Variate Regression Analyses suggest that most of the previously mentioned physical factors significantly contribute to the FP [*tilt*]{}. Again, this points towards an “hybrid” solution of the FP problem; - the differential [*tilt*]{} in the $V$- and $K$-band should be ascribed to the entangled variation of structure and stellar population between the two bands. This variation reflects the mass dependent color gradients ($V-K$) likely originated by metallicity effects; - the [*tilt*]{} and the small scatter around the FP are likely originated by the conspiracy between structure and stellar population, through the relation $\log(M^*) - \log(n) - \log(M*/L)$ that works as a fine-tuning mechanism preserving the tightness of the FP plane along the whole sequence of masses. According to this relation, at increasing the stellar mass, ETGs become (on average) ’older’ and more centrally concentrated. Such twofold behaviour is due to the fact that the old stellar populations tend to be more centrally concentrated than the young ones, this fact being also responsible of the differential FP tilt between the V- and K-band; - the different slope of the dynamical $M/L - M$ relation at low and high redshifts must be used with care to derive the amount of the “downsizing” mechanism, since selection effects acting on the galaxy samples ( the Malmquist bias) mimic a differential rate of the $M/L$ evolution for small and massive ETGs. Acknowledgments {#acknowledgments .unnumbered} =============== We warmly thank the anonymous referee for his comments and suggestions that have greatly helped us to improve the paper. We also acknowledge the partial financial support by contract INAF/PRIN 2011 ÒGalaxy Evolution with the VLT Survey Tele- scope (VST)Ó and PRIN-MIUR 2009 “Nature and evolution of superdense galaxies”. Credits are finally due to the $SDSS$ and $ATLAS3D$ collaborations for the data used in this work. [99]{} Abazajian, K. Adelman-McCarthy, J. K., Agüeros, M. A. et al. 2009, ApJS, 182, 543 Allen P.D., Driver S.P., Graham A.W., Cameron E., Liske J., De Propris R., 2006, MNRAS, 371, 2 Almeida C., Baugh C. M., Lacey C. G., 2007, MNRAS, 376, 1711 Bell, E.F., & de Jong, R.S. 2001, ApJ 550, 212 Bender R., Burstein D., Faber S.M., 1992, ApJ 399, 462 Bernardi M., Sheth R.K., Annis J., Burles S., Eisenstein D.J., et al. 2003, ApJ, 125, 1866 Bertelli G., Bressan A., Chiosi C., Fagotto F., & Nasi E. 1994, A&AS, 106, 275 Bertin G., Ciotti L., Del Principe M., 2002, MNRAS, 386, 149 Blanton M.R., Lupton R.H., Schlegel D.J., Strauss M.A., Brinkmann J., Fukugita M., Loveday J., 2003, ApJ, 594, 186 Blanton M.R., Roweis S., 2007, AJ, 133, 734 Bezanson R., van Dokkum P.G., Tal T., Marchesini D., Kriek M., Franx M., Coppi P., 2009, ApJ 697, 1290 Bruzual G., Charlot S., 2003, MNRAS, 344, 1000 Bindoni D., et al., A&A in preparation Binney J., 1978, MNRAS, 183, 501 Binney J., Tremaine S.D., 1987, Galactic Dynamics. Princeton Univ. Press, NJ Borriello A., Salucci P., Danese L., 2003, MNRAS, 341, 1109 Bolton A.S., Treu T,, Koopmans L.V. E., et al., 2008, ApJ, 684, 248 Bruzual, G.; Charlot, S. 2003, MNRAS, 344, 1000 Burkert A., 1993, A&A, 278, 23 Busarello G., Capaccioli M., Longo G., Puddu E., 1997, in: The Second Stromlo Symposium “The nature of Elliptical Galaxies”, ASP Conference Series, 166, 184 Caon N., Capaccioli M., D’Onofrio M., 1993, MNRAS, 265, 1013 Capaccioli M., 1987, in Structure and dynamics of elliptical galaxies, ed. P.T. de Zeeuw (Reidel, Dordrecht), p. 47 Capaccioli M., 1989, in The world of galaxies, ed. H.G. Corwin & L. Bottinelli (Springer-Verlag, Berlin), p. 208 Cappellari M., Bacon R., Bureau M., et al., 2006, MNRAS, 366, 1126 Cappellari M., 2008, MNRAS, 390, 71 Cappellari M., Emsellem E., Bacon R., et al., 2007, MNRAS, 379, 418 Cappellari M., Emsellem E., Krajnović D., McDermid R.M., Scott N., Verdoes Kleijn G. A., 2011, MNRAS 413, 813 Cappellari M., McDermid R.M., Alatalo K., et al., 2012, \[arXiv:1202.3308v1\] Cappellari M., McDermid R.M., Alatalo K., et al., 2012, \[arXiv:1208.3523\] Cappellari M., Scott, N., Alatalo K., et al., 2013, \[arXiv:1208.3522\], in press Cava A., Bettoni D., Poggianti B. M., Couch W. J., et al. 2009, A&A 495, 707 Chevance M., Weijmans A.-M., Damjanov I., Abraham R.G., Simard L., van den Bergh S., Caris E., Glazebrook K., 2012, ApJL 754, L24 Chiosi C., Bressan A., Portinari L., Tantalo R., 1998, A&A, 339, 355 Chiosi C., Carraro G., 2002, MNRAS, 335, 335 Cimatti, A., Mignoli M., Daddi E., et al., 2002, A&A, 392, 395 Cimatti A., Cassata P., Pozzetti L., Kurk J., Mignoli M., Renzini A., Daddi E., 2008, A&A 482, 21 Ciotti L., Lanzoni B., Renzini A., 1996, MNRAS, 282, 1 Daddi E., Renzini A., Pirzkal N., Cimatti A., Malhotra S., Stiavelli M., Xu C., 2005, ApJ 626, 680 Damjanov I., McCarthy P.J., Abraham R.G.; Glazebrook K., Yan H., Mentuch E., ApJ 695, 101 Davies R.L., Efstathiou G., Fall S.M., Illingworth G., Schechter P. L., 1983, ApJ, 266, 41 de Carvalho, R.R., da Costa L.N., 1988, ApJS, 68, 173 de la Rosa, I.G., de Carvalho, R.R., Zepf, S.E., 2001, AJ 122, 93 De Lucia G., Springel V., White S. D. M., Croton D., Kau?mann G., 2006, MNRAS, 366, 499 De Lucia G., Fontanot F., Wilman D., Monaco P., 2011, MNRAS, 414, 1439 Dekel A., Cox T.J., 2006, MNRAS 370, 1445 Desroches L.B., et al. 2007, MNRAS, 377, 402 de Vaucouleurs, G., de Vaucouleurs, A., Corwin, H. G., Jr., Buta, R. J., Paturel, G., & Fouqué, P. 1991, Third Reference Catalogue of Bright Galaxies, Springer, New York, NY (USA) di Serego Alighieri S., Vernet J., Cimatti A., et al., 2005, A&A, 442, 125 Djorgovski S., Davis M., 1987 ApJ 313, 59 Djorgovski S., De Carvalho R., Han S.M., 1988, ASPC, 4, 329 Djorgovski S., Santiago B.X., 1993, in “Structure, Dynamics, and Chemical Evolution of Early Type Galaxies”, Danziger I.J., Zeilinger W.W., Kj[ä]{}r K., eds. p. 59 D’Onofrio M., Fasano G., Varela J., Bettoni D., Moles M., Kj[æ]{}rgaard P., Pignatelli E., Poggianti B., Dressler A., Cava A., and 3 co-authors, 2008, ApJ 685, 875 D’Onofrio M., Valentinuzzi T., Fasano G., et al., 2011, ApJL 727, 6 Dressler A., Lynden-Bell D., Burstein D., Davies R.L., Faber S.M., Terlevich R.J., Wegner G., 1987, ApJ 313, 42 Emsellem E., Monnet G., Bacon R., 1994, A&A 285, 723 Emsellem E., Cappellari M., Krajnović D., Alatalo K., Blitz L., Bois M., et al., 2011, MNRAS 414, 888 Faber S.M., Dressler A., Davies R., Burstein D., Lynden-Bell D., 1987, in: Nearly normal galaxies: From the Planck time to the present; Proceedings of the Eighth Santa Cruz Summer Workshop in Astronomy and Astrophysics, Santa Cruz, CA, July 21-Aug. 1, 1986 (A88-18401 05-90). New York, Springer-Verlag, 1987, p. 175-183 Fasano G., Marmo C., Varela J., D’Onofrio M., et al. 2006, A&A 445, 805 Fasano G., Vanzella, E., Dressler A., et al., 2012, MNRAS 420, 926 Forbes D.A., Ponman T.J., Brown R.J.N., 1998, ApJ 508, L43 Fritz J., Poggianti B. M., Bettoni D., et al. 2007, A&A, 470, 137 Fritz J., Poggianti B. M., Cava A., et al. A&A 526, 45 Gallazzi A., et al. 2006, MNRAS 370, 1106 Gargiulo A., 2009, MNRAS, 397, 75 Gerhard O., Kronawitter A., Saglia R.P., Bender R., 2001, AJ, 121, 1936 Gonzalez J. E., Lacey C. G., Baugh C. M., Frenk C. S., Benson A. J., 2009, MNRAS, 397, 1254 Graham A., Colless M., 1997, MNRAS, 287, 221 Graves G.J., Faber S.M., Schiavon R.P., 2009, ApJ, 698, 1590 Hansson A., et al. 2012, A&A submitted Hjorth J., Madsen J., 1995, ApJ, 445, 55 Hyde J.B., Bernardi M., 2009, MNRAS, 396, 1171 Hopkins Ph.F., Cox T.J., Hernquist L., 2008, ApJ, 689, 17 Hopkins P. F., Hernquist L., Cox T. J., Keres D., Wuyts S., 2009, ApJ, 691, 1424 Kauffmann, G., Heckman, T. M., White, S. et al. (2003), MNRAS 341, 33 Kelson, D.D., Illingworth, G.D., van Dokkum, P.G., Franx, M., 2000, ApJ 531, 137 Kelson, D.D., Illingworth, G.D., van Dokkum, P.G., Franx, M., 2000, ApJ 531, 137 Kochanek, C.S., Falco, E.E., Impey, C.D., Lehár, J., McLeod, B.A., et al. 2000, ApJ 543, 131 Kroupa P., 2001, MNRAS, 322, 231 Kroupa P., 2012, \[arXiv:1210:1211\] Jorgensen Franx, Kjaergaard 1996 Jorgensen I., Chiboucas K., Flint K., Bergmann M., Barr J., Davies R., 2006, ApJ, 639, L9 La Barbera F., Busarello G., Capaccioli M., 2000, A&A, 362, 851 La Barbera F., de Carvalho R. R., de La Rosa I. G., Lopes P. A. A., 2010 MNRAS 408, 1335 La Barbera F., De Carvalho R. R., De La Rosa I. G., et al., 2010 ApJ, 140, 1528 Lupton R., 2005, [*http://www.sdss.org/dr5/algorithms/sdssUBVRITransform.html*]{} McCarthy I. G., Font A. S., Crain R. A., Deason A. J., Schaye J., Theuns, T., 2012, MNRAS, 420, 2245 Magoulas C., et al. 2012, MNRAS 427, 245 Merlin E., Chiosi C., 2006, A&A 457, 437 Merlin E., Chiosi C., 2007, A&A 473, 733 Merlin E., Chiosi C., Piovan L., Grassi T., Buonomo U., 2012, MNRAS in press \[ArXiv:1204.5118\] Michard R., 1985, A&AS, 59, 205 Naab T., Johansson P.H., Ostriker J.P., 2009, ApJL 699, 178 Nipoti C., Londrillo P, Ciotti L., 2006, MNRAS, 370, 681 Nipoti C., Treu T., Bolton S., 2009, ApJ 703, 1731 O[ñ]{}orbe J., Domínguez-Tenreiro R., Sáiz A., Serna A., Artal H., 2005, ApJ 632, L57 Oser L., Naab T., Ostriker J. P., Johansson P. H., 2012, ApJ, 744, 63 Padmanabhan N., Seljak U., Strauss, M.A., et al. 2004, New Astronomy, 9, 329 Pahre, M.A., Djorgovski, S.G., de Carvalho, R,R., 1998, AJ, 116, 1591 Pahre M.A., De Carvalho R.R., Djorgovski S.G., 1998, AJ, 116, 1606 Parry O. H., Eke V. R., Frenk C. S., 2009, MNRAS, 396, 1972 Paturel G., Petit C., Prugniel P., Theureau G., Rousseau J., Brouty M., Dubois P., Cambr[é]{}sy L., 2003, A&A, 412, 45 Pignatelli E., Fasano G., Cassata P., 2006, A&A, 446, 373 Poggianti B.M., Calvi R., Bindoni D., D’Onofrio M., Moretti A.,Valentinuzzi T., 2012, ApJ in press \[ArXiv:1211.1005\] Poveda A., 1958, Boletín de los Observatorios de Tonantzintla y Tacubaya Vol. 2, Num. 17, pp. 3-7 Press W.H., Teukolsky S.A., Vetterling W.T., Flannery B.P., 1992, [*Numerical Recipes in Fortran: the art of scientific computing*]{}, Cambridge University Press, Second Edition. Prugniel Ph., Simien F., 1994, A&A, 282, L1 Prugniel Ph., Simien F., 1997, A&A, 321, 111 Renzini A., Ciotti L., 1993, ApJ, 416, L49 Renzini A., 1995, in “Stellar Populations”, Gilmore G., van der Kruit P. eds., p. 325 Robertson B., Cox T.J., Hernquist L., et al., 2006, ApJ, 641, 21 Salim, S. Rich, R. M,, Charlot, S. et al. (2007), ApJS, 173, 267 Salpeter E.E., 1955, ApJ 121, 161 Saracco P., Longhetti M., Gargiulo A., 2012, MNRAS 422, 3107 Schombert J.M., 1986, ApJS, 60, 603 Scodeggio M., Gavazzi G., Belsole E., Pierini D., Boselli A., 1998, MNRAS, 301, 1001 Sersic J.L., 1968, Atlas de galaxias australes, Observatorio Astronomico de Cordoba Smith R.J., Hudson M.J., Nelan J. E., et al., 2004, AJ, 128, 1558 Terlevich A.I., Forbes D.A., 2002, MNRAS, 330, 547 Tortora C., Napolitano N. R., Romanowsky A. J., Capaccioli M., Covone G., 2009, MNRAS, 396, 1132 Tortora C., Napolitano N. R., Cardone V. F., Capaccioli M., Jetzer Ph., Molinaro R., 2010, MNRAS, 407, 144 Treu T., Ellis R.K., Liao T.X., van Dokkum P.G., 2005, ApJ, 622, L5 Trujillo I., Burkert A., Bell E.F., 2004, ApJ, 127, 1917 Trujillo I., Feulner G., Goranova Y., Hopp U., Longhetti M., Saracco P., 2006, MNRAS 373, L36 Valentinuzzi T., Woods D., Fasano G., Riello M., D’Onofrio M., Varela J., Bettoni D., Cava A., Couch W.J., Dressler A., and 5 coauthors, 2009, A&A, 501, 851 Valentinuzzi T., Poggianti B. M., Saglia, R. P., et al., 2010, ApJ 721, 19 Valentinuzzi T., Fritz J., Poggianti B. M., et al., 2010, ApJ 712, 226 van der Wel A., Franx M., van Dokkum P. G., et al., ApJ, 631, 145 van der Wel A., Holden B.P., Zirm A.W., Franx M., Rettura A., Illingworth G.D., Ford H.C., 2008, ApJ 688, 48 van Dokkum G., Stanford S.A., 2003, ApJ 585, 78 van Dokkum P.G., Franx M., Kriek M., Holden B., Illingworth G.D., Magee D., Bouwens R., 2008, ApJL 677, L5 Varela J., D’Onofrio M., Marmo C., et al., 2009, A&A, 497, 667 Vulcani B., Poggianti B. M., Aragón-Salamanca A., et al. 2011, MNRAS, 412, 246 Wuyts S., Forster S., Natascha M., et al. 2011, ApJ 742, 96 Young C.K. & Currie M.J., 1994, MNRAS, 268, 11 \[lastpage\] [^1]: E-mail:[email protected]
{ "pile_set_name": "ArXiv" }
--- abstract: 'Pulsar timing arrays act to detect gravitational waves by observing the small, correlated effect the waves have on pulse arrival times at Earth. This effect has conventionally been evaluated assuming the gravitational wave phasefronts are planar across the array, an assumption that is valid only for sources at distances $R\gg2\pi{}L^2/\lambda$, where $L$ is physical extent of the array and $\lambda$ the radiation wavelength. In the case of pulsar timing arrays (PTAs) the array size is of order the pulsar-Earth distance (kpc) and $\lambda$ is of order pc. Correspondingly, for point gravitational wave sources closer than $\sim100$ Mpc the PTA response is sensitive to the source parallax across the pulsar-Earth baseline. Here we evaluate the PTA response to gravitational wave point sources including the important wavefront curvature effects. Taking the wavefront curvature into account the relative amplitude and phase of the timing residuals associated with a collection of pulsars allows us to measure the distance to, and sky position of, the source.' author: - | Xihao Deng${}^1$ and Lee Samuel Finn${}^{1,2}$\ ${}^{1}$Department of Physics, The Pennsylvania State University, University Park PA 16802\ ${}^{2}$Department of Astronomy and Astrophysics, The Pennsylvania State University, University Park PA 16802 title: Pulsar timing array observations of gravitational wave source timing parallax --- \[firstpage\] gravitational waves – methods: observational Introduction ============ It has been just over thirty years since @sazhin:1978:ofd and @detweiler:1979:ptm showed how passing gravitational waves cause disturbances in pulsar pulse arrival times that could be used to detect or limit gravitational wave signal strength, and twenty years since @foster:1990:cpt proposed using correlated timing residuals of a collection of pulsars — i.e., a *pulsar timing array* (PTA) — to achieve greater sensitivity. Thought of as a gravitational wave detector a PTA is large: its size $L$ (on order kpc) is much greater than the gravitational radiation wavelength scale $\lambda$ (on order pc) that we use it to probe. The conventional approximation of gravitational waves as plane-fronted applies only for sources at distances $R\gg{}L^2/\lambda$: for less distant sources the curvature of the gravitational radiation phasefronts contribute significantly to the detector response. Here we evaluate the response of a PTA to discrete gravitational wave sources at distances close enough that gravitational wave phasefront curvature is important and describe how PTA observations of such sources may be used to measure the source distance and location on the sky. Pulsar timing currently provides the best observational evidence for gravitational waves: as of this writing the orbital decay of the binary pulsar system PSR B$1913+16$, induced by gravitational wave emission and measured by timing measurements, is the most conclusive observational evidence for gravitational waves [@taylor:1979:mog; @weisberg:2005:rbp]. This evidence is often described, without any intent of disparagement, as *indirect* since it is the effect of gravitational wave *emission* — i.e., the effect of the emission on the emitter — that is observed.[^1] The effect of the gravitational radiation, independent of its source, on another system — a *detector* — itself remains to be observed. As proposed by @sazhin:1978:ofd and @detweiler:1979:ptm precision pulsar timing can also be used to detect gravitational waves *directly.* Gravitational waves passing between the pulsar and Earth— i.e., temporal variations in the space-time curvature along the pulsar-Earth null geodesics — change the time required by successive pulses to travel the path from pulsar to Earth. Since successive pulses are emitted from the pulsar at regular intervals the effect of passing gravitational waves are apparent as irregularities in the observed pulse arrival times. Through the use of precision timing a collection of especially stable pulsars may thus be turned into a galactic scale gravitational wave detector. Timing of individual pulsars has been used to bound the gravitational wave power associated with stochastic gravitational waves of cosmological origin [@romani:1983:ulo; @kaspi:1994:hto; @lommen:2002:nlo]; timing residual correlations among pulsar pairs have been used to bound the power in an isotropic stochastic gravitational wave background [@jenet:2006:ubo]; the absence of evidence for gravitational waves in the timing of a single pulsar has been used to rule-out the presence of a proposed supermassive black hole binary in 3C 66B [@jenet:2004:cpo]; observations from multiple pulsars have been used to place limits on periodic gravitational waves from Sag $\mathrm{A}^*$ [@lommen:2001:upt]; and timing analyses of individual pulsars have been combined to place a sky-averaged constraint on the merger rate of nearby black-hole binaries in the early phases of coalescence [@yardley:2010:sop]. Additionally, analyses of pulsar timing data have been proposed to search for the gravitational wave “memory” effect [@van-haasteren:2010:gma; @seto:2009:sfm; @pshirkov:2010:ogw] and more general gravitational wave bursts [@finn:2010:dla], which might be generated during, e.g., the periapsis passage of a highly eccentric binary system [@sesana:2010:scm] or by cosmic string cusps or kinks [@damour:2001:gwb; @siemens:2007:gsb; @leblond:2009:gwf]. Analyses focused on discovering or bounding the intensity of a stochastic gravitational wave “background” may rightly resolve the radiation into a superposition of plane waves with density $\partial^3\widetilde{\mathbf{h}}/\partial f\partial^2\Omega$, where $\mathbf{h}(t,\vec{x})$ is the gravitational wave strain and $(f,\Omega)$ denotes a plane wave component at frequency $f$ and propagating in direction $\Omega$. Analyses focused on the discovery of gravitational wave point sources — e.g., supermassive black hole binaries — must take proper account of the curvature of the radiation phase-front, owing to the finite distance between the source and the detector, over the detector’s extent. When detector extent, measured in units of the radiation wavelength, is greater than the distance to the source, measured in units of detector extent, then curvature of the radiation wavefront over the detector contributes significantly to the detector response to the incident radiation. In §\[sec:resp\] we evaluate the pulse arrive-time disturbance associated with a spherically-fronted gravitational wave traversing a pulsar-Earth line-of-sight and compare it with the response to the same wave in the plane-wave approximation. In §\[sec:disc\] we describe how, owing to their sensitivity to gravitational radiation phasefront curvature, PTA observations of point sources can be used to measure or place lower bounds on the distance to the source distance and, more surprisingly, the distance to the array pulsars. Finally we summarize our conclusions in §\[sec:concl\]. Pulsar timing residuals from spherically-fronted gravitational waves {#sec:resp} ==================================================================== Timing residuals {#sec:residuals} ---------------- Focus attention on the electromagnetic field associated with the pulsed emission of a pulsar and denote the field phase, at the pulsar, as $\phi_0(t)$. We are interested in the time-dependent phase $\phi_0(t)$ of the electromagnetic field associated with the pulsed emission measured at an Earth-based radio telescope, which we write as $$\phi(t) = \phi_0[t-L-\tau_0(t)-\tau_{\mathrm{GW}}(t)]$$ where $$\begin{aligned} \tau_0 &= \left(\begin{array}{l} \text{Corrections owing exclusively to the spatial motion of the Earth}\\ \text{within the solar system, the solar system with respect to the pulsar,}\\ \text{and electromagnetic wave propagation the interstellar medium} \end{array}\right)\\ \tau_{\mathrm{GW}} &= \left(\begin{array}{l} \text{Corrections owing exclusively to $\mathbf{h}(t,\mathcal{X})$} \end{array}\right)\\ L &= \left(\text{Earth-pulsar distance}\right). \end{aligned}$$ (Note that we work in units where $c=G=1$.) In the absence of gravitational waves $\tau_{\mathrm{GW}}$ vanishes and the phase front $\phi_0(t)$ arrives at Earth at time $t_\oplus(t) = t+L+\tau_0(t)$. In the presence of a gravitational wave signal the phase front arrives at time $t_\oplus(t)+\tau_{\mathrm{GW}}(t)$; thus, $\tau_{\mathrm{GW}}$ is the gravitational wave timing residual. Following @finn:2009:roi Eqs. (3.26) and (3.12e), the arrival time correction $\tau_{\mathrm{GW}}(t)$ is \[eq:tauGW\] $$\begin{aligned} \tau_{\mathrm{GW}}(t) &= -\frac{1}{2}\hat{n}^l\hat{n}^m\mathcal{H}_{lm}(t)\end{aligned}$$ where $\mathcal{H}_{lm}$ is the integral of the transverse-traceless metric perturbation over the null geodesic ranging from the pulsar to Earth: $$\begin{aligned} \mathcal{H}_{lm}(t) &= L\int_{-1}^0 h_{lm}\left(t+L\xi,\mathcal{P}-L(1+\xi)\hat{n}\right)\,d\xi \label{eq:scrH}\\ \mathcal{P} &= \left(\text{Pulsar location}\right)\\ \hat{n} &= \left(\text{Unit vector pointing from Earth to pulsar}\right).\end{aligned}$$ Gravitational waves from a compact source {#sec:gravWaves} ----------------------------------------- At a point $\mathcal{X}$ in the perturbative regime far from the source the transverse-traceless gauge gravitational wave metric perturbation associated with radiation from a compact source may be written [@finn:1985:gwf] \[eq:sphH\] $$\begin{aligned} h_{lm}(t,\mathcal{X}) &= \int_{-\infty}^{\infty} \widetilde{h}_{lm}(f,\hat{k}) e^{-2\pi ift}df\\ &=\frac{1}{\left|\mathcal{X}-\mathcal{S}\right|}\int_{-\infty}^{\infty} \left[\widetilde{A}_{lm}(f,\hat{k}) e^{2\pi if|\mathcal{X}-\mathcal{S}|}\right]e^{-2\pi ift}df\end{aligned}$$ where $$\begin{aligned} \mathcal{S} &= \left(\text{source location}\right)\\ \hat{k} &= \frac{\mathcal{X}-\mathcal{S}}{\left|\mathcal{X}-\mathcal{S}\right|} = \left(\text{unit vector in direction of wave propagation at $\mathcal{X}$}\right)\end{aligned}$$ The Fourier coefficient functions $\widetilde{A}_{lm}(f,\hat{k})$ are everywhere symmetric, traceless and transverse with respect to $\mathcal{X}-\mathcal{S}$: i.e., $$\begin{aligned} \widetilde{A}_{lm}(f,\hat{k}) &= \widetilde{A}_{+}(f,\hat{k})\mathbf{e}^{(+)}_{lm} + \widetilde{A}_{\times}(f,\hat{k})\mathbf{e}^{(\times)}_{lm}\end{aligned}$$ where $\mathbf{e}^{(+)}_{lm}(\hat{k})$ and $\mathbf{e}^{(\times)}_{lm}(\hat{k})$ are the usual transverse-traceless gravitational wave polarization tensors for waves traveling in direction $\hat{k}$. Response function ----------------- With expression \[eq:sphH\] for $h_{lm}$ we can evaluate $\mathcal{H}_{lm}$ in the Fourier domain: \[eq:tildeScrH\] $$\begin{aligned} \widetilde{\mathcal{H}}_{lm}(f) &= L\int_{-1}^0 \widetilde{h}_{lm}\left(f,\mathcal{P}-L(1+\xi)\hat{n}\right) e^{-2\pi i fL\xi} \,d\xi\\ &= L\int_{-1}^0 \frac{\widetilde{A}_{lm}\left(f,\hat{k}(\xi)\right)}{\left|\mathcal{P}-L(1+\xi)\hat{n}-\mathcal{S}\right|} \exp\left[{2\pi i f\left(-L\xi+|\mathcal{P}-L(1+\xi)\hat{n}-\mathcal{S}|\right)}\right] \,d\xi\end{aligned}$$ Figure \[fig:geometry\] describes the Earth-pulsar-source geometry: note we define $R$ to be the Earth-source distance and $\theta$ the angle between the Earth-pulsar and Earth-source lines-of-sight. Along the path traversed by the pulsar’s electromagnetic pulse phase fronts ($\mathcal{P}-L\xi\hat{n}$ for $-1\leq\xi\leq0$) we may write $$\begin{aligned} \hat{k}(\xi) &= \frac{\mathcal{P}-L(1+\xi)\hat{n}-\mathcal{S}}{\left|\mathcal{P}-L(1+\xi)\hat{n}-\mathcal{S}\right|}\\ r(\xi) &= \left(\text{distance between gravitational wave phase front and source}\right)\nonumber\\ &= \left|\mathcal{P}-L(1+\xi)\hat{n}-\mathcal{S}\right| = R\left[1+2\frac{L\xi}{R}\cos\theta+\left(\frac{L\xi}{R}\right)^2\right]^{1/2}. \label{eq:r}\end{aligned}$$ From Equations \[eq:tildeScrH\] and \[eq:r\] we note that the time of arrival disturbance owing to a passing gravitational wave depends on the dimensionless quantities $\pi f L$ and $L/R$. PTA pulsar distances $L$ are all on-order kpc. Gravitational waves detectable via pulsar timing observations have frequencies greater than the inverse duration of the observational data set and less than the typical sampling period: i.e., $10^{-5}~Hz\gtrsim f\gtrsim10^{-9}$ Hz. Strong sources of gravitational waves (e.g., supermassive black hole binaries and triplets) in this band are all expected to be extragalactic, with significant numbers within on order 100 Mpc [@sesana:2010:gwa]. The scales of interest are thus either very large or very small: \[eq:numerology\] $$\begin{aligned} \epsilon &= 10^{-5}\left(\frac{L}{1\,\text{kpc}}\right)\left(\frac{100\,\text{Mpc}}{R}\right)\\ \pi fL &= 10^4\,\left(\frac{f}{1\,\text{yr}^{-1}}\right)\left(\frac{L}{1\,\text{kpc}}\right)\\ \pi fL\epsilon &= 10^{-1}\,\left(\frac{f}{1\,\text{yr}^{-1}}\right)\left(\frac{L}{1\,\text{kpc}}\right)^2\left(\frac{100\,\text{Mpc}}{R}\right).\end{aligned}$$ Taking advantage of these scales the Fourier transform $\widetilde{\mathcal{H}}_{lm}(f)$ may be expressed $$\begin{aligned} \widetilde{\mathcal{H}}_{lm}(f) &= L\int_{-1}^0 \widetilde{h}_{lm}\left(f,\mathcal{P}-L(1+\xi)\hat{n}\right) e^{-2\pi i fL\xi} \,d\xi\\ &= L\int_{-1}^0 \frac{\widetilde{A}_{lm}\left(f,\hat{k}(\xi)\right)}{\left|\mathcal{P}-L(1+\xi)\hat{n}-\mathcal{S}\right|} e^{2\pi i f\left(-L\xi+|\mathcal{P}-L(1+\xi)\hat{n}-\mathcal{S}|\right)} \,d\xi\\ &= \widetilde{A}_{lm}(f,\hat{k}_0) \frac{\exp\left[\pi i fR\left(2-\frac{1-\cos\theta}{1+\cos\theta}\right)\right]}{\sqrt{fR}\sin\theta} \frac{e^{-i\pi/4}}{2} \left. \operatorname{erf}\left[e^{i\pi/4}x\right] \right|^{\sqrt{\pi fR}\frac{\epsilon(1+\cos\theta)+1}{1+\cos\theta}\sin\theta}_{x=\frac{\sqrt{\pi fR}\sin\theta}{1+\cos\theta}} \nonumber\\ &\qquad{} \times\left[1+ \mathcal{O}(\epsilon) +\mathcal{O}\left(\epsilon\frac{\partial\log\tilde{A}}{\partial{\hat{k}}}\cdot\hat{n}\right) \right]\end{aligned}$$ Except when $\theta\leq\left(\pi fR\right)^{-1/2}\ll1$ the error function arguments are always large in magnitude. Taking advantage of the asymptotic expansion of $\operatorname{erf}(z)$ about the point at infinity, $$\begin{aligned} \operatorname{erf}(z) \sim 1 - \frac{\exp\left(-z^2\right)}{z\sqrt{\pi}}\end{aligned}$$ we may thus write \[eq:tau\] $$\begin{aligned} \widetilde{\tau}_{\text{gw}} &= -\frac{1}{2}\hat{n}^l\hat{n}^m\widetilde{\mathcal{H}}_{lm}(f) \\ &= \widetilde{\tau}_{\mathrm{pw}} + \widetilde{\tau}_{\mathrm{cr}}\end{aligned}$$ where $$\begin{aligned} \widetilde{\tau}_{\mathrm{pw}} &= \left( \text{plane wave approximation timing residual} \right)\nonumber\\ &= 2\widetilde{\mathcal{A}}(f,\hat{k}_0) \exp\left[2\pi ifL\sin^2\frac{\theta}{2}\right] \operatorname{sinc}\left(2\pi fL\sin^2\frac{\theta}{2}\right) \\ \widetilde{\tau}_{\mathrm{cr}} &= \left( \text{phasefront curvature correction to plane wave timing residual} \right)\nonumber\\ &= \epsilon\left(1+\cos\theta\right)\widetilde{\mathcal{A}}(f,\hat{k}_0) \exp\left[ \pi ifL\left(4\sin^2\frac{\theta}{2}+\frac{\epsilon}{2}\sin^2\theta\right) \right] \operatorname{sinc}\left[\frac{\pi fL\epsilon}{2}\sin^2\theta\right]\\ \widetilde{\mathcal{A}}(f,\hat{k}_0) &= \frac{i\epsilon}{4}\left[\hat{n}^l\hat{n}^m\widetilde{A}_{lm}(f,\hat{k}_0)\right]\exp\left[2\pi i fR\right]\end{aligned}$$ and $$\begin{aligned} \operatorname{sinc}(x) &=\left(\begin{array}{l}\text{Unnormalized}\\\text{sinc function}\end{array}\right) = \frac{\sin(x)}{x}.\end{aligned}$$ These expressions are valid for all $\pi fL\gg1$. ![Gravitational wave source (depicted as a binary system), Earth and pulsar. The gravitational wave phasefronts intersect the pulsar-Earth path, traveled by the electromagnetic pulses, in a manner that depends on the distance to the source and the source-Earth-pulsar angle $\theta$.[]{data-label="fig:geometry"}](fig1){width="4in"} Discussion {#sec:disc} ========== The Fourier amplitude and phase of the gravitational wave timing residual $\widetilde{\tau}_{\text{GW}}$ for any particular pulsar depends on the pulsar’s position and distance relative to the position and distance of the gravitational wave source. Correspondingly, observation in three or more pulsars of the gravitational wave timing residuals associated with a single source can, in principle, be used to infer the three-dimensional source position and location on the sky. In this section we discuss the conditions under which sufficiently accurate measurements of the $\widetilde{\tau}_{\text{GW}}$ are possible and provide a proof-of-principle demonstration of how they may be combined to determine the sky location and distance to a gravitational wave source. Plane wave approximation timing residuals {#sec:planeWave} ----------------------------------------- Focus attention first on $\widetilde{\tau}_{\text{pw}}$: the plane-wave approximation to $\widetilde{\tau}_{\text{gw}}$. In order to relate the angular location of a gravitational wave source to the relative amplitude and phase of the $\widetilde{\tau}_{\text{pw}}$ for two or more PTA pulsars the uncertainties $\sigma_L$ in the Earth-pulsar distances, and $\sigma_{\theta}$ in pulsar sky locations, must satisfy \[eq:sigmaLT\] $$\begin{aligned} \sigma_L&\lesssim 2\,\text{pc} \left(\frac{0.1\,\text{yr}^{-1}}{f}\right) \left(\frac{1/2}{\sin^2(\theta/2)}\right)\\ \sigma_\theta&\lesssim {2.5'}\left(\frac{0.1\,\text{yr}^{-1}}{f}\right)\left(\frac{\text{kpc}}{L}\right)\left(\frac{\pi/4}{|\sin\theta|}\right). $$ The constraint on pulsar angular position accuracy is not a severe one; however, the constraint on pulsar distance accuracy is quite significant. For this reason previous pulsar timing searches for periodic gravitational waves [@yardley:2010:sop; @jenet:2004:cpo; @lommen:2001:upt; @lommen:2001:pmt] have sought only to identify an anomalous periodic contribution to pulse arrival times, avoiding explicit use of the amplitude of $\widetilde{\tau}_{\text{gw}}$. Improvements in instrumentation and timing techniques are making possible pulsar distance measurements of the accuracy and precision necessary to relate the $\widetilde{\tau}_{\text{pw}}$ to the gravitational wave source angular location. For example, in recent years a succession of VLBI observations have established the distance to J$0437-4715$ with an accuracy of $\pm1.3\,\text{pc}$ [@verbiest:2008:pto; @deller:2008:ehp; @deller:2009:pva]; correspondingly, for J$0437-4715$ predictions of the amplitude and phase of $\widetilde{\tau}_{\text{pw}}$ can be made and take part in the analysis of timing data to search for gravitational waves at frequencies $f\lesssim0.1\,\text{yr}$. Accurate pulsar distance measurements have been identified as a crucial element of a continuing program to use precision pulsar timing and interferometry to test gravity and measure neutron star properties [@cordes:2009:tog]; correspondingly, we can reasonably expect that the number of pulsars with high-precision distances will increase rapidly over the next several years. Looking toward the future, observations with the SKA [@smits:2009:psa] using existing timing techniques are expected to be capable of measuring timing parallax distances to millisecond pulsars at 20 kpc with an accuracy of 20% [@tingray:2010:har]. When measured by VLBI or timing parallax the fractional uncertainty in the pulsar distance is equal to the fractional uncertainty in the semi-annual pulse arrival time variation owing to the curvature of the pulsar’s electromagnetic phasefronts: i.e., $$\begin{aligned} \frac{\sigma_L}{L} &= \frac{\sigma_{\tau}}{\Delta\tau}\end{aligned}$$ where $\sigma^2_{\tau}$ is the timing noise power in the bandwidth $T^{-1}$ (for $T$ the observation duration) about a frequency of $2/yr$ and $$\begin{aligned} \Delta\tau &= \left(\frac{\text{au}}{2c}\right)\left(\frac{\text{au}}{L}\right) = 1.2\,\mu\text{s}\,\left(\frac{\text{kpc}}{L}\right)\end{aligned}$$ is the peak-to-peak variation in the pulse arrival time residual owing to the curvature of the pulsar’s electromagnetic phasefronts. Correspondingly, a 20% accuracy measurement of the distance to a pulsar at 20 kpc corresponds to an 0.5% accuracy measurement of the distance to a pulsar of the $\sigma_{\tau}$ at 500 pc: i.e., 2.5 pc. This estimate of $\sigma_L=2.5$ pc for a pulsar at 500 pc assumes constant signal-to-noise and, correspondingly, constant timing precision $\sigma_{\tau}$. Closer pulsars will have greater fluxes and, correspondingly, better timing precision; consequently we may regard this as an upper bound on the distance measurement error $\sigma_L$. If we assume that $\sigma_{\tau}^{-2}$ is proportional to the pulsar energy flux then \[eq:sigmaL\] $$\begin{aligned} \sigma_{\tau} &= 0.6\,\text{ns}\,\left(\frac{L}{\text{kpc}}\right)\\ \sigma_L &= 0.5\,\text{pc}\,\left(\frac{L}{\text{kpc}}\right)^3; \end{aligned}$$ i.e., if the only barrier to improved timing precision is signal-to-noise then sub-parsec precision distance measurements should be possible in the SKA era for pulsars within a kpc. In practice other contributions to the noise budget will eventually limit the timing precision. Among these the most important are intrinsic pulsar timing noise, pulse broadening owing to scattering in the ISM, tropospheric scintillation. Recent improvements in pulsar timing techniques [@lyne:2010:smr] suggest that intrinsic pulsar timing noise need not limit timing precision measurements in either present or future instrumentation. Unresolved pulse broadening owing to scattering in the ISM will limit the attainable $\sigma_{\tau}$ and, thus, $\sigma_L$. For nearby ($L<1$ kpc) pulsars the *measured* pulse broadening at $\nu=1$ GHz is typically in the 1–10 ns range [@manchester:2005:atn]. Since pulse broadening is scales with observation frequency as $\nu^{-4.4}$ (though observations suggest the index may be closer to $-3.9$ [@bhat:2004:moo]) scattering in the ISM also need not limit our ability to measure the distance to nearby pulsars. Finally, recent estimates by @jenet:2010:pts indicate that, for gravitational wave frequencies $\lesssim\,\text{yr}^{-1}$, tropospheric scintillation will contribute to the noise budget at no more than the 0.03 ns rms level. Consequently, over the next decade $\sigma_L\lesssim1\,\text{pc}$ is thus a reasonable expectation for millisecond pulsars at distances up to a kpc. Correction to plane wave approximation -------------------------------------- Turn now to the ratio of Fourier amplitudes $$\begin{aligned} \rho(f,\hat{k}_0) &= \frac{\widetilde{\tau}_{\text{cr}}}{\widetilde{\tau}_{\text{pw}}}\nonumber\\ &= \exp\left[ \pi ifL\left(2\sin^2\frac{\theta}{2}+\frac{\epsilon}{2}\sin^2\theta\right) \right]\nonumber\\ &\qquad{}\times \frac{ \sin\left[\frac{1}{2}\pi fL\epsilon\sin^2\theta\right] }{ \sin\left(2\pi fL\sin^2\frac{\theta}{2}\right) }. \label{eq:rho}\end{aligned}$$ When $\pi fL\epsilon\sin^2\theta\geq1$ each of the $\sin$ functions whose ratio determines the $|\rho|$ has order-unity magnitude and $|\rho|$ oscillates rapidly from $0$ to $\infty$. For a typical PTA pulsar (distance $L\simeq$ kpc) this will be the case for source frequencies greater than or of order $\text{yr}^{-1}$ and distances $R$ less than or of order 100 Mpc (cf. Eqs. \[eq:numerology\]). In this regime the gravitational wave phasefront curvature plays an essential role in determining the response of a pulsar timing array. As an example consider the residuals in the pulse arrival times for PSR J$1939+2134$ owing to gravitational waves from a supermassive black hole binary system like that proposed by @sudou:2003:omi in the radio galaxy 3C 66B at RA 2h23m11.4s, Dec $42^{\text{o}}$59’31” and luminosity distance $R=85.8$ Mpc [@ned:2010:ned]. The binary, ruled-out by the subsequent analysis of @jenet:2004:cpo, was supposed to have a total mass of $5.4\times10^{10}\mathrm{M_{\odot}}$, a mass ratio of 0.1, and a period of 1.05 yr, corresponding to a gravitational wave frequency of $1.9\,\text{yr}^{-1}$. Assuming these binary parameters to be exact and taking the distance and location of J$1939+2134$ as provided by the ATNF pulsar catalog [@manchester:2005:atn] yields $\rho=1.1\exp[-0.86\pi i]$: i.e., the magnitude of the “correction” $\tau_{\text{cr}}$ is greater than the “leading-order” plane-wave approximation contribution, the magnitude of $\tau_{\text{gw}}$ is 46% of that predicted by the plane wave approximation, and the phase of $\tau_{\text{gw}}$ is retarded by $\pi/2$ radians relative to $\tau_{\text{pw}}$.[^2] Gravitational wave source distance and location on sky {#sec:distances} ------------------------------------------------------ The magnitude and phase of the plane-wave approximation Fourier coefficient function $\widetilde{\tau}_{\text{pw}}$ depends on the distance to the associated pulsar and the pulsar-Earth-source angle. The correction depends on these and, in addition, the source distance. Measurements of $\widetilde{\tau}_{\text{gw}}$ in a collection of pulsars whose angular location and distance are known sufficiently accurately can thus be used to measure the curvature of the gravitational radiation phase-fronts and, thus, the luminosity distance and direction to the gravitational wave source. Consider a periodic source of gravitational waves observed in an array of pulsars whose distances and relative locations are known to high precision. The gravitational wave contribution to the pulse arrival times for array pulsar $\ell$ may be represented by an amplitude and phase, which will depend on the different $\epsilon_\ell=L_\ell/R$, $\pi fL_\ell$ and $\theta$ for each pulsar $\ell$ in the array. Requiring that these phases and amplitudes all be consistent is a powerful constraint on the source distance and location on the celestial sphere. As a crude but effective proof-of-principle demonstration of distance measurement by gravitational wave timing parallax consider observations made with a selection of IPTA pulsars with distances whose precision has been projected, as described above, into the SKA era. Focus attention on the example source in 3C 66B. For this source $\left(\pi f\right)^{-1} = 0.051\,\text{pc}$. Following Equation \[eq:sigmaL\], four pulsars currently monitored as part of the International Pulsar Timing Array are close enough that we may expect their distance to be measured with a one-sigma uncertainty of less than 0.025 pc. Table \[tbl:ipta\] lists these pulsars and their currently measured distances. For these four pulsars, 1. Calculate the four $\widetilde{\tau}_{\text{gw}}$ for the example source in 3C 66B described above. These complex amplitudes, corresponding to periodic timing residual amplitude and phase, are our “observations”. Denote the $\widetilde{\tau}_{\text{gw}}$ for pulsar $k$ by $\widetilde{\tau}_k$.\[step:one\] 2. Adopt approximate distances to each of these four pulsars consistent with normally-distributed measurement error with standard deviation as given in Equation \[eq:sigmaL\]. \[step:two\] 3. Using these approximate distances evaluate $$\begin{aligned} \psi^2(r) &= \sum_k \left|\log\left(\frac{\widetilde{\tau}'_{k}(r)}{\widetilde{\tau}_{k}}\right)\right|^2\end{aligned}$$ where $\widetilde{\tau}'_k(r)$ is the expected $\widetilde{\tau}_{\text{gw}}$ for pulsar $k$ *assuming the source at distance $r$ and pulsar $k$ at the approximate distance found in step* \[step:two\]. The quantity $\psi^2(r)$ is a measure of the misfit between the observations $\widetilde{\tau}_k$ and the prediction $\widetilde{\tau}'_k$ assuming pulsars at the approximate distances and source at distance $r$. The $r$ that minimizes $\psi^2(r)$ is thus an estimator for the distance to the source. Figure \[fig:pp\] shows a set of three scatter plots of the estimated $r$, declination $\theta$, and right ascension $\phi$ relative to their actual values over $10^4$ realizations of pulsar distance errors, for this example. Table \[tbl:stats\] provides quantitative descriptive statistics for the distribution of errors. Clearly, if the pulsar distances can be sufficiently accurately determined, the distance and location of a detected periodic gravitational wave source can be accurately measured. J-Name [kpc]{} -------------- --------- J$0437-4715$ 0.16 J$1741+1300$ 0.19 J$0030+0451$ 0.30 J$2124-3358$ 0.25 mean median std. dev. ---------- ------------------------ ------------------------ ----------------------- $r$ 85.72 Mpc 85.79 Mpc 0.99 Mpc $\theta$ $42^{\text{o}}59'31''$ $42^{\text{o}}59'31''$ $0^{\text{o}}0'5.3''$ $\phi$ 2h23m11.0s 2h23m11.0s 0h0m0.3s ![Scatter plots of the timing parallax estimates of the distance, declination and right ascension of a simulated gravitational wave source in 3C 66B (i.e., $r=87.8$ Mpc, dec=$2^{\text{o}}59'31''$ and right ascension 2h23m11). See §\[sec:distances\] for details.[]{data-label="fig:pp"}](fig2){width="3.5in"} We emphasize that this is no more than a proof-of-principle demonstration. It is naive in its treatment of uncertainties in the $L_k$, ignores pulsar timing noise in the measured $\widetilde{\tau}_k$, presumes a sufficiently accurate knowledge of pulsar locations $\theta_k$, and makes use of an ad hoc misfit statistic $\psi^2(r)$. Nevertheless, taken as a proof-of-principle demonstration it shows that the potential exists for pulsar timing array observations to determine the distance to periodic gravitational wave sources if the pulse arrival time dependence on the passing waves is properly modeled. Conclusion {#sec:concl} ========== Gravitational waves crossing the pulsar-Earth line-of-sight lead to a disturbance in the pulsar pulse arrival time. All previous derivations of this disturbance have assumed planar gravitational wave phasefronts. The gravitational wave phasefronts from point gravitational wave sources — e.g., coalescing supermassive black hole binary systems — are curved, with curvature radius equal to the source luminosity distance. The approximation of planar wavefronts is thus valid only at distances $R\gg2\pi fL^2/c$, where $R$ is the Earth-source distance, $L$ the Earth-pulsar distance, and $f$ the gravitational radiation frequency. For typical pulsars distances (kpc) and relevant gravitational wave frequencies ($f\lesssim\,\text{yr}^{-1}$) the correction to the disturbance magnitude and phase owing to phasefront curvature is thus significant for sources within $\sim100\,\text{Mpc}$. Here we have derived the curved wavefront corrections to the pulsar timing response for such “nearby” sources, described when they are important, and shown that future gravitational wave observations using pulsar timing arrays may be capable of measuring luminosity distances to supermassive black hole binary systems, and other periodic gravitational wave sources, approaching or exceeding 100 Mpc. The gravitational wave source distance measurement described here is properly considered a parallax distance measurement. The baselines over which the parallax is measured are, in this case, the timing array pulsar-Earth baselines. Crucial to the ability to observe the effects of gravitational wave phasefront curvature is a knowledge of the pulsar-Earth distance $L$ with an uncertainty $\sigma_L\lesssim\left(\pi f\right)^{-1}\sim2\,\text{pc}\,(0.1\text{yr}^{-1}/f)$. We argue that pulsar distance measurements of this accuracy, while beyond present capabilities (except in the case of J$0437-4715$, which is exceptionally bright and close), are within the capability of SKA-era observations for pulsars at distances $L\lesssim\,\mathrm{kpc}$. When considered against the cosmic distance ladder, the distance measurement described here involves three rungs to reach distances greater than tens of Mpc: first, the determination of the astronomical unit; second, the distance to the array pulsars; finally, the distance to the gravitational wave source. Since the method described here also provides precise source angular position the likelihood of identifying an electromagnetic counterpart (e.g., host galaxy) to the gravitational wave source is great, raising the possibility of a precise measurement of both the redshift and luminosity distance to a single object at cosmological distances, with the obvious consequences for independent verification of the parameters describing our expanding universe. Even in the absence of an electromagnetic counterpart measurement of the luminosity distance and angular location of a supermassive black hole binary system — the most likely source of periodic gravitational waves — enables the determination of the system’s so-called ”chirp mass”, or $(m_1 m_2)^{3/5}/(m_1+m_2)^{1/5}$, without the need to observe the binary’s evolution. We have only begun to plumb the potential of gravitational wave observations as a tool of astronomical discovery. While this potential will not be realized until gravitational waves are detected and observations become, if not routine, more than occasional, explorations like these will position us to more readily exploit the opportunities that future observations present. Acknowledgments {#acknowledgments .unnumbered} =============== We gratefully acknowledge helpful discussions with Jim Cordes, Rick Jenet, Andrea Lommen, Duncan Lorimer, David Nice, Scott Ransom, Alberto Sesana, Dan Stinebring, and Ben Stappers. This research has made use of the ATNF Pulsar Catalogue [@manchester:2005:atn] and the NASA/IPAC Extragalactic Database (NED). NED is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. This work was supported in part by National Science Foundation grants PHY 06-53462 (LSF) and the Penn State Physics Department (XD). N. D. R., [Cordes]{} J. M., [Camilo]{} F., [Nice]{} D. J., [Lorimer]{} D. R., 2004, ApJ, 605, 759 J., [Arzoumanian]{} Z., [Brisken]{} W., [Freire]{} P., [Kramer]{} M., [Lai]{} D., [Lasio]{} J., [McLaughlin]{} M., [Nice]{} D., [Stairs]{} I., [Weisberg]{} J., 2009, in astro2010: The Astronomy and Astrophysics Decadal Survey Vol. 2010, Tests of gravity and neutron star properties from precision pulsar timing and interferometry. pp 56–+ Damour T., Vilenkin A., 2001, Phys. Rev. D, 64, 064008 A. T., 2009, ArXiv e-prints, 0902.1000 A. T., [Verbiest]{} J. P. W., [Tingay]{} S. J., [Bailes]{} M., 2008, ApJL, 685, L67 S., 1979, ApJ, 234, 1100 Finn L. S., 1985, Class. Quantum Grav., 2, 381 Finn L. S., 2009, Phys. Rev. D, 79 L. S., [Lommen]{} A. N., 2010, ApJ, 718, 1400 Finn L. S., Sutton P. J., 2002, Phys. Rev. D, 65, 044022 Foster R. S., Backer D. C., 1990, ApJ, 361, 300 Jenet F. A., Armstrong J. W., Tinto M., 2010, Pulsar Timing Sensitivity to Very-Low-Frequency Gravitational Waves, In preparation F. A., [Hobbs]{} G. B., [van Straten]{} W., [Manchester]{} R. N., [Bailes]{} M., [Verbiest]{} J. P. W., [Edwards]{} R. T., [Hotan]{} A. W., [Sarkissian]{} J. M., [Ord]{} S. M., 2006, ApJ, 653, 1571 Jenet F. A., Lommen A., Larson S. L., Wen L., 2004, ApJ, 606, 799 Kaspi V. M., Taylor J. H., Ryba M., 1994, ApJ, 428, 713 Leblond L., Shlaer B., Siemens X., 2009, Phys. Rev. D, 79, 123519 Lommen A. N., 2001, PhD thesis, University of California, Berkeley, Berkeley, CA, U.S.A. A. N., 2002, in [Becker]{} W., [Lesch]{} H., [Tr[ü]{}mper]{} J., eds, Neutron Stars, Pulsars, and Supernova Remnants New limits on gravitational radiation using pulsars. pp 114–+ A. N., [Backer]{} D. C., 2001, ApJ, 562, 297 Lyne A., Hobbs G., Kramer M., Stairs I., Stappers B., 2010, Science, 329, 408 R. N., [Hobbs]{} G. B., [Teoh]{} A., [Hobbs]{} M., 2005, AJ, 129, 1993 NASA/JPL, 2010, NASA/IPAC Extragalactic Database M. S., [Baskaran]{} D., [Postnov]{} K. A., 2010, [MNRAS]{}, 402, 417 Romani R. W., Taylor J. H., 1983, ApJL, 265, L35 M. V., 1978, Soviet Astronomy, 22, 36 Sesana A., 2010, arXiv, 1006.0730v1 A., [Vecchio]{} A., 2010, Class. Quantum Grav., 27, 084016 N., 2009, [MNRAS]{}, 400, L38 X., [Mandic]{} V., [Creighton]{} J., 2007, Phys. Rev. Lett., 98, 111101 R., [Kramer]{} M., [Stappers]{} B., [Lorimer]{} D. R., [Cordes]{} J., [Faulkner]{} A., 2009, A&A, 493, 1161 H., [Iguchi]{} S., [Murata]{} Y., [Taniguchi]{} Y., 2003, Science, 300, 1263 Sutton P. J., Finn L. S., 2002, Class. Quantum Grav., 19, 1355 Taylor J. H., Fowler L. A., McCulloch P. M., 1979, Nature, 277, 437 Tingray S., Bignall H., Colegate T., Aben G., Nicolls J., Weston S., 2010, in Proceedings of SKA2010 - International SKA Science and Engineering Meeting The high angular resolution component of the [SKA]{}. University of Manchester, UK R., [Levin]{} Y., 2010, [MNRAS]{}, 401, 2372 J. P. W., [Bailes]{} M., [van Straten]{} W., [Hobbs]{} G. B., [Edwards]{} R. T., [Manchester]{} R. N., [Bhat]{} N. D. R., [Sarkissian]{} J. M., [Jacoby]{} B. A., [Kulkarni]{} S. R., 2008, ApJ, 679, 675 J. M., [Taylor]{} J. H., 2005, in Rasio F., Stairs I. H., eds, Binary Radio Pulsars The relativistic binary pulsar b1913+16. Astronomical Society of the Pacific, San Francisco, pp 25–31 D. R. B., [Hobbs]{} G. B., [Jenet]{} F. A., [Verbiest]{} J. P. W., [Wen]{} Z. L., [Manchester]{} R. N., [Coles]{} W. A., [van Straten]{} W., [Bailes]{} M., [Bhat]{} N. D. R., [Burke-Spolaor]{} S., [Champion]{} D. J., [Hotan]{} A. W., [Sarkissian]{} J. M., 2010, ArXiv e-prints, 1005.1667 \[lastpage\] [^1]: Indirect observation is enough to inaugurate the field of gravitational wave astronomy: i.e., the study of the physical universe through the application of the laws and theories to astronomical observations involving gravitational waves. See, for example, @finn:2002:bmo [@sutton:2002:bgm], which use the effect of gravitational wave emission on several binary pulsar systems to bound the mass of the graviton. [^2]: This comparison is meant only to illustrate the importance of $\tau_{\text{cr}}$ in estimating $\tau_{\text{gw}}$. In this case the distance to J$1939+2134$ is not known accurately enough allow a prediction for an accurate estimate of either $\tau_{\text{pw}}$ or $\tau_{\text{gw}}$.
{ "pile_set_name": "ArXiv" }