diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzmzco" "b/data_all_eng_slimpj/shuffled/split2/finalzzmzco" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzmzco" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction} \\label{intro}\nLattice regularization of chiral gauge theories has remained a long standing\nproblem of nonperturbative investigation of quantum field theory. Lack of\nchiral gauge invariance in $L\\chi GT$ proposals is responsible for\nthe longitudinal gauge degrees of freedom ({\\em dof}) coupling to\nfermionic {\\em dof} and eventually spoiling the chiral nature of the theory.\nThe well-known example is the Smit-Swift proposal of $L\\chi GT$ \\cite{smit}.\nAlthough in a recent development using a Dirac operator that\nsatisfies the Ginsparg-Wilson relation, it was possible to formulate a\n$L\\chi GT$ without violating gauge-invariance or locality \\cite{luscher}, an\nexplicit model for nonperturbative numerical studies is still not \navailable.\n \nIn this paper, we follow the gauge fixing approach to $L\\chi GT$\n\\cite{golter1}. The obvious remedy to control the longitudinal gauge {\\em dof}\nis to gauge fix with a target theory in mind. The Roma proposal \\cite{roma}\ninvolving gauge fixing passed perturbative tests but does not address the\nproblem of gauge fixing of compact gauge fields and the associated problem of\nlattice artifact Gribov copies. The formal problem is that\nfor compact\ngauge fixing a BRST-invariant partition function as well as (unnormalized)\nexpectation values of BRST invariant operators vanish as a consequence of\nlattice Gribov copies \\cite{neuberger}. Shamir and Golterman \\cite{golter1}\nhave proposed to\nkeep the gauge fixing part of the action BRST noninvariant and tune\ncounterterms to recover BRST in the continuum. In their formalism, the\ncontinuum limit is to be taken from within the broken ferromagnetic (FM) phase\napproaching another broken phase which is called ferromagnetic directional\n(FMD) phase, with the mass of the gauge field vanishing at the FM-FMD\ntransition. This was tried out in a $U(1)$ Smit-Swift model and so\nfar the results show that in the pure gauge sector, QED is recovered in\nthe continuum limit \\cite{recent} and in the {\\em reduced} model limit \n(to be defined below) free chiral fermions in the appropriate chiral \nrepresentation are obtained \\cite{bock1}. Tuning with counterterms has \nalso not posed any practical problem, actually very little tuning is \nnecessary. Efforts are currently underway to extend this gauge fixing \nproposal to include nonabelian gauge groups \\cite{nab}. \n \nWithout gauge fixing the longitudinal gauge {\\em dof}, which are \nradially frozen scalar fields, are rough and nonperturbative even if the \ntransverse gauge coupling may be weak (this is because with the standard \nlattice measure, each point on the gauge orbit has equal weight). \nThe theory in the continuum limit, taken at the transition \nbetween the broken symmetry ferromagnetic (FM) phase and the \nsymmetric paramagnetic (PM) phase, displays undesired nonperturbative \neffects of the scalar-fermion coupling that usually spells disaster for \nthe chiral theory. The job of the gauge fixing is to introduce a new \ncontinuous phase transition, from the FM phase to a new broken symmetry \nphase (FMD), at which the gauge symmetry is recovered and at the same \ntime the gauge fields become smooth.\n \nThe problem can be cleanly studied in the reduced model as explained in the\nfollowing. When one gauge transforms a gauge non-invariant theory, one picks\nup the longitudinal gauge degrees of freedom (radially frozen scalars)\nexplicitly in the action. The reduced model is then obtained \nby making the lattice gauge field unity for \nall links, {\\em i.e.}, by switching off the transverse gauge coupling. \nThe action becomes that of a chiral Yukawa theory with interaction \nbetween the fermions and the longitudinal gauge {\\em dof}. \nThe reduced model would have a phase structure similar to the full \ntheory, {\\em e.g.}, the gauge fixed theory in the reduced limit \nwill have a FM-FMD transition in addition to the FM-PM transition. \nNow for the gauge fixing proposal to work, the scalars need \nto decouple from the fermions at the FM-FMD transition leaving the \nfermions free in the appropriate chiral representation. Passing the \nreduced model test is an important first step for any $L\\chi GT$ \nproposal that breaks gauge invariance. \n\nIn the reduced model derived from the gauge fixed theory the \nscalar fields become smooth and expandable in a perturbative \nseries as $1+{\\cal O}(\\mbox{coupling constant})$ at the FM-FMD \ntransition. If continuum limit can be taken near the point in the \ncoupling parameter space around which this perturbative expansion is \ndefined, the scalar fields will decouple from the theory. \nThe parameterization of the gauge fixing action turns out to be a good \none, because this continuum limit can be taken ($i$) very easily by \napproaching the FM-FMD transition almost perpendicularly by tuning \nessentially one counterterm, and ($ii$) at a point on this transition \nline which is reasonably far away from the expansion point. This has \nbeen possible in \\cite{bock1} and again in the present work. \n\nA central claim of the gauge fixing proposal \nis that it is universal, {\\em i.e.}, it should work with any lattice \nfermion action that has the correct classical continuum limit. This is \nbecause the central idea as discussed above is independent of the \nparticular lattice fermion regularization. In the \npresent paper we want to confirm the universality claim by applying the \nproposal to domain wall fermions \\cite{kaplan} with $U(1)$ gauge group. \nFor this purpose we have chosen the waveguide formulation \\cite{kaplan2} \nof the domain wall fermion and investigate in the reduced model. This \nmodel was investigated before without gauge fixing and the free domain \nwall spectrum was not obtained in the reduced limit \\cite{golter2}. \nMirror chiral modes were found at the waveguide boundaries in addition \nto the chiral modes at the domain wall or anti-domain wall.\n \nIn section II we present the gauge-fixed domain wall fermion action for a\n$U(1)$ chiral gauge theory and then go to the so-called reduced\nmodel by switching off the transverse gauge coupling. In section III we\nperform a weak coupling perturbation theory in the reduced model for the\nfermion propagators and mass matrix to 1-loop. However, in sections\n\\ref{ferm_m} and \\ref{ferm_m1} we have used special boundary conditions\n(instead of the actual Kaplan boundary conditions) to arrive at explicit\nexpressions for the overlap of the opposite chiral modes. Our numerical \nresults for \nthe quenched phase diagram and chiral fermion propagators at the domain \nwall and anti-domain wall and at the waveguide boundaries are presented \nand compared with the perturbative results in section IV. We summarize \nin the concluding section V. In Appendix A, we describe the special \nboundary conditions used in sections \\ref{ferm_m} and \\ref{ferm_m1}.\nIn Appendix B we schematically discuss how using Kaplan boundary conditions\none can arrive at the same qualitative conclusion about the 1-loop\noverlap of the opposite chiral modes.\n \n\\section{Gauge-fixed Domain Wall Action} \\label{gfdwa}\n\nKaplan's free domain wall fermion action \\cite{kaplan} on a\n$4+1$-dimensional lattice is given by (lattice constant is taken to be\nunity throughout this paper),\n\\begin{equation}\nS_F = \\sum_{XY} \\overline{\\psi}_X \\left[ \\partial\\!\\!\\!\/_5 - w_5 +\n{\\bf M}\\right]_{XY} \\psi_Y \\label{dwact}\n\\end{equation}\nwhere $\\overline{\\psi}$ and $\\psi$ are the fermion fields, and \n$\\partial\\!\\!\\!\/_5$ and $w_5$ are respectively the 5-dimensional\nDirac operator and the Wilson term,\n\\begin{eqnarray}\n(\\partial\\!\\!\\!\/_5)_{XY} &=&\n\\frac{1}{2}\\sum_{\\alpha=1}^5 \\gamma_\\alpha \\left( \n\\delta_{X+\\hat{\\alpha},Y} - \\delta_{X-\\hat{\\alpha},Y} \\right), \n\\nonumber \\\\ (w_5)_{XY} &=& \\frac{r}{2} \\sum_{\\alpha=1}^5\n\\left( \\delta_{X+\\hat{\\alpha},Y} + \\delta_{X-\\hat{\\alpha},Y} - 2\\delta_{XY}\n\\right), \\label{par5}\n\\end{eqnarray}\nThe $\\gamma_\\alpha$'s are the five hermitian euclidean gamma matrices,\n$r$ is the Wilson parameter, $X=(x,s),~Y=(y,t)$ label the sites of\nthe $L^4 L_s$ lattice and $L_s$ is the extent of the 5th dimension:\n$0 \\leq s,t \\leq L_s-1$. We are interested in\ntaking the continuum limit in the 4 space-time dimensions only. It is\nconvenient to look at the 5-th dimension as a flavor space.\n\n\n\nWith periodic boundary conditions in the 5th or\n$s$-direction ($s,t = L_s \\Rightarrow s,t=0$) and the domain wall mass\n${\\bf M}$ taken as\n\\begin{equation}\n{\\bf M}_{XY}=m(s)\\delta_{XY},\\;\\; {\\rm where},\n\\end{equation}\n\\begin{equation}\nm(s) = \\begin{array}{rl}\n-m_0, & 0 0 \\\\\n c_{t}^{2} & \\equiv \\frac{\\omega_{4}}{\\omega_{1}} = \\frac{1 - \\frac{1}{2}\\bar{H}^2\\tilde{x}^2 - \\frac{1}{8}\\bar{H}^4 \\tilde{x}^4 + 2 \\xi \\bar{H}^3 \\tilde{x}^2 (\\bar{H} \\tilde{x})' + 2 \\xi \\bar{H}^5 \\tilde{x}^4 (\\bar{H}\\tilde{x})'}{1 + \\frac{1}{2} \\bar{H}^2 \\tilde{x}^2 + \\frac{3}{8} \\bar{H}^4 \\tilde{x}^4 + 2 \\xi \\bar{H}^4 \\tilde{x}^3 + 2\\xi \\bar{H}^6 \\tilde{x}^5} \\geq 0 \n\\end{align}\n\nIn particular, we see that the gravitational wave speed depends on $\\tilde{x}$, and tends to 1 when $\\tilde{x} \\rightarrow 0$ :\n\\begin{align}\n4 Q_t & \\simeq 1 + 2(\\bar{H} \\tilde{x})^2 \\\\\n c_{t} & \\simeq 1 - \\frac{1}{2} (\\bar{H} \\tilde{x})^2\n\\end{align}\nGiven the very tight constraint on the speed of gravitational waves, equal to the speed of light up to a $\\sim 10^{-15}$ difference \\cite{LIGOScientific:2017zic,Creminelli:2017sry,Ezquiaga:2017ekz}, this justifies \\textit{a posteriori} the relevance of the non-relativistic limit where $\\tilde{x} \\ll 1$. Moreover, we see that tensorial perturbations are stable in this limit since $Q_t > 0$.\n\n\\subsection*{Scalar stability conditions}\n\nSimilar stability conditions apply to the scalar degrees of freedom, here including the scalar perturbations of matter components:\n\\begin{align}\n Q_{s} & \\equiv \\frac{\\omega_{1} \\left( 4\\omega_{1}\\omega_{3} + 9\\omega_{2}^{2} \\right)}{3\\omega_{2}^{2}} > 0 \\label{eq:Qs} \\\\\n c_{s}^{2} & \\equiv \\frac{3 \\left( 2\\omega_{1}^{2}\\omega_{2}H - \\omega_{2}^{2}\\omega_{4} + 4\\omega_{1}\\omega_{2}\\dot{\\omega}_{1} - 2\\omega_{1}^{2}\\dot{\\omega}_{2} \\right) - 6\\omega_{1}^{2} \\sum \\left( 1+w_{i} \\right) \\rho_{i}}{\\omega_{1} \\left( 4\\omega_{1}\\omega_{3} + 9\\omega_{2}^{2} \\right)} \\geq 0\n\\end{align}\nwhere $w_{i}$ and $\\rho_{i}$ are respectively the equation of state parameter and the energy density of the fluid $i$, and the sum runs over all the components of the Universe (here only pressureless matter and radiation). At the lowest order in $\\tilde{x}$, we get:\n\\begin{align}\nQ_s & \\simeq \\frac{3}{2} (\\Omega_\\Lambda^0 - \\bar{H}^2) \\tilde{x}^2 \\\\\nc_{s}^{2} & \\simeq 1 + \\frac{2 \\left( \\eta \\bar{H}^{2}\\tilde{x}' - \\bar{H}\\h' - 2\\xi\\bar{H}^{4}\\tilde{x}' \\right)}{3 \\left( \\Omega_{\\Lambda}^{0} - \\bar{H}^{2} \\right)}\n\\end{align}\n\nWith $\\tilde{x} \\ll 1$, a fit of the DBI-Galileon model to data leads to cosmological parameters close to the standard model ones: $\\Omega_m^0 \\approx 0.3$ and $\\Omega_\\Lambda^0 \\approx 0.7$ \\cite{Planck2018}. Therefore, from the first Friedmann equation, we get $\\Omega_\\Lambda^0 < \\bar{H}^2$ for all relevant models in agreement with cosmological observations. As $Q_s \\leq 0$, the DBI-Galileon model contains scalar instabilities unless it reduces to GR. One way to avoid this would be to add a spatial curvature to the metric, but with a strong energy density (at least $\\sim 0.3$) which is also excluded by observations \\cite{Planck2018}. \n\n\n\\section{Discussion}\\label{sec:discussion}\n\n\\subsection*{Physical interpretation}\n\nFrom the definition \\eqref{eq:Qs}, we see that the dominant terms come $\\G2$ (giving the $\\Omega_\\Lambda^0$ term) and $\\G4$ (giving the $\\bar{H}^2$ term). The competition between the two terms leads to the ghost-like behaviour in a cosmological setting: $Q_s \\leq 0$. In other words, it is the result of the competition between the DBI and the Einstein-Hilbert terms. The DBI action will have the effect of stretching the brane towards an extremal surface, whereas the Einstein-Hilbert term on the brane will tend to make the brane contract on itself from the effect of curvature. However, in the non-relativistic limit of the DBI-Galileon, the Einstein-Hilbert term destabilizes the scalar field perturbations and the stretching effect from the cosmological constant is not strong enough to counterbalance, leading to an instantaneous decay of the vacuum state.\n\nBecause the DBI-Galileon action is the most general one can find of a 4D probe brane in a 5D bulk, we expect this statement to be quite general for all such theories studied in the current context. Indeed, this is true in a standard cosmological setting which is realised with an FLRW slicing of the bulk space-time (equivalent to an FLRW background on the brane). Therefore, the only way to evade this ghostly behaviour in cosmology is to include the full relativistic dynamics of the theory ($\\tilde{x} \\sim 1$). We have seen that, in this case, we expect significant deviations of the speed of gravitational waves $c_t$ from $c$. This is not a definitive impossibility though if the full DBI-Galileon is viewed as an effective theory valid only at cosmological scales for which the speed of gravitational waves has not been probed \\cite{deRham:2018red,Ezquiaga:2018btd}. Indeed, the constraint on the gravitational speed from the observation of GW170817 in coincidence with GRB170817A \\cite{LIGOScientific:2017zic} is only valid on small scales probed by LIGO and Virgo. A modification of the dispersion relation of gravitational waves at small scales from operators present in the UV complete theory could allow $c_{t} \\neq c$ on cosmological scales while being compatible with current astrophysical observations. Waiting for the next generation of gravitational wave interferometers, in particular LISA, which will be able to probe this relation at larger scales \\cite{Barausse:2020rsu}, this possibility remains open.\n\n\\subsection*{Direct coupling to matter}\n\nIn the context of cosmology, where standard model matter is present, there might be direct coupling to the scalar field. In that case, the metric $\\tilde{q}$ to which matter is sensitive is different than the space-time metric $q$:\n\\begin{equation}\\label{eq:action_galdbi-qtilde}\n \\mathcal{S} = \\int dx^{4} \\sqrt{-q} \\left(\\mathcal{L}_f + \\mathcal{L}_K + \\mathcal{L}_R + \\mathcal{L}_{\\mathcal{K}_{GB}}\\right) + \\int dx^{4} \\sqrt{-\\tilde q} \\mathcal{L}_{m} \\left(\\tilde q_{\\mu\\nu}, \\psi_{m} \\right).\n\\end{equation}\nIt has been shown in \\cite{Bekenstein:1992pj} that the two metrics are related by a disformal transformation of the following form:\n\\begin{equation}\n q_{\\mu\\nu} = A \\left( \\varphi, \\tilde{X} \\right) \\tilde{q}_{\\mu\\nu} + B \\left( \\varphi, \\tilde{X} \\right) \\frac{\\partial_{\\mu}\\varphi \\partial_{\\nu}\\varphi}{f^{4}} \\label{eq:disformal transformation}\n\\end{equation}\nwhere $A$ and $B$ are arbitrary functions of the scalar field and $\\tilde{X} = -\\tilde{q}_{\\mu\\nu}\\partial_{\\mu}\\varphi\\partial_{\\nu}\\varphi\/2f^{4}$. For simplicity and following the treatment of the covariant Galileon \\cite{Stephen-Appleby_2012}, we assume that $A$ and $B$ are constant parameters. This can be further justified by the fact that, a dependency on $X$ would introduce, in general, higher order terms which would go beyond the framework of Horndeski theories \\cite{Bettoni:2013diz}, and a dependency on $\\varphi$ would, in general, break the shift symmetry followed by the scalar field $\\varphi$ in the probe brane context. Note that, when $A = -B$, matter is coupled to the induced metric on the brane.\n\nContrary to the covariant Galileon, the DBI-Galileon action is not invariant by such a change of reference frame. However, new terms that can not be absorbed into a redefinition of the parameters arise only at higher order in $\\tilde{X}$. Therefore, the non-relativistic dynamics is not change by the introduction of a direct coupling between the scalar field and matter of the form \\eqref{eq:disformal transformation} with constant parameters. In particular, this does not prevent the perturbations around the FLRW background from showing ghost instabilities.\n\\subsection*{Generalization}\n\nThe DBI-Galileon is a particular example of the more general class of shift-symmetric Horndeski theories. These are subclass of Horndeski theories which are invariant under a shift symmetry of the scalar field $\\varphi \\rightarrow \\varphi + c$ \\cite{Sotiriou:2013qea,Sotiriou:2014pfa}. In these theories, the arbitrary Horndeski functions are restricted to be functions of $X$ alone. In order to make the non-relativistic limit apparent, we Taylor expand these arbitrary functions around GR:\n\\begin{align}\n G_{2} & \\equiv \\Lambda + \\sum_{n=1}^{+\\infty} g_{2}^{\\left( n \\right)} X^{n} \\\\\n G_{3} & \\equiv \\sum_{n=1}^{+\\infty} g_{3}^{\\left( n \\right)} X^{n} \\\\\n G_{4} & \\equiv \\frac{M_{P}^{2}}{2} + \\sum_{n=1}^{+\\infty} g_{4}^{\\left( n \\right)} X^{n} \\\\\n G_{5} & \\equiv \\sum_{n=1}^{+\\infty} g_{5}^{\\left( n \\right)} X^{n}\n\\end{align}\n\nThe constant terms in $G_{3}$ and $G_{5}$ do not appear in the expansion as they lead to total derivative terms. Because the Horndeski functions depend only on $X$, the $\\omega$ functions that determine the stability conditions reduce to:\n\\begin{eqnarray}\n \\omega_{1} & \\equiv & 2 \\G4 - 2X\\left( 2\\G{4,X} + \\dot{\\phi}H\\G{5,X} \\right) \\\\\n \\omega_{2} & \\equiv & 4 H\\G4 - 2X \\left( \\dot{\\phi}\\G{3,X} + 8 H\\G{4,X} + 5 \\dot{\\phi}H^{2} \\G{5,X} \\right) - 4X^{2}H \\left( 4 \\G{4,XX} + \\dot{\\phi}H\\G{5,XX} \\right) \\\\\n \\omega_{3} & \\equiv & -18H^{2}\\G4 + 3X \\left( \\G{2,X} + 12 \\dot{\\phi}H\\G{3,X} + 42H^{2} \\G{4,X} + 30 \\dot{\\phi}H^{3} \\G{5,X} \\right) \\nonumber \\\\\n && + 6X^{2} \\left( \\G{2,XX} + 3\\dot{\\phi}H\\G{3,XX} + 48 H^{2}\\G{4, XX} + 13H^{3}\\dot{\\phi}\\G{5,XX} \\right) \\nonumber \\\\\n && + 12 X^{3}H^{2} \\left( 6\\G{4,XXX} + H\\dot{\\phi}\\G{5,XXX} \\right) \\\\\n \\omega_{4} & \\equiv & 2\\G4 - 2X\\ddot{\\phi} \\G{5,X}\n\\end{eqnarray}\n\nFrom these, we can compute the quantity $Q_{s}$ up to first order in $X$:\n\\begin{equation}\n Q_{s} = \\frac{X}{H^{2}} \\left( \\g2^{\\left( 1 \\right)} + 6H^{2} \\g4^{\\left( 1 \\right)} \\right) + O \\left( X^{\\frac{3}{2}} \\right)\n\\end{equation}\n\nThis leads to a very simple formulation of the no-ghost condition, independent of $X$, in the context of Shift-Symmetric Horndeski theories in the non-relativistic limit:\n\\begin{equation}\n \\g2^{\\left( 1 \\right)} + 6H^{2} \\g4^{\\left( 1 \\right)} > 0\n\\end{equation}\n\nIn the context of the brane galileon, where $\\g4^{\\left( 1 \\right)} = -M_{P}^{2}\/2$ and $\\g2^{\\left( 1 \\right)} = \\Lambda$, this is equivalent to the inequality which is never fulfilled in flat space:\n\\begin{equation}\n \\Lambda - 3M_{P}^{2}H^{2} > 0 \\quad \\Leftrightarrow \\quad \\Omega_{\\Lambda}^{0} > \\bar{H}^{2}\n\\end{equation}\n\nOther stability conditions are given by:\n\\begin{align}\n c_{s}^{2} & \\simeq 1 + \\frac{2\\Ddot{\\phi} \\g3^{\\left( 1 \\right)} + 4\\Dot{H}\\g4^{\\left( 1 \\right)} + 2\\Ddot{\\phi}H^{2}\\g5^{\\left( 1 \\right)}}{\\g2^{\\left( 1 \\right)} + 6H^{2} \\g4^{\\left( 1 \\right)}} > 0\\\\\n Q_t & \\simeq \\frac{M_{P}^{2}}{4} \\\\\n c_t^{2} & \\simeq 1\n\\end{align}\nwhere we expressed these quantities at the lowest order. The two tensorial conditions are, thus, automatically satisfied in this context. On the other hand, the stability conditions for scalar perturbations at the lowest order give a simple inequality involving the parameters of the Taylor expansion, that can be easily checked at the background level.\n\n\n\\section{Conclusion}\\label{sec:conclusion}\n\nWe described the DBI-Galileon theory of a four-dimensional brane evolving in a 5D bulk space-time in the non-relativistic limit where its local kinetic energy is small compared to its tension. This model belongs to the class of shift-symmetric Horndeski theories, themselves being a subclass of the more general family of Horndeski theories. From the construction of the DBI-Galileon model, the free parameters of the model acquire a physical meaning. In particular, the interpretation of the cosmological constant is linked to the brane tension energy density. We derived the equations driving the evolution of the late-time Universe around a spatially flat FLRW cosmological background and studied the stability of scalar and tensorial perturbations. This model reduces to an expansion around standard GR, and therefore around standard $\\Lambda$CDM in the cosmological context. As such, it is naturally compatible with data in the non-relativistic limit provided the effect of the scalar field is small enough, even considering the speed of the gravitational waves. However, it revealed fatal ghostly behaviour for scalar perturbations around the FLRW background. From there, we derived the corresponding stability conditions for shift-symmetric Horndeski theories in the non-relativistic limit in the cosmological context and found very simple formulations for these conditions.\n\n\\section*{Acknowledgements}\n\nWe would like to thank Marc Besan\u00e7on, Arnaud de Mattia and Vanina Ruhlmann-Kleider for their comments on the present paper. We also want to thank David Langlois for useful and interesting comments and suggestions.\n\n\n\\section*{References}\n\\bibliographystyle{iopart-num}\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\\label{sec1}\n\nThe supersymmetry technique is a powerful method in random matrix theory and disordered systems. For a long time it was thought to be applicable for Gaussian probability densities only \\cite{Efe83,VerZir85,VWZ85,Efe97}. Due to universality on the local scale of the mean level spacing \\cite{BreZee93a,BreZee93b,HacWei95,GMW98}, this restriction was not a limitation for calculating in quantum chaos and disordered systems. Indeed, results of Gaussian ensembles are identical for large matrix dimension with other invariant matrix ensembles on this scale. In the Wigner--Dyson theory \\cite{Bee97} and its corrections for systems with diffusive dynamics \\cite{Mir00}, Gaussian ensembles are sufficient. Furthermore, universality was found on large scale, too \\cite{AJM90}. This is of paramount importance when investigating matrix models in high--energy physics.\n\nThere are, however, situations in which one can not simply resort to Gaussian random matrix ensembles. The level densities in high--energy physics \\cite{BIPZ78} and finance \\cite{LCBP99} are needed for non-Gaussian ensembles. But these one--point functions strongly depend on the matrix ensemble. Other examples are bound--trace and fixed--trace ensembles \\cite{Meh04}, which are both norm--dependent ensembles \\cite{Guh06}, as well as ensembles derived from a non-extensive entropy principle \\cite{TVT04,BBP04,Abu04}. In all these cases one is interested in the non-universal behavior on special scales.\n\nRecently, the supersymmetry method was extended to general rotation invariant probability densities \\cite{Guh06,Som07,LSZ07,KGG08}. There are two approaches. The first one is the generalized Hubbard--Stratonovich transformation \\cite{Guh06,KGG08}. With help of a proper Dirac--distribution in superspace an integral over rectangular supermatrices was mapped to a supermatrix integral with non-compact domain in the Fermion--Fermion block. The second approach is the superbosonization formula \\cite{Som07,LSZ07} mapping the same integral over rectangular matrices as before to a supermatrix integral with compact domain in the Fermion--Fermion block.\n\nIn this work, we prove the equivalence of the generalized Hubbard--Stratonovich transformation with the superbosonization formula. The proof is based on integral identities between supersymmetric Wishart--matrices and quadratic supermatrices. The orthogonal, unitary and unitary-symplectic classes are dealt with in a unifying way.\n\nThe article is organized as follows. In Sec. \\ref{sec2}, we give a motivation and introduce our notation. In Sec. \\ref{sec3}, we define rectangular supermatrices and the supersymmetric version of Wishart-matrices built up by supervectors. We also give a helpful corollary for the case of arbitrary matrix dimension discussed in Sec. \\ref{sec7}. In Secs. \\ref{sec4} and \\ref{sec5}, we present and further generalize the superbosonization formula and the generalized Hubbard--Stratonovich transformation, respectively. The theorem stating the equivalence of both approaches is given in Sec. \\ref{sec6} including a clarification of their mutual connection. In Sec. \\ref{sec7}, we extend both theorems given in Secs. \\ref{sec4} and \\ref{sec5} to arbitrary matrix dimension. Details of the proofs are given in the appendices.\n\n\\section{Ratios of characteristic polynomials}\\label{sec2}\n\nWe employ the notation defined in Refs. \\cite{KKG08,KGG08}. ${\\rm Herm\\,}(\\beta,N)$ is either the set of $N\\times N$ real symmetric ($\\beta=1$), $N\\times N$ hermitian ($\\beta=2$) or $2N\\times 2N$ self-dual ($\\beta=4$) matrices, according to the Dyson--index $\\beta$. We use the complex representation of the quaternionic numbers $\\mathbb{H}$. Also, we define\n\\begin{equation}\\label{2.0}\n \\gamma_1=\\left\\{\\begin{array}{ll} 1 & ,\\ \\beta\\in\\{2,4\\} \\\\ 2 & ,\\ \\beta=1 \\end{array}\\right. \\quad ,\\quad \\gamma_2=\\left\\{\\begin{array}{ll} 1 & ,\\ \\beta\\in\\{1,2\\} \\\\ 2 & ,\\ \\beta=4 \\end{array}\\right.\n\\end{equation}\nand $\\tilde{\\gamma}=\\gamma_1\\gamma_2$.\n\nThe central objects in many applications of supersymmetry are averages over ratios of characteristic polynomials \\cite{AkeFyo03,AkePot04,BorStr05}\n\\begin{eqnarray}\n \\fl Z_{k_1k_2}(E^-) & = & \\int\\limits_{{\\rm Herm\\,}(\\beta,N)}P(H)\\frac{\\prod\\limits_{n=1}^{k_2}\\det\\left(H-(E_{n2}-\\imath\\varepsilon)\\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_2N}\\right)}{\\prod\\limits_{n=1}^{k_1}\\det\\left(H-(E_{n1}-\\imath\\varepsilon)\\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_2N}\\right)}d[H]\\nonumber\\\\\n & = & \\int\\limits_{{\\rm Herm\\,}(\\beta,N)}P(H){\\rm Sdet\\,}^{-1\/\\tilde{\\gamma}}\\left(H\\otimes \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\tilde{\\gamma}(k_1+k_2)}- \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_2N}\\otimes E^-\\right)d[H]\\label{2.1}\n\\end{eqnarray}\nwhere $P$ is a sufficiently integrable probability density on the matrix set ${\\rm Herm\\,}(\\beta,N)$ invariant under the group\n\\begin{equation}\\label{2.2}\n {\\rm U\\,}^{(\\beta)}(N)=\\left\\{\\begin{array}{ll} \n \\Or(N)\t& ,\\ \\beta=1\\\\\n\t {\\rm U\\,}(N)\t& ,\\ \\beta=2\\\\\n\t {\\rm USp\\,}(2N)\t& ,\\ \\beta=4\n \\end{array}\\right. .\n\\end{equation}\nHere, we assume that $P$ is analytic in its real independent variables. We use the same measure for $d[H]$ as in Ref. \\cite{KKG08} which is the product over all real independent differentials, see also Eq.~\\eref{t1.4}. Also, we define $E={\\rm diag\\,}(E_{11},\\ldots,E_{k_11},E_{12},\\ldots,E_{k_22})\\otimes \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\tilde{\\gamma}}$ and $E^-=E-\\imath\\varepsilon \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\tilde{\\gamma}(k_1+k_2)}$.\n\nThe generating function of the $k$--point correlation function \\cite{BreHik00,Zir06,Guh06,KGG08}\n\\begin{equation}\\label{2.3}\n R_{k}(x)=\\gamma_2^{-k}\\int\\limits_{{\\rm Herm\\,}(\\beta,N)}P(H)\\prod\\limits_{p=1}^k\\tr\\delta(x_p-H)d[H]\n\\end{equation}\nis one application and can be computed starting from the matrix Green function and Eq.~\\eref{2.1} with $k_1=k_2=k$. Another example is the $n$--th moment of the characteristic polynomial \\cite{MehNor01,Fyo02,Zir06}\n\\begin{equation}\\label{2.4}\n \\widehat{Z}_{n}(x,\\mu)=\\int\\limits_{{\\rm Herm\\,}(\\beta,N)}P(H)\\Theta(H){\\det}^n\\left(H-E \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_2k}\\right)d[H],\n\\end{equation}\nwhere the Heavyside--function for matrices $\\Theta(H)$ is unity if $H$ is positive definite and zero otherwise. \\cite{KGG08} \n\nWith help of Gaussian integrals, we get an integral expression for the determinants in Eq.~\\eref{2.1}. Let $\\Lambda_j$ be the Grassmann space of $j$--forms. We consider a complex Grassmann algebra \\cite{Ber87} $\\Lambda=\\bigoplus\\limits_{j=0}^{2\\gamma_2Nk_2}\\Lambda_j$ with $\\gamma_2Nk_2$ pairs $\\{\\zeta_{jn},\\zeta_{jn}^*\\}$, $1\\leq n\\leq k_2,\\ 1\\leq j\\leq \\gamma_2N$, of Grassmann variables and use the conventions of Ref. \\cite{KKG08} for integrations over Grassmann variables. Due to the $\\mathbb{Z}_2$--grading, $\\Lambda$ is a direct sum of the set of commuting variables $\\Lambda^0$ and of anticommuting variables $\\Lambda^1$. The body of an element in $\\Lambda$ lies in $\\Lambda_0$ while the Grassmann generators are elements in $\\Lambda_1$.\n\nLet $\\imath$ be the imaginary unit. We take $\\gamma_2Nk_1$ pairs $\\{z_{jn},z_{jn}^*\\}$, $1\\leq n\\leq k_1,\\ 1\\leq j\\leq \\gamma_2N$, of complex numbers and find for Eq.~\\eref{2.1}\n\\begin{equation}\\label{2.5}\n \\fl Z_{k_1k_2}(E^-)= (2\\pi)^{\\gamma_2N(k_2-k_1)}\\imath^{\\gamma_2Nk_2}\\int\\limits_{\\mathfrak{C}} \\mathcal{F}P(K)\\exp\\left(-\\imath{\\rm Str\\,} BE^-\\right)d[\\zeta]d[z]\n\\end{equation}\nwhere $d[z]=\\prod\\limits_{p=1}^{k_1}\\prod\\limits_{j=1}^{\\gamma_2N}dz_{jp}dz_{jp}^*$ , $d[\\zeta]=\\prod\\limits_{p=1}^{k_2}\\prod\\limits_{j=1}^{\\gamma_2N}(d\\zeta_{jp}d\\zeta_{jp}^*)$ and $\\mathfrak{C}=\\mathbb{C}^{\\gamma_2k_1N}\\times\\Lambda_{2\\gamma_2Nk_2}$. The characteristic function appearing in \\eref{2.5} is defined as\n\\begin{equation}\\label{2.6}\n \\mathcal{F}P(K)=\\int\\limits_{{\\rm Herm\\,}(\\beta,N)}P(H)\\exp\\left(\\imath\\tr HK\\right)d[H] .\n\\end{equation}\nThe two matrices\n\\begin{equation}\\label{2.7}\n K = \\frac{1}{\\tilde{\\gamma}}V^\\dagger V\\qquad{\\rm and}\\qquad\n B = \\frac{1}{\\tilde{\\gamma}}VV^\\dagger\n\\end{equation}\nare crucial for the duality between ordinary and superspace. While $K$ is a $\\gamma_2N\\times\\gamma_2N$ ordinary matrix whose entries have nilpotent parts, $B$ is a $\\tilde{\\gamma}(k_1+k_2)\\times\\tilde{\\gamma}(k_1+k_2)$ supermatrix. They are composed of the rectangular $\\gamma_2N\\times\\tilde{\\gamma}(k_1+k_2)$ supermatrix\n\\begin{eqnarray}\n V^\\dagger|_{\\beta\\neq2} & = & (z_1,\\ldots,z_{k_1},Yz_1^*,\\ldots,Yz_{k_1}^*,\\zeta_1,\\ldots,\\zeta_{k_2},Y\\zeta_1^*,\\ldots,Y\\zeta_{k_2}^*) ,\\nonumber\\\\\n V|_{\\beta\\neq2} & = & (z_1^*,\\ldots,z_{k_1}^*,Yz_1,\\ldots,Yz_{k_1},-\\zeta_1^*,\\ldots,-\\zeta_{k_2}^*,Y\\zeta_1,\\ldots,Y\\zeta_{k_2})^T ,\\nonumber\\\\\n V^\\dagger|_{\\beta=2} & = & (z_1,\\ldots,z_{k_1},\\zeta_1,\\ldots,\\zeta_{k_2}) ,\\nonumber\\\\\n V|_{\\beta=2} & = & (z_1^*,\\ldots,z_{k_1}^*,-\\zeta_1^*,\\ldots,-\\zeta_{k_2}^*)^T .\\label{2.9}\n\\end{eqnarray}\nThe transposition ``$T$'' is the ordinary transposition and is not the supersymmetric one. However, the adjoint ``$\\dagger$'' is the complex conjugation with the supersymmetric transposition ``$T_{\\rm S}$''\n\\begin{equation}\\label{2.9b}\n \\sigma^{T_{\\rm S}}=\\left[\\begin{array}{cc} \\sigma_{11} & \\sigma_{12} \\\\ \\sigma_{21} & \\sigma_{22} \\end{array}\\right]^{T_{\\rm S}}=\\left[\\begin{array}{cc} \\sigma_{11}^T & \\sigma_{21}^T \\\\ -\\sigma_{12}^T & \\sigma_{22}^T \\end{array}\\right],\n\\end{equation}\nwhere $\\sigma$ is an arbitrary rectangular supermatrix. We introduce the constant $\\gamma_2N\\times\\gamma_2N$ matrix\n\\begin{equation}\\label{2.10}\n Y=\\left\\{\\begin{array}{ll}\n \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_N & ,\\ \\beta=1\\\\\n\t Y_s^T\\otimes \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_N & ,\\ \\beta=4\n \\end{array}\\right.\\ ,\\ \\ Y_s=\\left[\\begin{array}{cc} 0 & 1 \\\\ -1 & 0 \\end{array}\\right] .\n\\end{equation}\nThe crucial duality relation \\cite{Guh06,KGG08}\n\\begin{equation}\\label{2.11}\n \\tr K^m={\\rm Str\\,} B^m\\ ,\\ \\ m\\in\\mathbb{N} ,\n\\end{equation}\nholds, connecting invariants in ordinary and superspace. As $\\mathcal{F}P$ inherits the rotation invariance of $P$, the duality relation \\eref{2.11} yields\n\\begin{equation}\\label{2.12}\n \\fl Z_{k_1k_2}(E^-)= (2\\pi)^{\\gamma_2N(k_2-k_1)}\\imath^{\\gamma_2Nk_2}\\int\\limits_{\\mathfrak{C}} \\Phi(B)\\exp\\left(-\\imath{\\rm Str\\,} BE^-\\right)d[\\zeta]d[z] .\n\\end{equation}\nHere, $\\Phi$ is a supersymmetric extension of a representation $\\mathcal{F}P_0$ of the characteristic function,\n\\begin{equation}\\label{2.13}\n \\Phi(B)=\\mathcal{F}P_0({\\rm Str\\,} B^m|m\\in\\mathbb{N})=\\mathcal{F}P_0(\\tr K^m|m\\in\\mathbb{N})=\\mathcal{F}P(K) .\n\\end{equation}\nThe representation $\\mathcal{F}P_0$ is not unique \\cite{BEKYZ07}. However, the integral \\eref{2.12} is independent of a particular choice \\cite{KGG08}.\n\nThe supermatrix $B$ fulfills the symmetry\n\\begin{equation}\\label{2.14}\n B^*=\\left\\{\\begin{array}{ll}\n \\widetilde{Y}B\\widetilde{Y}^T &,\\ \\beta\\in\\{1,4\\}, \\\\\n\t\t\\widetilde{Y}B^*\\widetilde{Y}^T &,\\ \\beta=2\n \\end{array}\\right.\n\\end{equation}\nwith the supermatrices\n\\begin{equation}\\label{2.15}\n \\fl\\widetilde{Y}|_{\\beta=1}=\\left[\\begin{array}{ccc} 0 & \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_1} & 0 \\\\ \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_1} & 0 & 0 \\\\ 0 & 0 & Y_s\\otimes \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_2} \\end{array}\\right]\\quad,\\quad\\widetilde{Y}|_{\\beta=4}=\\left[\\begin{array}{ccc} Y_s\\otimes \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_1} & 0 & 0 \\\\ 0 & 0 & \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_2} \\\\ 0 & \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_2} & 0 \\end{array}\\right]\n\\end{equation}\nand $\\widetilde{Y}|_{\\beta=2}= \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_1+k_2}$ and is self-adjoint for every $\\beta$. Using the $\\pi\/4$--rotations\n\\begin{equation}\\label{2.16}\n \\fl U|_{\\beta=1}=\\frac{1}{\\sqrt{2}}\\left[\\begin{array}{ccc} \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_1} & \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_1} & 0 \\\\ -\\imath \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_1} & \\imath \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_1} & 0 \\\\ 0 & 0 & \\sqrt{2}\\ \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{2k_2} \\end{array}\\right],\\ U|_{\\beta=4}=\\frac{1}{\\sqrt{2}}\\left[\\begin{array}{ccc} \\sqrt{2}\\ \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{2k_1} & 0 & 0 \\\\ 0 & \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_2} & \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_2}\\\\ 0 & -\\imath \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_2} & \\imath \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_2} \\end{array}\\right]\n\\end{equation}\nand $U|_{\\beta=2}= \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{k_1+k_2}$, $\\widehat{B}=UBU^\\dagger$ lies in the well-known symmetric superspaces \\cite{Zir96},\n\\begin{eqnarray}\n \\fl\\widetilde{\\Sigma}_{\\beta,\\gamma_1k_1,\\gamma_2k_2}^{(\\dagger)} & = &\\Biggl\\{\\sigma\\in{\\rm Mat}(\\tilde{\\gamma}k_1\/\\tilde{\\gamma}k_2)\\Biggl|\\sigma^\\dagger=\\sigma,\\nonumber\\\\\n & & \\left.\\sigma^*=\\left\\{\\begin{array}[c]{ll}\n \\widehat{Y}_{\\gamma_1k_1,\\gamma_2k_2}\\sigma\\widehat{Y}_{\\gamma_1k_1,\\gamma_2k_2}^{T} & ,\\ \\beta\\in\\{1,4\\} \\\\\n\t\t\\widehat{Y}_{k_1k_2}\\sigma^*\\widehat{Y}_{k_1k_2}^{T} & ,\\ \\beta=2\n \\end{array}\\right\\}\\right\\}\\label{2.17}\n\\end{eqnarray}\nwhere\n\\begin{equation}\\label{2.18}\n \\fl\\left.\\widehat{Y}_{pq}\\right|_{\\beta=1}=\\left[\\begin{array}{cc} \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{p} & 0 \\\\ 0 & Y_s\\otimes \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{q} \\end{array}\\right]\\ ,\\ \\ \\left.\\widehat{Y}_{pq}\\right|_{\\beta=2}= \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{p+q}\\ \\ \\ {\\rm and}\\ \\ \\ \\left.\\widehat{Y}_{pq}\\right|_{\\beta=4}=\\left[\\begin{array}{cc} Y_s\\otimes \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{p} & 0 \\\\ 0 & \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{q} \\end{array}\\right].\n\\end{equation}\nThe set ${\\rm Mat}(p\/q)$ is the set of $(p+q)\\times(p+q)$ supermatrices on the complex Grassmann algebra $\\bigoplus\\limits_{j=0}^{2pq}\\Lambda_j$. The entries of the diagonal blocks of an element in ${\\rm Mat}(p\/q)$ lie in $\\Lambda^0$ whereas the entries of the off-diagonal block are elements in $\\Lambda^1$.\n\nThe rectangular supermatrix $\\widehat{V}^\\dagger=V^\\dagger U^\\dagger$ is composed of real, complex or quaternionic supervectors whose adjoints form the rows. They are given by\n\\begin{equation}\\label{2.19}\n \\fl\\Psi_j^\\dagger=\\left\\{\\begin{array}{ll}\n \\left(\\left\\{\\sqrt{2}{\\rm Re\\,} z_{jn},\\sqrt{2}{\\rm Im\\,} z_{jn}\\right\\}_{1\\leq n\\leq k_1},\\left\\{\\zeta_{jn},\\zeta^*_{jn}\\right\\}_{1\\leq n\\leq k_2}\\right) & ,\\ \\beta=1,\\\\\n \\left(\\left\\{z_{jn}\\right\\}_{1\\leq n\\leq k_1},\\left\\{\\zeta_{jn}\\right\\}_{1\\leq n\\leq k_2}\\right) & ,\\ \\beta=2,\\\\\n \\left(\\left\\{\\begin{array}{cc} z_{jn} & -z_{j+N,n}^* \\\\ z_{j+N,n} & z_{jn}^* \\end{array}\\right\\}_{1\\leq n\\leq k_1},\\displaystyle\\left\\{\\begin{array}{cc} \\zeta_{jn}^{(-)} & \\zeta_{jn}^{(+)} \\\\ \\zeta_{jn}^{(-)*} & \\zeta_{jn}^{(+)*} \\end{array}\\right\\}_{1\\leq n\\leq k_2}\\right) & ,\\ \\beta=4,\n \\end{array}\\right.\n\\end{equation}\nrespectively, where $\\zeta_{jn}^{(\\pm)}=\\imath^{(1\\pm1)\/2}(\\zeta_{jn}\\pm\\zeta_{j+N,n}^*)\/\\sqrt{2}$. Then, the supermatrix $\\widehat{B}$ acquires the form\n\\begin{equation}\\label{2.20}\n \\widehat{B}=\\frac{1}{\\tilde{\\gamma}}\\sum\\limits_{j=1}^N\\Psi_j\\Psi_j^\\dagger .\n\\end{equation}\nThe integrand in Eq.~\\eref{2.12}\n\\begin{equation}\\label{2.21}\n F\\left(\\widehat{B}\\right)=\\Phi\\left(\\widehat{B}\\right)\\exp\\left(-\\imath{\\rm Str\\,} E\\widehat{B}\\right)\n\\end{equation}\ncomprises a symmetry breaking term,\n\\begin{equation}\\label{2.22}\n \\exists\\ U\\in{\\rm U\\,}^{(\\beta)}(\\gamma_1k_1\/\\gamma_2k_2)\\ {\\rm\\ that\\ \\ }F\\left(\\widehat{B}\\right)\\neq F\\left(U\\widehat{B}U^\\dagger\\right) ,\n\\end{equation}\naccording to the supergroup\n\\begin{equation}\\label{2.23}\n {\\rm U\\,}^{(\\beta)}(\\gamma_1k_1\/\\gamma_2k_2)=\\left\\{\\begin{array}{ll}\n {\\rm UOSp\\,}^{(+)}(2k_1\/2k_2) & ,\\ \\beta=1\\\\\n {\\rm U\\,}(k_1\/k_2) & ,\\ \\beta=2\\\\\n {\\rm UOSp\\,}^{(-)}(2k_1\/2k_2) & ,\\ \\beta=4\n \\end{array}\\right. .\n\\end{equation}\nWe use the notation of Refs.~\\cite{KohGuh05,KKG08} for the representations ${\\rm UOSp\\,}^{(\\pm)}$ of the supergroup ${\\rm UOSp\\,}$. These representations are related to the classification of Riemannian symmetric superspaces by Zirnbauer \\cite{Zir96}. The index ``$+$'' in Eq.~\\eref{2.23} refers to real entries in the Boson--Boson block and to quaternionic entries in the Fermion--Fermion block and ``$-$'' indicates the other way around.\n\n\\section{Supersymmetric Wishart--matrices and some of their properties}\\label{sec3}\n\nWe generalize the integrand \\eref{2.21} to arbitrary sufficiently integrable superfunctions on rectangular $(\\gamma_2c+\\gamma_1d)\\times(\\gamma_2a+\\gamma_1b)$ supermatrices $\\widehat{V}$ on the complex Grassmann--algebra $\\Lambda=\\bigoplus\\limits_{j=0}^{2(ad+bc)}\\Lambda_j$. Such a supermatrix\n\\begin{equation}\\label{3.1}\n \\fl\\widehat{V}=\\left(\\Psi_{11}^{({\\rm C})},\\ldots,\\Psi_{a1}^{({\\rm C})},\\Psi_{12}^{({\\rm C})},\\ldots\\Psi_{b2}^{({\\rm C})}\\right)=\\left(\\Psi_{11}^{({\\rm R})*}\\ldots,\\Psi_{c1}^{({\\rm R})*},\\Psi_{12}^{({\\rm R})*},\\ldots\\Psi_{d2}^{({\\rm R})*}\\right)^{T_{\\rm S}}\n\\end{equation}\nis defined by its columns\n\\begin{eqnarray}\n \\Psi_{j1}^{({\\rm C})\\dagger} & = & \\left\\{\\begin{array}{ll}\n \\left(\\left\\{x_{jn}\\right\\}_{1\\leq n\\leq c},\\left\\{\\chi_{jn},\\chi_{jn}^*\\right\\}_{1\\leq n\\leq d}\\right) & ,\\ \\beta=1,\\\\\n \\left(\\left\\{z_{jn}\\right\\}_{1\\leq n\\leq c},\\left\\{\\chi_{jn}\\right\\}_{1\\leq n\\leq d}\\right) & ,\\ \\beta=2,\\\\\n \\left(\\left\\{\\begin{array}{cc} z_{jn1} & -z_{jn2}^* \\\\ z_{jn2} & z_{jn1}^* \\end{array}\\right\\}_{1\\leq n\\leq c},\\left\\{\\begin{array}{c} \\chi_{jn} \\\\ \\chi_{jn}^* \\end{array}\\right\\}_{1\\leq n\\leq d}\\right) & ,\\ \\beta=4,\n \\end{array}\\right.\\label{3.2a}\\\\\n \\Psi_{j2}^{({\\rm C})\\dagger} & = & \\left\\{\\begin{array}{ll}\n \\left(\\left\\{\\begin{array}{c} \\zeta_{jn} \\\\ \\zeta_{jn}^* \\end{array}\\right\\}_{1\\leq n\\leq c},\\left\\{\\begin{array}{cc} \\tilde{z}_{jn1} & -\\tilde{z}_{jn2}^* \\\\ \\tilde{z}_{jn2} & \\tilde{z}_{jn1}^* \\end{array}\\right\\}_{1\\leq n\\leq d}\\right) & ,\\ \\beta=1,\\\\\n \\left(\\left\\{\\zeta_{jn}\\right\\}_{1\\leq n\\leq c},\\left\\{\\tilde{z}_{jn}\\right\\}_{1\\leq n\\leq d}\\right) & ,\\ \\beta=2,\\\\\n \\left(\\left\\{\\zeta_{jn},\\zeta^*_{jn}\\right\\}_{1\\leq n\\leq c},\\left\\{y_{jn}\\right\\}_{1\\leq n\\leq d}\\right) & ,\\ \\beta=4,\n \\end{array}\\right.\\label{3.2b}\n\\end{eqnarray}\nor by its rows\n\\begin{eqnarray}\n \\Psi_{j1}^{({\\rm R})\\dagger} & = & \\left\\{\\begin{array}{ll}\n \\left(\\left\\{x_{nj}\\right\\}_{1\\leq n\\leq a},\\left\\{ \\zeta_{nj}^*, -\\zeta_{nj} \\right\\}_{1\\leq n\\leq b}\\right) & ,\\ \\beta=1,\\\\\n \\left(\\left\\{z_{nj}^*\\right\\}_{1\\leq n\\leq a},\\left\\{\\zeta_{nj}^*\\right\\}_{1\\leq n\\leq b}\\right) & ,\\ \\beta=2,\\\\\n \\left(\\left\\{\\begin{array}{cc} z_{nj1}^* & z_{nj2}^* \\\\ -z_{nj2} & z_{nj1} \\end{array}\\right\\}_{1\\leq n\\leq a},\\left\\{\\begin{array}{c} \\zeta_{nj}^* \\\\ -\\zeta_{nj} \\end{array}\\right\\}_{1\\leq n\\leq b}\\right) & ,\\ \\beta=4,\n \\end{array}\\right.\\label{3.3a}\\\\\n \\Psi_{j2}^{({\\rm R})\\dagger} & = & \\left\\{\\begin{array}{ll}\n \\left(\\left\\{\\begin{array}{c} -\\chi_{nj}^* \\\\ \\chi_{nj} \\end{array}\\right\\}_{1\\leq n\\leq a},\\left\\{\\begin{array}{cc} \\tilde{z}_{nj1}^* & \\tilde{z}_{nj2}^* \\\\ -\\tilde{z}_{nj2} & \\tilde{z}_{nj1} \\end{array}\\right\\}_{1\\leq n\\leq b}\\right) & ,\\ \\beta=1,\\\\\n \\left(\\left\\{-\\chi_{nj}^*\\right\\}_{1\\leq n\\leq a},\\left\\{\\tilde{z}_{nj}^*\\right\\}_{1\\leq n\\leq b}\\right) & ,\\ \\beta=2,\\\\\n \\left(\\left\\{-\\chi_{nj}^*,\\chi_{nj}\\right\\}_{1\\leq n\\leq a},\\left\\{y_{nj}\\right\\}_{1\\leq n\\leq b}\\right) & ,\\ \\beta=4\n \\end{array}\\right.\\label{3.3b}\n\\end{eqnarray}\nwhich are real, complex and quaternionic supervectors. We use the complex Grassmann variables $\\chi_{mn}$ and $\\zeta_{mn}$ and the real numbers $x_{mn}$ and $y_{mn}$. Also, we introduce the complex numbers $z_{mn}$, $\\tilde{z}_{mn}$, $z_{mnl}$ and $\\tilde{z}_{mnl}$. The $(\\gamma_2c+\\gamma_1d)\\times(\\gamma_2c+\\gamma_1d)$ supermatrix $\\widehat{B}=\\tilde{\\gamma}^{-1}\\widehat{V}\\widehat{V}^\\dagger$ can be written in the columns of $\\widehat{V}$ as in Eq.~\\eref{2.20}. As this supermatrix has a form similar to the ordinary Wishart--matrices, we refer to it as supersymmetric Wishart--matrix. The rectangular supermatrix above fulfills the property\n\\begin{equation}\\label{3.4}\n \\widehat{V}^*=\\widehat{Y}_{cd}\\widehat{V}\\widehat{Y}_{ab}^T .\n\\end{equation}\nThe corresponding generating function \\eref{2.1} is an integral over a rotation invariant superfunction $P$ on a superspace, which is sufficiently convergent and analytic in its real independent variables,\n\\begin{equation}\\label{3.5}\n Z_{cd}^{ab}(E^-) =\\int\\limits_{\\widetilde{\\Sigma}_{\\beta,ab}^{(-\\psi)}} P(\\sigma) {\\rm Sdet\\,}^{-1\/\\tilde{\\gamma}}\\left(\\sigma\\otimes\\widehat{\\Pi}_{2\\psi}^{({\\rm C})}- \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_2a+\\gamma_1b}\\otimes E^-\\right)d[\\sigma],\n\\end{equation}\nwhere\n\\begin{equation}\\label{3.5b}\n \\fl E^-={\\rm diag\\,}\\left(E_{11}\\otimes \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_2},\\ldots,E_{c1}\\otimes \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_2},E_{12}\\otimes \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_1},\\ldots,E_{d2}\\otimes \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_1}\\right)-\\imath\\varepsilon \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_2c+\\gamma_1d} .\n\\end{equation}\nLet $\\widetilde{\\Sigma}_{\\beta,ab}^{0(\\dagger)}$ be a subset of $\\widetilde{\\Sigma}_{\\beta,ab}^{(\\dagger)}$. The entries of elements in $\\widetilde{\\Sigma}_{\\beta,ab}^{0(\\dagger)}$ lie in $\\Lambda_0$ and $\\Lambda_1$. The set $\\widetilde{\\Sigma}_{\\beta,ab}^{(-\\psi)}=\\widehat{\\Pi}_{-\\psi}^{({\\rm R})}\\widetilde{\\Sigma}_{\\beta,ab}^{0(\\dagger)}\\widehat{\\Pi}_{-\\psi}^{({\\rm R})}$ is the Wick--rotated set of $\\widetilde{\\Sigma}_{\\beta,ab}^{0(\\dagger)}$ by the generalized Wick--rotation $\\widehat{\\Pi}_{-\\psi}^{({\\rm R})}={\\rm diag\\,}( \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_2a},e^{-\\imath\\psi\/2} \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_1b})$. As in Ref.~\\cite{KGG08}, we introduce such a rotation for the convergence of the integral \\eref{3.5}. The matrix $\\widehat{\\Pi}_{2\\psi}^{({\\rm C})}={\\rm diag\\,}( \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_2c},e^{\\imath\\psi} \\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_{\\gamma_1d})$ is also a Wick--rotation.\n\nIn the rest of our work, we restrict the calculations to a class of superfunctions. These superfunctions has a Wick--rotation such that the integrals are convergent. We have not explicitly analysed the class of such functions. However, this class is very large and sufficient for physical interests. We consider the probability distribution\n\\begin{equation}\\label{3.5c}\n P(\\sigma)=f(\\sigma)\\exp(-{\\rm Str\\,}\\sigma^{2m}),\n\\end{equation}\nwhere $m\\in\\mathbb{N}$ and $f$ is a superfunction which does not increase so fast as $\\exp({\\rm Str\\,}\\sigma^{2m})$ in the infinty, in particular\n\\begin{equation}\\label{3.5d}\n \\underset{\\epsilon\\to\\infty}{\\lim}P\\left(\\epsilon e^{\\imath\\alpha}\\sigma\\right)=0 \\Leftrightarrow \\underset{\\epsilon\\to\\infty}{\\lim}\\exp\\left(-\\epsilon e^{\\imath\\alpha}{\\rm Str\\,}\\sigma^{2m}\\right)=0\n\\end{equation}\nfor every angle $\\alpha\\in[0,2\\pi]$. Then, a Wick--rotation exists for $P$.\n\nTo guarantee the convergence of the integrals below, let $\\widehat{V}_{\\psi}=\\widehat{\\Pi}_{\\psi}^{({\\rm C})}\\widehat{V}$, $\\widehat{V}_{-\\psi}^\\dagger=\\widehat{V}^\\dagger\\widehat{\\Pi}_{\\psi}^{({\\rm C})}$ and $\\widehat{B}_{\\psi}=\\widehat{\\Pi}_{\\psi}^{({\\rm C})}\\widehat{B}\\widehat{\\Pi}_{\\psi}^{({\\rm C})}$. Considering a function $f$ on the set of supersymmetric Wishart--matrices, we give a lemma and a corollary which are of equal importance for the superbosonization formula and the generalized Hubbard--Startonovich transformation. For $b=0$, the lemma presents the duality relation between the ordinary and superspace \\eref{2.11} which is crucial for the calculation of \\eref{2.1}. This lemma was proven in Ref.~\\cite{LSZ07} by representation theory. Here, we only state it.\n\\begin{lemma}\\label{c1}\\ \\\\\n Let $f$ be a superfunction on rectangular supermatrices of the form \\eref{3.1} and invariant under\n \\begin{equation}\\label{c1.1}\n f(\\widehat{V}_{\\psi},\\widehat{V}_{-\\psi}^\\dagger)=f\\left(\\widehat{V}_{\\psi}U^\\dagger,U\\widehat{V}_{-\\psi}^\\dagger\\right)\\ ,\n \\end{equation}\n for all $\\widehat{V}$ and $U\\in{\\rm U\\,}^{(\\beta)}(a\/b)$. Then, there is a superfunction $F$ on the ${\\rm U\\,}^{(\\beta)}(c\/d)$--symmetric supermatrices with\n \\begin{equation}\\label{c1.2}\n F(\\widehat{B}_{\\psi})=f(\\widehat{V}_{\\psi},\\widehat{V}_{-\\psi}^\\dagger)\\ .\n \\end{equation}\n\\end{lemma}\n\nThe ${\\rm U\\,}^{(\\beta)}(c\/d)$--symmetric supermatrices are elements of $\\widetilde{\\Sigma}_{\\beta,ab}^{(\\dagger)}$. The invariance condition \\eref{c1.1} implies that $f$ only depends on the rows of $\\widehat{V}_{\\psi}$ by $\\Psi_{nr}^{({\\rm R})\\dagger}\\Psi_{ms}^{({\\rm R})}$ for arbitrary $n,m,r$ and $s$. These scalar products are the entries of the supermatrix $\\widehat{V}_{\\psi}\\widehat{V}_{-\\psi}^\\dagger$ which leads to the statement.\n\nThe corollary below is an application of integral theorems by Wegner \\cite{Weg83} worked out in Refs.~\\cite{Con88,ConGro89} and of the Theorems III.1, III.2 and III.3 in Ref.~\\cite{KKG08}. It states that an integration over supersymmetric Wishart--matrices can be reduced to integrations over supersymmetric Wishart--matrices comprising a lower dimensional rectangular supermatrix. In particular for the generating function, it reflects the equivalence of the integral \\eref{3.5} with an integration over smaller supermatrices \\cite{KKG08}. We assume that $\\tilde{a}=a-2(b-\\tilde{b})\/\\beta\\geq0$ with\n\\begin{equation}\\label{3.6}\n \\tilde{b}=\\left\\{\\begin{array}{ll}\n\t\t1 & ,\\ \\beta=4\\ {\\rm and}\\ b\\in2\\mathbb{N}_0+1\\\\\n\t\t0 & ,\\ {\\rm else}\n\t \\end{array}\\right. .\n\\end{equation}\n\\begin{corollary}\\label{c2}\\ \\\\\n Let $F$ be the superfunction of Lemma \\ref{c1}, real analytic in its real independent entries and a Schwartz--function. Then, we find\n \\begin{equation}\\label{c2.1}\n \\displaystyle\\int\\limits_{\\mathfrak{R}}F(\\widehat{B}_{\\psi})d[\\widehat{V}]=C\\int\\limits_{\\widetilde{\\mathfrak{R}}}F(\\widetilde{B}_{\\psi})d[\\widetilde{V}]\n \\end{equation}\n where $\\widetilde{B}=\\tilde{\\gamma}^{-1}\\widetilde{V}\\widetilde{V}$. The sets are $\\mathfrak{R}=\\mathbb{R}^{\\beta ac+4bd\/\\beta}\\times\\Lambda_{2(ad+bc)}$ and $\\widetilde{\\mathfrak{R}}=\\mathbb{R}^{\\beta\\tilde{a}c+4\\tilde{b}d\/\\beta}\\times\\Lambda_{2(\\tilde{a}d+\\tilde{b}c)}$, the constant is\n \\begin{equation}\\label{c2.2}\n C=\\left[-\\frac{\\gamma_1}{2}\\right]^{(b-\\tilde{b})c}\\left[\\frac{\\gamma_2}{2}\\right]^{(a-\\tilde{a})d}\n \\end{equation}\n and the measure\n\\begin{equation}\\label{c2.3}\n d[\\widehat{V}]=\\underset{1\\leq l\\leq \\beta}{\\underset{1\\leq n\\leq c}{\\underset{1\\leq m\\leq a}{\\prod}}}dx_{mnl}\\underset{1\\leq l\\leq 4\/\\beta}{\\underset{1\\leq n\\leq d}{\\underset{1\\leq m\\leq b}{\\prod}}}dy_{mnl}\\underset{1\\leq n\\leq c}{\\underset{1\\leq m\\leq b}{\\prod }}d\\zeta_{mn}d\\zeta_{mn}^*\\underset{1\\leq n\\leq d}{\\underset{1\\leq m\\leq a}{\\prod }}d\\chi_{mn}d\\chi_{mn}^*\\ .\n\\end{equation}\nThe $(\\gamma_2c+\\gamma_1d)\\times(\\gamma_2\\tilde{a}+\\gamma_1\\tilde{b})$ supermatrix $\\widetilde{V}$ and its measure $d[\\widetilde{V}]$ is defined analogous to $\\widehat{V}$ and $d[\\widehat{V}]$, respectively. Here, $x_{mna}$ and $y_{mna}$ are the independent real components of the real, complex and quaternionic numbers of the supervectors $\\Psi_{j1}^{({\\rm R})}$ and $\\Psi_{j2}^{({\\rm R})}$, respectively.\n\\end{corollary}\n\\textbf{Proof:}\\\\\nWe integrate $F$ over all supervectors $\\Psi_{j1}^{({\\rm R})}$ and $\\Psi_{j2}^{({\\rm R})}$ except $\\Psi_{11}^{({\\rm R})}$. Then, \n \\begin{equation}\\label{c2.4}\n \\displaystyle\\int\\limits_{\\mathfrak{R}^\\prime}F(V_{\\psi}V_{-\\psi}^\\dagger)d[\\widehat{V}_{\\neq11}]\n \\end{equation}\nonly depends on $\\Psi_{11}^{({\\rm R})\\dagger}\\Psi_{11}^{({\\rm R})}$. The integration set is $\\mathfrak{R}^\\prime=\\mathbb{R}^{\\beta a(c-1)+4bd\/\\beta}\\times\\Lambda_{2(ad+b(c-1))}$ and the measure $d[\\widehat{V}_{\\neq11}]$ is $d[\\widehat{V}]$ without the measure for the supervector $\\Psi_{11}^{({\\rm R})}$. With help of the Theorems in Ref.~\\cite{Weg83,Con88,ConGro89,KKG08}, the integration over $\\Psi_{11}^{({\\rm R})}$ is up to a constant equivalent to an integration over a supervector $\\widetilde{\\Psi}_{11}^{({\\rm R})}$. This supervector is equal to $\\Psi_{11}^{({\\rm R})}$ in the first $\\tilde{a}$--th entries and else zero. We repeat this procedure for all other supervectors reminding that we only need the invariance under the supergroup action ${\\rm U\\,}^{(\\beta)}\\left(b-\\tilde{b}\/b-\\tilde{b}\\right)$ on $f$ as in Eq.~\\eref{c1.1} embedded in ${\\rm U\\,}^{(\\beta)}(a\/b)$. This invariance is preserved in each step due to the zero entries in the new supervectors. \\hfill$\\square$\n\nThis corollary allows us to restrict our calculation on supermatrices with $b=1$ only to $\\beta=4$ and $b=0$ for all $\\beta$. Only the latter case is of physical interest. Thus, we give the computation for $b=0$ in the following sections and consider the case $b=1$ in Sec. \\ref{sec7}. For $b=0$ we omit the Wick--rotation for $\\widehat{B}$ as it is done in Refs.~\\cite{Guh06,KGG08} due to the convergence of the integral \\eref{3.5}.\n\n\\section{The superbosonization formula}\\label{sec4}\n\nWe need for the following theorem the definition of the sets\n\\begin{eqnarray}\n \\fl\\Sigma_{1,pq} = \\left\\{\\left.\\sigma=\\left[\\begin{array}{ccc} \\sigma_1 & \\eta & \\eta^* \\\\ -\\eta^\\dagger & \\sigma_{21} & \\sigma_{22}^{(1)} \\\\ \\eta^T & \\sigma_{22}^{(2)} & \\sigma_{21}^T \\end{array}\\right]\\in{\\rm Mat}(p\/2q)\\right|\\sigma_1^\\dagger=\\sigma_1^*=\\sigma_1\\ {\\rm with\\ positive}\\right.\\nonumber\\\\\n \\left.{\\rm definite\\ body},\\ \\sigma_{22}^{(1)T}=-\\sigma_{22}^{(1)},\\ \\sigma_{22}^{(2)T}=-\\sigma_{22}^{(2)}\\right\\},\\label{4.1}\\\\\n \\fl\\Sigma_{2,pq} = \\left\\{\\left.\\sigma=\\left[\\begin{array}{cc} \\sigma_1 & \\eta \\\\ -\\eta^\\dagger & \\sigma_2 \\end{array}\\right]\\in{\\rm Mat}(p\/q)\\right|\\sigma_1^\\dagger=\\sigma_1\\ {\\rm with\\ positive\\ definite\\ body}\\right\\},\\label{4.2}\\\\\n \\fl\\Sigma_{4,pq} = \\left\\{\\sigma=\\left[\\begin{array}{ccc} \\sigma_{11} & \\sigma_{12} & \\eta \\\\ -\\sigma_{12}^* & \\sigma_{11}^* & \\eta^* \\\\ -\\eta^\\dagger & \\eta^T & \\sigma_2 \\end{array}\\right]\\in{\\rm Mat}(2p\/q)\\right|\\sigma_1^\\dagger=\\sigma_1=\\left[\\begin{array}{cc} \\sigma_{11} & \\sigma_{12} \\\\ -\\sigma_{12}^* & \\sigma_{11}^* \\end{array}\\right]\\nonumber\\\\\n {\\rm with\\ positive\\ definite\\ body},\\ \\sigma_2=\\sigma_2^T\\Biggl\\}.\\label{4.3}\n\\end{eqnarray}\nAlso, we will use the sets\n\\begin{equation}\\label{4.4}\n \\Sigma_{\\beta,pq}^{(\\dagger)} = \\left\\{\\left.\\sigma\\in\\Sigma_{\\beta,pq}\\right|\\sigma_2^\\dagger=\\sigma_2\\right\\}=\\widetilde{\\Sigma}_{\\beta,pq}^{(\\dagger)}\\cap\\Sigma_{\\beta,pq}\n\\end{equation}\nand\n\\begin{equation}\\label{4.5}\n \\Sigma_{\\beta,pq}^{({\\rm c})} = \\left\\{\\left.\\sigma\\in\\Sigma_{\\beta,pq}\\right|\\sigma_2\\in{\\rm CU\\,}^{(4\/\\beta)}\\left(q\\right)\\right\\}\n\\end{equation}\nwhere ${\\rm CU\\,}^{(\\beta)}\\left(q\\right)$ is the set of the circular orthogonal (COE, $\\beta=1$), unitary (CUE, $\\beta=2$) or unitary-symplectic (CSE, $\\beta=4$) ensembles,\n\\begin{equation}\\label{4.6}\n \\fl{\\rm CU\\,}^{(\\beta)}\\left(q\\right)=\\left\\{A\\in{\\rm Gl}(\\gamma_2q,\\mathbb{C})\\left| \\begin{array}{ll} A=A^T\\in{\\rm U\\,}^{(2)}(q) & ,\\ \\beta=1 \\\\ A\\in{\\rm U\\,}^{(2)}(q) & ,\\ \\beta=2 \\\\ A=(Y_s\\otimes\\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_q)A^T(Y_s^T\\otimes\\leavevmode\\hbox{\\small1\\kern-3.8pt\\normalsize1}_q)\\in{\\rm U\\,}^{(2)}(2q) & ,\\ \\beta=4 \\end{array}\\right.\\right\\}\n\\end{equation}\nThe index ``$\\dagger$'' in Eq.~\\eref{4.4} refers to the self-adjointness of the supermatrices and the index ``${\\rm c}$'' indicates the relation to the circular ensembles. We notice that the set classes presented above differ in the Fermion--Fermion block. In Sec. \\ref{sec6}, we show that this is the crucial difference between both methods. Due to the nilpotence of $B$'s Fermion--Fermion block, we can change the set in this block for the Fourier--transformation. The sets of matrices in the sets above with entries in $\\Lambda_0$ and $\\Lambda_1$ are denoted by $\\Sigma_{\\beta,pq}^{0}$, $\\Sigma_{\\beta,pq}^{0(\\dagger)}$ and $\\Sigma_{\\beta,pq}^{0({\\rm c})}$, respectively.\n\nThe proof of the superbosonization formula \\cite{Som07,LSZ07} given below is based on the proofs of the superbosonization formula for arbitrary superfunctions on real supersymmetric Wishart--matrices in Ref.~\\cite{Som07} and for Gaussian functions on real, complex and quaternionic Wishart--matrices in Ref.~\\cite{Som08}. This theorem extends the superbosonization formula of Ref.~\\cite{LSZ07} to averages of square roots of determinants over unitary-symplectically invariant ensembles, i.e. $\\beta=4$, $b=c=0$ and $d$ odd in Eq.~\\eref{3.5}. The proof of this theorem is given in \\ref{app1}.\n\\begin{theorem}[Superbosonization formula]\\label{t1}\\ \\\\\n Let $F$ be a conveniently integrable and analytic superfunction on the set of $\\left(\\gamma_2c+\\gamma_1d\\right)\\times\\left(\\gamma_2c+\\gamma_1d\\right)$ supermatrices and\n\\begin{equation}\\label{t1.3}\n \\kappa=\\frac{a-c+1}{\\gamma_1}+\\frac{d-1}{\\gamma_2} .\n\\end{equation}\nWith\n\\begin{equation}\\label{t1.0}\n a\\geq c\\ ,\n\\end{equation}\nwe find\n \\begin{equation}\\label{t1.1}\n \\fl\\int\\limits_{\\mathfrak{R}}F(\\widehat{B})\\exp\\left(-\\varepsilon{\\rm Str\\,} \\widehat{B}\\right)d[\\widehat{V}]=C_{acd}^{(\\beta)}\\int\\limits_{\\Sigma_{\\beta,cd}^{0({\\rm c})}}F(\\rho)\\exp\\left(-\\varepsilon{\\rm Str\\,}\\rho\\right){\\rm Sdet\\,}\\rho^{\\kappa}d[\\rho] ,\n \\end{equation}\n where the constant is\n\\begin{eqnarray}\n \\fl C_{acd}^{(\\beta)}\n = \\left(-2\\pi\\gamma_1\\right)^{-ad}\\left(-\\frac{2\\pi}{\\gamma_2}\\right)^{cd}2^{-c}\\tilde{\\gamma}^{\\beta ac\/2}\\frac{{\\rm Vol}\\left({\\rm U\\,}^{(\\beta)}(a)\\right)}{{\\rm Vol}\\left({\\rm U\\,}^{(\\beta)}(a-c)\\right)}\\times\\nonumber\\\\\n \\times\\prod\\limits_{n=1}^d\\frac{\\Gamma\\left(\\gamma_1\\kappa+2(n-d)\/\\beta\\right)}{\\imath^{4(n-1)\/\\beta}\\pi^{2(n-1)\/\\beta}} .\\label{t1.2}\n\\end{eqnarray}\nWe define the measure $d[\\widehat{V}]$ as in Corollary \\ref{c2} and the measure on the right hand side is $d[\\rho]=d[\\rho_1]d[\\rho_2]d[\\eta]$ where\n\\begin{eqnarray}\n \\fl d[\\rho_1] & = & \\prod\\limits_{n=1}^{c}d\\rho_{nn1}\\times\\left\\{\\begin{array}{ll}\n \\prod\\limits_{1\\leq n< m\\leq c}d\\rho_{nm1} & ,\\ \\beta=1,\\\\\n \\prod\\limits_{1\\leq n< m\\leq c}d{\\rm Re\\,}\\rho_{nm1}d{\\rm Im\\,}\\rho_{nm1} & ,\\ \\beta=2,\\\\\n \\prod\\limits_{1\\leq n< m\\leq c}d{\\rm Re\\,}\\rho_{nm11}d{\\rm Im\\,}\\rho_{nm11}d{\\rm Re\\,}\\rho_{nm12}d{\\rm Im\\,}\\rho_{nm12} & ,\\ \\beta=4,\n \\end{array}\\right.\\label{t1.4}\\\\\n \\fl d[\\rho_2] & = & {\\rm FU}_{d}^{(4\/\\beta)}|\\Delta_{d}(e^{\\imath\\varphi_j})|^{4\/\\beta}\\prod\\limits_{n=1}^{d}\\frac{de^{\\imath\\varphi_n}}{2\\pi\\imath}d\\mu(U)\\label{t1.5} ,\\\\\n \\fl d[\\eta] & = & \\prod\\limits_{n=1}^{c}\\prod\\limits_{m=1}^{d}(d\\eta_{nm}d\\eta_{nm}^*)\\label{t1.6}.\n\\end{eqnarray}\nHere, $\\rho_2=U{\\rm diag\\,}\\left(e^{\\imath\\varphi_1},\\ldots,e^{\\imath\\varphi_{d}}\\right)U^\\dagger$, $U\\in{\\rm U\\,}^{(4\/\\beta)}\\left(d\\right)$ and $d\\mu(U)$ is the normalized Haar-measure of ${\\rm U\\,}^{(4\/\\beta)}\\left(d\\right)$. We introduce the volumes of the rotation groups\n\\begin{equation}\\label{t1.7}\n {\\rm Vol}\\left({\\rm U\\,}^{(\\beta)}(n)\\right)=\\prod\\limits_{j=1}^n\\frac{2\\pi^{\\beta j\/2}}{\\Gamma\\left(\\beta j\/2\\right)}\n\\end{equation}\nand the ratio of volumes of the group flag manifold and the permutation group\n\\begin{equation}\\label{t1.8}\n {\\rm FU}_{d}^{(4\/\\beta)}=\\frac{1}{d!}\\prod\\limits_{j=1}^d\\frac{\\pi^{2(j-1)\/\\beta}\\Gamma(2\/\\beta)}{\\Gamma(2j\/\\beta)}\\ .\n\\end{equation}\nThe absolute value of the Vandermonde determinant $\\Delta_{d}(e^{\\imath\\varphi_j})=\\prod\\limits_{1\\leq n10$ for a $40\\,$mas binary observed with a $100\\,$m baseline at $1.7\\,\\mu$m (H band). Such a low spectral resolution is available in most modern interferometric beam combiners. In the case of spatially filtered beam combiners, a similar effect may occur because of baseline smearing, in the case where the telescope size cannot be neglected in Eq.~\\ref{eq:1}. For the $1.8$-m Auxiliary Telescopes, this corresponds to angular separations of about $175\\,$mas for H-band observations. For $10$-m class telescopes, this corresponds to angular separations of about $45\\,$mas. In practice, these limitations are not very relevant to our study because AO-assisted spare aperture-masking imaging on 10-m class telescopes becomes more efficient than long-baseline interferometry for separations larger than about $40\\,$mas \\citep[see e.g.][]{Kraus08,Lacour:2011}.\n\nFinally, Eq.~\\ref{eq:1} assumes that the angular diameters of the binary components are unresolved by the interferometric baselines. Resolving the diameter of the faint component indeed appears unrealistic, although maybe not for the central star, especially in the case of bright late-type giants or very nearby stars. In this latter situation, Eq.~\\ref{eq:1} underestimates the closure phase signal, which peaks for a fully resolved primary star. This effect, referred to as \\emph{closure phase nulling}, can lead to larger magnification factors than the limit $m<149\\deg$ presented in Sect.~\\ref{sec:signal}. This specific observing technique is discussed in detail in \\citet{Chelli:2009}, while on-sky applications can be found in \\citet{Monnier:2006}, \\citet{Lacour08}, \\citet{Zhao08}, and \\citet{Duvert:2010}.\n\n\n\\section{Optical interferometric array}\n\\label{sec:array}\n\nWe use the formalism introduced previously to compute and compare the capabilities of various four-telescope interferometric configurations.\n\n\\subsection{Performances of VLTI configurations}\n\n\\begin{figure}\n \\centering \n \\includegraphics[scale=0.58]{visa}\n \\caption{Interferometric configuration offered at VLTI. The configurations using the four relocatable Auxiliary Telescopes are represented by colours. The configuration using the four fixed Unit Telescopes is represented in black.}\n \\label{fig:visa}\n\\end{figure}\n\n\\begin{figure*}\n \\centering \n \\includegraphics[scale=0.52]{A1G1K0I1_rhoangle1}\\hspace{0.7cm}\n \\includegraphics[scale=0.49]{A1G1K0I1_distance1}\n \\includegraphics[scale=0.52]{A1G1K0I1_rhoangle3}\\hspace{0.7cm}\n \\includegraphics[scale=0.49]{A1G1K0I1_distance3}\n \\includegraphics[scale=0.52]{A1G1K0I1_rhoangle5}\\hspace{0.7cm}\n \\includegraphics[scale=0.49]{A1G1K0I1_distance5}\n \\caption{{\\it Left:} Map of the 3-$\\sigma{}$ sensitivity with the A1-K0-G1-I1 configuration from VLTI. Radial zones are the bins in separation used for the plots in the right panel. {\\it Right:} Sensitivity as a function of angular distance, for various completeness levels. {\\it From top to botom:} Simulations for a single snapshot pointing (top), and for three pointing (middle), and for five pointing (bottom). The contrast axes can be scaled for any accuracy on the closure phase (here $\\sigma=0.25\\deg$).}\n \\label{fig:det_map}\n\\end{figure*}\n\n\\begin{figure*}\n \\centering \n \\includegraphics[scale=0.55]{all_contdist}\\hspace{2cm}\n \\includegraphics[scale=0.55]{all_compcont}\n \\caption{{\\it Left:} 3-$\\sigma{}$ sensitivity versus separation for a completeness level of 80\\%. {\\it Right:} Completeness in the separation range $6-40\\,$mas versus contrast. Simulations are for a single snapshot pointing (dashed lines) and for five pointings separated by one hour (solid lines). Colours are for the four configurations of VLTI displayed in Fig.~\\ref{fig:visa}. The contrast axes can be scaled for any given accuracy of the closure phase measurements (here $\\sigma=0.25\\deg$).}\n \\label{fig:cont_sep}\n \\label{fig:comp_cont}\n\\end{figure*}\n\nWe compute the map of the 3-$\\sigma{}$ sensitivity limit using all the VLTI configurations displayed in Fig.~\\ref{fig:visa}, assuming a target at declination $-35\\,$deg. The limits are computed considering data sets consisting of respectively one pointing at an hour angle $\\mathrm{HA}=-2\\,$h, three pointings at hour angles $\\mathrm{HA}=-2\\,$h, $-1\\,$h and $0\\,$h, and five pointings at hour angles $\\mathrm{HA}=-2\\,$h, $-1\\,$h, $0\\,$h, $+1\\,$h, and $+2\\,$h. A closure phase accuracy of $0.25\\deg$ is assumed (see Sect.~\\ref{sec:accuracy}). We consider a maximum binary separation of $40\\,$mas, which is the separation where AO-assisted spare aperture-masking imaging on 10-m class telescopes becomes more efficient than long-baseline interferometry.\n\nDetailed results of the widest AT configuration (A1-K0-G1-I1) are displayed in Fig.~\\ref{fig:det_map} for illustration. The figure shows that the detection completeness for a given sensitivity level increases drastically with the number of pointings. This is clearly illustrated by the decreasing number of ``blind spots'' (white zones) in the left-hand side plots. The sensitivity level is mostly flat for angular separations larger than 2\\,mas, while the detection performance drops considerably within this inner working angle (IWA). The median sensitivity levels in the region 2--40 mas are respectively about $6\\times 10^{-3}$, $4.5\\times 10^{-3}$ and $4\\times 10^{-3}$ for the three considered data sets. For more than five pointings, the median sensitivity level would continue to improve slightly, but the shape of the sensitivity curve would no longer significantly change.\n\nThe relative performances of the different configurations are illustrated in Fig.~\\ref{fig:cont_sep} (left: sensitivity versus separation, right: completeness level versus contrast). All configurations provide flat performances for large separations, down to their respective IWA where the performances dramatically drop. The IWA are respectively $2\\,$mas for the A1-K0-G1-I1 and U1-U2-U3-U4 configurations, $3\\,$mas for the D0-H0-I1-G1 configuration, and $6\\,$mas for the A1-B2-C1-D0 configuration. They correspond to the spatial resolution of the smallest baseline of the array. Given the similarity of the results for all configurations, we discuss them together in two different regimes: (i) the close-companion and (ii) the wide-companion regimes.\n\n\\subsection{Close-companion regime}\n\nClose companions are defined here as companions with angular separations that are not fully resolved by \\emph{at least} one baseline (that is $\\vec{B}\\cdot\\vec{\\Delta}\/\\lambda<1$). In this regime, the achievable contrast for a given completeness follows a power law of the angular separation $\\ensuremath{\\rho} \\propto \\Delta^{-3}$, as predicted by Eq.~\\ref{eq:close2}. The exact factor entering into this law depends on the array geometry but, as expected, the longest arrays provide the highest spatial resolution and are thus able to detect both the deepest and the closest binaries.\n\nThis well-known result was discussed by \\citet{Lachaume03} in the context of partially resolved interferometric observations. We emphasize that our study additionally provides a quantitative estimation. As a typical example, we now detail the case of a faint companion with a contrast of $5\\times10^{-3}$. We first consider the companion to be located at 2\\,mas from the central star. A closure-phase accuracy of $0.25\\deg$ results in a detection efficiency of about 50\\% using three pointings with the configuration A0-K0-G1-I1, according to Fig.~\\ref{fig:det_map} (middle-right plot). However, if we now consider the companion to be located at $1\\,$mas from the central star, the closure-phase accuracy should be $0.025\\deg$ to reach the same efficiency. Since the angular separation is only marginally resolved by the interferometer, the lack of spatial resolution has to be compensated for by an increase in the accuracy on the signal (super-resolution effect).\n\n\\subsection{Wide-companion regime}\n\nWe note that those companions are \\emph{wide} only in the interferometric sense, corresponding to separations larger than about 4\\,mas for the typical $\\sim100\\,$m baselines available in modern interferometric facilities. \n\nIn this regime, the detection efficiency becomes independent of the companion separation. Interestingly, all arrays have the same efficiency. In other words, as long as the companion is expected to be resolved by the interferometric baselines, the choice of array configuration does not matter. We conclude that there is no reason to favour a given VLTI configuration when looking for faint unknown companion with separations in the range $6-40\\,$mas. More quantitatively, Fig.~\\ref{fig:cont_sep} (right) displays the detection efficiency in this annular region for the four VLTI configurations versus the companion contrast, and for two observing scenarios (snapshot and long integration). The combination of five observations separated by one hour provides a detection efficiency higher than 95\\% for companion contrasts of $10^{-2}$, assuming a realistic closure phase accuracy of $0.25\\deg$. \n\nWe note that the curve of completeness versus contrast become significantly sharper when accumulating observations. As shown by the solid lines in the right panel of Fig.~\\ref{fig:cont_sep}, when accumulating five pointings, the efficiency drops from 80\\% for a contrast of $5\\times10^{-3}$ to less than 10\\% for a contrast of $3\\times10^{-3}$. The constraints provided by this dataset can thus be presented as a sensitivity limit and an inner working-angle, as for a classical imaging observation.\n\nQuantitatively, when accumulating several pointings, these detection limits computed from the derivation of Sect.~\\ref{sec:theory} are compatible with the blind-test analyses presented by \\citet[Fig.~5]{Absil:2011} and the Monte-Carlo simulations of \\citet[Fig.~4 and Eq.~2]{Lacour:2011}.\n\n\\subsection{Performances of standard configurations}\n\\label{sec:fake}\n\nTo add generality to the results presented in the previous section, we now study configurations that are not specifically linked to any existing interferometric array. We select the configurations presented in Fig.~\\ref{fig:fake}, which is a non-redundant linear configuration, a fully redundant linear configuration and a Y-shaped configuration. All configurations have their longest baseline of the same size. The results are the following:\n\n\\begin{enumerate}\n\\item The linear non-redundant and Y-shaped configurations have similar detection limits as the currently offered (irregular) VLTI configurations for snapshot observations. Surprisingly, they also have similar performances when accumulating several pointings, while we may have expected that the Y-shape would unveil faster the remaining blind-spots.\n\\item For snapshot observations, the linear redundant configuration favours the \\emph{dynamic range} with respect to the \\emph{completeness}: it has a fainter detection limit for completeness levels below 50\\%, but becomes significantly worse for higher completeness levels. When accumulating several pointings, both the highest completeness and largest dynamic range are reached, although the gain is never higher than 20\\%.\n\\item Y-shaped arrays have smaller inner working angles than linear configurations of identical maximum baseline length, even when considering the accumulation of several pointings. The gain is almost a factor of two in terms of angular resolution.\n\\end{enumerate}\n\n\\begin{figure}\n \\centering \n \\includegraphics[scale=0.58]{fake}\n \\caption{Fake VLTI interferometric configurations used in this paper: non-redundant linear D0-E0-H0-K0 (red), redundant linear D9-G2-H9-K9 (green, fake stations), and Y-shaped E0-J3-J2-H0 (blue).}\n \\label{fig:fake}\n\\end{figure}\n\n\\begin{figure*}\n \\centering \n \\includegraphics[scale=0.55]{all_contdist_fake}\\hspace{2cm}\n \\includegraphics[scale=0.55]{all_compcont_fake}\n \\caption{{\\it Left:} 3-$\\sigma{}$ sensitivity versus separation for a completeness level of 80\\%. {\\it Right:} Completeness in the separation range $6-40\\,$mas versus contrast. Simulations are for a single snapshot pointing (dashed lines) and for five pointing separated by one hour (solid lines). Colours are for the four configurations displayed in Fig.~\\ref{fig:fake}. The black curves are for the U1-U2-U3-U4 configuration displayed in Fig.~\\ref{fig:visa}. The contrast axes can be scaled for any accuracy on the closure phase (here $\\sigma=0.25\\deg$). }\n \\label{fig:cont_sep_fake}\n \\label{fig:comp_cont_fake}\n\\end{figure*}\n\n\n\n\n\\section{Closure phase accuracy and achievable dynamic range}\n\\label{sec:accuracy}\n\n\\subsection{Photon and piston noises}\nWe consider the theoretical photon noise limit for an observation of 1\\,h on a star of sixth magnitude using one-metre class telescopes (e.g., the $1.8\\,$m auxiliary telescopes of the VLTI). The choice of a sixth-magnitude star is driven by the current sensitivity limit of most interferometric instruments world-wide. A crude estimation of the photon noise is given by\n\\begin{equation}\n\\sigma_{phot} \\approx \\frac{360\\deg}{\\sqrt{N}} \\; ,\n\\end{equation}\nwhere $N$ is the total number of detected photons. If we consider that the 1\\,h observing time should include the overheads and the observation of a calibration star, the effective integration time on target will be of the order of 20\\,min. Assuming a realistic total transmission of 2\\%, including both the instrumental and atmospheric contributions, the number of detected photons is $N\\approx 10^8$ and the resulting photon noise is $\\sigma_{phot} \\approx 0.1\\deg$. For a $1^\\mathrm{mag}$ star, the resulting photon noise would be $\\sigma_{phot} \\approx 0.01\\deg$.\n\nIn theory, closure phase is a robust observable against the telescope phase errors \\citep{Monnier03}. However, in the context of non-zero exposure times and the presence of atmospheric turbulence, closure phase measurements are also affected by piston noise. A proper estimation of its amplitude is beyond the scope of this paper but it is still possible to provide a rough upper limit. Since piston noise is independent of the number of incident photons, it is expected to dominate the final statistical uncertainty for very bright stars. As an example, in the case of the PIONIER instrument at VLTI, the statistical uncertainty for bright stars is typically of the order of $0.25\\deg$ to $2.5\\deg$ for an integration time of 1\\,min. This uncertainty depends on the atmospheric conditions as expected for piston noise. In decent atmospheric conditions, integrating over 20\\,min allows the piston noise contribution to be reduced below $0.2\\deg$, as it decreases with the square root of the integration time.\n\n\n\t\\subsection{Calibration accuracy}\n\nIt is interesting to compare these fundamental limits to published accuracies that include the calibration of the instrumental closure phase (also called the transfer function):\n\\begin{description}\n\\item[VLTI\/AMBER:] \\citet{Absil:2010} reported calibration errors of between $0.20\\deg$ and $0.37\\deg$ depending on the night, using this three-telescope combiner in its medium spectral resolution mode ($R=1\\,500$). With the low spectral resolution mode ($R=35$), typical calibration errors range from one to a few degrees \\citep[see for instance][]{kraus:2009apr,le-bouquin:2009mar}.\n\\item[VLTI\/PIONIER:] Typical calibration errors range from $0.25\\deg$ to $1\\deg$ \\citep{LeBouquin:2011,Absil:2011} for this four-telescope combiner. Sequences with closure phases stable down to $0.1\\deg$ have been recorded. Systematic discrepancies have been noted when calibration stars were separated by more than $10\\deg$ on the sky.\n\\item[CHARA\/MIRC:] The typical accuracy obtained with this four-telescope combiner is between $0.1$ and $0.2\\deg$, which makes this instrument the most accurate of the currently available suite. Calibration uncertainties dominate the final accuracy at this level \\citep{Zhao08,Zhao:2010,Zhao11}.\n\\end{description}\n\nAltogether, a typical noise floor of $\\sim0.25\\deg$ seems to appear for the calibration of the closure phases in long baseline interferometry. Two results indicate that the major cause is probably longitudinal dispersion: (i) that the accuracy depends on the spectral resolution and (ii) the dependence on position of the calibration star on the sky. This is also the finding of \\citet{Zhao11}, who proposed an elaborate calibration scheme for MIRC. Although this is clearly a very promising way of characterizing already known substellar companions, this method is probably not suited to surveying a large number of stars with a standard calibration procedure. Interestingly, $0.25\\deg$ is also the noise floor reported by \\citet{Lacour:2011} for the calibration of the closure phase of the spare aperture masking mode of NACO at VLT. This calibration noise floor of $0.25\\deg$ theoretically does not prevent us from reaching very high dynamic ranges, by accumulating a large number of observations and\/or baselines, as for instance in sparse aperture masking or spectrally dispersed observations. Care should however be taken to ensure that individual closure phase measurements are statistically independent. In particular, one should avoid repeating the same systematic errors in individual data sets, e.g., by choosing different calibrator stars, instrumental setups, etc., in order not to reach a true noise floor in the observations.\n\nConcerning future instruments, the announced accuracy on the closure phases is $1\\deg$ for the K-band four-telescope combiner GRAVITY \\citep[document \\mbox{VLT-SPE-ESO-15880-4853} and][]{Gillessen:2010} and from $1\\deg$ to $5\\deg$ for the L-band four-telescope combiner MATISSE \\citep[Florentin Millour, private communication and][]{Lopez:2008}. Although these performances may be conservative, we conclude that the next generation of VLTI instruments is unlikely to break the $0.25\\deg$ limit.\n\n\n\t\\subsection{Discussion}\n\nIt appears realistic to reach a closure phase accuracy of $0.25\\deg$ within less than one hour on one-metre class telescopes for stars of magnitude six and brighter. According to the results of Sect.~\\ref{sec:array}, such performances allow a dynamic range of $5\\times 10^{-3}$ ($\\Delta \\mathrm{mag}=5.75$) to be reached with 80\\% completeness when five pointings are obtained with a four-telescope interferometer. The same performance would probably be reached within a snapshot using an interferometric instrument combining six telescopes or more at a time.\n\nThis result can be compared with the survey for stellar and sub-stellar companions on the ten-metre Keck and five-metre Palomar telescopes using aperture masking techniques in the K band \\citep{Kraus08,Kraus:2011}. The achieved dynamic range is $\\Delta K\\approx5.5$ for separation as small as $25\\,$mas (slightly worse for Palomar). A similar dynamic range and inner working angle have been achieved within an ongoing survey of massive stars using aperture masking at VLT\/NACO in the H band, e.g. $\\Delta H\\approx5$ down to $25\\,$mas with this eight-metre diameter telescope (Hugues Sana, private communication). All close companions presented in these near-infrared surveys would have been detected by interferometry with an efficiency higher than 90\\%, provided that they reside within the interferometric field-of-view. In addition, this efficiency would have been achieved down to about 2\\,mas. At longer wavelengths, the aperture masking technique has a dynamic range of about $\\Delta L\\approx 7.5$ \\citep{Hinkley:2011}, although the inner working angle in that case is only $70\\,$mas. There is currently no L-band interferometric beam-combiner with closure phase capabilities to which these performances could be compared.\n\nSeveral observing programs related to faint companion detection would benefit significantly from the capabilities of closure phase measurements on long-baseline interferometric instruments. An example is the search for low-mass (sub)-stellar companions around main-sequence stars residing in nearby young associations. Considering associations with ages between 10\\,Myr and 200\\,Myr, and a limiting magnitude $K=6$ for the instrument, one could survey stars up to about 15-20\\,pc for stellar type M0V, 40\\,pc for type G0V, and 120\\,pc for type A0V. With an estimated median dynamic range of $\\Delta K \\simeq 6$, we computed the masses of the faintest companions that could be detected within a survey of nearby moving groups, using the (sub-)stellar cooling models of \\citet{Baraffe98,Baraffe03}. The results are given in Table~\\ref{tab:associations}, showing that the $13 M_{\\rm Jup}$ ($=0.012 M_{\\odot}$) limit between the brown dwarf and planetary regimes can be reached for young late-type dwarfs. In the case of A-type stars, one would be sensitive to companions in the range M3V-M7V depending on the age. For even younger stars, located in nearby star forming regions, closure phase measurements have the potential to reveal the formation of planetary-mass objects, as suggested by \\citet{Kraus12}.\n\n\\begin{table}[t]\n\\caption{Sensitivity limits in terms of companion masses around young main-sequence stars.}\n\\centering\n\\begin{tabular}{c c c c}\n\\hline\\hline\nAge & A0V & G0V & M0V \\\\\n\\hline\n10 Myr & $0.09 M_{\\odot}$ & $0.017 M_{\\odot}$ & $0.012 M_{\\odot}$ \\\\\n50 Myr & $0.22 M_{\\odot}$ & $0.043 M_{\\odot}$ & $0.013 M_{\\odot}$ \\\\\n200 Myr & $0.35 M_{\\odot}$ & $0.08 M_{\\odot}$ & $0.030 M_{\\odot}$ \\\\\n\\hline\n\\end{tabular}\n\\label{tab:associations}\n\\end{table}\n\nAnother program is the determination of the binary fraction for massive stars. The interest is that despite the preponderance of multiple stars, the mechanism that produces multiple stars rather than single stars is still uncertain. The measurement of the mass distribution, and how it evolves with the mass of the primary, is an appropriate tool for disentangling between \\emph{capture} and \\emph{fragmentation} models. Stellar companions to B-type stars have been investigated using AO \\citep[e.g. ][]{Roberts:2007}, although the stars observed typically have a large range of distances, limiting the statistical significance of the results. Radial velocity measurement of massive stars is challenging owing to the lack of suitable spectral lines and their intrinsic broadening. With the limiting magnitude $K=6$ presented in this paper, it is possible to observe interferometrically all the B stars within a distance of 75\\,pc ($\\sim$$100$ objects for the southern hemisphere), providing the first comprehensive study of massive binaries in the $0.25-5\\,$AU separation range. \n\nLast but not least, one of the main selling argument for high-precision closure phases in optical interferometry has been the direct detection of hot extrasolar giant planets (EGP). Several hot EGP host stars are indeed bright enough to be observed with state-of-the-art interferometric instruments. For mature planetary systems ($>10$\\,Myr), the expected contrast between the planet and the star is however generally too low ($<10^{-3}$) to be currently accessible with closure phase measurements \\citep{Zhao08,Zhao11}. To routinely reach the hot EGP regime ($\\Delta K \\simeq 8 - 10$), a gain of two to four magnitudes is required in the dynamic range, which would translate into a noise floor of between $0.04\\deg$ and $0.006\\deg$ on the closure phase. Achieving such an accuracy would probably require a significant breakthrough in the instrumental domain.\n\n\n\\section{Conclusions}\n\nIn summary, optical interferometric surveys designed to detect faint companions have the following properties:\n\\begin{enumerate}\n\\item The observable (closure phase) is robust against unstable atmospheric seeing conditions \\citep{Monnier03}. Integrating over $20\\,$min is sufficient to consistently reduce the photon and atmospheric noises below $0.25\\deg$, which appears as a hard limit for the calibration of current instruments.\n\\item A single snapshot with four telescopes provides a 80\\% detection efficiency at \\mbox{$\\Delta\\mathrm{mag}=4.5$} as soon as the binary separation is fully resolved. The only requirement of the interferometric array is to use baselines as long as possible to improve the inner working angle, which is typically of the order of a few milli-arcseconds.\n\\item Accumulating more observations (several pointing and\/or recombining more telescopes) allows a dynamic range \\mbox{$\\Delta\\mathrm{mag}=6$} to be reached, which appears to be a realistic limit in respect to published performances. Going deeper would require us to break the current limit of $0.25\\deg$ on the closure phase accuracy, or to massively increase the number of observations.\n\\item The achievable dynamic range scales linearly with the closure phase accuracy.\n\\end{enumerate}\n\nIn conclusion, interferometric closure phase surveys would be well-suited as filler programs for service-mode interferometric facilities, such as the VLTI. They can be considered as a useful complement to the AO-assisted imaging surveys currently carried out on ten-metre class telescopes. In particular, the search space of long-baseline interferometry bridges the gap between the wide companions found in direct imaging and the close companions detected by RV measurements. Moreover, interferometry could nicely complement RV studies in the particular cases where RV measurements are quite inappropriate. Young stars, for instance, are especially promising targets since their (sub)stellar companions are supposed to be relatively bright compared to their host stars.\n\n\\begin{acknowledgements} \nThe authors thank the referee for his helpful comments. This work has made use of the Smithsonian\/NASA Astrophysics Data System (ADS) and of the Centre de Donnees astronomiques de Strasbourg (CDS). All calculations and graphics were performed with the freeware \\texttt{Yorick}\\footnote{\\texttt{http:\/\/yorick.sourceforge.net}}.\n\\end{acknowledgements}\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\nDiscovering new particles would entail the Standard Model (SM) being falsified in Popper's sense~\\cite{Popper:1934} and force us to extend it. Absent such discovery, \nthe SM is still falsifiable upon finding new forces among the known particles. Because the SM has a characteristic energy scale of 100 GeV (the mass of the Higgs boson at $m_h=125$ GeV, as well as the vacuum constant $v=246$ GeV exemplify it), but no new particles below 1000 GeV have been found, there is a scale separation that begs the use of Effective Field Theory.\n\nThe popular SMEFT extension of the SM Electroweak Symmetry Breaking Sector (EWSBS) adds to it operators classified by their mass dimension,\n\\begin{equation}\n \\mathcal{L}_{\\rm SMEFT} =\n \\mathcal{L}_{\\rm SM} + \n \\sum_{n=5}^{\\infty}\n \\sum_i\n \\frac{c_i^{(n)}}{\\Lambda^{n-4}} \\mathcal{O}_i^{(n)}(H) \\ .\n \\label{SMEFTL}\n \\end{equation}\nThese operators $\\mathcal{O}_i^{(n)}(H)$, whose intensity is controlled by the Wilson coefficients $c_i^{(n)}$ for a given reference scale $\\Lambda$, are the potentials of those new forces being sought. \nA nonzero $c_i^{(n)}$ would signal departure from the SM, that then would need to be extended,\n{perhaps by new resonances~\\cite{Dobado:2017lwg}}.\n\nBut it is easy to ask oneself how the whole framework of SMEFT can be tested. \nEffective theories include all the possible interactions that are compatible with the known particle content and symmetries believed to hold. Would it not be that any separation from the SM could be recast in SMEFT form? In that case, absent some new light particle, any phenomena could be described by adding an operator with a parameter to the SM. This is not so, as we will detail. \n\nThe particle content of the electroweak sector is packaged in a Higgs doublet field in the SM as well as in SMEFT\n\\begin{equation}\nH = \n\\frac{1}{\\sqrt{2}} \\begin{pmatrix} \\varphi_1+i\\varphi_2 \\\\ \\varphi_0 + i\\varphi_3\n\\end{pmatrix} \n= \nU(\\boldsymbol{\\omega}) \\begin{pmatrix} 0 \\\\ (v+h_{\\rm SMEFT})\/\\sqrt{2}\n\\end{pmatrix} \\, ,\n\\end{equation} \nwhere the Cartesian coordinates $\\varphi_a$ can be rearranged to the polar decomposition in terms the $\\omega_i$ Goldstone bosons (which set the orientation of $H$ through the unitary matrix $U(\\boldsymbol{\\omega})$) and the radial coordinate $h_{\\rm SMEFT}$ (with $|H|=(v+h_{\\rm SMEFT})\/\\sqrt{2}$). \n\n\n\n\\begin{figure}[!b] \n\\centering\n \\begin{tikzpicture}[scale=1]\n \\draw[decoration={aspect=0, segment length=1.8mm, amplitude=0.7mm,coil},decorate] (-1,1) -- (0,0)-- (-1,-1);\n \\draw[] (0,0)-- (1,1);\n \\draw[] (1.25,1) node {$h_1$};\n \\draw[] (1.25,-1) node {$h_n$};\n \\draw[] (0,0)-- (.98,0.6);\n \\draw[] (1.25,0.6) node {$h_2$};\n \\draw[] (.75,-.15) node {.};\n \\draw[] (.75,0.) node {.};\n \\draw[] (.75,-0.3) node {.};\n \\draw[] (0,0)-- (1,-1);\n \\draw[] (3.25,0) node {$\\displaystyle{ \\, =\\, -\\frac{n!a_n}{2v^n} \\, s }$}; \n\\end{tikzpicture}\n\\caption{\\label{fig:Feynman}\\small\nThe $\\omega\\omega\\to nh$ processes can be the key to disentangling the nature of the EWSBS. They give direct access to the $a_i$ coefficients of the flare function $\\mathcal{F}$, and hence to their correlations, as listed in Table~\\ref{tab:correlations}.\n}\n\\end{figure}\n\n\n\nAn additional non-linear redefinition of $h_{\\rm SMEFT}$ allows us to rearrange the SMEFT Lagrangian in the form of a more general theory, HEFT: \n\\begin{align} \\label{HEFTL}\n{\\cal L}_{\\rm HEFT} = \\frac{1}{2}\\partial_\\mu h_{\\rm HEFT}\\partial^\\mu h_{\\rm HEFT}-V(h_{\\rm HEFT}) + \\nonumber \\\\ \n\\frac{1}{2}\\!\\mathcal{F}(h_{\\rm HEFT})\n\\partial_\\mu \\omega^i \\partial^\\mu \\omega^j\\!\\left(\\!\\delta_{ij}\\!+\\!\\frac{\\omega^i\\!\\omega^j}{v^2\\!-\\!\\boldsymbol{\\omega}^2}\\!\\right)\\ .\n\\end{align}\nOf current focus therein is the flare function~\\cite{Alonso:2016oah, Grinstein:2007iv} \n\\begin{equation} \\label{Fexpansion}\n {\\mathcal F}(h_{\\rm HEFT})=1+\\sum_{n=1}^{\\infty}{a_n}\\Big(\\frac{h_{\\rm HEFT}}{v}\\Big)^n \\,,\n\\end{equation}\nwhich amounts to a radial ``scale'' (think of $a(t)$ in a Friedmann-Robertson-Walker cosmology) in the field space of the $(h,\\omega_i)$ electroweak bosons (with $\\omega_i$ analogous to the spatial coordinates). \nWhat we call attention to in this letter is that the Taylor-series coefficients of $\\mathcal{F}$ as defined in Eq.~(\\ref{Fexpansion}) must satisfy experimental correlations or constraints as given in Table~\\ref{tab:correlations} and in the companion extended manuscript~\\cite{Gomez-Ambrosio:2022giw} \n{\\it if SMEFT is a valid description}. It is clear that an experimental program aimed at these correlations\nvia the key process to access $\\mathcal{F}$, $\\omega\\omega\\to nh$ as sketched in Figure~\\ref{fig:Feynman}, or $mh\\to nh$ to access $V(h_{\\rm HEFT})$, can test the validity of SMEFT itself, and not only its parameters.\n\n\n\n\n\\begin{figure}[!t]\n \\centering\n \\includegraphics[width=0.8\\columnwidth]{Bandasa1a2.png}\n \\caption{\\small The correlation $a_2=2a_1-3$ that SMEFT predicts at order $1\/\\Lambda^2$ is plotted against the current 95\\% confidence intervals for these two HEFT parameters~\\cite{ATLAS:2020qdt,CMS:2022cpr}.}\n \\label{fig:a1a2}\n\\end{figure}\n\n\n\n\\begin{table}[!t] \n\\setlength{\\arrayrulewidth}{0.3mm}\n\\setlength{\\tabcolsep}{0.2cm} \n\\renewcommand{\\arraystretch}{1.4}\n \\caption{\\small Correlations between the $a_i$ HEFT coefficients necessary for SMEFT to exist, at order $\\Lambda^{-2}$ (first and second columns, with the numbers in the second consistent with 95\\% confidence-level experimental bounds $a_1\/2 \\in [0.97,1.09]$~\\cite{ATLAS:2020qdt}).\n The right column provides the corresponding numerical values at the next order~\\cite{Gomez-Ambrosio:2022giw}. \n \n They are quoted in terms of $\\Delta a_1:=a_1-2$ and $\\Delta a_2:=a_2-1$, so that all objects in the table vanish in the Standard Model, with all the equalities becoming $0=0$. \n \\label{tab:correlations}}\n \\centering\n \\begin{tabular}{|c|c|c|}\\hline\n $\\mathcal{O}(1\/\\Lambda^2)$ & $\\mathcal{O}(1\/\\Lambda^2)$ & $\\mathcal{O}(1\/\\Lambda^4)$ \\\\ \\hline \n $\\Delta a_2=2\\Delta a_1$ & $\\Delta a_2\\in [-0.12,0.36] $ & \\\\\n $a_3=\\frac{4}{3} \\Delta a_1$ & $a_3\\in[-0.08,0.24]$ & $a_3\\in [-3.1,1.7]$ \\\\ \n $a_4=\\frac{1}{3} \\Delta a_1$ & $a_4\\in [-0.02,0.06] $ & $a_4\\in [-3.3,1.5]$ \\\\\n $a_5=0$ & & $a_5 \\in[-1.5,0.6]$ \\\\\n $a_6=0$ & & $ a_6=a_5$ \\\\ \n \\hline\n \\end{tabular}\n\\end{table}\n\n\n\n\nAs an example, the correlation predicted by SMEFT among $a_1$ and $a_2$ is shown in Figure~\\ref{fig:a1a2}.\n\n\n\n\n\n\nThere are several reasons why the experimental tests of those coefficients need large energies and statistics, at the limit of what is possible today at the LHC and beyond.\nFirst, the $\\mathcal{F}$ function multiplies terms with derivatives of the Goldstone bosons $\\partial_\\mu \\omega_i\\to q_\\mu \\omega_i$ that yield couplings proportional to their four-momenta, and become more relevant at higher energies. \nSecond, \n the equivalence theorem~\\cite{Veltman:1989ud,Dobado:1993dg} tells us that the scattering of longitudinal gauge bosons is related to scattering of Goldstone boson ($\\omega_i \\sim W^\\pm_L,\\ Z_L$): the EW gauging of the HEFT Lagrangian~(\\ref{HEFTL}) leads to the $WW\\to n h$ interaction $\\Delta\\mathcal{L}_{\\rm HEFT}^{WW\\to n h} = \\left(\\frac{1}{2}m_Z^2 Z_\\mu Z^\\mu + m_W^2 W_\\mu^+ W_\\mu^-\\right) \\mathcal{F}(h_{\\rm HEFT})$, which for longitudinal gauge bosons clearly dominates over the non-derivative interactions from $V$ only at high energies~\\cite{Dicus:1987ez,Kallianpur:1988cs}. \nAnd third, an increasing number of Higgs bosons (necessary to access each $h^n$ order of $\\mathcal{F}$, the Higgs-flare function) requires an ample phase space and, thus, high energy. \n\nHowever, as we shortly show after\nEq.~(\\ref{changeofvariable}) below, the correlations arise from the need for\nconsistency of the SMEFT formulation when a change of variable $h_{\\rm HEFT}\\to h_{\\rm SMEFT}$ is performed. This change affects any other piece of the Lagrangian that involves the Higgs bosons, such as the Yukawa couplings to fermions (that we leave for future works) but also the interactions among Higgs bosons themselves. Such interactions do not require derivative couplings and first appear even in the renormalizable SM, \n\\begin{equation}\n{\\mathcal{L}}_{\\rm SM} = |\\partial H|^2 - \\underbrace{\\left( \\mu^2 |H|^2 +\\lambda|H|^4\\right)}_{V(H)} \\, ,\n\\end{equation}\nin the much discussed $V(H)$ \nHiggs-potential, accessible already at low $\\sqrt{s}$. In HEFT, this potential has additional non-renormalizable couplings and can be organized in a power-series expansion\n\\begin{align}\\label{expandV}\n V_{\\rm HEFT}= \\frac{m_h^2 v^2}{2} \\Bigg[& \\left(\\frac{h_{\\rm HEFT}}{v}\\right)^2 + v_3 \\left(\\frac{h_{\\rm HEFT}}{v}\\right)^3 \\nonumber+ \\\\ &+ v_4 \\left(\\frac{h_{\\rm HEFT}}{v}\\right)^4 + \\dots \\Bigg]\\,,\n\\end{align}\nwith $v_3=1$, $v_4=1\/4$ and $v_{n\\geq 5}=0$ in the SM. \nIts coefficients also need to satisfy constraints that are exposed in Table~\\ref{tab:corV} and Figure~\\ref{fig:a1a22} if and when SMEFT applies.\n\n\nLet us then see, very briefly, how these correlations come about. Instead of relying on the powerful geometric methods of~\\cite{Alonso:2016oah,Alonso:2015fsp,Alonso:2016btr,Alonso:2021rac,Alonso:2022ffe} we use the more pedestrian coordinate-dependent approach, more familiar to phenomenologists working on LHC physics.\nThe goal is to see when is it possible to cast \nEq.~(\\ref{HEFTL}) into the specific SMEFT one, Eq.~(\\ref{SMEFTL}).\nThis we write as \n\\begin{align}\n \\mathcal{L}_{\\rm SMEFT}= |\\partial H|^2-V(|H|^2)\n+\\frac{1}{2} B(|H|^2)(\\partial (|H|^2))^2 +\\dots \n\\label{eq:SMEFTL-kin+V+B}\n\\end{align}\nwhere the non-derivative and derivative terms, respectively given by $V$ and $B$, \ncollect typical SMEFT operators of Eq.~(\\ref{SMEFTL}) such as, at lowest $1\/\\Lambda^2$ order,\n\\begin{equation}\\label{operatorsSMEFT}\n \\mathcal{O}_H := (H^\\dagger H)^3 \\, , \\ \\ \\ \n \\mathcal{O}_{H \\Box} := (H^\\dagger H) \\Box (H^\\dagger H)\\ .\n\\end{equation} \n There are also other operators, such as, e.g., \n$ \\mathcal{O}_{HD} = (H^\\dagger D_{\\mu} H)^* (H^\\dagger D^{\\mu} H) $, but they break custodial symmetry, and LEP studies suggest that the $SU(2)\\times SU(2)\\to SU(2)$ electroweak symmetry breaking mechanism is the appropriate pattern, leaving the residual custodial $SU(2)$ as a good approximate global symmetry of the scalar sector. \nThe additional $A(H)$ structure pointed out in~\\cite{Cohen:2020xca} for Lagrangian~(\\ref{eq:SMEFTL-kin+V+B}) can be eliminated through partial integration and the use of the equations of motion~\\cite{Gomez-Ambrosio:2022giw}. \n\n \n \n\n\n\\begin{figure}[!t]\n \\centering\n \\includegraphics[width=0.8\\columnwidth]{Bandasv3v4.png}\n \\caption{\\small The correlation $v_4=\\frac{3}{2} v_3 -\\frac{5}{4} -\\frac{1}{6}\\Delta a_1$ that SMEFT predicts at $\\mathcal{O}(1\/\\Lambda^2)$ is plotted making use of current 95\\% confidence interval for $v_3\\in[-2.5,5.7]$~\\cite{ATLAS:2021jki}. The experimental $a_1$ uncertainty~\\cite{ATLAS:2020qdt,CMS:2022cpr}, $a_1\/2\\in[0.97,1.09]$, is numerically negligible and allows to predict a SMEFT band given by the solid black line. An experimental measurement for $v_4$ is still missing. }\n \\label{fig:a1a22}\n\\end{figure}\n\\begin{table}[!t]\n\\setlength{\\arrayrulewidth}{0.3mm} \n\\setlength{\\tabcolsep}{0.2cm} \n\\renewcommand{\\arraystretch}{1.4} \n\\caption{\\small Correlations among the coefficients $\\Delta v_3:=v_3-1$, $\\Delta v_4:=v_4-1\/4$, $v_5$ and $v_6$ of the HEFT Higgs potential expansion in Eq.~(\\ref{expandV}) that need to hold, at $\\mathcal{O}(1\/\\Lambda^2)$, if SMEFT is a valid description of the electroweak sector.\nBased on the current bound $\\Delta v_3\\in[-2.5,5.7]$ in Ref.~\\cite{ATLAS:2021jki}, $\\mathcal{O}(1\/\\Lambda^2)$ SMEFT predicts the coefficient intervals in the last column, testable in few-Higgs final states. A coupling \n$c_{H\\Box}\\neq 0$ induces the correction $\\Delta a_1\\propto c_{H\\Box}$, nevertheless numerically negligible since $v_3$ experimental uncertainties much exceed those of $a_1$.}\n\\label{tab:corV}\n\\begin{center}\n \\begin{tabular}{|c|c|} \\hline \n $\\Delta v_4=\\frac{3}{2}\\Delta v_3 -\\frac{1}{6}\\Delta a_1$ & $\\Delta v_4\\in[-3.8,8.6]$\\\\[2ex]\n $ v_5=6v_6=\\frac{3}{4}\\Delta v_3 -\\frac{1}{8}\\Delta a_1 $ \n & $v_5=6 v_6 \\in[-1.9,4.3]$\n \\\\\n \\hline \n \\end{tabular}\n\\end{center} \n\\end{table}\n \nTo proceed, we need to perform the following conversion to pass from SMEFT to HEFT and viceversa: \n \\begin{align}\n& |\\partial H|^2 + \n\\frac{1}{2} B(|H|^2)(\\partial (|H|^2))^2 \\longleftrightarrow \\nonumber \\\\\n&\\frac{v^2}{4} \\mathcal{F}(h_{\\rm HEFT}) \\, {\\rm Tr}\\{ \\partial_\\mu U^\\dagger \\partial^\\mu U\\} \n+\n\\frac{1}{2} (\\partial h_{\\rm HEFT})^2 \\,,\n\\end{align} \nThe change from SMEFT to HEFT is straightforward and always possible, with the canonical, nonlinear change of variables given in differential form as\n\\begin{equation} \\label{changeofvariable} \ndh_{\\rm HEFT}\\, =\\, \\sqrt{1+(v+h_{\\rm SMEFT})^2 B(h_{\\rm SMEFT})}\\,\\, dh_{\\rm SMEFT} \\,,\n\\end{equation}\nwhere the flare-function is provided by the relation \n\\begin{equation}\n\\mathcal{F}(h_{\\rm HEFT}) \\,=\\, \\left(1+h_{\\rm SMEFT}\/v\\right)^2\\, .\n\\end{equation} \nHowever, \nthe reverse conversion from HEFT to SMEFT, \n\\begin{equation}\nh_{\\rm HEFT} \\,=\\, \\mathcal{F}^{-1}\\left((1+h_{\\rm SMEFT}\/v)^2\\right) \\,,\n\\end{equation}\nruns into difficulty. \nThis is because of the need to reconstruct squared operators of the Higgs doublet field $H$ that is the basis of SMEFT, such as\n\\begin{eqnarray}\n|H|^2 &=& \\frac{(v+h_{\\rm SMEFT})^2}{2}\\, ,\n\\nonumber \\\\\n(\\partial|H|^2)^2 &=& (v+h_{\\rm SMEFT})^2 \\, (\\partial h_{\\rm SMEFT})^2 \\nonumber \\\\ &=&\\, 2 |H|^2 \\, \n(\\partial h_{\\rm SMEFT})^2 \\, .\n\\end{eqnarray} \nThe extra $|H|^2$ on the right hand side of the second equation \nends in a denominator\n\\begin{align}\n&\\mathcal{L}_{\\rm SMEFT} = \\underbrace{ |\\partial H|^2}_{= \\mathcal{L}_{\\rm SM}} \\quad +\\quad\n\\label{eq:HEFT2SMEFT}\n\\\\ \\nonumber \n&\\underbrace{ \\frac{1}{2} \\bigg[ \\frac{8|H|^2}{v^2}\\bigg( (\\mathcal{F}^{-1})'\\left(2| H |^2\/v^2\\right) \\bigg)^2 \\,\\,\\,-\\,\\,\\, 1\\bigg] \\, \\frac{(\\partial| H |^2)^2}{2| H |^2} }_{=\\Delta \\mathcal{L}_{\\rm BSM}}\\, .\n\\end{align}\nAs SMEFT is assumed to have the analytical power expansion\nin Eq.~(\\ref{SMEFTL}), such singularity precludes its existence and needs to be cancelled by the preceding bracket in the second line of Eq.~(\\ref{eq:HEFT2SMEFT}). \n\nThe result is the same as that obtained by geometric methods~\\cite{Cohen:2020xca}, there must be a double zero of $\\mathcal{F}$, a symmetric point with respect to the global $SU(2)\\times SU(2)$ group so that the SMEFT expansion can be performed. Furthermore,\nanalyticity requires that all its odd derivatives vanish at the symmetric point.\n\nThe particular case of the SM is given by $\\mathcal{F}=(1+h_{\\rm SMEFT}\/v)^2 $. As already pointed out,\nat higher orders in $h\/v$, the existence of SMEFT requires that the odd derivatives of $\\mathcal{F}$ at the symmetric point $h_\\ast$ vanish. \n\nThe correlations from Table~\\ref{tab:correlations} can then be obtained by matching the Taylor expansion of $\\mathcal{F}$ around such symmetric point \n$ h_{\\rm HEFT}= h_\\ast$ with the expansion around our physical vacuum $h_{\\rm HEFT}=0$.\nInstead of that matching, one can also obtain the correlations by eliminating the SMEFT Wilson coefficients order by order. For example, at $\\mathcal{O}(1\/\\Lambda^2)$ there is only one operator, $\\mathcal{O}_{H\\Box}$, in Eq.~(\\ref{operatorsSMEFT}), that controls all the HEFT coefficients of $\\mathcal{F}$: \n\\begin{align}\n&&\na_1 = 2a=2\\left(1 + v^2\\frac{c_{H\\Box}}{\\Lambda^2}\\right)\\,, \\quad \na_2 = b=1+{4v^2}\\frac{c_{H\\Box}}{\\Lambda^2} \\,, \n\\nonumber \\\\ \n&&\na_3 = \\frac{8v^2}{3}\\frac{c_{H\\Box}}{\\Lambda^2}\\,, \\qquad \na_4 = \\frac{2v^2}{3}\\frac{c_{H\\Box}}{\\Lambda^2}\\,,\\qquad \na_{n\\geq 5} = 0 \\, .\n\\end{align}\nThe potential $V(h_{\\rm HEFT})$ is in turn also affected by $\\mathcal{O}_H$,\n\\begin{eqnarray}\n&&\nv_3 =1 + \\frac{3v^2 c_{H\\Box}}{\\Lambda^2} +\\epsilon_{c_H}\n\\,, \\,\\,\\, \nv_4 = \\frac{1}{4} + \\frac{25v^2c_{H\\Box} }{6\\Lambda^2} +\\frac{3}{2}\\epsilon_{c_H}\n\\,,\n\\nonumber\\\\\n&& \nv_5 = \\frac{2v^2c_{H\\Box} }{\\Lambda^2} +\\frac{3}{4}\\epsilon_{c_H}\n\\,,\\,\\,\\, \nv_6 = \\frac{v^2c_{H\\Box} }{3\\Lambda^2} +\\frac{1}{8}\\epsilon_{c_H}\n\\,,\n\\nonumber\\\\\n&& \nv_{n\\geq 7} = 0 \\,, \\qquad\\qquad\\qquad \n\\end{eqnarray}\nwith $m_h^2=\\, -2\\mu^2 \\left(1+\\frac{2 c_{H\\Box}v^2}{\\Lambda^2}+ \\frac{3}{4}\\epsilon_{c_H}\\right)$, $2\\langle |H|^2\\rangle =v^2= -\\frac{\\mu^2}{\\lambda}\\left(1-\\frac{3}{4}\\epsilon_{c_H}\\right) $ and $\\epsilon_{c_H}= - \\frac{2 v^4 c_H}{m_h^2\\Lambda^2}= \\frac{\\mu^2 c_H}{\\lambda^2 \\Lambda^2}$. \n\nVarious authors, see {\\it e.g.}~\\cite{Brivio:2016fzo} have pointed out to differences between the SMEFT and HEFT formulations~\\cite{Dobado:2019fxe}.\nFor example, in SMEFT the Goldstone\n $\\omega_i$ and Higgs $h_{\\rm SMEFT}$ bosons are arranged in a left-$SU(2)$ doublet while in HEFT\n$h_{\\rm HEFT}$ is an $SU(2)\\times SU(2)$ singlet, independent of the Goldstone triplet $\\omega_i$. \nAlso, in SMEFT the Higgs field always appears in the combination $ (h_{\\rm SMEFT} + v)$ \nand thus, HEFT deploys more independent higher-dimension effective operators (in exchange, it is less model dependent).\nThis means that SMEFT is natural when $h_{\\rm SMEFT}$ is a fundamental field while HEFT is typical for composite models of the EWSBS (such as those with $h_{\\rm HEFT}$ as a Goldstone boson).\nAnd finally, the counting of SMEFT is based in a cutoff $\\Lambda$ expansion taking the canonical operator dimensions, $\\mathcal{O}(d)\/\\Lambda^{d-4}$ (independently of $N_{\\rm loops}$) whereas HEFT is a derivative expansion (independently of $N_{\\rm particles})$ like the older Electroweak Chiral Lagrangian, with $\\mathcal{F}(h)$ inserted in the derivative Goldstone term.\n\n\n\nAmong the two types of correlations that we have presented in tables~\\ref{tab:correlations} and~\\ref{tab:corV}, the first ones for the coefficients of $\\mathcal{F}$ are more interesting for large values of the energy $\\sqrt{s}\\gg m_h\\sim m_W \\sim m_Z$ whereas the second, that do not involve Goldstone bosons, are therefore of greater interest at low energies, when and additionally \nthe potential competes with the derivative operators on equal ground, as $\\sqrt{s} \\sim m_i$. \n\\\\\nNevertheless, a lot of this is cosmetic and can be reorganized by changing variables $h_{\\rm SMEFT}\\leftrightarrow h_{\\rm HEFT}$. What is key is the San Diego criterion~\\cite{Alonso:2015fsp,Alonso:2016oah}: \n$\\mathcal{F}(h_{\\rm HEFT})$ must have a point $h_\\ast$ symmetric under the global $SU(2)\\times SU(2)$ group and due to its existence and convergence in the $h$ field space, SMEFT is deployable if and only if (which is a statement about the HEFT Lagrangian)\n\\begin{itemize}\n\\item $\\exists h_\\ast \\in\\mathbb{R}$ where $\\mathcal{F} (h_\\ast)=0$, and \n\\item Because of the need for $\\mathcal{L}_{\\rm SMEFT}$ analyticity, $\\mathcal{F}$ is analytic between our vacuum $h=0$ and $h_\\ast$, particularly around $h_\\ast$. Moreover its odd derivatives vanish.\n\\end{itemize}\nWe have presented new relations that implement this criterion \nup to $\\mathcal{O}(1\/\\Lambda^2)$ and $\\mathcal{O}(1\/\\Lambda^4)$ in the $1\/\\Lambda$ counting;\nmore precision is unnecessary until (if) separations from the Standard Model are found. Then only, with the scale $\\Lambda$ at hand, out of separations of EFT coefficients from the SM, can we decide how relevant the corrections due to the higher orders are expected to be, and whether further work is warranted.\n\nIn conclusion, we have newly translated these conditions into correlations among HEFT coefficients \nwhose violation falsifies SMEFT. \nMoreover, since many extensions of the Standard Model incorporating supersymmetry, supergravity, or other possibilities, can be cast as a SMEFT, they can be likewise simultaneously falsified.\n\n\nFor the time being, \n{no separations from the SM have been found~\\cite{Eboli:2021unw} and} one can only infer direct experimental bounds on the first terms, $a_1$ and, perhaps, $a_2$, so we have to wait for data with a larger number of Higgs boson before assessing them. But when this will be done, the correlations will allow to falsify SMEFT in experiment\neven without new particles. We believe that this possibility improves the standing of SMEFT as a scientific theory.\n\n\n\\vspace{.2cm}\n\\begin{acknowledgments}\nSupported by spanish MICINN PID2019-108655GB-I00 grant, and Universidad Complutense de Madrid under research group 910309 and the IPARCOS institute; \nERC Starting Grant REINVENT-714788; UCM CT42\/18-CT43\/18;\nthe Fondazione Cariplo and Regione Lombardia, grant 2017-2070.\n\\end{acknowledgments}\n\\vspace*{-0.5cm}\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}}