Skip to main content
Log in

Supersymmetric Hyperbolic \(\sigma \)-Models and Bounds on Correlations in Two Dimensions

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

In this paper we study a family of nonlinear \(\sigma \)-models in which the target space is the super manifold \(\mathbb {H}^{2|2N}\). These models generalize Zirnbauer’s \(\mathbb {H}^{2|2}\) nonlinear \(\sigma \)-model (Zirnbauer in Commun Math Phys 141(3):503–522, 1991). The latter model has a number of special features which aid in its analysis: through a remarkable technique from symplectic geometry colloquial known as supersymmetric localization, the partition function of the \(\mathbb {H}^{2|2}\) model is equal to one independent of the coupling constants. Our main technical observation is to generalize this fact to \(\mathbb {H}^{2|2N}\) models as follows: the partition function is a multivariate polynomial of degree \(n=N-1\), increasing in each variable. As an application, these facts provide estimates on the Fourier and Laplace transforms of the ’t-field’ when we specialize to \(\mathbb {Z}^2\). We show that this field has fluctuations which are at least those of a massless free field. In addition we show that small fractional moments of \(e^{t_v-t_0}\) decay at least polynomially fast in the distance of v to 0.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Angel, O., Crawford, N., Kozma, G.: Localization for linearly edge reinforced random walks. Duke Math. J. 163(5), 889–921 (2014)

    Article  MathSciNet  Google Scholar 

  2. Bauerschmidt, R., Crawford, N., Helmuth, T., Swan, A.: Random spanning forests and hyperbolic symmetry. In preparation (2019)

  3. Bauerschmidt, R., Helmuth, T., Swan, A.: Dynkin isomorphism and Mermin–Wagner theorems for hyperbolic sigma models and recurrence of the two-dimensional vertex-reinforced jump process. Preprint. arXiv:1802.02077. To appear in Ann. Probab.

  4. Berezin, F.A., Kirillov, A.A., Niederle, J., Koteckỳ, R.: Introduction to Superanalysis. Mathematical Physics and Applied Mathematics, Springer, Berlin (1987)

    Book  Google Scholar 

  5. Coppersmith, D., Diaconis, P.: Random walks with reinforcement. Unpublished manuscript (1986)

  6. Caracciolo, S., Sokal, A.D., Sportiello, A.: Grassmann integral representation for spanning hyperforests. J. Phys. A 40(46), 13799–13835 (2007)

    Article  ADS  MathSciNet  Google Scholar 

  7. Duistermaat, J.J., Heckman, G.J.: On the variation in the cohomology of the symplectic form of the reduced phase space. Inventiones mathematicae 69(2), 259–268 (1982)

    Article  ADS  MathSciNet  Google Scholar 

  8. Disertori, M., Spencer, T.: Anderson localization for a supersymmetric sigma model. Commun. Math. Phys. 300(3), 659–671 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  9. Disertori, M., Spencer, T., Zirnbauer, M.R.: Quasi-diffusion in a 3D supersymmetric hyperbolic sigma model. Commun. Math. Phys. 300(2), 435–486 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  10. Davis, B., Volkov, S.E.: Continuous time vertex-reinforced jump processes. Report Eurandom. Technische Universiteit Eindhoven (2001)

  11. Grimmett, G.: The Random-Cluster Model, vol. 110, p. 06. Springer, Berlin (2002)

    Google Scholar 

  12. Kozma, G., Peled, R.: Power-law decay of weights and recurrence of the two-dimensional vrjp (2019)

  13. Merkl, F., Rolles, S.W.W.: Recurrence of edge-reinforced random walk on a two-dimensional graph. Ann. Probab. 37(5), 1679–1714 (2009)

    Article  MathSciNet  Google Scholar 

  14. McBryan, O.A., Spencer, T.: On the decay of correlations in \({\rm so}(n)\)-symmetric ferromagnets. Commun. Math. Phys. 53(3), 299–302 (1977)

    Article  ADS  MathSciNet  Google Scholar 

  15. Sabot, C.: Polynomial localization of the 2D-vertex reinforced jump rocess. arXiv e-prints, page arXiv:1907.07949 (2019)

  16. Sabot, C., Tarrès, P.: Edge-reinforced random walk, vertex-reinforced jump process and the supersymmetric hyperbolic sigma model. J. Eur. Math. Soc. 17(9), 2353–2378 (2015)

    Article  MathSciNet  Google Scholar 

  17. Sabot, C., Tarrès, P., Zeng, X.: The vertex reinforced jump process and a random Schrödinger operator on finite graphs. Ann. Probab. 45(6A), 3967–3986 (2017)

    Article  MathSciNet  Google Scholar 

  18. Spencer, T., Zirnbauer, M.R.: Spontaneous symmetry breaking of a hyperbolic sigma model in three dimensions. Commun. Math. Phys. 252, 167–187 (2004)

    Article  ADS  MathSciNet  Google Scholar 

  19. Sabot, C., Zeng, X.: A random Schrödinger operator associated with the vertex reinforced jump process on infinite graphs. J. Am. Math. Soc. 32(2), 311–349 (2019)

    Article  Google Scholar 

  20. Tong, D.: Lecture notes on statistical field theory. http://www.damtp.cam.ac.uk/user/tong/sft.html. Accessed 01 Sept 2019

  21. Zirnbauer, M.R.: Fourier analysis on a hyperbolic supermanifold with constant curvature. Commun. Math. Phys. 141(3), 503–522 (1991)

    Article  ADS  MathSciNet  Google Scholar 

Download references

Acknowledgements

I thank Roland Bauerschmidt for explaining to me the results in [3], which motivated the present work. I also thank Bauerschmidt, Tyler Helmuth and Andrew Swan for numerous discussions on related topics throughout the preparation of this manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nicholas Crawford.

Additional information

Communicated by Alessandro Giuliani.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supported by Israel Science Foundation Grant Number 1692/17.

SUSY for \(a\in \pmb {{\mathbb {N}}}+1/2\)

SUSY for \(a\in \pmb {{\mathbb {N}}}+1/2\)

In this appendix we sketch the development of the \(\mathbb {H}^{2|2N}\) nonlinear \(\sigma \)-model from which the measure \(d\mu _{{\mathcal {G}}, a, J, \varepsilon } (t)\), defined at (1.4) is derived. Let us first introduce the \(\mathbb {H}^{2|2N}\) \(\sigma \)-models. We consider the general case where the model is defined over a finite graph \({\mathcal {G}}=(V, E)\).

For each \(j \in V\) we introduce a supervector \(u_j \in \mathbb {R}^{3|2N}\),

$$\begin{aligned} u_j = (z_j, {x}_j, y_j, {\xi }_j, {\eta }_j). \end{aligned}$$
(A.1)

The variables \({\xi }_j, \eta _j\,\) are N-tuples of odd generators for a Grassmann algebra over j. Subscript indices denote locations in \({\mathcal {G}}\) whereas supercript indices will denote internal components of the Grassmann vectors: \(\xi ^{(\ell )}_i\) is the \(\ell \)’th component of \(\xi _i\). We then define a Minkowski signature bilinear form on \(\mathbb {R}^{3|2N}\) by

$$\begin{aligned} (u,u') = - z z' +x x'+y y' + \xi \cdot \eta ' - \eta \cdot \xi ' \end{aligned}$$
(A.2)

where

$$\begin{aligned} \xi \cdot \eta '= \sum _{\ell =1}^N \xi ^{\ell } {\eta '}^{\ell }. \end{aligned}$$
(A.3)

To define the \(\mathbb {H}^{2|2N}\) \(\sigma \)-model, the \(u_j\)’s are constrained to satisfy the quadratic equation

$$\begin{aligned} (u_j\, , u_j) = - 1. \end{aligned}$$
(A.4)

Note that the solutions to this equation form a two-sheeted hyperboloid in \(\mathbb {R}^{3|2N}\). If, for each j, we restrict attention to spins lying in the sheet with positive square root, we arrive at \(\mathbb {H}^{2|2N}\). This is a super manifold. Geometrically, it is an infinitesimal extension by Grassmann variables of the \(d=2\) hyperbolic plane, parametrized by 2 real variables \(x_j, y_j\) and 2N Grassmman variables \(\xi _j, \eta _j\).

On the product space \((\mathbb {H}^{2|2N})^{V}\) we introduce a ‘measure’ (more accurately, a Berezin superintegration form)

$$\begin{aligned} D\mu _V = \prod _{j\in V} \text {d}x \text {d}y \prod _{j\in V, \ell } \partial _{\xi ^{\ell }_j} \partial _{\eta ^{\ell }_j} \circ (z_j)^{-1/2} \;. \end{aligned}$$
(A.5)

Note that \(z_j^2=1+x_j^2+y_j^2 + 2\xi _j\cdot \eta _j\) - it is an element of the Grassmann algebra, not a real variable. The Gibbs state is then proportional to the Grassmann integration form \(D\mu _\Lambda \,\text {e}^{-A_{J, \varepsilon }}\) with

$$\begin{aligned} A_{J,\varepsilon } = \frac{1}{2}\mathop \sum \limits _{(jj')\in E} J_{jj'}\,(u_j-u_j'\,,u_j-u_j') + \mathop \sum \limits _{i \in V} \varepsilon _i(z_i-1) \end{aligned}$$
(A.6)

The coupling constants \(J_{jj'} >0\) if \(jj'\) are nearest neighbors in \({\mathcal {G}}\) and \(J_{jj'} = 0\) otherwise.

1.1 Horospherical Coordinates and \(d\mu _{{\mathcal {G}}, a, J, \varepsilon } (t)\)

To connect the \(\mathbb {H}^{2|2N}\) model to \(d\mu _{{\mathcal {G}}, a, J, \varepsilon } (t)\) with \(a=N-1/2\) we need to introduce a change of variables called horospherical (or Iwasawa) coordinates. The next coordinates are \((t, s, {\bar{\theta }}, \theta )\) defined by

$$\begin{aligned} x = \sinh t - \text {e}^t \left( {\textstyle {\frac{1}{2}}} {{s}}^2 + \sum _{i=1}^N{{\bar{\theta }}}^{\ell }\theta ^{\ell } \right) \ , \quad y = \text {e}^t s\ ,\quad \xi ^{\ell } = \text {e}^t {{\bar{\theta }}}^{\ell }\;, \quad \eta ^{\ell } = \text {e} ^t \theta ^{\ell }\;, \end{aligned}$$
(A.7)

where \(t\in {\mathbb {R}}\) and \(s\in {\mathbb {R}}\) range over the real numbers. In the Poincaré disc and given a point p, t represents the signed distance of the horocycle tangent to 1 from 0 and containing p whereas s is represents the (normalized) location of p on the horocycle). In these coordinates, the expression for the action becomes

$$\begin{aligned} A_{J,\varepsilon } = \mathop \sum \limits _{(ij)} J_{ij} (S_{ij}-1) + \mathop \sum \nolimits _{k\in \Lambda } \varepsilon _k (z_k - 1) \;, \end{aligned}$$
(A.8)

where \((jj')\) are NN pairs and

$$\begin{aligned} S_{jj'}&=\,B_{jj'} + ({{\bar{\theta }}}_j - {{\bar{\theta }}}_{j'})\cdot (\theta _j - \theta _{j'})\, \text {e}^{t_j + t_{j'}}\;, \end{aligned}$$
(A.9)
$$\begin{aligned} B_{jj'}&= \cosh (t_j-t_{j'}) + {\textstyle {\frac{1}{2}}} (s_j - s_{j'})^2 \, \text {e}^{t_j + t_{j'}}\;, \end{aligned}$$
(A.10)
$$\begin{aligned} z_{j}&=\, \cosh t_j +\left( {\textstyle {\frac{1}{2}}}s_j^2+ {{\bar{\theta }}}_j \cdot \theta _j\right) \text {e}^{t_j} \;. \end{aligned}$$
(A.11)

By applying Berezin’s transformation formula [4] for changing variables in a (super-)integral, one finds that

$$\begin{aligned} D\mu _{V} = \prod \nolimits _{j\in V} \text {e}^{-t_j[2N-1]} \text {d}t_j \text {d} s_j \, \prod _{\ell } \partial _{{\bar{\theta }}^{\ell }_j} \partial _{\theta ^{\ell }_j} \; \end{aligned}$$
(A.12)

For any function f of the field variables \(\{ t_j\,, s_j\,, {{\bar{\theta }}}_j\,,\theta _j\}_{j\in \Lambda }\) we now define its expectation as

$$\begin{aligned} \left\langle f \right\rangle _{{\mathcal {G}}, J,\varepsilon } = \frac{ \int D\mu _V \, \text {e}^{-A_{J,\varepsilon }} f}{\int D\mu _V \, \text {e}^{-A_{J,\varepsilon }} } \;, \end{aligned}$$
(A.13)

whenever the numerator integral exists.

There is no notational conflict with earlier definitions due to the following. Observe the beautiful feature that all coordinates, besides the t coordinate, are Gaussian in this representation. Thus if f depends on the \(t_j\)’s alone, we may easily integrate the other variables out. We find (\(a=N-1/2\))

$$\begin{aligned}&\int D\mu _V \, \text {e}^{-A_{J,\varepsilon }}=Z_{{\mathcal {G}}, a, J, \varepsilon }&\left\langle f \right\rangle _{{\mathcal {G}}, J,\varepsilon }= \frac{\int f d\mu _{V }^{J, \varepsilon } (t) }{Z_{{\mathcal {G}}, a, J, \varepsilon }}=\left\langle f \right\rangle _{{\mathcal {G}}, a, J,\varepsilon }. \end{aligned}$$

1.2 SUSY Localization

Section 2 highlighted the importance of controlling partition function ratios

$$\begin{aligned} R_{J', J}:=\frac{Z_{{\mathcal {G}}, a, J', \varepsilon }}{Z_{{\mathcal {G}}, a, J, \varepsilon }} \end{aligned}$$

for two choices of coupling constants \(J, J'\). For the \(\sigma \)-models taking values in \(\mathbb {H}^{2|2}\), this ratio is 1 by SUSY localization [9]. On \(\mathbb {H}^{2|2N}\) this is not true, but the localization argument is still extremely useful.

We now sketch the localization computation from [9] as it applies to the \(\mathbb {H}^{2|2N}\) models, the reader may consult that paper for the missing details. The computation is most easily explained by first considering the special case that \({\mathcal {G}}\) is just a single vertex. Let H be the quadratic polynomial

$$\begin{aligned} H = x^2 +y^2 + 2 \xi \cdot \eta . \end{aligned}$$

Let us isolate one pair of Grassmann components \((\xi ^1, \eta ^1)\). With respect to this pair, let q be the distinguished first-order differential operator defined by

$$\begin{aligned} q = x \partial _{\eta ^{(1)} } - y \partial _{\xi ^{(1)} } + \xi ^{1} \partial _{x} + \eta ^{1} \partial _{y} \;. \end{aligned}$$
(A.14)

Note that q annihilates H,

In this notation, the a priori superintegration form is

$$\begin{aligned} D\mu = \text {d}{ x} \text {d} y \, \prod _{\ell } \partial _{\xi ^\ell } \partial _{\eta ^\ell } \circ (1 + H)^{-1/2} \;. \end{aligned}$$

Lemma A.1

The Berezin superintegration form \(D\mu \) is q-invariant, i.e.,

$$\begin{aligned} \int D\mu \; q \cdot f = 0 \end{aligned}$$

for any compactly supported smooth superfunction f.

Corollary A.2

Suppose \(q\cdot f=0\). Then for any \(\tau >0\),

$$\begin{aligned} \int D\mu \; f= \int D\mu \; e^{-\tau H} f. \end{aligned}$$

Every superfunction f can be expanded over the Grassmann variables \((\xi ^{\ell }, \eta ^{\ell })_{\ell \ge 2}\) as

$$\begin{aligned} f=\sum _{I,J\subseteq \{2, \dotsc , N\}} f_{I,J}(x,y, \xi ^{1}, \eta ^{1}) \xi ^{I}\eta ^{J} \end{aligned}$$

where \(\xi ^I=\prod _{\ell \in I} \xi ^\ell \) and simialrly for \(\eta \). Let \(f_0\) be the superfunction obtained by setting \(x = y=\xi ^{1} = \eta ^{1}=0\),

$$\begin{aligned} f_0=\sum _{I,J\subseteq \{2, \dotsc , N\}} f_{I,J}(o) \xi ^{I}\eta ^{J}, \end{aligned}$$

so that the coefficient functions are evaluated as \( f_{I,J}(o)\). Thus superfunction is a constant coefficient polynomial in \((\xi ^{\ell }, \eta ^{\ell })_{\ell \ge 2}\). Let \(n=N-1\) and let \({\bar{\psi }}, {\psi }\) denote the n-tuples defined by the last n components of \(\xi , \eta \), so \({\bar{\psi }}^{\ell } =\xi ^{\ell +1}, \psi ^{\ell }=\eta ^{\ell +1}\) for \(\ell \in \{1, \dotsc , n\}\) and let \(\sigma =(1 + 2{\bar{\psi }}\cdot {\psi })^{1/2}\). The variables \(\sigma , {\bar{\psi }}, \psi \) live in a degenerate hyperboloid \(\mathbb {H}^{0|2n}\subset {\mathbb {R}}^{1|2n}\).

$$\begin{aligned} D\mu _0 = \prod _{\ell } \partial _{{\bar{\psi }}^{\ell }} \partial _{{\psi }^{\ell }} \circ \sigma ^{-1} \;. \end{aligned}$$

Lemma A.3

(Localization from \(\mathbb {H}^{2|2N}\) to \(\mathbb {H}^{0|2(N-1)}\)) Let f be a smooth superfunction which satisfies the invariance condition \(q f = 0\) and decreases sufficiently fast at infinity in order for the integral \(\int D\mu \, f\) to exist. Then

$$\begin{aligned} \int D\mu \, f = \int D\mu _0 f_0. \end{aligned}$$

We now generalize this last lemma to \({\mathbb {H}^{2|2N}}^V\) for general finite graphs. Let \(\text {n}_j=(\sigma _j, {\bar{\psi }}_{j}, \psi _{ j'})\) and set

$$\begin{aligned} S_{J,\varepsilon } = \frac{1}{2}\mathop \sum \limits _{(jj')\in E} J_{jj'}\,(\text {n}_j-\text {n}_{j'}\,,\text {n}_j-\text {n}_{j'}) + \mathop \sum \limits _{i \in \Lambda } \varepsilon _i(\sigma _i-1) \end{aligned}$$
(A.15)

where the bilinear form is

$$\begin{aligned} (\text {n}, \text {n}')=-\sigma \sigma ' + {\bar{\psi }}\cdot \psi '+ {\bar{\psi }}'\cdot \psi . \end{aligned}$$

Let \(\left\langle \cdot \right\rangle ^f_{{\mathcal {G}}, n, J, \epsilon }\) denote the corresponding Gibbs state, in particular

$$\begin{aligned} Z^f_{{\mathcal {G}}, n, J, \varepsilon }:=\int \prod _{j\in V} D\mu _{0}({\bar{\psi }}_j, \psi _j) e^{-S_{{\mathcal {G}}, n, J, \epsilon }} \end{aligned}$$

Let \(q_{V}=\sum _{j \in V} q_j\)

Lemma A.4

Let \(N\in {\mathbb {N}}\), \(a=N-1/2\), \(n=N-1\). Then the \(\mathbb {H}^{2|2N}\) \(\sigma \)-model is equivalent to a purely fermionic \(\mathbb {H}^{0|2n}\) \(\sigma \)-model in the sense that for any function \(F(\underline{z}, (\underline{\xi }^\ell )_{\ell =2}^N,(\underline{\eta }^\ell )_{\ell =2}^N)\) which decays sufficiently fast and such that \(q_{V} \cdot F=0\)

$$\begin{aligned} \left\langle F(\underline{z}, (\underline{\xi }^\ell )_{\ell =2}^N,(\underline{\eta }^\ell )_{\ell =2}^N) \right\rangle _{{\mathcal {G}}, N, J, \epsilon }= \left\langle F (\sigma , {\bar{\psi }}, \psi ) \right\rangle ^f_{{\mathcal {G}}, n, J, \epsilon }. \end{aligned}$$
(A.16)

In particular

$$\begin{aligned} Z_{{\mathcal {G}}, a, J, \epsilon }=Z^f_{{\mathcal {G}}, n, J, \varepsilon }\end{aligned}$$

1.3 Residual SUSY After Localization; Connection with [6]

The action \(S_{J, \varepsilon }\) can be interpreted as a nonlinear \(\sigma \)-model with respect to the target space \({\mathbb {R}}^{1|2n}\), with even coordinate \(\sigma \) and odd generators \(({\bar{\psi }}^{\ell }, \psi ^{\ell })_{\ell =1}^n\). There are two natural quadratic forms we can put on this space:

$$\begin{aligned} -\sigma ^2 +2{\bar{\psi }}\cdot \psi&\text { Lorentizian}, \end{aligned}$$
(A.17)
$$\begin{aligned} \sigma ^2+ 2{\bar{\psi }}\cdot \psi&\text { Euclidean}. \end{aligned}$$
(A.18)

Constraining spins to be one when evaluated by the quadratic form Appendix A.17 gives a nonlinear \(\sigma \)-model with target space one sheet of the degenerate hyperboloid \(\sigma ^2- 2{\bar{\psi }}\cdot \psi =1\) whereas if we use the quadratic form Appendix A.18, the spins are interpreted as taking values in the ’upper hemisphere’ of the degenerate sphere \(\sigma ^2+2{\bar{\psi }}\cdot \psi =1\). The latter situation was discussed in [6]. However, these two superspaces are the same via a change of fermionic coordinates. As such, the discussion of Sect. 7 in [6] provides useful insight here. Let us adapt and generalize that discussion for the sake of completeness.

We begin by introducing, at each vertex \(i \in V\), a superfield \(v_i := (\sigma _i,\psi _i,{\bar{\psi }}_i)\) consisting of a single bosonic variable \(\sigma _i\) and 2n Grassmann variables \((\psi ^{\ell }_i, {\bar{\psi }}^{\ell }_i)_{\ell =1}^{n}\). We equip \({\mathbb {R}}^{1|2n}\) with the Lorentzian scalar product

$$\begin{aligned} ( v_i, v_j) \; := \; - \sigma _i \sigma _j +({\bar{\psi }}_i \cdot \psi _j - \psi _i \cdot {\bar{\psi }}_j) \;, \end{aligned}$$
(A.19)

There are two types of symmetries which preserve this bilinear form. The first type is a symplectic linear transformation mapping the Grassmann variables into themselves and fixing the bosonic component. That is, if M is an invertible 2n-by-2n matrix preserving the blilinear form induced by

$$\begin{aligned} J:=\left( \begin{array}{cc}0 &{} I_n \\ -I_n &{} 0\end{array}\right) \end{aligned}$$

and if \(u_i =(\sigma _i, M\cdot [\psi _i, {\bar{\psi }}_i])\), then \((u_i, u_j)=(v_i, v_j)\).

The second type of transformation is supersymmetric, mixing \(\sigma \) with the \(\psi , {{\bar{\psi }}}\)’s. There are, in the general case, n noncommuting SUSY transformations, parametrized by fermionic (Grassmann-odd) global parameters \((\epsilon ^{\ell },{\bar{\epsilon }}^{\ell })_{\ell =1}^n\):

$$\begin{aligned} \delta \sigma _i= & {} ({\bar{\epsilon }}^{\ell } \psi _i^{\ell } + {\bar{\psi }}_i^{\ell } \epsilon ^{\ell }) \end{aligned}$$
(A.20)
$$\begin{aligned} \delta \psi _i^{\ell }= & {} \epsilon ^{\ell } \,\sigma _i \end{aligned}$$
(A.21)
$$\begin{aligned} \delta {\bar{\psi }}_i^{\ell }= & {} {\bar{\epsilon }}^{\ell } \,\sigma _i \end{aligned}$$
(A.22)

To check that these transformations leave eq. (A.19) invariant, we compute

$$\begin{aligned} \delta ( v_i\cdot v_j )= & {} (\delta \sigma _i) \sigma _j + \sigma _i (\delta \sigma _j) + \big [ (\delta {\bar{\psi }}_i)\cdot \psi _j + {\bar{\psi }}_i\cdot (\delta \psi _j) -(\delta \psi _i) \cdot {\bar{\psi }}_j - \psi _i\cdot (\delta {\bar{\psi }}_j) \big ]\nonumber \\ \end{aligned}$$
(A.23)
$$\begin{aligned}= & {} -({\bar{\epsilon }}^{\ell } \psi _i + {\bar{\psi }}_i \epsilon ^{\ell }) \sigma _j - ({\bar{\epsilon }}^{\ell } \psi _j + {\bar{\psi }}_j \epsilon ^{\ell }) \sigma _i \nonumber \\&+\, \big [ {\bar{\epsilon }}^{\ell } \psi ^{\ell }_j \sigma _i + {\bar{\psi }}^{\ell }_i \epsilon ^{\ell } \sigma _j -\epsilon ^{\ell } {\bar{\psi }}^{\ell }_j \sigma _i - \psi ^{\ell }_i {\bar{\epsilon }}^{\ell } \sigma _j \big ] \end{aligned}$$
(A.24)
$$\begin{aligned}= & {} 0 \;. \end{aligned}$$
(A.25)

Now let us consider a \(\sigma \)-model in which the superfields \(v_i\) are constrained to lie on the upper sheet of the hyperboloid \({\mathbb {R}}^{1|2n}\), \( \sigma _i^2 -2 {\bar{\psi }}_i \cdot \psi _i \;=\; 1 \) This constraint is solved by writing

$$\begin{aligned} \sigma _i \;=\; \pm (1 + 2 {\bar{\psi }}_i\cdot \psi _i)^{1/2}, \end{aligned}$$

so that \(\sigma _i\) is an even invertible element of the Grassmann algebra. We take only the \(+\) sign in (A.3) and denote the corresponding unit vector by \(\text {n}_i\).

The \({\mathfrak {sp}}(2n)\) transformations continue to act as above while the SUSY transformations act via

$$\begin{aligned} \delta \psi ^{\ell }_i= & {} \epsilon ^{\ell } \sigma _i \\ \delta {\bar{\psi }}^{\ell }_i= & {} {{\bar{\epsilon }}}^{\ell } \sigma _i \end{aligned}$$

These transformations leave invariant the scalar product \( \text {n}_i\cdot \text {n}_j\). and the corresponding generators \(Q^{\ell }_\pm \) are defined as

$$\begin{aligned} Q^{\ell }_+= & {} \sum _{i\in V} \sigma _i \partial ^{\ell }_i \\ Q^{\ell }_-= & {} \sum _{i\in V} \sigma _i {\bar{\partial }}_i \end{aligned}$$

where \(\partial ^{\ell }_i = \partial _{\psi _i^{\ell }}\) and \({\bar{\partial }}^{\ell }_i = \partial _{{\bar{\psi }}_i^{\ell }}\)

$$\begin{aligned}&Q^{\ell }_{+}{ {\psi }}_i^\ell =Q^{\ell }_{-}{ {{\bar{\psi }}}}_i^\ell = \sigma _i,&Q^{\ell }_{\pm }[ \text {n}_i\cdot \text {n}_j]=0. \end{aligned}$$
(A.26)

From this identity, \(Q^{\ell }_{\pm } S_{J, \varepsilon }=Q^{\ell }_{\pm } e^{-S_{J, \varepsilon }}=0\). Note also that a priori integration form \({\mathcal {D}}_0({\bar{\psi }}, \psi )\) is invariant with respect to the \(Q^{\ell }_{\pm }\)’s:

$$\begin{aligned} \int {\mathcal {D}}_0({\bar{\psi }}, \psi ) Q^{\ell }_{\pm } F(\psi , {\bar{\psi }})=0. \end{aligned}$$

From these facts, (4.6) follows immediately.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Crawford, N. Supersymmetric Hyperbolic \(\sigma \)-Models and Bounds on Correlations in Two Dimensions. J Stat Phys 184, 32 (2021). https://doi.org/10.1007/s10955-021-02817-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10955-021-02817-y

Keywords

Navigation