This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Replica theory for disorder-free spin-lattice glass transition on a tree -like simplex network

Kota Mitsumoto Institute of Industrial Science, University of Tokyo, 4-6-1 Komaba, Meguro-ku, Tokyo 153-8505, Japan    Hajime Yoshino Cybermedia Center, Osaka University, Toyonaka, Osaka 560-0043, Japan Graduate School of Science, Osaka University, Toyonaka, Osaka 560-0043, Japan
Abstract

A class of pyrochlore oxides, A2A_{2}Mo2O7 (A=A= Ho, Y, Dy, Tb) with magnetic ions on corner-sharing tetrahedra is known to exhibit spin-glass transitions without appreciable amount of quenched disorder. Recently a disorder-free theoretical model for such a system has been proposed which takes into account not only spins but also lattice distortions as dynamical variables [K. Mitsumoto, C. Hotta and H. Yoshino, Phys. Rev. Lett. 124, 087201 (2020)]. In the present paper we develop and analyze an exactly solvable disorder-free mean-field model which is a higher-dimensional counterpart of the model. We find the system exhibit complex free-energy landscape accompanying replica symmetry breaking through the spin-lattice coupling.

I Introduction

Glass formation is a generic phenomenon that occurs in various systems, including structural glasses Berthier and Biroli (2011); Tarjus (2011); Parisi et al. (2020), spin glasses Mézard et al. (1987); Mydosh (1993); Kawamura and Taniguchi (2015), orbital glasses Fichtl et al. (2005); Thygesen et al. (2017); Mitsumoto et al. (2020), and charge glasses Kagawa et al. (2013). While the possibility of genuine thermodynamic glass transition is highly debated in the case of structural glassesBerthier and Biroli (2011), thermodynamic spin glass transitions in magnetic systems with strong quenched disorder is well established experimentally Mydosh (1993). The thermodynamic spin-glass transition is most unambiguously demonstrated by the measurements of the negatively diverging non-linear susceptibility with critical scaling at the spin-glass transition temperature Chikazawa et al. (1980); Taniguchi et al. (1985); Levy and Ogielski (1986). On the theoretical side, Edwards and Anderson proposed that the origin of the spin glass with quenched disorder is randomness and frustration Edwards and Anderson (1975), which is now widely accepted through analysis of the Edwards-Anderson (EA) model by mean-field theories exact in the large dimensional limit Sherrington and Kirkpatrick (1975); de Almeida and Thouless (1978); Parisi (1979) and by numerical simulations at 3 dimensions on both Ising Ogielski (1985); Bhatt and Young (1988); Kawashima and Young (1996) and Heisenberg Hukushima and Kawamura (2000); Lee and Young (2007); Ogawa et al. (2020) spins. Remarkably thermodynamic spin-glass transitions have been established experimentally also in a class of systems without quenched disorder, pyrochlore antiferromagnets Greedan et al. (1986); Gaulin et al. (1992); Dunsiger et al. (1996); Gingras et al. (1997); Gardner et al. (1999); Hanasaki et al. (2007); Gardner et al. (2010); Thygesen et al. (2017); Mitsumoto et al. (2020) through experiments including, in particular, the measurements of the negatively diverging non-linear susceptibility Gingras et al. (1997). On the theoretical side, the explanation for the sharp spin-glass transition without quenched disorder remained a challenging open problem.

A common feature in the glassy systems mentioned above is frustration: strong many-body interactions competing with each other. The resultant glassiness is often viewed in a phenomenological complex free-energy landscape picture: free-energy landscape with multiple nearly degenerate minima separated by high free-energy barriers. One important aspect of such a free-energy landscape is the near degeneracy of the energy minima or flatness. Frustration can realize such flat energy landscapes. Indeed, as a primary approximation, the pyrochlore antiferromagnets can be modeled by the classical Heisenberg spin model with the nearest-neighbor antiferromagnetic interaction on the pyrochlore lattice which is a three-dimensional corner-sharing network of tetrahedra (Fig. 1 (a)). Because of the extremely strong geometrical frustration, the low energy landscape of the model exhibits flatness and the system remains in a classical spin liquid state down to T=0T=0 Reimers et al. (1991); Moessner and Chalker (1998); Gardner et al. (2010). However, apparently, something is missing in this simplest model as it fails to capture the spin-glass transition found experimentally. In short, it does not realize high free-energy barriers indispensable for glassiness. In the case of spin glasses with strong quenched disorder, the energy landscape is distorted randomly by the quenched disorder leading to high free-energy barriers. Then a simple solution is to add some quenched disorder onto the antiferromagnetic model Saunders and Chalker (2007); Shinaoka et al. (2011) but such a disorder has never been observed in the pyrochlore oxides. How a lugged free-energy landscape with high barriers can be created by distorting a flat energy landscape without the help of such quenched disorder is a non-trivial problem.

Recently, a microscopic theoretical model Mitsumoto et al. (2020) aimed to capture the emergence of disorder-free spin-glass transition on the pyrochlore oxides was proposed. It was argued that the model realizes a complex free-energy landscape by coupling two distinct dynamical variables which are strongly frustrated among themselves. One is the set of classical Heisenberg spins on the vertices (molybdenum ions) which are interacting anti-ferromagnetically with the nearest neighbors. As just mentioned above this anti-ferromagnetic spin system on the pyrochlore lattice is highly frustrated giving rise to the flat energy landscape. The other degree of freedom is the set of spatial displacements of the vertices, the molybdenum ions. Following the suggestion by a recent experiment Thygesen et al. (2017), they are assumed to be displaced either towards (inin) or away from (outout) the center of the tetrahedron following two-inin-two-outout ice rule Pauling (1935) at each tetrahedron. This is equivalent to the so-called ‘spin ice’ which is another highly frustrated system on the pyrochlore lattice with a flat energy landscape (only low enough barriers) that allows liquid state down to T=0T=0 Gardner et al. (2010); Bramwell and Harris (2020); Udagawa and Jaubert (2021). In the model Mitsumoto et al. (2020) these two distinct degrees of freedom are coupled through the Goodenough-Kanamori mechanism Goodenough (1955); Kanamori (1959); Solovyev (2003); Smerald and Jackeli (2019); Mitsumoto et al. (2020). Since the two degrees of freedom are totally unrelated to each other in the absence of the coupling, the coupling can distort their flat energy landscapes. Extensive Monte Carlo simulation Mitsumoto et al. (2020) in the presence of the coupling suggested a glass transition where both the spins and lattice displacements become cooperatively frozen simultaneously. This was demonstrated by two key observations. One is the critical slowing down of the two degrees of freedom approaching a common freezing temperature TcT_{\rm c}. The other is an observation that the nonlinear magnetic and dielectric susceptibilities, which are associated with the freezing of the spins and lattice displacements respectively, grow negatively large in the vicinity of TcT_{\rm c}.

Refer to caption
Figure 1: (a) Pyrochlore lattice and (b) kagomé lattice. (c) (k1)(k-1)-dimensional simplices in the cases of k=3,4k=3,4 and 5. (d, e) The corner-sharing (k1k-1)-simplices network in the cases of (d) k=3k=3 and c=2c=2 and (e) k=4k=4 and c=3c=3.
Refer to caption
Figure 2: Schematic picture of the free-energy landscape of the mean-field theory. (a, b) In the absence of the spin-lattice coupling (δ=0)(\delta=0), the (a) spin and (b) lattice degrees of freedom independently exhibit low temperature phases where global rotational symmetries are broken. They are characterized by marginally stable replica symmetric solutions which imply flat landscapes of their own. (c) Switching on the spin-lattice coupling (δ>0)(\delta>0), at low enough temperatures, the system exhibits a glass phase where both the spin and lattice degrees of freedom are frozen. The replica symmetry is broken there implying a complex free-energy landscape.

To obtain further insights into the problem, we develop an exactly solvable, disorder-free mean-field model which can be regarded as a variant of the model Mitsumoto et al. (2020) in a high-dimensional limit dd\rightarrow\infty in the present paper. We analyze the model using the replica method. Recently the replica approach, which was originally developed for systems with quenched disorder Edwards and Anderson (1975); Sherrington and Kirkpatrick (1975); Parisi (1979), was applied successfully to describe glassy systems without quenched disorder Charbonneau et al. (2014); Yoshino (2018). In the nutshell, one considers the effect of infinitesimal random pinning field which is switched off after taking the thermodynamic limit NN\rightarrow\infty Parisi and Virasoro (1989); Monasson (1995); yos . The replica approach established existence of genuine thermodynamic glass transitions in high-dimensional limits in the absence of quenched disorder: structural glasses Charbonneau et al. (2014) and disorder-free pp-body interacting spin systems Yoshino (2018). The mean-field approach is useful to elucidate relevant order parameters, which are overlaps of dynamical degrees of freedom belonging to different replicas in the case of glasses, and to extract mean-field phase diagrams, which serves as useful guidelines. It is particularly helpful in understanding the formation of glasses Mézard and Parisi (1999); Parisi et al. (2020) as it derives microscopically the complex free-energy landscape, which has been anticipated phenomenologically, based on the notion of replica symmetry breaking (RSB) de Almeida and Thouless (1978); Parisi (1979); Mézard et al. (1987). On the other hand, one should always keep in mind that the mean-field approaches are exact only at high enough dimensions. In general phase transitions found at high dimensions disappear at low enough dimensions because of thermal fluctuations. More seriously some of the predictions made by mean-field theory can be problematic such that they only hold strictly in high dimensional limits Fisher (1991).

We define our mean-field model for the spin-lattice coupled systems on a tree-like simplex lattice Husimi (1950); ESSAM and FISHER (1970); Rieger and Kirkpatrick (1992); Chandra and Doucot (1994); Udagawa and Moessner (2019); Cugliandolo et al. (2020) which is a higher-dimensional extension of the pyrochlore lattice and kagomé lattice (Figs. 1 (a) and (b)), and solve it exactly by the replica method in a dense connectivity limit corresponding to the large dimensional limit. Here, a simplex means a geometrical unit in which all vertices are fully connected, and a simplex with kk vertices is called a (k1)(k-1)-dimensional simplex. For example, triangle, tetrahedron and 55-cell are (k1)(k-1)-dimensional simplex with k=3,4k=3,4 and 55, respectively, as shown in Fig. 1 (c). For simplicity, both the spins and lattice displacements are modeled by continuous variables subjected to spherical constraints. We derive exact free-energy functional in terms of glass order parameters for the spins and lattice displacements following the approach of Yoshino (2018). Then we analyze the problem within the replica symmetric (RS) ansatz and examine the stability of the solution to detect instability toward RSB.

In Fig. 2, we show the schematic free-energy landscape predicted by the theory. In the absence of the spin-lattice coupling by the Goodenough-Kanamori mechanism (δ=0)(\delta=0) (Figs. 2 (a) and (b)), the spin and lattice degrees of freedom independently exhibit low temperature phases where their global rotational symmetries are broken. The low temperature phases are characterized by marginally stable replica symmetric solutions which imply flat landscapes of their own. The independent flat landscapes may be related to the those of the corresponding 3 dimensional systems: the antiferromagnet model Reimers et al. (1991); Moessner and Chalker (1998); Gardner et al. (2010) and the spin-ice system Gardner et al. (2010); Bramwell and Harris (2020); Udagawa and Jaubert (2021) on the pyrochlore lattice. However, we believe that the breaking of the global rotational symmetry is an artifact in the high dimensional limit. In the presence of the coupling (δ>0)(\delta>0) (Fig. 2 (c)), we find a glass phase at low enough temperatures where both spins and lattice degrees of freedom are frozen cooperatively. We find the replica symmetric solution is unstable there suggesting spontaneous replica symmetry breaking (RSB). This strongly implies the emergence of a complex free-energy landscape.

The rest of this paper is organized as follows. In Sec. II, we introduce our mean-field model. We present the development of the replica theory for our model in Sec. III. In Sec. IV, we present the analysis of the replica theory based on the replica symmetric (RS) ansatz: the phase diagram (Sec. IV.1) including the spin-lattice decoupled case (Sec. IV.2), the analysis of the stability condition of RS ansatz, the asymptotic behavior of the glass susceptibilities near the glass transition temperatures, and the behavior of the internal energy and the heat capacity for both separated (Sec. IV.3) and simultaneous (Sec. IV.4) glass transition cases. A summary and discussion of the results are given in Sec. V, including a comparison with the three-dimensional pyrochlore system. The details of the derivation of the free-energy functional, the analysis of the stability condition, and the computation of glass susceptibility are presented in Appendix. A, B and C, respectively. Finally in Appendix. D the definition of the internal energy and the heat capacity with the RS ansatz are presented.

II Model

II.1 Tree-like simplex network

Our model consists of classical spins and lattice distortions on a corner-sharing (k1)(k-1)-dimensional simplices network with connectivity cc which is the number of tetrahedra sharing one lattice site. The pyrochlore lattice and kagomé lattice correspond to the cases of (c,k)=(2,4)(c,k)=(2,4) and (c,k)=(2,3)(c,k)=(2,3), respectively, as shown in Figs. 1 (a) and (b). In order to develop an exactly solvable mean-field model, we consider a tree-like graph of simplices with connectivity cc, as shown in Figs. 1 (d) and (e) so that the presence of loops can be neglected except for the local triangles within the simplices. More precisely we consider dense limit Nc1N\gg c\gg 1Yoshino (2018) which greatly simplifies the theory.

II.2 Spin-lattice coupled model

We assume that spins are MM-component classical vector spins 𝑺i=(Si1,Si2,,SiM)\bm{S}_{i}=(S_{i}^{1},S_{i}^{2},...,S_{i}^{M}) (i=1,2,,N)(i=1,2,...,N) subjected to a spherical constraint i=1N|𝑺i|2=NM\sum_{i=1}^{N}|\bm{S}_{i}|^{2}=NM. We assume that the lattice distortions are cc-component vectors 𝝈=(σi1,σi2,,σic)\bm{\sigma}=(\sigma_{i}^{1},\sigma_{i}^{2},...,\sigma_{i}^{c}) subjected to a spherical constraint i=1N|𝝈i|2=Nc\sum_{i=1}^{N}|\bm{\sigma}_{i}|^{2}=Nc. Here cc is the connectivity of the lattice which plays the role of spacial dimension within our model. In order to obtain a non-trivial theory in the dense limit Nc1N\gg c\gg 1, we also assume M1M\gg 1 with,

α=cM\alpha=\frac{c}{M} (1)

being fixed to a value of O(1)O(1).

The Hamiltonian is given by,

H\displaystyle H =1c(k1)i,j[Jσi,σj𝑺i𝑺j+ϵ𝝈i𝝈j],\displaystyle=\frac{1}{\sqrt{c(k-1)}}\sum_{\langle i,j\rangle}\quantity[J_{\sigma_{i},\sigma_{j}}\bm{S}_{i}\cdot\bm{S}_{j}+\epsilon\bm{\sigma}_{i}\cdot\bm{\sigma}_{j}], (2)
Jσi,σj\displaystyle J_{\sigma_{i},\sigma_{j}} =J[1+δ(𝝈i𝒓^ij+𝝈j𝒓^ji)],(J>0)\displaystyle=J[1+\delta(\bm{\sigma}_{i}\cdot\hat{\bm{r}}_{ij}+\bm{\sigma}_{j}\cdot\hat{\bm{r}}_{ji})],~{}~{}~{}(J>0) (3)

where i,j\langle i,j\rangle denotes the summation over the nearest-neighboring pairs, 𝒓^ij\hat{\bm{r}}_{ij} is the unit vector from the ii-th site to the jj-th site, JJ and ϵ\epsilon are energy scales of spin-spin and lattice-lattice interactions, respectively. The parameter ϵ\epsilon reflects the elastic energy cost of lattice distortion Mitsumoto et al. (2022) so that we call it rigidity in the following. The spin-exchange interaction is antiferromagnetic without lattice distortion and is modified by the lattice distortion via the Goodenough-Kanamori mechanism Goodenough (1955); Kanamori (1959). The parameter δ\delta controls the strength of spin-lattice coupling.

Let us note that the Hamiltonian is invariant under the global reflection of the spins 𝑺i𝑺i\bm{S}_{i}\rightarrow-\bm{S}_{i} for i\forall i. On the other hand, it is not invariant under the global reflection of the lattice distortions 𝝈i𝝈i\bm{\sigma}_{i}\rightarrow-\bm{\sigma}_{i} for i\forall i in general except in the absence of the spin-lattice coupling δ=0\delta=0.

The present model can be regarded as a high dimensional variant of the model defined in Ref. Mitsumoto et al. (2020) for the pyrochlore system. In the pyrochlore system, if the spin-lattice coupling is switched off δ=0\delta=0, both spin and lattice distortion remain in liquid (paramagnetic) phase down to T=0T=0 due to frustration. In the presence of the coupling δ>0\delta>0, if the energy scale of lattice-lattice interactions is not too small, numerical simulations suggest the system exhibits a simultaneous spin and lattice glass transition at a critical temperature Mitsumoto et al. (2020).

On the corner-sharing (k1)(k-1)-simplices network, we can formally rewrite the Hamiltonian Eq. (2) as the summation of the potential energy running over all simplices =1,2,,N\triangle=1,2,\ldots,N_{\triangle},

H==1NV(𝑺(),σ())H=\sum_{\triangle=1}^{N_{\triangle}}V(\vec{\bm{S}}_{(\triangle)},\vec{\sigma}_{(\triangle)}) (4)

with

V(𝑺(),σ())\displaystyle V(\vec{\bm{S}}_{(\triangle)},\vec{\sigma}_{(\triangle)})
=1c(k1)i,j[Jσi,σj𝑺i𝑺j+ϵ𝝈i𝝈j],\displaystyle=\frac{1}{\sqrt{c(k-1)}}\sum_{\langle i,j\rangle\in\triangle}[J_{\sigma_{i},\sigma_{j}}\bm{S}_{i}\cdot\bm{S}_{j}+\epsilon\bm{\sigma}_{i}\cdot\bm{\sigma}_{j}], (5)

where

N=NckN_{\triangle}=N\frac{c}{k} (6)

is the number of simplices in the whole system, 𝑺()=(𝑺1(),𝑺2(),,𝑺k())\vec{\bm{S}}_{(\triangle)}=(\bm{S}_{1(\triangle)},\bm{S}_{2(\triangle)},...,\bm{S}_{k(\triangle)}) and 𝝈()=(𝝈1(),𝝈2(),,𝝈k())\vec{\bm{\sigma}}_{(\triangle)}=(\bm{\sigma}_{1(\triangle)},\bm{\sigma}_{2(\triangle)},...,\bm{\sigma}_{k(\triangle)}) represent the set of spin variables and lattice displacements belonging to a given simplex \triangle. The summation i,j\sum_{\langle i,j\rangle\in\triangle} represents summation over nearest neighbor pairs within a simplex \triangle.

The free energy of the system is given by βF=logZ-\beta F=\log Z, where β\beta is the inverse temperature and ZZ represents partition function defined as, Z=Tr𝑺Tr𝝈eβHZ=\text{Tr}_{\bm{S}}\text{Tr}_{\bm{\sigma}}e^{-\beta H}, where Tr𝑺\text{Tr}_{\bm{S}} and Tr𝝈\text{Tr}_{\bm{\sigma}} represent traces over the spin space of the spin 𝑺i\bm{S}_{i} and the lattice displacement 𝝈i\bm{\sigma}_{i} for all ii with spherical constraint,

Tr𝑺\displaystyle\text{Tr}_{\bm{S}} =i=1Nμ=1M(\bigintssssdSiμ)δ(11NMi=1N|𝑺i|2),\displaystyle=\prod_{i=1}^{N}\prod_{\mu=1}^{M}\quantity(\bigintssss_{-\infty}^{\infty}dS_{i}^{\mu})\delta\quantity(1-\frac{1}{NM}\sum_{i=1}^{N}|\bm{S}_{i}|^{2}), (7)
Tr𝝈\displaystyle\text{Tr}_{\bm{\sigma}} =i=1Nν=1c(\bigintssssdσiν)δ(11Nci=1N|𝝈i|2).\displaystyle=\prod_{i=1}^{N}\prod_{\nu=1}^{c}\quantity(\bigintssss_{-\infty}^{\infty}d\sigma_{i}^{\nu})\delta\quantity(1-\frac{1}{Nc}\sum_{i=1}^{N}|\bm{\sigma}_{i}|^{2}). (8)

II.3 Ground state manifold in the absence of spin-lattice coupling

In the absence of the coupling δ=0\delta=0, we have Jσi,σjJJ_{\sigma_{i},\sigma_{j}}\rightarrow J so that the system decouples into two independent subsystems (see Eq. (2), Eq. (3)). Within each of the subsystem the vectorial variables simplex network interact with each other anti-ferromagnetically. In the other representation of the hamiltonian Eq. (4) with Eq. (5) we find,

=1Nc(k1)V(𝑺,σ)\displaystyle\sum_{\triangle=1}^{N_{\triangle}}\sqrt{c(k-1)}V(\vec{\bm{S}_{\triangle}},\vec{\sigma}_{\triangle})
δ0\displaystyle\xrightarrow{\delta\rightarrow 0} =1N[J2(i𝑺i)2+ϵ2(i𝝈i)2]\displaystyle\sum_{\triangle=1}^{N_{\triangle}}\quantity[\frac{J}{2}\left(\sum_{i\in\triangle}\bm{S}_{i}\right)^{2}+\frac{\epsilon}{2}\left(\sum_{i\in\triangle}\bm{\sigma}_{i}\right)^{2}]
Nc2(JM+ϵc),\displaystyle-\frac{Nc}{2}(JM+\epsilon c), (9)

where we used i|𝑺i|2=NM\sum_{i}|\bm{S}_{i}|^{2}=NM and i|𝝈i|2=Nc\sum_{i}|\bm{\sigma}_{i}|^{2}=Nc. Thus the local Hamiltonian can be minimized by requiring

i𝑺i=0i𝝈i=0.\sum_{i\in\triangle}\bm{S}_{i}=0\qquad\sum_{i\in\triangle}\bm{\sigma}_{i}=0. (10)

Furthermore, since we are considering tree-like network of simplices, the configurations which satisfy the local sum rules Eq. (10) in all simplices \triangle are the ground states of the total system. Then a Maxwellian counting argument Reimers et al. (1991); Moessner and Chalker (1998) can be made for the ground state manifold. For the spin sector, we have NMNM degrees of freedom (disregarding the small correction due to the spin normalization condition) while the sum rule Eq. (10) imposes NM=NM(c/k)N_{\triangle}M=NM(c/k) conditions. Then NM(1c/k)NM(1-c/k) degrees of freedom remain in the ground state manifold of spins. Similarly, αNM(1c/k)\alpha NM(1-c/k) degrees of freedom remain in the ground state manifold of of the lattice distortions. Thus if we have,

0<ck<10<\frac{c}{k}<1 (11)

there is a macroscopic degeneracy of the ground states in the absence of spin-lattice coupling.

II.4 Microscopic background of the model

The microscopic background of the spin-lattice coupled model is explained in Mitsumoto et al. (2020). Let us summarize it here and add some further remarks. The lattice distortion is a dynamical Jahn-Teller distortion caused by the orbital degeneracy of the molybdenum ions Mitsumoto et al. (2022) and favors 2-inin-2-outout patterns following the ice rule. The lattice distortion changes the angle between an oxygen ion O2- and two magnetic ions Mo4+, modifying the superexchange interaction between spins by the Goodenough-Kanamori mechanism Goodenough (1955); Kanamori (1959); Solovyev (2003); Smerald and Jackeli (2019); Mitsumoto et al. (2020). Our numerical model on a pyrochlore lattice incorporates the above two features: lattice distortions favor the ice rule, and the interaction between classical Heisenberg spins is modified by the lattice distortion patterns as antiferromagnetic for inin-inin and inin-outout bonds and ferromagnetic for outout-outout bond.

In Ref. Mitsumoto et al. (2020), the dynamical Jahn-Teller effect was modeled phenomenologically by a simple spin-ice Hamiltonian with only nearest-neighbor interactions between lattice distortions. This was motivated by the experiment Thygesen et al. (2017) which suggested 2-inin-2-outout patterns of the lattice distortions. However in Ref. Mitsumoto et al. (2022), it was found based on a detailed microscopic analysis of the Jahn-Teller effect, that the lattice model should be extended to include also second and third nearest-neighbor interactions. It was found that the ground state of this extended lattice model obeys the 2-inin-2-outout rule but it exhibits a specific periodically alternating pattern, i. e. a long-ranged order. However, quite surprisingly, numerical simulations of the extended model showed that the system avoids the ordering transition even under very slow cooling and remains in a supercooled, disordered ice state down to the lowest temperature. This result suggests that we can assume an ice-like liquid state for the lattice degrees of freedom. Thus for the sake of simplicity, we just assume the nearest neighbor interactions like the simple spin-ice Hamiltonian for the lattice distortions and do not include further neighbor interactions in the present mean-field model.

III Replica theory

III.1 Free-energy functional

In this section, we introduce the replica method, derive the replicated free-energy as a functional of the order parameters and obtain the replica symmetric (RS) solution. The details of the derivation of the free-energy functional are given in Appendix. A.

The free energy of the system can be obtained formally as,

βF=logZ=limn0nZn.-\beta F=\log Z=\lim_{n\rightarrow 0}\partial_{n}Z^{n}. (12)

where ZnZ^{n} is the partition function of the replicated system a=1,2,,na=1,2,\ldots,n,

Zn=a(Tr𝑺aTr𝝈a)eβHnZ^{n}=\prod_{a}\quantity(\text{Tr}_{\bm{S}^{a}}\text{Tr}_{\bm{\sigma}^{a}})e^{-\beta H_{n}} (13)

Here HnH_{n} is the Hamiltonian of the replicated system,

βHn=a=1nβV(𝑺()a,σ()a).\beta H_{n}=\sum_{a=1}^{n}\sum_{\triangle}\beta V(\vec{\bm{S}}_{(\triangle)}^{a},\vec{\sigma}_{(\triangle)}^{a}). (14)

To investigate the glass transition, we introduce the glass order parameter matrices of spins and lattice distortions as overlaps between different replicas aa and bb,

Qab\displaystyle Q_{ab} =limN1NMi=1N𝑺ia𝑺ib,\displaystyle=\lim_{N\rightarrow\infty}\frac{1}{NM}\sum_{i=1}^{N}\expectationvalue{\bm{S}^{a}_{i}\cdot\bm{S}^{b}_{i}}, (15)
qab\displaystyle q_{ab} =limN1Nci=1N𝝈ia𝝈ib.\displaystyle=\lim_{N\rightarrow\infty}\frac{1}{Nc}\sum_{i=1}^{N}\expectationvalue{\bm{\sigma}^{a}_{i}\cdot\bm{\sigma}^{b}_{i}}. (16)

For convenience, we extend the matrices to include the diagonal elements, Qaa=qaa=1Q_{aa}=q_{aa}=1, to take into account the spherical constraints.

As explained in Appendix. A, the replicated partition function can be expressed in the dense limit Nc1N\gg c\gg 1 as,

Znab(𝑑Qab𝑑qab)e(Nc)sn[Q^,q^],Z^{n}\propto\prod_{a\leq b}\quantity(\int_{-\infty}^{\infty}dQ_{ab}\int_{-\infty}^{\infty}dq_{ab})e^{(Nc)s_{n}[\hat{Q},\hat{q}]}, (17)

where we defined,

sn[Q^,q^]=sent(s)[Q^]+sent(σ)[q^]+int[Q^,q^].s_{n}[\hat{Q},\hat{q}]=s_{\rm ent}^{(s)}[\hat{Q}]+s_{\rm ent}^{(\sigma)}[\hat{q}]+\mathcal{F}_{\rm int}[\hat{Q},\hat{q}]. (18)

with

sent(s)[Q^]\displaystyle s_{\rm ent}^{(s)}[\hat{Q}] =1α[n2(1+log(2π))+12log(detQ^)],\displaystyle=\frac{1}{\alpha}\quantity[\frac{n}{2}(1+\log(2\pi))+\frac{1}{2}\log(\det\hat{Q})], (19)
sent(σ)[q^]\displaystyle s_{\rm ent}^{(\sigma)}[\hat{q}] =n2(1+log(2π))+12log(detq^),\displaystyle=\frac{n}{2}(1+\log(2\pi))+\frac{1}{2}\log(\text{det}\hat{q}), (20)
int[Q^,q^]\displaystyle\mathcal{F}_{\rm int}[\hat{Q},\hat{q}] =β24a,b[J2α(1+2δ2qab)Qab2+ϵ2qab2].\displaystyle=\frac{\beta^{2}}{4}\sum_{a,b}\quantity[\frac{J^{2}}{\alpha}\quantity(1+2\delta^{2}q_{ab})Q_{ab}^{2}+\epsilon^{2}q_{ab}^{2}]. (21)

Note that the partition function doesn’t depend on the simplex dimension kk thanks to the scaling factor 1/k11/\sqrt{k-1} we introduced in the Hamiltonian. The resultant free energy functional becomes,

βf[Q^,q^]=βF[Q^,q^]Nc=sn[Q^,q^]n|n=0-\beta f[\hat{Q}^{*},\hat{q}^{*}]=\frac{-\beta F[\hat{Q}^{*},\hat{q}^{*}]}{Nc}=\evaluated{\partialderivative{s_{n}[\hat{Q}^{*},\hat{q}^{*}]}{n}}_{n=0} (22)

where QQ^{*} and qq^{*} are the saddle points which verify the saddle point equations,

0=sn[Q^,q^]Qab|Q^=Q^,q^=q^=sn[Q^,q^]qab|Q^=Q^,q^=q^.\displaystyle 0=\evaluated{\partialderivative{s_{n}[\hat{Q},\hat{q}]}{Q_{ab}}}_{\hat{Q}=\hat{Q}^{*},\hat{q}=\hat{q}^{*}}=\evaluated{\partialderivative{s_{n}[\hat{Q},\hat{q}]}{q_{ab}}}_{\hat{Q}=\hat{Q}^{*},\hat{q}=\hat{q}^{*}}. (23)

The stability condition is also required, namely the eigenvalues of the n(n1)×n(n1)n(n-1)\times n(n-1) Hessian matrix,

^=[HQQHQqHqQHqq],\hat{\mathcal{H}}=\matrixquantity[H_{QQ}&H_{Qq}\\ H_{qQ}&H_{qq}], (24)

where the components of the n(n1)2×n(n1)2\frac{n(n-1)}{2}\times\frac{n(n-1)}{2} block matrices HQQ,HQq=HqQH_{QQ},H_{Qq}=H_{qQ} and HqqH_{qq} are defined as

HQab,Qcd\displaystyle H_{Q_{ab},Q_{cd}} 2sn[Q^,q^]QabQcd,HQab,qcd2sn[Q^,q^]Qabqcd,\displaystyle\equiv-\partialderivative{s_{n}[\hat{Q},\hat{q}]}{Q_{ab}}{Q_{cd}},~{}~{}H_{Q_{ab},q_{cd}}\equiv-\partialderivative{s_{n}[\hat{Q},\hat{q}]}{Q_{ab}}{q_{cd}},
Hqab,qcd\displaystyle H_{q_{ab},q_{cd}} 2sn[Q^,q^]qabqcd\displaystyle\equiv-\partialderivative{s_{n}[\hat{Q},\hat{q}]}{q_{ab}}{q_{cd}} (25)

with a<b,c<da<b,c<d, must be non-negative in the n0n\rightarrow 0 limit. The glass susceptibility, which is associated with the nonlinear magnetic or dielectric susceptibilities, is given by the inverse matrix of the Hessian matrix,

χ^=[χQQχQqχqQχqq]=[HQQHQqHqQHqq]1=^1.\hat{\chi}=\matrixquantity[\chi_{QQ}&\chi_{Qq}\\ \chi_{qQ}&\chi_{qq}]=\matrixquantity[H_{QQ}&H_{Qq}\\ H_{qQ}&H_{qq}]^{-1}=\hat{\mathcal{H}}^{-1}. (26)

Hereafter, we call χQQ,χqq\chi_{QQ},\chi_{qq} and χQq\chi_{Qq} as the spin glass susceptibility, the lattice glass susceptibility, the cross glass susceptibility, respectively.

III.2 Spin-lattice decoupled case

In the absence of the spin-lattice coupling, i.e δ=0\delta=0, the free-energy (see Eq. (18)-Eq. (22)) become decoupled into the spin part and lattice part. The free energy associated with them is essentially the same as that of the spherical Sherrington-Kirkpatrick (SK) model Kosterlitz et al. (1976). It is known that the latter model has a low temperature phase where the global rotational symmetry is broken. However, the replica symmetry (RS) survives marginally in the sense that the stability of the RS solution (see sec III.3) is marginally stable.

Thus in the absence of the spin-lattice coupling, the spin and lattice degrees of freedom exhibit ’pseudo glass transitions’ accompanying breaking of the global rotational symmetry but without RSB. The marginally stable RS solutions imply that the free-energy landscapes in these low temperature phases are flat as shown schematically in Fig. 2 (a) and (b).

III.3 Replica symmetric ansatz

Now, our task is to solve the above problem with the simplest ansatz called the replica symmetric (RS) ansatz. In the RS ansatz we assume that the overlap matrices QabQ_{ab} and qabq_{ab} are parametrized as,

Qab=(1Q)δab+Q,qab=(1q)δab+q,\begin{split}Q_{ab}&=(1-Q)\delta_{ab}+Q,\\ q_{ab}&=(1-q)\delta_{ab}+q,\end{split} (27)

where δab\delta_{ab} is the Kronecker’s delta. Then the free energy within RS ansatz is obtained as,

βfRS\displaystyle-\beta f^{\rm RS} =12α(Q1Q+log(1Q))\displaystyle=\frac{1}{2\alpha}\quantity(\frac{Q}{1-Q}+\log(1-Q))
+12(q1q+log(1q))\displaystyle+\frac{1}{2}\quantity(\frac{q}{1-q}+\log(1-q))
+β24{J2α(1+2δ2)+ϵ2\displaystyle+\frac{\beta^{2}}{4}\Biggl{\{}\frac{J^{2}}{\alpha}(1+2\delta^{2})+\epsilon^{2}
[J2α(1+2δ2q)Q2+ϵ2q2]}+const.\displaystyle-\quantity[\frac{J^{2}}{\alpha}(1+2\delta^{2}q)Q^{2}+\epsilon^{2}q^{2}]\Biggr{\}}+\text{const.} (28)

The saddle point equations with respect to the order parameters QQ and qq are obtained as,

0\displaystyle 0 =Q[1(1Q)2(βJ)2(1+2δ2q)]\displaystyle=Q\quantity[\frac{1}{(1-Q)^{2}}-(\beta J)^{2}(1+2\delta^{2}q)] (29)
0\displaystyle 0 =q(1q)2(βϵ)2q(βJδ)2αQ2.\displaystyle=\frac{q}{(1-q)^{2}}-(\beta\epsilon)^{2}q-\frac{(\beta J\delta)^{2}}{\alpha}Q^{2}. (30)

We see that Q=q=0Q=q=0 is always a solution which represents the high temperature disordered (paramagnetic or liquid) state. Below the glass transition temperatures (See Appendix. B) this solution becomes unstable.

IV Results

IV.1 Phase diagram

Refer to caption
Figure 3: Mean-field phase diagrams for δ=0\delta=0 and δ>0\delta>0.

We show in Fig. 3 the typical phase diagram, which is derived from the RS saddle point equations Eq. (29) and Eq. (30). In the absence of the spin-lattice coupling δ=0\delta=0, we find the spin and lattice degrees of freedom exhibit ’pseudo glass transitions’ without RSB as discussed already in sec. IV.2. We show the decoupled phase diagrams in panels (a) and (b). In the presence of the spin-lattice coupling δ>0\delta>0 we find a new phase which we call as ’spin-lattice glass phase’ which accompanies RSB. As shown in panels (c),(d),(e), we always find the spin-lattice glass phase at low enough temperatures as long as δ>0\delta>0. In rigid regime ϵ/J>1\epsilon/J>1, the spin-lattice glass phase emerges within the pseudo lattice glass. On the other hand, direct transition from the high temperature phase (paramagnetic and lattice liquid phase) to the spin-lattice glass phase happens if the rigidity is low ϵ/J<1\epsilon/J<1.

IV.2 Spin-lattice decoupled case δ=0\delta=0

In the absence of the spin-lattice coupling, i.e δ=0\delta=0, the free-energy Eq. (28) as well as the saddle point equations Eq. (30) become decoupled into the spin part and lattice part. As noted in sec IV.2, the two decoupled systems are essentially the same as the spherical SK model Kosterlitz et al. (1976).

The pseudo lattice glass transition occurs at,

T(ϵ)=ϵ.T^{*}(\epsilon)=\epsilon. (31)

and below TT^{*}, q>0q>0 solution is obtained as,

q=1T/ϵ=t,q=1-T/\epsilon=t^{*}, (32)

where

t=(TT)/Tt^{*}=(T^{*}-T)/T^{*} (33)

is a dimensionless temperature that measures the distance of the temperature to the pseudo lattice glass transition temperature TT^{*}. From the analysis of the Hessian matrix Eq. (25), we find that the q>0q>0 RS solution is marginally stable, i.e., the minimum eigenvalue of the Hessian matrix is strictly zero in the pseudo lattice glass phase. Correspondingly, the glass susceptibility of the lattice distortion χqq\chi_{qq}, given in Eq. (5), diverges approaching TT^{*} from above with the power law |t|1|t^{*}|^{-1}, and remain divergent within the lattice glass phase. The detailed analysis of the Hessian matrix and the glass susceptibility are presented in Appendix B and Appendix. C, respectively.

The same analysis can be repeated for the spin degrees of freedom and one easily obtain the phase diagrams as shown in Fig. 3 (a)(b). In the following we turn to the cases of finite spin-lattice coupling δ>0\delta>0.

IV.3 High rigidity case

Let us first discuss the case of high rigidity, i. e. the lattice-lattice coupling is strong ϵ/J>1\epsilon/J>1. Lowering the temperature, we first pass over the pseudo lattice glass transition temperature T(ϵ)T^{*}(\epsilon) discussed above, which does not depend on the strength of the spin-lattice coupling δ\delta. The lattice order parameter qq linearly grows lowering the temperature below TT^{*} while the spin glass order parameter QQ still remains zero at high enough temperatures, as shown in Fig. 4 (a). As shown in Fig. 5 (a), χqq\chi_{qq} diverges passing TT^{*} while χQQ\chi_{QQ} and χQq\chi_{Qq} do not exhibit any singularity as expected. We also show in the figure the behaviour of the order parameters below TcT_{\rm c} using the RS solution in broken lines. Note that the RS solution is unstable below TcT_{\rm c} due to RSB.

By lowering the temperature further, the spin glass transition occurs at,

TcJ=Jϵ(δ4+(ϵJ)2(2δ2+1)δ2).\frac{T_{\rm c}}{J}=\frac{J}{\epsilon}\quantity(\sqrt{\delta^{4}+\quantity(\frac{\epsilon}{J})^{2}(2\delta^{2}+1)}-\delta^{2}). (34)

below which Q>0Q>0 solution exist. It can be seen that the phase boundary given by Eq. (34) is shifted to higher temperatures by increasing δ\delta, and it converges to Tc=T=ϵT_{\rm c}=T^{*}=\epsilon in the limit δ\delta\rightarrow\infty, i.e., the pseudo lattice glass phase disappears in this limit.

Refer to caption
Figure 4: Temperature dependencies of the order parameter QQ and qq in the RS ansatz for (a) high rigidity case (ϵ/J=2,α=1,δ=1.5\epsilon/J=2,\alpha=1,\delta=1.5) and (b) low rigidity case (ϵ/J=0.5,α=1,δ=1.5\epsilon/J=0.5,\alpha=1,\delta=1.5). The inset shows the behavior around TcT_{\rm c}^{\prime}. The broken lines represent the RS solution where it is unstable.

Below the spin glass transition temperature TcT_{\rm c}, the solutions of the order parameters are obtained from the saddle point equations Eq. (29) and Eq. (30) which can be written as,

Q\displaystyle Q =1TJ1+2δ2q,\displaystyle=1-\frac{T}{J\sqrt{1+2\delta^{2}q}}, (35)
Q\displaystyle Q =αTJδq(1q)2(ϵT)2q,\displaystyle=\frac{\sqrt{\alpha}T}{J\delta}\sqrt{\frac{q}{(1-q)^{2}}-\quantity(\frac{\epsilon}{T})^{2}q}, (36)

The equations can be solved numerically as shown in Fig. 4 (a). Just below TcT_{\rm c}, the spin glass order parameter continuously increases as

Q=(1+βcJ2δ2/ϵ)t+𝒪(t2),Q=(1+\beta_{c}J^{2}\delta^{2}/\epsilon)t+\mathcal{O}(t^{2}), (37)

where βc=1/Tc\beta_{c}=1/T_{\rm c} and

t=(TcT)/Tct=(T_{\rm c}-T)/T_{\rm c} (38)

measures the distance of the temperature to the spin glass transition temperature TcT_{\rm c}. On the other hand, the lattice glass order parameter is also modified reflecting the spin glass transition,

q\displaystyle q =t+u,\displaystyle=t^{*}+u, (39)
u\displaystyle u =(Jδϵ)2Q22α(βcϵ1)t2.\displaystyle=\quantity(\frac{J\delta}{\epsilon})^{2}\frac{Q^{2}}{2\alpha(\beta_{c}\epsilon-1)}\sim t^{2}. (40)
Refer to caption
Figure 5: Schematic picture of the temperature dependence of the glass susceptibilities. t=(TcT)/Tc,t=(TT)/Tct=(T_{\rm c}-T)/T_{\rm c},t^{*}=(T^{*}-T)/T_{\rm c} and t=(TcT)/Tct^{\prime}=(T_{\rm c}^{\prime}-T)/T_{\rm c}^{\prime} are the dimensionless temperatures normalized by the glass transition temperatures. The broken lines represent the RS solution where it is unstable.

Interestingly, the minimum eigenvalue of the Hessian matrix becomes negative below TcT_{\rm c},

λmin=2(βJδ)2Qt.\lambda_{\rm min}=-2(\beta J\delta)^{2}Q\sim-t. (41)

This means replica symmetry breaking must occur below TcT_{\rm c}. Most probably full RSB occurs. Note that the eigenvalue is quadratically proportional to the strength of spin-lattice coupling δ\delta. Thus, spin-lattice coupling is essential for this model to exhibit the RSB.

Approaching TcT_{\rm c} from above, the spin glass susceptibility χQQ\chi_{QQ} diverges with power law |t|1|t|^{-1}. Below TcT_{\rm c}, the spin glass susceptibility χQQ\chi_{QQ} as well as the lattice glass susceptibility decay with |t|1|t|^{-1} within the RS ansatz. The cross glass susceptibility χQq\chi_{Qq} doesn’t grow above TcT_{\rm c}, but it suddenly diverges at TcT_{\rm c} and decays with |t|1|t|^{-1} below TcT_{\rm c}.

Refer to caption
Figure 6: (a), (b): Internal energy and heat capacity in the high rigidity case (ϵ/J=2,α=1,δ=1.5\epsilon/J=2,\alpha=1,\delta=1.5). (c), (d): Internal energy and heat capacity in the low rigidity case (ϵ/J=0.5,α=1,δ=1.5\epsilon/J=0.5,\alpha=1,\delta=1.5). T,TcT^{*},T_{\rm c} and TcT_{\rm c}^{\prime} are the glass transition temperatures obtained above. The broken lines represent the RS solution where it is unstable.

Fig.6 (a) and (c) shows the internal energy and the heat capacity for ϵ/J=2,α=1\epsilon/J=2,\alpha=1 and δ=1.5\delta=1.5 obtained by the RS solution (see Appendix. D). Note that the heat capacity exhibits a kink at TT^{*} and a cusp at TcT_{\rm c}.

IV.4 Low rigidity case

If the rigidity is low ϵ/J<1\epsilon/J<1, the situation changes: a simultaneous spin and lattice glass transition takes place at a common transition temperature,

Tc/J=1,T_{\rm c}^{\prime}/J=1, (42)

which doesn’t depend on the elastic energy scale ϵ\epsilon nor the amplitude δ\delta of the spin-lattice coupling. The latter implies that the spin degrees of freedom play the dominant role in this case.

Below TcT_{\rm c}^{\prime}, the non-zero order parameters q,Q>0q,Q>0 are obtained by solving Eqs. (35) and (36) numerically as shown in Fig. 4 (b). We find continuous growth of the spin and lattice glass order parameters passing TcT_{\rm c}^{\prime} lowering the temperature. Note that both of the glass order parameters exhibit jumps slightly below TcT_{\rm c}^{\prime}. Using the dimensionless temperature,

t=(TcT)/Tct^{\prime}=(T_{\rm c}^{\prime}-T)/T_{\rm c}^{\prime} (43)

which measures the distance of the temperature to TcT_{\rm c}^{\prime}, the order parameters QQ and qq slightly below TcT_{\rm c}^{\prime} can be obtained as,

Q\displaystyle Q =1T/J+δ4α(1(ϵ/J)2)t2+𝒪(t3)\displaystyle=1-T/J+\frac{\delta^{4}}{\alpha(1-(\epsilon/J)^{2})}t^{\prime 2}+\mathcal{O}(t^{\prime 3})
=t+δ2q+𝒪(t3),\displaystyle=t^{\prime}+\delta^{2}q+\mathcal{O}(t^{\prime 3}), (44)
q\displaystyle q =δ2α(1(ϵ/J)2)t2+𝒪(t3).\displaystyle=\frac{\delta^{2}}{\alpha(1-(\epsilon/J)^{2})}t^{\prime 2}+\mathcal{O}(t^{\prime 3}). (45)

Comparing the two order parameters, we notice that the growth of the lattice glass order parameter qq is weaker than that of the spin glass order parameter QQ. More precisely, the critical exponents are different: qq is proportional to t2t^{2} while QQ is proportional to tt. If we switch off the spin-lattice coupling, δ0\delta\rightarrow 0, the lattice glass order parameter qq becomes zero, and the behavior of the spin glass order parameter QQ agrees with the spherical SK model Kosterlitz et al. (1976), i.e., Q=tQ=t^{\prime}. The spin glass order parameter QQ is weakly modified by the lattice glass order parameter qq at order O(t2)O(t^{\prime 2}).

From the analysis of the Hessian matrix, we find that the RS solution becomes unstable below TcT_{\rm c}^{\prime}. Slightly below TcT_{\rm c}^{\prime}, the minimum eigenvalue behaves as,

λmin=4(βJδ)4Q21(βϵ)2(t)2,\lambda_{\rm min}=-\frac{4(\beta J\delta)^{4}Q^{2}}{1-(\beta\epsilon)^{2}}\sim-(t^{\prime})^{2}, (46)

which also becomes zero if δ=0\delta=0. The spin glass susceptibility χQQ\chi_{QQ} exhibits a diverging feature approaching TcT_{\rm c} from above with power law |t|1|t^{\prime}|^{-1} while it decays as |t|2|t|^{-2} below TcT_{\rm c}, reflecting λmin\lambda_{\rm min}, as shown in Fig. 5 (b). In sharp contrast, the lattice glass susceptibility does not diverge at TcT_{\rm c}. Similarly to the case of the separated glass transition ϵ/J<1\epsilon/J<1, the cross glass susceptibility χQq\chi_{Qq} does not grow above TcT_{\rm c}, suddenly diverges at TcT_{\rm c} and decays as |t|1|t^{\prime}|^{-1} below TcT_{\rm c}^{\prime}.

Fig.6 (b) and (d) shows the internal energy and the heat capacity for ϵ/J=0.5,α=1\epsilon/J=0.5,\alpha=1 and δ=1.5\delta=1.5 using the RS solution. Reflecting the jumps of the order parameters shown in Fig. 4 (b), the energy exhibits a non-monotonic behavior, and the heat capacity diverges negatively. We believe that the latter unphysical behavior is due to the failure of the RS solution below TcT_{\rm c}^{\prime}.

V Discussion and Summary

In summary, we have constructed a disorder-free mean-field model of the spin-lattice coupled system on the tree-like (k1)(k-1)-simplices network. Using the replica method, we solved the model exactly in the dense limit Nc1N\gg c\gg 1 corresponding to the large dimensional limit dd\rightarrow\infty. We have found the spin-lattice glass phase where both the spin and lattice degrees of freedom are frozen. The replica symmetry is broken there through the spin-lattice coupling yielding a complex and hierarchical free-energy landscape.

One may wander why a spin glass transition without the pseudo lattice glass transition does not happen while the pseudo lattice glass transition without a spin glass transition happens at T(ϵ)T^{*}(\epsilon). This is due to the spin-lattice coupling of the form ((𝝈i𝒓^ij)𝑺i𝑺j+(𝝈j𝒓^ji)𝑺i𝑺j)((\bm{\sigma}_{i}\cdot\hat{\bm{r}}_{ij})\bm{S}_{i}\cdot\bm{S}_{j}+(\bm{\sigma}_{j}\cdot\hat{\bm{r}}_{ji})\bm{S}_{i}\cdot\bm{S}_{j}) (see Eq. (3)). When the lattice distortions become frozen, the lattice degrees of freedom play the role of bond randomness for the spins, while when the spins become frozen, the spin degrees of freedom play the role of random fields for the lattice distortions. The spin-glass transition does not take place until the effective coupling between spins become sufficiently strong. On the other hand, when a non-zero random field sets-in it inevitably acts as a polarizing field. Thus when the spins become frozen, the glass order parameter of the lattice distortions necessarily becomes finite but not vice versa.

Now let us compare the results obtained in the mean-field model and the numerical observations in the three-dimensional model on the pyrochlore lattice Mitsumoto et al. (2020). In the low rigidity regime ϵ/J<1\epsilon/J<1, the mean-field model undergoes simultaneous spin and lattice glass transitions. This point would appear to be consistent with the numerical observation in the three-dimensional system Mitsumoto et al. (2020). However the crucial difference is that in the mean-field case the glass susceptibility associated with the lattice degrees of freedom does not exhibit a singularity in this regime. This may be related to the fact that the mechanism to enable the non-zero lattice glass order parameter, in this regime, is the random field effect mentioned above.

Next let us turn to the high rigidity regime ϵ/J>1\epsilon/J>1. Here it is interesting to compare first the heat capacity. In the three-dimensional model, the heat capacity exhibits two peaks at a high temperature and a low temperature (see Supplementary material Fig. S3 of Ref. Mitsumoto et al. (2020)). Interestingly the heat capacity of the present mean-field model also exhibits two anomalies at different temperatures in the case of high rigidity ϵ/J>1\epsilon/J>1: there is a kink at TT^{*} and peak at TcT_{\rm c} (see Fig. 6 (c) ). In the three-dimensional model, the peak of the heat capacity at the higher temperature reflects a crossover from simple liquid at higher temperatures to ice-like liquid at lower temperatures where the 2-in-2-out ice rule is satisfied nearly perfectly. On the other hand, the peak at the lower temperature reflects a glass transition where both the spin and lattice degrees of exhibit a simultaneous glass transition accompanying diverging behavior of non-linear magnetic and dielectric susceptibilities associated with the spin and lattice degrees of freedom. On the other hand, in the mean-field model, the two anomalies in the heat capacity reflect the two separated glass transitions. The (pseudo) lattice glass transition temperature T=ϵT^{*}=\epsilon does not depend on the amplitude of the spin-lattice coupling δ\delta, and it can occur without the spin-lattice coupling δ=0\delta=0. We speculate that this (pseudo) lattice glass transition is an artifact of the mean-field model Fisher (1991) which is replaced by a smooth crossover in three-dimensions: TT^{*} is the crossover temperature below which the lattice-lattice interactions become significant. In the three-dimensional model, the heat capacity peak at the lower temperature shifts to the higher temperatures by increasing ϵ\epsilon, but it saturates at the higher ϵ\epsilon region. Very similar ϵ\epsilon dependence can be seen in the glass transition temperature TcT_{\rm c} of the mean-field theory (See Fig. 3). Indeed TcT_{\rm c} given in Eq. (34) converges to a constant Tc/J(2δ2+1)T_{\rm c}/J\rightarrow\sqrt{(2\delta^{2}+1)} in ϵ\epsilon\rightarrow\infty limit where lattice degrees of freedom should be completely frozen. This is consistent with the fact that in the usual spin glass models with quenched disorder, the spin glass transition is proportional to the amplitude of the quenched disorder Saunders and Chalker (2007). Based on the above observations, we speculate that the mean-field model in the high rigidity regime ϵ/J>1\epsilon/J>1 compares better with the numerical observations made in the three-dimensional system.

Acknowledgments

We warmly thank Chisa Hotta for the collaborations closely related to this work and many useful discussions. This work was supported by KANENHI (No. 19H01812) (No. 21K18146) from MEXT, Japan.

Appendix A Derivation of the free energy functional

In order to observe the glass transition with spontaneous replica symmetry breaking, we explicitly put the coupling between replicas into the Hamiltonian asParisi and Virasoro (1989); Monasson (1995); Mézard and Parisi (1999); Yoshino (2018),

βH\displaystyle\beta H =a=1nβV(𝑺()a,σ()a)\displaystyle=\sum_{a=1}^{n}\sum_{\triangle}\beta V(\vec{\bm{S}}_{(\triangle)}^{a},\vec{\sigma}_{(\triangle)}^{a})
a<bi[Λabμ(Sa)iμ(Sb)iμ+λabν(σa)iν(σb)iν],\displaystyle-\sum_{a<b}\sum_{i}\quantity[\Lambda_{ab}\sum_{\mu}(S^{a})_{i}^{\mu}(S^{b})_{i}^{\mu}+\lambda_{ab}\sum_{\nu}(\sigma^{a})_{i}^{\nu}(\sigma^{b})_{i}^{\nu}], (47)

and then study the behavior of the glass order parameter matrices of spins and lattice distortions under the coupling fields,

Qab\displaystyle Q_{ab} =limΛab,λab0limN1NMi=1Nμ=1M(Sa)iμ(Sb)iμΛ,λ,\displaystyle=\lim_{\Lambda_{ab},\lambda_{ab}\rightarrow 0}\lim_{N\rightarrow\infty}\frac{1}{NM}\sum_{i=1}^{N}\sum_{\mu=1}^{M}\expectationvalue{(S^{a})_{i}^{\mu}(S^{b})_{i}^{\mu}}_{\Lambda,\lambda}, (48)
qab\displaystyle q_{ab} =limΛab,λab0limN1Nci=1Nν=1c(σa)iν(σb)iνΛ,λ.\displaystyle=\lim_{\Lambda_{ab},\lambda_{ab}\rightarrow 0}\lim_{N\rightarrow\infty}\frac{1}{Nc}\sum_{i=1}^{N}\sum_{\nu=1}^{c}\expectationvalue{(\sigma^{a})_{i}^{\nu}(\sigma^{b})_{i}^{\nu}}_{\Lambda,\lambda}. (49)

Here Λ,λ\expectationvalue{...}_{\Lambda,\lambda} represents the thermal average in the presence of the symmetry breaking fields λ,Λ\lambda,\Lambda, and these become the same as Eq. (15) and Eq. (16) after taking the limit λ,Λ0\lambda,\Lambda\rightarrow 0. Below we closely follow the steps used in Yoshino (2018).

To obtain a free-energy functional in terms of the order parameters, we perform the Legendre transformation. This can be done in practice by making use of the following identities,

1\displaystyle 1 =\bigintssssdQabδ(Qab1NMi=1Nμ=1M(Sa)iμ(Sb)iμ),\displaystyle=\bigintssss_{-\infty}^{\infty}dQ_{ab}\delta\quantity(Q_{ab}-\frac{1}{NM}\sum_{i=1}^{N}\sum_{\mu=1}^{M}(S^{a})_{i}^{\mu}(S^{b})_{i}^{\mu}), (50)
1\displaystyle 1 =\bigintssssdqabδ(qab1Nci=1Nν=1c(σa)iν(σb)iν),\displaystyle=\bigintssss_{-\infty}^{\infty}dq_{ab}\delta\quantity(q_{ab}-\frac{1}{Nc}\sum_{i=1}^{N}\sum_{\nu=1}^{c}(\sigma^{a})_{i}^{\nu}(\sigma^{b})_{i}^{\nu}), (51)

for a<ba<b.

Then the traces of spin can be expressed formally as,

a=1nTr𝑺a[]=12nab(NM\bigintssssiidΛab2πi\bigintssssdQab)\displaystyle\prod_{a^{\prime}=1}^{n}\text{Tr}_{\bm{S}^{a^{\prime}}}[\cdots]=\frac{1}{2^{n}}\prod_{a\leq b}\quantity(NM\bigintssss_{-i\infty}^{i\infty}\frac{d\Lambda_{ab}}{2\pi i}\bigintssss_{-\infty}^{\infty}dQ_{ab})
×exp[NM(12a,b=1nΛabQab+logZΛ)]i,μi,μ,\displaystyle\times\exp\quantity[NM\quantity(\frac{1}{2}\sum_{a,b=1}^{n}\Lambda_{ab}Q_{ab}+\log Z_{\Lambda})]\prod_{i,\mu}\expectationvalue{\cdots}_{i,\mu}, (52)

where ZΛ=((2π)n/det(Λ^))1/2Z_{\Lambda}=((2\pi)^{n}/\text{det}(\hat{\Lambda}))^{1/2},

i,μ\displaystyle\expectationvalue{\cdots}_{i,\mu} =1ZΛa=1n(d(Sia)μ)\displaystyle=\frac{1}{Z_{\Lambda}}\prod_{a^{\prime}=1}^{n}\quantity(\int_{-\infty}^{\infty}d(S_{i}^{a^{\prime}})^{\mu})
×exp[12a,b=1nΛab(Sia)μ(Sib)μ][],\displaystyle\times\exp\quantity[{-\frac{1}{2}\sum_{a,b=1}^{n}\Lambda_{ab}(S_{i}^{a})^{\mu}(S_{i}^{b})^{\mu}}][\cdots], (53)

Λab\Lambda_{ab} is Lagrange multiplier of Eq. (7) and Eq. (50), and i,μi,μ\prod_{i,\mu}\expectationvalue{\cdots}_{i,\mu} means averages for all sites ii and components μ\mu. In the limit NN\rightarrow\infty, we can use a saddle-point method to integrate out Λab\Lambda_{ab}. and the saddle point equation which determines the saddle point Λab(Q^)\Lambda_{ab}^{*}(\hat{Q}) is given by,

Qab=(Λ)ab1Q_{ab}=(\Lambda^{*})_{ab}^{-1} (54)

and we find,

a=1nTr𝑺a[]ab(𝑑Qab)eNcsent(s)[Q^]i,μi,μ,\prod_{a^{\prime}=1}^{n}\text{Tr}_{\bm{S}^{a^{\prime}}}[\cdots]\propto\prod_{a\leq b}\quantity(\int_{-\infty}^{\infty}dQ_{ab})e^{Ncs_{\rm ent}^{(s)}[\hat{Q}]}\prod_{i,\mu}\expectationvalue{\cdots}_{i,\mu}, (55)

where sent(s)[Q^]s_{\rm ent}^{(s)}[\hat{Q}] represents the entropic contribution of spins to the free energy, which is given by Eq. (19).

Similarly, the trace of lattice displacement is formally obtained as,

a=1nTr𝝈a[]=12nab(Nc\bigintssssiidλab2πi\bigintssssdqab)\displaystyle\prod_{a^{\prime}=1}^{n}\text{Tr}_{\bm{\sigma}^{a^{\prime}}}[\cdots]=\frac{1}{2^{n}}\prod_{a\leq b}\quantity(Nc\bigintssss_{-i\infty}^{i\infty}\frac{d\lambda_{ab}}{2\pi i}\bigintssss_{-\infty}^{\infty}dq_{ab})
×exp[Nc(12a,b=1nλabqab+logZλ)]i,νi,ν,\displaystyle\times\exp\quantity[Nc\quantity(\frac{1}{2}\sum_{a,b=1}^{n}\lambda_{ab}q_{ab}+\log Z_{\lambda})]\prod_{i,\nu}\expectationvalue{\cdots}_{i,\nu}, (56)

where Zλ=((2π)n/det(λ^))1/2Z_{\lambda}=((2\pi)^{n}/\text{det}(\hat{\lambda}))^{1/2},

i,ν\displaystyle\expectationvalue{\cdots}_{i,\nu} =1Zλa=1n(d(σia)ν)\displaystyle=\frac{1}{Z_{\lambda}}\prod_{a^{\prime}=1}^{n}\quantity(\int_{-\infty}^{\infty}d(\sigma_{i}^{a^{\prime}})^{\nu})
×exp[12a,b=1nλab(σia)ν(σib)ν][],\displaystyle\times\exp\quantity[-\frac{1}{2}\sum_{a,b=1}^{n}\lambda_{ab}(\sigma_{i}^{a})^{\nu}(\sigma_{i}^{b})^{\nu}][\cdots], (57)

λab\lambda_{ab} is the Lagrange multiplier of Eq. (8) and Eq. (51), and i,νi,ν\prod_{i,\nu}\expectationvalue{\cdots}_{i,\nu} represents averages for all sites ii and components ν\nu. In the limit NN\rightarrow\infty, we use the saddle point method to integrate out λab\lambda_{ab}. The saddle point equation is given by,

qab=(λ)ab1.\displaystyle q_{ab}=(\lambda^{*})_{ab}^{-1}. (58)

We obtain,

a=1nTr𝝈a[]ab(𝑑qab)eNcsent(σ)[q^]i,νi,ν,\prod_{a^{\prime}=1}^{n}\text{Tr}_{\bm{\sigma}^{a^{\prime}}}[\cdots]\propto\prod_{a\leq b}\quantity(\int_{-\infty}^{\infty}dq_{ab})e^{Ncs_{\rm ent}^{(\sigma)}[\hat{q}]}\prod_{i,\nu}\expectationvalue{\cdots}_{i,\nu}, (59)

where sent(σ)[q^]s_{\rm ent}^{(\sigma)}[\hat{q}] represents the entropic contribution of spins to the free energy, which is given by Eq. (20).

Supposing the translational and rotational symmetry, where all sites ii and components μ,ν\mu,\nu are equivalent to each other, we find

i,μ(Sia)μi,μ=0,i,μ(Sia)μ(Sib)μi,μ=Qab,i,ν(σia)νi,ν=0,i,ν(σia)ν(σb)νi,ν=qab.\displaystyle\begin{aligned} &\prod_{i,\mu}\expectationvalue{(S_{i}^{a})^{\mu^{\prime}}}_{i,\mu}=0,\\ &\prod_{i,\mu}\expectationvalue{(S_{i}^{a})^{\mu^{\prime}}(S_{i}^{b})^{\mu^{\prime}}}_{i,\mu}=Q_{ab},\\ &\prod_{i,\nu}\expectationvalue{(\sigma_{i}^{a})^{\nu^{\prime}}}_{i,\nu}=0,\\ &\prod_{i,\nu}\expectationvalue{(\sigma_{i}^{a})^{\nu^{\prime}}(\sigma^{b})^{\nu^{\prime}}}_{i,\nu}=q_{ab}.\end{aligned} (60)

These relations are very useful for performing the cumulant expansion, and we can obtain the exact form of the replicated free energy. Now the replicated partition function can be rewritten as,

Zn\displaystyle Z^{n} ab(𝑑Qab𝑑qab)\displaystyle\propto\prod_{a\leq b}\quantity(\int_{-\infty}^{\infty}dQ_{ab}\int_{-\infty}^{\infty}dq_{ab})
×exp[Nc(sent(s)[Q^]+sent(σ)[q^]+int[Q^,q^])],\displaystyle\times\exp\quantity[Nc\quantity(s_{\rm ent}^{(s)}[\hat{Q}]+s_{\rm ent}^{(\sigma)}[\hat{q}]+\mathcal{F}_{\rm int}[\hat{Q},\hat{q}])], (61)

where

Ncint\displaystyle Nc\mathcal{F}_{\rm int}
=\displaystyle= log(i,μ,νea=1n=1NβV(𝑺()a,σ()a)i,μ,ν)\displaystyle\log\quantity(\prod_{i,\mu,\nu}\expectationvalue{e^{-\sum_{a=1}^{n}\sum_{\triangle=1}^{N_{\triangle}}\beta V(\vec{\bm{S}}^{a}_{(\triangle)},\vec{\sigma}^{a}_{(\triangle)})}}_{i,\mu,\nu})
=\displaystyle= log(i,μ,νeβc(k1)a=1n=1Ni<jkAij()ai,μ,ν).\displaystyle\log\quantity(\prod_{i,\mu,\nu}\expectationvalue{e^{\frac{-\beta}{\sqrt{c(k-1)}}\sum_{a=1}^{n}\sum_{\triangle=1}^{N_{\triangle}}\sum_{i<j}^{k}A_{ij(\triangle)}^{a}}}_{i,\mu,\nu}). (62)

with

Aija\displaystyle A_{ij}^{a} =J(1+δ(𝝈ia𝒓^ij+𝝈ja𝒓^ji))μ=1M(Sia)μ(Sja)μ\displaystyle=J\quantity(1+\delta(\bm{\sigma}^{a}_{i}\cdot\hat{\bm{r}}_{ij}+\bm{\sigma}^{a}_{j}\cdot\hat{\bm{r}}_{ji}))\sum_{\mu=1}^{M}(S_{i}^{a})^{\mu}(S_{j}^{a})^{\mu}
+ϵν=1c(σia)ν(σja)ν.\displaystyle+\epsilon\sum_{\nu=1}^{c}(\sigma_{i}^{a})^{\nu}(\sigma_{j}^{a})^{\nu}. (63)

i,μ,νi,μ,ν\prod_{i,\mu,\nu}\expectationvalue{\cdots}_{i,\mu,\nu} means averages for all i,μ,νi,\mu,\nu. Noting that the average can be decoupled with respect to different site i()i~{}(\triangle), spin component μ\mu, and distortion component ν\nu, we can evaluate int\mathcal{F}_{\rm int} using the cumulant expansion,

log(i,μ,νeβc(k1)ai<jAij()ai,μ,ν)\displaystyle\log\quantity(\prod_{i,\mu,\nu}\expectationvalue{e^{\frac{-\beta}{\sqrt{c(k-1)}}\sum_{a}\sum_{\triangle}\sum_{i<j}A_{ij(\triangle)}^{a}}}_{i,\mu,\nu})
=βc(k1)N(k2)aAija\displaystyle=\frac{-\beta}{\sqrt{c(k-1)}}N_{\triangle}\matrixquantity(k\\ 2)\sum_{a}\expectationvalue{A_{ij}^{a}}
+12!β2c(k1)N(k2)a,b(AijaAijbAijaAijb)+\displaystyle+\frac{1}{2!}\frac{\beta^{2}}{c(k-1)}N_{\triangle}\matrixquantity(k\\ 2)\sum_{a,b}\quantity(\expectationvalue{A_{ij}^{a}A_{ij}^{b}}-\expectationvalue{A_{ij}^{a}}\expectationvalue{A_{ij}^{b}})+\cdots (64)

with

Aija\displaystyle\expectationvalue{A_{ij}^{a}} =0,\displaystyle=0, (65)
AijaAijb\displaystyle\expectationvalue{A_{ij}^{a}A_{ij}^{b}} =J2(1+2δ2qab)Qab2+ϵ2qab2.\displaystyle=J^{2}\quantity(1+2\delta^{2}q_{ab})Q_{ab}^{2}+\epsilon^{2}q_{ab}^{2}. (66)

Here, we used ν=1c(r^ijν)2=1\sum_{\nu=1}^{c}(\hat{r}_{ij}^{\nu})^{2}=1 and Eq. (60). Higher order cumulants vanish in the dense limitYoshino (2018), i.e. limc\lim_{c\rightarrow\infty} after limN\lim_{N\rightarrow\infty}. Finally, we obtain Eq. (21) using Eq. (6).

Appendix B Stability of the RS solution

Now, let us investigate the stability of the RS solution, which is associated with the divergence of the glass susceptibility, shown in Appendix. C. The necessary condition of the RS solution to be stable is that the eigenvalues of Hessian matrix \mathcal{H} given by,

^=[HQQHQqHqQHqq],\hat{\mathcal{H}}=\matrixquantity[H_{QQ}&H_{Qq}\\ H_{qQ}&H_{qq}], (67)

are all non-negative, where the submatrices HQQH_{QQ}, HQq=HqQH_{Qq}=H_{qQ} and HqqH_{qq} are given by Eq. (25). To analyze these matrices we need the first and second derivatives of the functional sn[Q^,q^]s_{n}[\hat{Q},\hat{q}], which are obtained as,

2snQabQcd\displaystyle\partialderivative{s_{n}}{Q_{ab}}{Q_{cd}} =1α(Qac1Qbd1+Qad1Qbc1)\displaystyle=-\frac{1}{\alpha}\quantity(Q_{ac}^{-1}Q_{bd}^{-1}+Q_{ad}^{-1}Q_{bc}^{-1})
+δacδbd(βJ)2α(1+2δ2qab)\displaystyle+\delta_{ac}\delta_{bd}\frac{(\beta J)^{2}}{\alpha}(1+2\delta^{2}q_{ab}) (68)
2snQabqcd\displaystyle\partialderivative{s_{n}}{Q_{ab}}{q_{cd}} =2δacδbd(βJδ)2Qab\displaystyle=2\delta_{ac}\delta_{bd}(\beta J\delta)^{2}Q_{ab} (69)
2snqabqcd\displaystyle\partialderivative{s_{n}}{q_{ab}}{q_{cd}} =[(qac1qbd1+qad1qbc1)+δacδbd(βϵ)2],\displaystyle=\quantity[-\quantity(q_{ac}^{-1}q_{bd}^{-1}+q_{ad}^{-1}q_{bc}^{-1})+\delta_{ac}\delta_{bd}(\beta\epsilon)^{2}], (70)

where a<b,c<da<b,c<d. In the replica symmetric case we have Qab=(1Q)δabQ_{ab}=(1-Q)\delta_{ab} and qab=(1q)δab+qq_{ab}=(1-q)\delta_{ab}+q (see Eq. (27)). Then submatrices of the Hessian matrix in the limit n0n\rightarrow 0 can be cast into the following form,

{HQQ}ab,cd\displaystyle\quantity{H_{QQ}}_{ab,cd} =M1Qδacδbd+δadδbc2\displaystyle=M_{1}^{Q}\frac{\delta_{ac}\delta_{bd}+\delta_{ad}\delta_{bc}}{2}
+M2Qδac+δad+δbc+δbd4+M3Q\displaystyle+M_{2}^{Q}\frac{\delta_{ac}+\delta_{ad}+\delta_{bc}+\delta_{bd}}{4}+M_{3}^{Q} (71)
{HQq}ab,cd\displaystyle\quantity{H_{Qq}}_{ab,cd} =γδacδbd\displaystyle=\gamma\delta_{ac}\delta_{bd} (72)
{Hqq}ab,cd\displaystyle\quantity{H_{qq}}_{ab,cd} =M1qδacδbd+δadδbc2\displaystyle=M_{1}^{q}\frac{\delta_{ac}\delta_{bd}+\delta_{ad}\delta_{bc}}{2}
+M2qδac+δad+δbc+δbd4+M3q,\displaystyle+M_{2}^{q}\frac{\delta_{ac}+\delta_{ad}+\delta_{bc}+\delta_{bd}}{4}+M_{3}^{q}, (73)

where

M1Q\displaystyle M_{1}^{Q} =\displaystyle= 2α(1Q)22(βJ)2(1+2δ2q)α\displaystyle\frac{2}{\alpha(1-Q)^{2}}-\frac{2(\beta J)^{2}(1+2\delta^{2}q)}{\alpha}
M1q\displaystyle M_{1}^{q} =\displaystyle= 2(1q)22(βϵ)2\displaystyle\frac{2}{(1-q)^{2}}-2(\beta\epsilon)^{2} (74)
M2Q\displaystyle M_{2}^{Q} =\displaystyle= 4Qα(1Q)3M2q=4q(1q)3\displaystyle\frac{-4Q}{\alpha(1-Q)^{3}}\qquad M_{2}^{q}=\frac{-4q}{(1-q)^{3}} (75)
M3Q\displaystyle M_{3}^{Q} =\displaystyle= 2Q2α(1Q)4M3q=2q2(1q)4\displaystyle\frac{-2Q^{2}}{\alpha(1-Q)^{4}}\qquad M_{3}^{q}=\frac{-2q^{2}}{(1-q)^{4}} (76)

and

γ=2(βJδ)2Q.\gamma=-2(\beta J\delta)^{2}Q. (77)

First let us analyze the eigenvalues of the submatrices HQQH_{QQ} and HqqH_{qq} before considering the eigenvalues of the total Hessian matrix ^\hat{\mathcal{H}}. From Eqs. (71) and (73), we find that both HQQH_{QQ} and HqqH_{qq} become diagonalized by the same orthogonal matrix PP and then the eigenvalues λkω\lambda_{k}^{\omega} for k=1,2,,n(n1)/2k=1,2,...,n(n-1)/2 and ω=Q,q\omega=Q,q are obtained as

λkω={λLω(k=1)λAω(2kn)λRω(n+1kn(n1)/2),\lambda_{k}^{\omega}=\begin{cases}\lambda_{L}^{\omega}~{}~{}~{}(k=1)\\ \lambda_{A}^{\omega}~{}~{}~{}(2\leq k\leq n)\\ \lambda_{R}^{\omega}~{}~{}~{}(n+1\leq k\leq n(n-1)/2),\end{cases} (78)

with

λLω\displaystyle\lambda_{L}^{\omega} =12M1ω+(n1)2M2ω+n(n1)2M3ω\displaystyle=\frac{1}{2}M_{1}^{\omega}+\frac{(n-1)}{2}M_{2}^{\omega}+\frac{n(n-1)}{2}M_{3}^{\omega}
n012(M1ωM2ω),\displaystyle\xrightarrow[n\rightarrow 0]{}\frac{1}{2}(M_{1}^{\omega}-M_{2}^{\omega}), (79)
λAω\displaystyle\lambda_{A}^{\omega} =12M1ω+n24M2ωn012(M1ωM2ω),\displaystyle=\frac{1}{2}M_{1}^{\omega}+\frac{n-2}{4}M_{2}^{\omega}\xrightarrow[n\rightarrow 0]{}\frac{1}{2}(M_{1}^{\omega}-M_{2}^{\omega}), (80)
λRω\displaystyle\lambda_{R}^{\omega} =12M1ω.\displaystyle=\frac{1}{2}M_{1}^{\omega}. (81)

λR\lambda_{R}, called as the replicon mode, is the minimum one among the three eigenvalues if the order parameters QQ and qq are positive. Therefore replica symmetry breaking occurs when the replicon mode λR\lambda_{R} becomes negative.

Now let us introduce a matrix Ξ^\hat{\Xi} of size n(n1)×n(n1)n(n-1)\times n(n-1),

where I^m\hat{I}_{m} is identity matrix of size m×mm\times m. Note that the diagonal submatricies P1HQQPP^{-1}H_{QQ}P and P1HqqPP^{-1}H_{qq}P are already diagonalized within themselves. The eigen values λ\lambda of the total Hessian matrix ^\hat{\mathcal{H}} must satisfy the following equation

det[Ξ^λI^n(n1)]=k[(λkQλ)(λkqλ)γ2]=0\det\quantity[\hat{\Xi}-\lambda\hat{I}_{n(n-1)}]=\prod_{k}\quantity[(\lambda_{k}^{Q}-\lambda)(\lambda_{k}^{q}-\lambda)-\gamma^{2}]=0 (82)

which yields the eigen values,

λk±=λkQ+λkq±(λkQλkq)2+4γ22\lambda_{k\pm}=\frac{\lambda_{k}^{Q}+\lambda_{k}^{q}\pm\sqrt{(\lambda_{k}^{Q}-\lambda_{k}^{q})^{2}+4\gamma^{2}}}{2} (83)

for k=1,2,,n(n1)/2k=1,2,...,n(n-1)/2. Note that λk±=λkQ,λkq\lambda_{k\pm}=\lambda_{k}^{Q},\lambda_{k}^{q} if γ=0\gamma=0, which means that each degree of freedom doesn’t affect the stability of another degree of freedom. Thus, in the paramagnetic phase Q=q=0Q=q=0 (see Fig. 3), all eigenvalues are positive,

λL=λA=λR={1(βϵ)2(ϵ>J),1(βJ)2α(ϵ<J).\displaystyle\lambda_{L-}=\lambda_{A-}=\lambda_{R-}=\begin{cases}1-(\beta\epsilon)^{2}~{}(\epsilon>J),\\ \frac{1-(\beta J)^{2}}{\alpha}~{}~{}(\epsilon<J).\end{cases} (84)

The Q=q=0Q=q=0 solution becomes unstable below the lattice glass transition temperature T=ϵT^{*}=\epsilon and the simultaneous glass transition temperature Tc=JT_{\rm c}^{\prime}=J.

B.1 Spin-lattice decoupled case δ=0\delta=0

In the absence of spin-lattice coupling, the two decoupled systems are essentially the same as the spherical SK model. In the pseudo spin glass and pseudo lattice glass phases, the minimum eigenvalues become

λRQ\displaystyle\lambda_{R}^{Q} =M1Q2=0,\displaystyle=\frac{M_{1}^{Q}}{2}=0, (85)
λRq\displaystyle\lambda_{R}^{q} =M1q2=0,\displaystyle=\frac{M_{1}^{q}}{2}=0, (86)

respectively, i. e., marginally stable.

B.2 High rigidity case

For ϵ>J\epsilon>J, the lattice glass transition occurs at the transition temperature T=ϵT^{*}=\epsilon where λk\lambda_{k-} for all kk become 0. below TT^{*}, i.e. in the lattice glass phase with q=1T/ϵq=1-T/\epsilon (see Eq. (32)), λL=λA\lambda_{L-}=\lambda_{A-} becomes positive again,

λL=λA=2q(1q)3>0\lambda_{\rm L-}=\lambda_{\rm A-}=\frac{2q}{(1-q)^{3}}>0 (87)

while the replicon mode remains marginal,

λR=M1q2=0\lambda_{R-}=\frac{M_{1}^{q}}{2}=0 (88)

It means that the system becomes marginally stable against RSB in the lattice glass phase similar to the spherical spin glass model Kosterlitz et al. (1976). There is another glass transition temperature Tc(<T)T_{\rm c}(<T^{*}) Eq. (34)below which Q>0Q>0 solution emerges. One can check that M1Q>0M_{1}^{Q}>0 for Q=0Q=0 above TcT_{\rm c} but vanishes at TcT_{\rm c}. where we can evaluate the replicon mode as,

λR=M1Q2=γ+O(t2)t.\lambda_{R-}=\frac{M_{1}^{Q}}{2}=\gamma+O(t^{2})\sim-t. (89)

where tt is the distance of the temperature to TcT_{\rm c} (see Eq. (38)). Therefore, the replica symmetry should be broken below TcT_{\rm c}.

B.3 Low rigidity case

For ϵ<J\epsilon<J, the simultaneous spin and lattice glass transition occurs at TcT_{\rm c}^{\prime} and λk\lambda_{k-} for all kk becomes 0. Below TcT_{\rm c}^{\prime}, λA=λL\lambda_{A-}=\lambda_{L-} becomes positive again,

λA=λL=2Qα(1Q)3,\lambda_{A-}=\lambda_{L-}=\frac{2Q}{\alpha(1-Q)^{3}}, (90)

and the replicon mode becomes

λR=γ2λRq=4(βJδ)4Q21(βϵ)2t2.\displaystyle\lambda_{R-}=-\frac{\gamma^{2}}{\lambda_{R}^{q}}=-\frac{4(\beta J\delta)^{4}Q^{2}}{1-(\beta\epsilon)^{2}}\sim-t^{{}^{\prime}2}.

Therefore, the continuous replica symmetry breaking takes place at TcT_{\rm c}^{\prime}.

Appendix C Glass susceptibility

The glass susceptibility is given by the inverse matrix of the Hessian matrix ^\hat{\mathcal{H}}, i.e.

χ^=[χQQχQqχqQχqq]=[HQQHQqHqQHqq]1.\hat{\chi}=\matrixquantity[\chi_{QQ}&\chi_{Qq}\\ \chi_{qQ}&\chi_{qq}]=\matrixquantity[H_{QQ}&H_{Qq}\\ H_{qQ}&H_{qq}]^{-1}. (91)

In order to obtain the n(n1)×n(n1)n(n-1)\times n(n-1) components χij\chi_{ij}, let us consider the matrix S^\hat{S} defined such that

where m=n(n1)/2m=n(n-1)/2. Using matrix S^\hat{S}, we find,

χij=lsilsjlλl\chi_{ij}=\sum_{l}\frac{s_{il}s_{jl}}{\lambda_{l}} (92)

where sijs_{ij} is the components of the matrix and λl\lambda_{l} is a short hand notation defined such that

λl={λl+(lm)λ(lm)(m<l).\lambda_{l}=\begin{cases}\lambda_{l+}~{}~{}~{}~{}~{}~{}~{}~{}(l\leq m)\\ \lambda_{(l-m)-}~{}~{}~{}(m<l).\end{cases} (93)

Using Eq. (LABEL:Xi), we notice that S^\hat{S} can be factorized as

S^=[POOP]U^,\hat{S}=\matrixquantity[P&O\\ O&P]\hat{U}, (94)

where U^\hat{U} is a matrix which diagonalize Ξ^\hat{\Xi} (Eq. (LABEL:Xi)), i.e.

U^1Ξ^U^=[λ1+λm+λ1λm],\hat{U}^{-1}\hat{\Xi}\hat{U}=\matrixquantity[\lambda_{1+}&&&&&\\ &\ddots&&&&\\ &&\lambda_{m+}&&&\\ &&&\lambda_{1-}&&\\ &&&&\ddots&\\ &&&&&\lambda_{m-}], (95)

where λkQ\lambda_{k}^{Q} and λkq\lambda_{k}^{q} are given by Eq. (78). Here, each eigenvector is normalized as

i=12muij2=1for all j,\sum_{i=1}^{2m}u_{ij}^{2}=1~{}~{}\text{for all $j$}, (96)

where uiju_{ij} is the component of the matrix U^\hat{U}. If γ=0\gamma=0, i.e. in the paramagnetic phase (Q=q=0Q=q=0) or the pseudo lattice glass phase (Q=0,q>0Q=0,q>0), U^\hat{U} becomes an identity matrix I^2m\hat{I}_{2m}. In the spin-lattice glass phase (Q,q>0Q,q>0), we obtain the elements of the matrix U^\hat{U} as,

ukj={u(k+m)j=0if λj=λk±λkqλjγu(k+m)jif λjλk±u_{kj}=\begin{dcases}u_{(k+m)j}=0~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\text{if }~{}\lambda_{j}=\lambda_{k\pm}\\ -\frac{\lambda_{k}^{q}-\lambda_{j}}{\gamma}u_{(k+m)j}~{}~{}~{}\text{if }~{}\lambda_{j}\neq\lambda_{k\pm}\end{dcases} (97)

for k=1,2,,mk=1,2,...,m.

Now we find the inverse matrix of Ξ^\hat{\Xi} is obtained as,

{Ξ^1}ij=luilujlλl,\quantity{\hat{\Xi}^{-1}}_{ij}=\sum_{l}\frac{u_{il}u_{jl}}{\lambda_{l}}, (98)

Then we obtain the glass susceptibility as,

Note that all components of PP are O(1)O(1). Here, it is convenient to decompose Ξ^1\hat{\Xi}^{-1} as,

Ξ^1=[ΞQQ1ΞQq1ΞqQ1Ξqq1],\hat{\Xi}^{-1}=\matrixquantity[\Xi^{-1}_{QQ}&\Xi^{-1}_{Qq}\\ \Xi^{-1}_{qQ}&\Xi^{-1}_{qq}], (99)

since the components of χαβ(α,β=Q,q)\chi_{\alpha\beta}~{}(\alpha,\beta=Q,q) are associated with only the components of Ξαβ\Xi_{\alpha\beta}.

We show the schematic picture of the temperature dependences of χαβ(α,β=Q,q)\chi_{\alpha\beta}~{}(\alpha,\beta=Q,q) in Fig. 5. In the following, we present the detail of the computation to obtain this.

C.1 Spin-lattice decoupled case δ=0\delta=0

In the absence of the spin-lattice coupling, the matrix U^\hat{U} becomes an identity matrix I^2m\hat{I}_{2m}. Thus, the components of χQQ\chi_{QQ} and χqq\chi_{qq} are proportional to 1/λkQ1/\lambda_{k}^{Q} and 1/λkq1/\lambda_{k}^{q}, respectively, which grow as 1/|t|1/|t^{*}| above TT^{*} and diverge at TT^{*}. In the pseudo glass phase, the replicon modes are always zero, i.e., the glass susceptibilities always diverge.

C.2 High rigidity case

For ϵ/J>1\epsilon/J>1, the system exhibits the lattice glass transition at TT^{*} and the spin glass transition at TcT_{\rm c} which is lower than TT^{*} (see Fig. 3). Slightly above TT^{*} the system is in the paramagnetic phase (Q=q=0Q=q=0) and 1/λkq1/\lambda_{k}^{q} for all kk proportional to 1/|t|1/|t^{*}| as discussed in sec. B.2. Thus we find all components in the sector χqq\chi_{qq} diverge at TT^{*},

{χqq}ijO(|t|1)(T>T),\quantity{\chi_{qq}}_{ij}\sim O(|t^{*}|^{-1})~{}~{}~{}~{}(T>T^{*}), (100)

while χQQ\chi_{QQ} and χQq\chi_{Qq} have no singularity since the spin-lattice coupling does not work as long as γQ=0\gamma\propto Q=0.

Slightly above TcT_{\rm c} the system is in the lattice glass phase (Q=0,q0Q=0,q\neq 0) and 1/λkQ1/\lambda_{k}^{Q} for all kk grows as 1/|t|\propto 1/|t| as t0t\rightarrow 0. The components in the sector χQQ\chi_{QQ} now diverge at TcT_{\rm c},

{χQQ}ijO(|t|1)(T>Tc),\quantity{\chi_{QQ}}_{ij}\sim O(|t|^{-1})~{}~{}~{}~{}(T>T_{\rm c}), (101)

while χQq\chi_{Qq} still has no singularity since Q=0Q=0.

In the spin-lattice glass phase, now γ\gamma becomes finite so that U^\hat{U} is no more an identity matrix. Let us consider the components of Ξ^1\hat{\Xi}^{-1} related to λR±\lambda_{R\pm} because they are relevant to the divergent features of the components of χ^\hat{\chi}. Using Eq. (97) and Eq. (98) and noting (λRqλR±)/γO(1)(\lambda_{R}^{q}-\lambda_{R\pm})/\gamma\sim O(1), we find the magnitude of the components of U^\hat{U} are,

U^=[O(1)O(1)O(1)O(1)].\hat{U}=\matrixquantity[O(1)&O(1)\\ O(1)&O(1)]. (102)

Here each block size is m×mm\times m. Since 1λR|t|1\frac{1}{\lambda_{R-}}\sim|t|^{-1}, the components of Ξ^ij1\hat{\Xi}^{-1}_{ij} become

{ΞQQ1}ij{ΞQq1}ij{Ξqq1}ijO(|t|1).\quantity{\Xi^{-1}_{QQ}}_{ij}\sim\quantity{\Xi^{-1}_{Qq}}_{ij}\sim\quantity{\Xi^{-1}_{qq}}_{ij}\sim O(|t|^{-1}). (103)

Therefore, the components of χ^\hat{\chi} slightly below TcT_{\rm c} are obtained as

{χQQ}ij{χQq}ij{χqq}ijO(|t|1)(T<Tc).\quantity{\chi_{QQ}}_{ij}\sim\quantity{\chi_{Qq}}_{ij}\sim\quantity{\chi_{qq}}_{ij}\sim O(|t|^{-1})~{}~{}~{}~{}(T<T_{\rm c}). (104)

C.3 Low rigidity case

For ϵ/J<1\epsilon/J<1, the system exhibits the simultaneous spin and lattice glass transitions at TcT_{\rm c}^{\prime} (see Fig. 3). Slightly above TcT_{\rm c}^{\prime} the system is in the paramagnetic phase (Q=q=0Q=q=0) and 1/λkQ1/\lambda_{k}^{Q} for all kk proportional to 1/|t|1/|t^{\prime}|. The components in the sector χQQ\chi_{QQ} diverge at TcT_{\rm c}^{\prime},

{χQQ}ijO(|t|1)(T>Tc),\quantity{\chi_{QQ}}_{ij}\sim O(|t^{\prime}|^{-1})~{}~{}~{}~{}(T>T_{\rm c}^{\prime}), (105)

while χQq\chi_{Qq} and χqq\chi_{qq} remain finite.

Similarly to the high rigidity case, in the spin-lattice glass phase, we consider the components of Ξ^1\hat{\Xi}^{-1} related to the only λR±\lambda_{R\pm}. Using Eq. (97) and Eq. (98) and noting (λRqλR)/γO(|t|1)(\lambda_{R}^{q}-\lambda_{R-})/\gamma\sim O(|t^{\prime}|^{-1}) and (λRqλR+)/γO(|t|)(\lambda_{R}^{q}-\lambda_{R+})/\gamma\sim O(|t^{\prime}|), we find the magnitude of the components of U^\hat{U} are evaluated as,

U^=[O(|t|)O(1)O(1)O(|t|)].\hat{U}=\matrixquantity[O(|t^{\prime}|)&O(1)\\ O(1)&O(|t^{\prime}|)]. (106)

Since 1/λR|t|21/\lambda_{R-}\sim|t|^{-2}, the components of Ξ^ij1\hat{\Xi}^{-1}_{ij} become

{ΞQQ1}ij\displaystyle\quantity{\Xi^{-1}_{QQ}}_{ij} O(|t|2),\displaystyle\sim O(|t^{\prime}|^{-2}), (107)
{ΞQq1}ij\displaystyle\quantity{\Xi^{-1}_{Qq}}_{ij} O(|t|1),\displaystyle\sim O(|t^{\prime}|^{-1}), (108)
{Ξqq1}ij\displaystyle\quantity{\Xi^{-1}_{qq}}_{ij} O(1).\displaystyle\sim O(1). (109)

Therefore, the components of χ^\hat{\chi} slightly below TcT_{\rm c} are obtained as

{χQQ}ij\displaystyle\quantity{\chi_{QQ}}_{ij} O(|t|2),\displaystyle\sim O(|t^{\prime}|^{-2}), (110)
{χQq}ij\displaystyle\quantity{\chi_{Qq}}_{ij} O(|t|1),\displaystyle\sim O(|t^{\prime}|^{-1}), (111)
{χqq}ij\displaystyle\quantity{\chi_{qq}}_{ij} O(1).\displaystyle\sim O(1). (112)

The notable fact is that the glass susceptibility of lattice distortions {χqq}ij\quantity{\chi_{qq}}_{ij} does not diverge by approaching the simultaneous glass transition temperature TcT_{\rm c}^{\prime} from both sides due to the pre-factor uiju_{ij} while the lattice glass order parameter qq becomes finite.

Appendix D Energy, Heat capacity

The internal energy and heat capacity within the RS ansatz are given by,

ENc\displaystyle\frac{\expectationvalue{E}}{Nc} =1Nc(βfRS)β\displaystyle=\frac{1}{Nc}\partialderivative{(\beta f^{\rm RS})}{\beta}
=β2{J2α(1+2δ2)+ϵ2\displaystyle=-\frac{\beta}{2}\Biggl{\{}\frac{J^{2}}{\alpha}(1+2\delta^{2})+\epsilon^{2}
[J2α(1+2δ2q)Q2+ϵ2q2]},\displaystyle-\quantity[\frac{J^{2}}{\alpha}(1+2\delta^{2}q)Q^{2}+\epsilon^{2}q^{2}]\Biggr{\}}, (113)
CNc\displaystyle\frac{C}{Nc} =1NcdEdT.\displaystyle=\frac{1}{Nc}\frac{d\expectationvalue{E}}{dT}. (114)

References

  • Berthier and Biroli (2011) Ludovic Berthier and Giulio Biroli, “Theoretical perspective on the glass transition and amorphous materials,” Reviews of Modern Physics 83, 587 (2011).
  • Tarjus (2011) Gilles Tarjus, “An overview of the theories of the glass transition,” Dynamical Heterogeneities in Glasses, Colloids, and Granular Media 150, 39 (2011).
  • Parisi et al. (2020) Giorgio Parisi, Pierfrancesco Urbani,  and Francesco Zamponi, Theory of Simple Glasses: Exact Solutions in Infinite Dimensions (Cambridge University Press, 2020).
  • Mézard et al. (1987) M. Mézard, G. Parisi,  and M. A. Virasoro, Spin glass theory and beyond: An Introduction to the Replica Method and Its Applications, Vol. 9 (World Scientific Publishing Company, 1987).
  • Mydosh (1993) John A Mydosh, Spin glasses: an experimental introduction (Taylor and Francis, 1993).
  • Kawamura and Taniguchi (2015) H. Kawamura and T. Taniguchi, “Spin glasses,” in Handbook of magnetic materials, Vol. 24 (Elsevier, 2015) pp. 1–137.
  • Fichtl et al. (2005) R. Fichtl, V. Tsurkan, P. Lunkenheimer, J. Hemberger, V. Fritsch, H.-A. Krug von Nidda, E.-W. Scheidt,  and A. Loidl, “Orbital freezing and orbital glass state in Fecr2s4\mathrm{F}\mathrm{e}{\mathrm{c}\mathrm{r}}_{2}{\mathrm{s}}_{4},” Phys. Rev. Lett. 94, 027601 (2005).
  • Thygesen et al. (2017) P. M. M. Thygesen, J. A. M. Paddison, R. Zhang, K. A. Beyer, K. W. Chapman, H. Y. Playford, M. G. Tucker, D. A. Keen, M. A. Hayward,  and A. L. Goodwin, “Orbital dimer model for the spin-glass state in y2mo2o7{\mathrm{y}}_{2}{\mathrm{mo}}_{2}{\mathrm{o}}_{7},” Phys. Rev. Lett. 118, 067201 (2017).
  • Mitsumoto et al. (2020) K. Mitsumoto, C. Hotta,  and H. Yoshino, “Spin-orbital glass transition in a model of a frustrated pyrochlore magnet without quenched disorder,” Phys. Rev. Lett. 124, 087201 (2020).
  • Kagawa et al. (2013) F. Kagawa, T. Sato, K. Miyagawa, K. Kanoda, Y. Tokura, K. Kobayashi, R. Kumai,  and Y. Murakami, “Charge-cluster glass in an organic conductor,” Nature Physics 9, 419–422 (2013).
  • Chikazawa et al. (1980) Susumu Chikazawa, Yuochunas YG,  and Yoshihito Miyako, “Nonlinear susceptibility of a spin glass compound (ti1- xvx) 2o3. i,” Journal of the Physical Society of Japan 49, 1276–1282 (1980).
  • Taniguchi et al. (1985) Toshifumi Taniguchi, Yoshihito Miyako,  and Tholence JL, “Nonlinear susceptibilities of the amorphous spin glass fe10ni70p20: critical exponents and gabay-toulouse line,” Journal of the Physical Society of Japan 54, 220–230 (1985).
  • Levy and Ogielski (1986) Laurent P Levy and Andrew T Ogielski, “Nonlinear dynamic susceptibilities at the spin-glass transition of ag: Mn,” Physical review letters 57, 3288 (1986).
  • Edwards and Anderson (1975) S F Edwards and P W Anderson, “Theory of spin glasses,” Journal of Physics F: Metal Physics 5, 965–974 (1975).
  • Sherrington and Kirkpatrick (1975) David Sherrington and Scott Kirkpatrick, “Solvable model of a spin-glass,” Phys. Rev. Lett. 35, 1792–1796 (1975).
  • de Almeida and Thouless (1978) J R L de Almeida and D J Thouless, “Stability of the sherrington-kirkpatrick solution of a spin glass model,” Journal of Physics A: Mathematical and General 11, 983–990 (1978).
  • Parisi (1979) G. Parisi, “Infinite number of order parameters for spin-glasses,” Phys. Rev. Lett. 43, 1754–1756 (1979).
  • Ogielski (1985) Andrew T. Ogielski, “Dynamics of three-dimensional ising spin glasses in thermal equilibrium,” Phys. Rev. B 32, 7384–7398 (1985).
  • Bhatt and Young (1988) R. N. Bhatt and A. P. Young, “Numerical studies of ising spin glasses in two, three, and four dimensions,” Phys. Rev. B 37, 5606–5614 (1988).
  • Kawashima and Young (1996) N. Kawashima and A. P. Young, “Phase transition in the three-dimensional ±j\pm{}j ising spin glass,” Phys. Rev. B 53, R484–R487 (1996).
  • Hukushima and Kawamura (2000) K. Hukushima and H. Kawamura, “Chiral-glass transition and replica symmetry breaking of a three-dimensional heisenberg spin glass,” Phys. Rev. E 61, R1008–R1011 (2000).
  • Lee and Young (2007) L. W. Lee and A. P. Young, “Large-scale monte carlo simulations of the isotropic three-dimensional heisenberg spin glass,” Phys. Rev. B 76, 024405 (2007).
  • Ogawa et al. (2020) Takumi Ogawa, Kazuki Uematsu,  and Hikaru Kawamura, “Monte carlo studies of the spin-chirality decoupling in the three-dimensional heisenberg spin glass,” Phys. Rev. B 101, 014434 (2020).
  • Greedan et al. (1986) J. E. Greedan, M. Sato, X. Yan,  and F. S. Razavi, “Spin-glass-like behavior in y2mo2o7, a concentrated, crystalline system with negligible apparent disorder,” Solid State Communications 59, 895–897 (1986).
  • Gaulin et al. (1992) B. D. Gaulin, J. N. Reimers, T. E. Mason, J. E. Greedan,  and Z. Tun, “Spin freezing in the geometrically frustrated pyrochlore antiferromagnet tb2{\mathrm{tb}}_{2}mo2{\mathrm{mo}}_{2}o7{\mathrm{o}}_{7},” Phys. Rev. Lett. 69, 3244–3247 (1992).
  • Dunsiger et al. (1996) S. R. Dunsiger, R. F. Kiefl, K. H. Chow, B. D. Gaulin, M. J. P. Gingras, J. E. Greedan, A. Keren, K. Kojima, G. M. Luke, W. A. MacFarlane, N. P. Raju, J. E. Sonier, Y. J. Uemura,  and W. D. Wu, “Muon spin relaxation investigation of the spin dynamics of geometrically frustrated antiferromagnets y2{\mathrm{y}}_{2}mo2{\mathrm{mo}}_{2}o7{\mathrm{o}}_{7} and tb2{\mathrm{tb}}_{2}mo2{\mathrm{mo}}_{2}o7{\mathrm{o}}_{7},” Phys. Rev. B 54, 9019–9022 (1996).
  • Gingras et al. (1997) M. J. P. Gingras, C. V. Stager, N. P. Raju, B. D. Gaulin,  and J. E. Greedan, “Static critical behavior of the spin-freezing transition in the geometrically frustrated pyrochlore antiferromagnet y2mo2o7{\mathrm{y}}_{2}{\mathrm{mo}}_{2}{\mathrm{o}}_{7},” Phys. Rev. Lett. 78, 947–950 (1997).
  • Gardner et al. (1999) J. S. Gardner, B. D. Gaulin, S.-H. Lee, C. Broholm, N. P. Raju,  and J. E. Greedan, “Glassy statics and dynamics in the chemically ordered pyrochlore antiferromagnet Y2mo2O7{Y}_{2}{\mathrm{mo}}_{2}{O}_{7},” Phys. Rev. Lett. 83, 211–214 (1999).
  • Hanasaki et al. (2007) N. Hanasaki, K. Watanabe, T. Ohtsuka, I. Kézsmárki, S. Iguchi, S. Miyasaka,  and Y. Tokura, “Nature of the transition between a ferromagnetic metal and a spin-glass insulator in pyrochlore molybdates,” Phys. Rev. Lett. 99, 086401 (2007).
  • Gardner et al. (2010) J. S. Gardner, M. J. P. Gingras,  and J. E. Greedan, “Magnetic pyrochlore oxides,” Rev. Mod. Phys. 82, 53–107 (2010).
  • Reimers et al. (1991) J. N. Reimers, A. J. Berlinsky,  and A.-C. Shi, “Mean-field approach to magnetic ordering in highly frustrated pyrochlores,” Phys. Rev. B 43, 865–878 (1991).
  • Moessner and Chalker (1998) R. Moessner and J. T. Chalker, “Properties of a classical spin liquid: The heisenberg pyrochlore antiferromagnet,” Phys. Rev. Lett. 80, 2929–2932 (1998).
  • Saunders and Chalker (2007) T. E. Saunders and J. T. Chalker, “Spin freezing in geometrically frustrated antiferromagnets with weak disorder,” Phys. Rev. Lett. 98, 157201 (2007).
  • Shinaoka et al. (2011) H. Shinaoka, Y. Tomita,  and Y. Motome, “Spin-glass transition in bond-disordered heisenberg antiferromagnets coupled with local lattice distortions on a pyrochlore lattice,” Phys. Rev. Lett. 107, 047204 (2011).
  • Pauling (1935) L. Pauling, “The structure and entropy of ice and of other crystals with some randomness of atomic arrangement,” Journal of the American Chemical Society 57, 2680–2684 (1935).
  • Bramwell and Harris (2020) S. T. Bramwell and M. J. Harris, “The history of spin ice,” Journal of Physics: Condensed Matter 32, 374010 (2020).
  • Udagawa and Jaubert (2021) Masafumi Udagawa and Ludovic Jaubert, Spin Ice (Springer, 2021).
  • Goodenough (1955) John B. Goodenough, “Theory of the role of covalence in the perovskite-type manganites [La,m(II)]Mno3[\mathrm{La},m(\mathrm{II})]\mathrm{Mn}{\mathrm{o}}_{3},” Phys. Rev. 100, 564–573 (1955).
  • Kanamori (1959) Junjiro Kanamori, “Superexchange interaction and symmetry properties of electron orbitals,” Journal of Physics and Chemistry of Solids 10, 87–98 (1959).
  • Solovyev (2003) I. V. Solovyev, “Effects of crystal structure and on-site coulomb interactions on the electronic and magnetic structure of A2mo2o7{A}_{2}{\mathrm{mo}}_{2}{\mathrm{o}}_{7} (a=Y,(a=\mathrm{Y}, gd, and nd) pyrochlores,” Phys. Rev. B 67, 174406 (2003).
  • Smerald and Jackeli (2019) A. Smerald and G. Jackeli, “Giant magnetoelastic-coupling driven spin-lattice liquid state in molybdate pyrochlores,” Phys. Rev. Lett. 122, 227202 (2019).
  • Charbonneau et al. (2014) Patrick Charbonneau, Jorge Kurchan, Giorgio Parisi, Pierfrancesco Urbani,  and Francesco Zamponi, “Fractal free energy landscapes in structural glasses,” Nature Communications 5, 3725 (2014).
  • Yoshino (2018) H. Yoshino, “Disorder-free spin glass transitions and jamming in exactly solvable mean-field models,” SciPost Phys. 4, 40 (2018).
  • Parisi and Virasoro (1989) Giorgio Parisi and Miguel Angel Virasoro, “On a mechanism for explicit replica symmetry breaking,” Journal de Physique 50, 3317–3329 (1989).
  • Monasson (1995) Rémi Monasson, “Structural glass transition and the entropy of the metastable states,” Physical review letters 75, 2847 (1995).
  • (46) H. Yoshino, in preparation.
  • Mézard and Parisi (1999) Marc Mézard and Giorgio Parisi, “A first-principle computation of the thermodynamics of glasses,” The Journal of Chemical Physics 111, 1076–1095 (1999).
  • Fisher (1991) Daniel S Fisher, “Pathologies of the infinite-n limit of random anisotropy: spin glass transition or local crossover?” Physica A: Statistical Mechanics and its Applications 177, 84–92 (1991).
  • Husimi (1950) Kodi Husimi, “Note on mayers’ theory of cluster integrals,” The Journal of Chemical Physics 18, 682–684 (1950)https://doi.org/10.1063/1.1747725 .
  • ESSAM and FISHER (1970) JOHN W. ESSAM and MICHAEL E. FISHER, “Some basic definitions in graph theory,” Rev. Mod. Phys. 42, 271–288 (1970).
  • Rieger and Kirkpatrick (1992) H. Rieger and T. R. Kirkpatrick, “Disordered p-spin interaction models on husimi trees,” Phys. Rev. B 45, 9772–9777 (1992).
  • Chandra and Doucot (1994) P Chandra and B Doucot, “Spin liquids on the husimi cactus,” Journal of Physics A: Mathematical and General 27, 1541–1556 (1994).
  • Udagawa and Moessner (2019) Masafumi Udagawa and Roderich Moessner, “Spectrum of itinerant fractional excitations in quantum spin ice,” Phys. Rev. Lett. 122, 117201 (2019).
  • Cugliandolo et al. (2020) L. F. Cugliandolo, L. Foini,  and M. Tarzia, “Mean-field phase diagram and spin-glass phase of the dipolar kagome ising antiferromagnet,” Phys. Rev. B 101, 144413 (2020).
  • Mitsumoto et al. (2022) Kota Mitsumoto, Chisa Hotta,  and Hajime Yoshino, “Supercooled jahn-teller ice,” Phys. Rev. Res. 4, 033157 (2022).
  • Kosterlitz et al. (1976) J. M. Kosterlitz, D. J. Thouless,  and Raymund C. Jones, “Spherical model of a spin-glass,” Phys. Rev. Lett. 36, 1217–1220 (1976).