Toric non-archimedean -entropy and thermodynamical structure
Abstract.
We study non-archimedean -entropy for toric variety as a further exploration of K-stability. We show the existence of optimizer of toric non-archimedean -entropy for and the uniqueness for . For the proof of existence, we establish a Rellich type compactness result for convex functions on simple polytope. We also reveal a thermodynamical structure on toric non-archimedean -entropy. This observation allows us to interpret the enigmatic parameter as temperature and non-archimedean -entropy as entropy of an infinite dimensional composite system.
1. Introduction
1.1. Main Results on existence
Canonical metric and K-stability are primary interests in Kähler geometry. There is a series of studies [21, 22, 23, 24, 25] exploring these topics from a unique perspective which motivates a minimization problem we study in this article. Compared to the general case [25], the problem in the toric case we discuss here can be described in a quite simple way by convex functions on convex polytope, for which we only need a few things to prepare. It would be simpler to begin with our convex setup and main existence result of this article, and explain its esoteric background and motivation for the problem afterwards.
1.1.1. Setup
We prepare some terminologies convenient for our arguments. Let be an -dimensional affine space over . A half space of is a subset of the form for some non-zero affine function . The face of a half space is the hyperplane .
We call a subset a polytope if it is a compact subset with non-empty interior and can be expressed as the intersection of finitely many half spaces. A subset is called a face of if it is the intersection of and the face of a half space which contains . A face is again a polytope of a lower dimensional affine subspace . The relative interior of a face is the interior in such . A facet of is a face of which gives a polytope of a codimension one affine subspace. A vertex of is a face of with one element. A simple polytope is a polytope for which every vertex can be written as the intersection of precisely -facets. In this article, we restrict our main interest to simple polytopes. In toric geometry, such (rational) polytopes represent toric orbifolds.
It is well-known that a complete parallel translation invariant measure on (with the usual -algebra) which puts positive finite mass on relatively compact open sets is a constant multiple of the Lebesgue measure under an affine identification . A flat measure on a polytope (resp. a face of a polytope) is the restriction of such translation invariant measure on (resp. such measure on the affine subspace which contains the face as a polytope). We note the topological boundary is the union of facets. A flat measure on is a measure whose restriction to each facet is flat.
For a polytope and , we consider the following spaces:
(1.1) | ||||
(1.2) |
which are independent of the choice of the flat measure on . Here lsc lc is short for lower semi-continuous log convex. Namely, we have for and . For two lsc convex functions , the condition almost everywhere with respect to is equivalent to the condition everywhere. In particular, the boundary value can be recovered from interior values .
We endow with the -topology and call its element (nonequilibrium) state. For , we can assign a state
(1.3) |
For , we have if and only if . Later we explain a Kähler geometric background of the space .
Now let us consider a triple of a simple polytope , a flat measure on and a flat measure on , which we call a simple non-archimedean Hamiltonian system or just a system. When we refer to with general non-simple polytope , we call it a general system.
For , we introduce
(1.4) | ||||
(1.5) |
which we call the (nonequilibrium) entropy and (non-archimedean nonequilibrium) internal -energy, respectively. For , which we call temperature, we further introduce
(1.6) |
We call this functional (non-archimedean nonequilibrium) free -energy. We will see implies and this definition makes the functional continuous along increasing sequence . As we explain later, there is a Kähler geometric background for these functionals.
In section 4, we newly unveil there is a stochastic thermodynamical interpretation of these functionals, which gives us a strong motivation to our terminologies of thermodynamical flavor. We note some references (cf. [3]) in the same field use similar terminologies like entropy and free energy, but these are not directly related to ours. We are mostly interested in K-unstable case, while the reference focuses on K-stable case.
1.1.2. Main results
Now we state our main result. Its output to Kähler geometry is explained later.
Theorem 1.1.
Let be an -dimensional system. Then for every , there exists which minimizes .
Moreover,
-
(1)
(Uniqueness) if , minimizer of is unique.
-
(2)
(Conditional uniqueness) if , there exists a unique minimizer of which maximizes among all minimizers.
-
(3)
(Regularity) if and , any minimizer of is bounded and continuous.
We call the unique minimizer in the above (1) and (2) the -canonical distribution of temperature . In Kähler geometric context, is also called the optimizer or optimal destabilizer of .
This main theorem summarizes Theorem 3.3, Theorem 3.5 and Theorem 3.18. To show the existence, we establish the following compactness result, which is proved in section 2.2. The proof works not only for lsc lc functions but for general non-negative convex functions.
Theorem 1.2.
Let be a sequence of non-negative convex functions with a uniform bound
Then all are in and after taking a subsequence, there exists a unique non-negative lsc convex function in such that converges to in -topology for every .
Remark 1.3.
The above theorem is reminiscent of Rellich’s compactness theorem. It is well-known by Sobolev embedding and trace theorem that we have
while what we prove is
It is natural to expect , but we do not pursue it in this article.
As an interest in Kähler geometry, we also prove the following. It summarizes Theorem 3.11 and Theorem 3.13. We explain terminologies in the next section.
Theorem 1.4.
Let be a toric variety and be the associated system we explain later. Let denote the linear map associated to . For , we have the following.
-
•
The toric variety is toric K-semistable if and only if minimizes .
-
•
If a toric manifold admits a -cscK metric with , then is the -canonical distribution of temperature .
We further prove some results in section 4 motivated by thermodynamics, using our existence result. To appreciate the results, we must recall the origin of -cscK metric and the enigmatic way our parameter is introduced in the theory. We explain these in section 4. Though we present our main result as the existence and uniqueness of minimizer for the free -energy , the observation in section 4 would be much deeper. It should be explored further in future study.
1.2. Why does it matter in Kähler geometry?
1.2.1. Toric variety and polytope
Let be an algebraic torus and be the maximal compact torus. For the character lattice , we put . We denote the dual vector space by and identify it with the Lie algebra of .
Ler be a polarized toric variety of dimension . As usual in toric geometry, we assign a general system as follows. We put
(1.7) |
Here denotes the eigenspace . The lattice defines a unique flat measure on by setting for the box
generated by a -basis . Since for , the flat measure is independent of the choice of . We denote its restriction to by . On the other hand, since the affine subspace spanned by a facet of is a rational subspace, the intersection gives a lattice spanning . Then we similarly define a unique flat measure on each facet compatible with the lattice and a flat measure on .
The flat measure on represents the anti-canonical divisor . In general, flat signed measures on are in one to one correspondence with -invariant -Weil divisors: each facet represents a -invariant prime divisor and a flat signed measure on represents the divisor . For the flat signed measure representing a -invariant -Weil divisor on , we put
(1.8) |
which represents the -divisor . From thermodynamical perspective, the measure plays the role of Hamiltonian of the system and its variation can be interpreted as thermodynamical work.
A triple of a toric variety , a -invariant -divisor with coefficient and a -equivariant polarization is called a polarized log toric variety. By the above construction, we can assign a general system for a polarized log toric variety.
The associated polytope is known to be simple if and only if has at most orbifold singularity. See [13, Theorem 3.1.19].
1.2.2. -cscK metric
Let be a compact Kähler manifold and be a Kähler class. We consider a pair of a Kähler metric (form) in and a smooth real valued function on . For , a pair is called -cscK metric if it satisfies the following two conditions:
-
(1)
The complex vector field is holomorphic.
-
(2)
The -scalar curvature is constant.
When the imaginary part is specified, we call the metric -cscK metric. Since , is the Hamiltonian potential of . The imaginary part preserves the metric , so it is a Killing vector field. This implies we can always take a closed torus acting on so that . We note there is a subtle but an important difference between -cscK metric and -cscK metric: when we refer to -cscK metric, the vector is not fixed.
Remark 1.5.
The above definition of -cscK metric is equivalent to what we call -cscK metric in [23].
The notion unifies two frameworks of canonical metrics: constant scalar curvature Kähler (cscK) metric and Kähler–Ricci soliton. Indeed, when is a Fano manifold (i.e. the anti-canonical class is positive), a Kähler metric in the anti-canonical Kähler class () is -cscK metric if and only if it is Kähler–Ricci soliton: . Kähler–Ricci soliton is defined only for Fano manifolds due to the a priori constraint
while -cscK metric makes sense for general polarized manifold.
A fundamental question on -cscK metric is the existence and uniqueness modulo translation by automorphisms. When we fix , the theory of -cscK metric is enclosed in the theory of weighted cscK metric. It is then proved by [28] that -cscK metric is unique modulo translation by aumorphisms preserving . As for the uniqueness of -cscK metric, we must show the uniqueness of (modulo translation) which admits a -cscK metric. We confirm this uniqueness for toric manifolds with .
Theorem 1.6.
Assume . On a toric manifold, -cscK metrics are unique modulo the action of automorphism group.
On the other hand, fixing , the existence of -cscK metric implies an algebro-geometric condition of called K-stability.
1.2.3. K-stability
The notion can be defined for general polarized scheme, but here for simplicity we only explain a toric version of the notion. It can be simply described in terms of convex functions.
Definition 1.7 (Toric K-stability).
Let be a polarized log toric variety and be the associated general system.
Let be a rational piecewise affine convex function on . Namely, is of the form for finitely many rational affine function on . Then for and , we define the -Futaki invariant by
(1.9) |
using the linear map . Here we put
We call toric K-semistable if for every rational piecewise affine function and toric K-polystable if moreover for non-affine .
Remark 1.8.
When , the invariant is independent of :
(1.10) |
It is called the Donaldson–Futaki invariant.
We note the -Futaki invariant is well-defined for general integrable convex function on general system . For a set of convex functions, we call toric K-semistable (resp. K-polystable) with respect to if for every (resp. if further for non-affine ). We call K-unstable if it is not K-semistable. We note even if is K-unstable, it may be K-semistable with respect to some and non-trivial .
Remark 1.9.
We define K-stability by rational piecewise affine convex functions because these convex functions correspond to toric test configurations. A (compactified ample) toric test configuration is a -equivariant flat family of polarized schemes over which is endowed with a -equivariant isomorphism from to over . Here acts on by and is an ample -line bundle over . We may assume is normal for our interest. Since is an -dimensional polarized toric variety, we can consider the associated polytope . We can show the polytope can be written as using a convex function on . Since is a rational polytope, is rational piecewise affine convex function.
For a polarized variety , we denote by the set of all (not necessarily -equivariant) test configurations of . Similarly for a rational polytope , we denote by the set of all rational piecewise affine convex functions on . Convex functions in represent -equivariant test configurations, which form a proper subset . We note every lower semi-continuous convex function has an increasing net which pointwisely converges to , thanks to Fenchel–Moreau theorem.
By the result of [1] (see also [27, 23]), it is known that a polarized manifold is K-polystable if it admits a -cscK metric. When , the case of cscK metric, it is also known that a uniform version of toric K-polystability implies the existence of cscK metric for toric manifold by [19, 10, 29]. As far as the author knows, this implication for general is still unkown. It is natural to expect the same method for would work. For general polarized manifold, it is still a conjecture even for cscK metric, which is known as YTD conjecture, but recently there are great progress (cf. [9, 29, 30, 7]). We do not pursue this direction in this article as we are rather interested in K-instability.
1.2.4. Perelman entropy
Now we return to -cscK metric. Originally, -cscK metric is introduced in [22] based on a moment map picture on Kähler–Ricci soliton observed in [21]. The moment map picture heuristically explains why the existence of -cscK metric is related to K-stability and moduli problem, so this perspective is important in stability aspect. We review this later in order to explain how the parameter appears.
On the other hand, it turns out later in [24] that -cscK metric is also characterized by Perelman’s entropy. This perspective illuminates “instability aspect” of the theory of -cscK metric and K-stability, which is the main interest of this article.
Now we introduce Perelman’s functionals. For a Kähler metric and a real valued function normalized as for , we put
(1.11) | ||||
(1.12) |
which we call Perelman’s -entropy and -entropy, respectively. We note the above convention differs from Perelman’s original one by sign and multiplication of constants, which is off course not essential. Perelman considered general Riemannian metric for , but we restrict our interest to Kähler metric in the context of -cscK metric, which is in turn essential for our variational result.
It is proved in [24] that we can characterize -cscK metrics as the critical points of . We note here we must restrict metrics to the space of Kähler metrics in . Moreover, for , we can also characterize of -cscK metrics as minimizers of . This characterization is nice as our differential geometric interest is in the metric and not so much in the function or the vector field .
Recall we fix vector to define K-stability, while now we have a vector-free characterization of -cscK metric. To match up, it is better to have an algebro-geometric criterion for which tells if could be K-semistable. However, in general, there may be no such vector. In such case, we are interested in finding “most destabilizing ” in a suitable sense. It is revealed in [25] that the nature of these two questions are actually the same. We can formalize and solve both questions in terms of toric non-archimedean -entropy, which is essentially the functional in our main theorem.
1.2.5. Toric non-archimedean -entropy
Let be the general system associated to a polarized toric variety . For a lsc convex function , we put
(1.13) | ||||
(1.14) |
For , we consider
(1.15) |
and call it the toric non-archimedean -entropy.
These are essentially the functionals we already introduced:
The differential of toric non-archimedean -entropy gives the minus of -Futaki invariant:
(1.16) |
This implies that if is maximized at a vector , then is toric K-semistable.
It is proved in [24] that there is an inequality between and for :
(1.17) |
Moreover if and there exists a -cscK metric , then the above equality is achieved by . This implies if there exists a -cscK metric for , then maximizes , which in particular gives another proof for the K-semistability of -cscK manifold with . It is conjectured the equality holds for even when there is no -cscK metric.
The non-archimedean -entropy is essentially introduced in [23] for and more generally for a test configuration of a non-toric polarized variety . The paper [25] introduces a completion of the space of test configuraitons in terms of non-archimedean pluripotential theory developed in [5]. The space consists of functions on the Berkovich analytification of , which is the reason for the word “non-archimedean”. In the toric setup, the -invariant part of the space corresponds to the intersection . It takes considerable pages to extend the non-archimedean -entropy to as an upper semi-continuous functional, which in our toric case turns extremely easy to show. In the toric case, is continuous along increasing sequence by the monotone convergence theorem, so we may replace in (1.17) with .
Theorem 1.1 confirms that there always exists a maximizer of , even when the toric variety is K-unstable. If is a bounded convex function, we can assign a filtration on the graded ring as in [25]. If further the filtration is finitely generated, which is the case for instance when , it produces an algebro-geometric degeneration of called polyhedral configuration in [25]. Then we can show that the central fibre of such degeneration is K-semistable with respect to the vector generated by the degeneration. See [25] for further detail.
1.2.6. Other works in the same spirit
One ultimate goal of our study is to construct a degeneration of K-unstable polarized variety to a K-semistable “space” (see Remark 3.17), which is often referred to as optimal degeneration. The non-archimedean -entropy is not the unique quantity measuring K-instability. We briefly review other works on K-instability as reference.
Optimal degeneration in the context of Kähler–Ricci soliton
The first pioneering work on optimal degeneration in the context of Kähler–Ricci soliton on Fano manifold would be Chen–Wang’s work [12]. They constructed a metric geometric degeneration along Kähler–Ricci flow. Chen–Sun–Wang [11] gives an algebro-geometric description of the degeneration, in which the idea of using filtration and two step degeneration is utilized. Dervan–Székelyhidi [14] introduced a quantity called -invariant and showed Chen–Sun–Wang’s filtration minimizes the -invariant. Han–Li [18] pursued an algebro-geometric aspect of -invariant and proved the uniqueness of Chen–Sun–Wang’s degeneration. Finally, Blum–Liu–Xu–Zhuang [4] constructed Chen–Sun–Wang’s degeneration in a purely algebro-geometric way, which allows them to extend the result to Fano variety with singularity. It is known by Wang–Zhu’s work [34] that every toric Fano manifold admits Kähler–Ricci soliton. This result can also be regarded as the explicit description of optimizer in toric case: the optimizer for -invariant is given by the vector associated to Kähler–Ricci soliton. In [25], it turns out that the non-archimedean -entropy also characterizes Chen–Sun–Wang’s degeneration.
Optimal degeneration in the context of extremal metric
Optimal degeneration problem of polarized variety would be firstly considered by Donaldson [16] in the context of extremal metric. The quantity for K-instability in this context is called normalized Donaldson–Futaki invariant. In toric case, Székelyhidi [33] proved the existence of optimal destabilizing -integrable convex function. Xia [35] pursued the problem from pluripotential theoretic perspective and proved the existence of an optimal destabilizing geodesic ray in a suitable completion of the space of Kähler metrics. It turns out by Li’s analysis [29] that the geodesic ray can be interpreted as a non-archimedean metric. Compared to Kähler–Ricci soliton, much less is known on the regularity of optimal destabilizer. We speculate this context can be treated as the limit of our framework.
Optimal degeneration in the context of Mabuchi soliton:
Optimal destabilization problem for Mabuchi soliton would be firstly considered by Hisamoto [20]. The quantity for K-instability in this context is called normalized Ding invariant. In toric case, Yao [36] constructed an optimal destabilizing convex function for normalized Ding invariant. He also proved the crucial regularity result: the optimizer is piecewise affine. This is much better than the current knowledge on the optimizer for normalized Donaldson–Futaki invariant and non-archimedean -entropy.
Still many other works…
There are still many other works on optimal degeneration. Among all, the structure of the above three frameworks is closest to that of our framework. For instance, we have Donaldson type inequality (cf. [16, 14, 20, 35, 24]) which can be interpreted as minimax theorem on some functional (cf. [24, 25]). Thus here we do not give further comments on those other works in the context of -invariant and the normalized volume.
Acknowledgments
This work is supported by RIKEN iTHEMS Program. We used Wolfram Mathematica to create some graphs.
2. Convex analysis on simple polytope
2.1. Local estimates
Here we prepare some estimates we use in the next section.
2.1.1. Mean value estimate
Proposition 2.1.
Let be a convex function. Then we have
for every . Here denotes the -dimensional Lebesgue measure.
Proof.
The claim is trivial if or for some . By supporting hyperplane theorem, for , we can take an affine function so that on and . If we put
we have on . Since is non-negative, we have , so that
Thus we get
Since , we get the desired inequality. ∎
We put
(2.1) |
Corollary 2.2.
Let be a convex function. For and for every , we have
Proof.
For , we have , so that
by the above proposition. Here we note when . ∎
2.1.2. Boundary mean value estimate
For , we put
(2.2) |
and consider the -dimensional Lebesgue measure .
Proposition 2.3.
Suppose . Let be a convex function. Then for , we have
for every . Here in runs over , excluding .
Proof.
The claim is trivial on . We assume . For
we have
Then by convexity,
When , we have , so that the claim follows by Corollary 2.2 applied to on . ∎
Corollary 2.4.
Suppose . There exists a function such that for every and
for every convex function and .
Proof.
By the above proposition, we have
Here in runs over except for . We compute
It follows that
We can easily check the function
enjoys the desired convexity and integrability. ∎
Proposition 2.5.
Suppose . Let be a convex function. Then we have
Proof.
For , we put
(2.3) |
Then we have
Now we consider
where in runs over except for and we put
By
we get a map from onto , which gives a homeomorphism outside the zero set . The Lebesgue measure is transformed into
A convex function on gives a convex function on for each . We also have
Since is convex, we compute
Here again in runs over except for . Since
on , we get
Taking the sum of this, we obtain the desired estimate. ∎
2.2. Rellich and Poincare type estimates
Here we establish a crucial compactness result.
2.2.1. Rellich type estimate
Proposition 2.6.
Let be a polytope. Using a flat measure , we put
Then for every . Moreover, for any non-negative convex function and , we have
Proof.
Let be the set of half spaces with . We identify with a sphere as topological space. Then the map is continuous, so that the infimum of is attained by some , so that we have on and on . The inequality follows by the same argument as in the proof of Proposition 2.1. ∎
Theorem 2.7.
Let be an -dimensional system. There exists a lsc convex function such that for every and
for every non-negative lsc convex function and .
Moreover, we can take such so that it is finite valued and continuous on and for each lying in the relative interior of a codimension face, there exists a neighbourhood such that for every .
Proof.
We put
This is clearly a lsc convex function on .
We firstly show that is finite valued and continuous on . For lying in the relative interior of a facet , we have
so that
Thus is finite for . For , we can take and so that . Then by the convexity, we have .
Since is a polytope, the upper semi-continuity of follows by the convexity. Indeed, as for the interior , it is well-known that any convex function is continuous on open set. The upper semi-continuity around can be seen as follows. Take a local neigbourhood of a point so that it is affine isomorphic to . Then for any convex function on we compute
When , we have
by the continuity of on and on . Thus we get
(2.4) |
which shows the upper semi-continuity.
Now it follows that for any compact set , there exists a constant such that . In particular, is for any .
It suffices to show the integrability around . We can reduce this task to show for every point there exists a neighbourhood of in and a function such that for every and .
For any boundary point lying in the relative interior of dimension face, we can take an affine map so that
-
•
,
-
•
The measure on is transformed into the Lebesgue measure on .
-
•
There exists such that gives a homeomorphism from onto an open neighbourhood of in .
-
•
For each , there exists a facet containing such that gives a homeomorphism from onto an open neighbourhood of in .
Since is affine, the measure on is transformed into on for some constant .
For a convex function on , we have
Then by Corollary 2.4, we have a function such that for every and
on for every convex . Now we put and , then we get the desired property. ∎
Corollary 2.8.
Let be a sequence of non-negative convex functions with a uniform bound
After taking a subsequence, there exists a unique lower semi-continuous convex function such that converges to in -topology for every . Moreover the convergence is uniform on each compact set .
Proof.
If we put
we have by the supporting hyperplane theorem and . Thus we may assume are lower semi-continuous. By the above theorem, we have
It follows by the continuity of on that for any compact set , is uniformly bounded. Then by a general argument, for any compact set , we get a uniform Lipschitz bound . Then by Arzelà–Ascoli theorem, we can find a subsequence and a function on so that converges uniformly to on every compact set . Since , we have in for every by the dominated convergence theorem.
Now we consider the lsc envelope
on . By the supporting hyperplane theorem, coincides with on . Then since is a zero set, we have in for every .
Since any convex function is continuous on open set, the -convergence characterizes on . Since restricted to a segment is automatically upper semi-continuous, the lowe semi-continuity of implies that is uniquely determined by , which shows the uniqueness. ∎
2.2.2. Poincare type estimate
Theorem 2.9.
Let be an -dimensional system. For every , there exists a constant such that
for every non-negative convex function .
Proof.
We may assume is lower semi-continuous. Since the measure is finite, it suffices to show the claim for . For each , we take an affine chart as in the above proof and cover by open sets and for some . Take a finite subcover so that it still covers . On , we have
By Proposition 2.5, we have
Here we put
Take the sum for , we get the result. ∎
Remark 2.10.
By inductive argument, we obtain the following: Let be an -dimensional simple polytope. If we endow a flat measure on , then for , there exists a constant such that
for every non-negative convex function . It would be interesting to find application of this fact to the higher regularity of minimizer of .
We use the following log Sobolev type estimate in the next section.
Corollary 2.11.
For , we have
for every non-negative convex function with .
Proof.
Since is a probability measure, we compute
by Jensen’s inequality on . The claim follows by applying the above theorem. ∎
3. Optimizers for non-archimedean -entropy
3.1. Existence and Uniqueness of optimizers
3.1.1. Existence of minimizer of
Here we show the existence part of main theorem. Let us firstly observe the lower semi-continuity of .
Proposition 3.1.
Suppose converges to in -topology, then
(3.1) |
If moreover the -norm of is bounded for some , then we have
(3.2) |
Proof.
We firstly note the following fact: for a sequence of convex functions on , if the restriction pointwisely converges to a convex function on , then for the lower semi-continuous extension on (see the proof of Corollary 2.8), we have
(3.3) |
for every point . This can be seen as follows. Take a point . For , we have
For , we have , so that we get
by the pointwise convergence on . As is lower semi-continuous, it is continuous on the segment . Thus by taking the limit , we get the claim.
Now since , the inequality (3.1) is a consequence of Fatou’s lemma. Only the pointwise convergence on is important for this.
On the other hand, by mean value theorem on , we compute
for and , by taking suitable . It follows by Cauchy–Schwarz theorem that
with a constant which depends only on and a uniform bound on -norm of . This proves the claim for . ∎
We prepare the following uniform estimate.
Lemma 3.2.
For any , there exists a constant depending only on such that the following holds: if a non-negative convex function with satisfies
then we have
Proof.
When , we can choose since we have by Jensen’s inequality on .
Now we prove our main existence theorem.
Theorem 3.3.
Let be an -dimensional system. For every , there exists a convex function which minimizes on .
Proof.
We recall
By the above lemma, for any constant , is uniformly bounded on the subset
Take a sequence so that . Thanks to the above uniform bound and the compactness in Corollary 2.8, after taking a subsequence we may assume converges in -topology () to a lsc log convex function . (Note log concavity is preserved by pointwise convergence. ) By the -convergence, we have , so that and hence . By the lower semi-continuity of with respect to -topology, we get , which shows the existence of minimizer. Finally, since by the above uniform bound, we conclude by Theorem 2.9. ∎
3.1.2. Uniqueness for
While has no convexity along the linear path in , has convexity along the linear path in , which corresponds to the ‘log linear exp’ path in .
Proposition 3.4.
For and , we put
Then we have and
-
•
For , is affine on .
-
•
For , is strictly convex when .
-
•
For , is strictly concave when .
Proof.
To see the log convexity of , we consider the space with the probability measure
and we compute
by the Hölder inequality, for every and .
Since , we obviously have
(3.4) |
On the other hand, since , the strict convexity of implies
(3.5) |
with the equality iff . This proves the claim. ∎
Now we obtain the following uniqueness.
Theorem 3.5.
For every , there exists a unique which minimizes , while for , there exists a unique which satisfies the following
-
•
minimizes and
-
•
.
Proof.
For , if we have two minimizers of , we get by the strict convexity, which contradicts to the assumption on .
Similarly, for , if satisfy the above two conditions, then satisfy the first condition by affinity of , while we have , which is a contradiction. The existence of satisfying the two conditions is another application of our compactness: by Corollary 2.8 and the lower semi-continuity of , the set
is compact in -topology, hence we can find a minimizer of which maximizes among all minimizers of thanks to the continuity of with respect to -topology. ∎
Let us adopt terminologies from physics for later discussion.
Definition 3.6 (-canonical distribution, ground states, optimizer).
We call the above the -canonical distribution of temperature . For , we call minimizers of ground states. We also call the optimizer for .
The following is a natural question.
Question 3.7.
Are all ground states are -canonical?
3.1.3. Family over
Let be the -canonical distribution for as in Theorem 3.5. We firstly observe the following generality.
Proposition 3.8.
Let be a general system. The function is increasing and concave, and the functions are increasing on . Moreover, we have
Proof.
The infimum of concave functions is concave, so that
is concave. Since , each is increasing, so that is increasing.
To see that is increasing, we compute
which shows
For , we compute
which shows the monotonicity for .
Now we note we have the following uniform bounds:
It then follows by monotonicity that the limits of these exist as tends to .
Since we have
(3.6) | ||||
we get
for every . It follows that
∎
Using the compactness we established, we further obtain the following.
Theorem 3.9.
Let be a system. The map
is continuous with respect to -topology for every with continuous . Furthermore, we have
Proof.
As we see in the above proof, we have , so by Corollary 2.8 the family is relatively compact in -topology for .
Take a convergent sequence and a subsequence so that in -topology to some . We show the limit is for and for , independent of the choice of subsequence.
Assume . By the lower semi-continuity, we get
For any , we compute
which shows the limit is a minimizer of . This shows for by the uniqueness of minimizer.
Suppose and . Then we have by the uniqueness of -canonical distribution. Take large so that for , which contradicts to the fact that is increasing. Thus we have when .
Suppose . By the above proposition, we have
This implies as is the maximum value of attained only at . Thus we proved the continuity of the family on .
Now it follows that
On the other hand, by (3.6), we have
which shows
Putting these together, we obtain
Since is concave, it is continuous. Since is continuous with respect to -topology, is continuous. Finally since , is continuous on , while the continuity at is already proved. ∎
As a sophisticated version of the extremal limit observation in [22, 24], we speculate the optimal destabilizer for normalized Donaldson–Futaki invariant appears in the rescaled limit.
Conjecture 3.10.
As , the convex function converges at least in -topology to an lsc convex function characterized as follows: minimizes the following normalized Donaldson–Futaki invariant among all -integrable convex functions on :
where .
3.2. Consequences on -cscK metric and K-stability
3.2.1. Uniqueness of -cscK metrics on toric manifolds for
Now we see -cscK metric determines optimizer. Note if we have a -cscK metric on toric manifold , we can take so that is a -cscK metric with .
Theorem 3.11.
Assume . If a toric manifold admits a -cscK metric with , then the linear map is the optimizer of . Here the metric is not necessarily a priori -invariant.
Proof.
Suppose we have a -cscK metric with . It is proved in [24] (see also [25]) that is a maximizer of for . Indeed, we have
and for we have
This proves the claim for .
To see that is the optimizer for , we perturb the -cscK metric to -cscK metrics for as in the construction in [22], which is just an application of implicit function theorem. We obviously have as by the construction. We already know are the optimizers. By Theorem 3.9, we conclude is also the optimizer. ∎
By the uniqueness of optimizers of for , we conclude the vectors associated to -cscK metrics are conjugate. Then thanks to the result [28] on the uniqueness of -cscK metric for fixed , we obtain Theorem 1.6. Here we mention it again.
Corollary 3.12.
Assume . On a toric manifold, -cscK metrics are unique modulo the action of automorphism group.
3.2.2. Toric K-semistability is -entropy maximization
Theorem 3.13.
A toric variety is toric K-semistable for if and only if the linear map maximizes .
Proof.
Thanks to (1.16), is K-semistable if maximizes .
Suppose is K-semistable for . Take a continuous convex function . For the log linear exp path
we compute
Since is a continuous function, its absolute value is bounded from above by a uniform constant, so that we can compute
By Fenchel–Moreau theorem, we can take an increasing sequence so that it converges to pointwiesly. By monotone convergence theorem, we get
Therefore, the K-semistability implies
Since is concave, we get for continuous .
For general , if , we obviously have . If , take an increasing sequence converging to pointwisely. Then by the monotone convergence theorem we get . This shows for general . ∎
In particular for , a toric variety can be K-semistable at most one . The following remark further implies a similar conclusion for under K-polystability assumption.
Remark 3.14.
For , we can also compute directly
for with .
Corollary 3.15.
For , if a toric variety is toric K-polystable with respect to , then ground states are unique and hence is the optimizer for .
Proof.
Suppose for . Since , we have . By the above remark, we have , which implies by our K-polystability assumption. ∎
3.3. Ground state in dimension 2 is bounded
As we have noted in Remark 1.9, a rational piecewise affine convex function on toric polytope realizes an algebro-geometric degeneration of toric variety: . In comparison with [4, 36], the best regularity we can expect for maximizer of would be piecewise affinity. As noted in [21, 4] in the context of Kähler–Ricci soliton, this conjecture is deeply related to moduli theory for polarized varieties. To appreciate its extreme importance, we refer to this conjecture as crystal conjecture.
Conjecture 3.16 (-entropy maximizer is crystalline).
Let be a system. Then for each , every maximizer of is piecewise affine.
Remark 3.17.
For a rational piecewise affine function , the base polytope of an affine component of (the image of a facet of contained in the roof along the projection ) is rational and realizes an irreducible component of the central fibre of the associated test configuration.
On the other hand, for an non-rational piecewise affine function , the base polytope of an affine component of might be not rational. (In this case, the filtration on the graded ring associated to as in [25] would be not finitely generated. ) This implicates the category of algebraic scheme might be insufficient to realize in a geometric way. The author speculates there would be a nice category of spaces extending that of algebraic schemes in which we can realize optimizer of in a geometric way. We do not pursue this further in this article, but we just refer to [26] as a hint for “extended toric geometry” on non-rational polytope/fan.
We have proved in the existence theorem that the exponential of a maximizer is -intergable. To approach crystal conjecture, we confirm the continuity of maximizer for in dimension 2.
Theorem 3.18.
Let be a -dimensional system. Then every maximizer of is bounded and continuous.
We firstly observe every maximizer of is equal to
(3.7) |
Indeed, recall firstly
for any lsc convex function by Fenchel–Moreau theorem. Since we obviously have , we get
and hence for any lsc convex function . If , we have and for small on a small ball . Then we compute
so that does not maximize .
In what follows, we consider convex function of the form . Take a vertex . For , we consider
(3.8) |
We will show if then we can find large and small so that for . For the computation, we observe a concrete description of .
Lemma 3.19.
Suppose as above. Let be the edges (dimension one face) containing . Assume for each the right differential diverges to as , which is the case for instance when .
Then for sufficiently colse to , there exists a unique affine function on such that , on and for some for each . Moreover, we have for any affine function on satisfying and .
Proof.
Observe for each , there exists a unique affine function on the edge such that , and for some point . Thanks to , we have for any other affine function on satisfying and .
Let be the affine function whose graph represents the plane spanned by the graphs of and . We clearly have . Let be the minimum of . By our assumption on , we have . Moreover, for with , we have
so that taking sufficiently close to , we may assume . (Compare the subsequent lemma. It implies under our assumption, so that as . ) This implies , so that we get . It follows that .
Let be an affine function on satisfying and , which in particular implies . By the above remark, we have , hence on . ∎
Since is closed, we have a unique point which is closest to in . For this point, we have the following.
Lemma 3.20.
Under the assumption in the above lemma, we have monotonically as .
Proof.
This is visually clear, but we show it logically in order to check how our assumption on works. For , we have
which implies the length of the segment is shorter than .
By the monotonicity, we have a limit point . Suppose . Then taking the limit of
we get
while
by convexity. This implies is affine on the segment , which contradicts to our assumption. ∎
Lemma 3.21.
Under the above assumption, we have
Proof.
For , take an affine function so that on and . The two segments and crosses at a point . Since and , we have . Then by
we get . By the lower semi-continuity, we have
Take an affine function on so that and . For , let be the intersection point with the line spanned by the segment . Since and , we have . It follows that . On the other hand, we have on as and . On the other hand, we obviously have . Thus we get for . ∎
Now we show the continuity. By [17] (or by an extension of the argument of (2.4)), every bounded convex function on a convex polytope is known to be upper semi-continuous. Since we assume the lower semi-continuity for , it suffices to show the boundedness.
Proof of Theorem 3.18.
Let be a maximizer of . We may normalize so that . Suppose is unbounded, then we can find a vertex with by an easy argument. For sufficiently large , we take and as above. We note
as , while
is constant.
We put
We consider a map from to given by
The measure is transformed to on and the measure is transformed to on .
Now we compute
Assume , we have for every , then we further compute
where the third inequality follows by
Fix large so that
Then for , we compute
which shows
for sufficiently small . This contradicts to the assumption that is a maximizer of . The case is similar. Thus must be bounded.
∎
4. Thermodynamical structure
To appreciate the results of this section, let us briefly recall how the parameter appeared in the beginning [22] of -cscK metric. The notion of -cscK metric was introduced in [22] based on the moment map picture on Kähler–Ricci soliton observed in [21]. Given a symplectic manifold with a Hamiltonian vector field , the space of -invariant almost complex structures possesses a symplectic structure and a moment map with respect to the action of -compatible Hamiltonian diffeomorphism group . These are dependent on . The Hamiltonian potential of defines an element of fixed by the coadjoint action, so that gives another moment map for the same symplectic structure . If we choose and (), the symplectic reduction can be interpreted as the moduli space of Kähler–Ricci soliton (cf. [21]).
At this point, there are at least two points of view to extend the theory of Kähler–Ricci soliton to general polarized manifold:
-
(1)
Determine the “best” for a given polarized manifold so that all things go well like Kähler–Ricci soliton. At least for a Fano manifold , it would be .
-
(2)
Regard as a free parameter in the theory and construct our theory for any as much as possible.
Nakagawa [31] takes the first stance, but it turns out in [22] that the latter perspective is more fruitful than the first one: the product -cscK metrics behaves well for the same , extremal metric appears in the limit and so on. Moreover, contrary to the theory on Kähler–Ricci soliton, the general theory works well for rather than . In any case, the parameter was just introduced at first as a free parameter which enriches the theory. Its geometric role was unclear.
Here we show our theory carries a thermodynamical sturcture for . It gives us a better geometric intuition on our enigmatic parameter and also implicates the geometry of -cscK metric and K-stability is of infinite dimensional nature: it is the geometry of a composition of the space of our interest and an infinite dimensional space working as heat bath. This is reminiscent of Perelman’s statistical mechanical heuristic argument [32] on his -entropy.
4.1. Optimizer for product
Mutual interaction of thermodynamical systems is a key concept in thermodynamics. Mathematically, it is described as a process (or its terminal state) on the product (or tensor product, in quantum setup) of two systems . Let us begin with the simplest case: mutual interaction of isothermal systems.
For two systems , , we consider the following composite system
where denotes the product measure and denotes the measure on given by
We call subsystems of this composite system.
Theorem 4.1.
For , the -canonical distribution of temperature on the composyte system is the product
of the -canonical distributions of the same temperature on the subsystems , respectively.
Proof.
For , we define a function on by
(4.1) | ||||
(4.2) |
These are non-zero log convex functions by Hölder’s inequality, lower semi-continuous by Fatou’s lemma and by Fubini’s theorem, so that and . If is of the form for and , we have and .
Now we compute
which shows
(4.3) |
This implies if minimizes , then minimizes for each .
On the other hand, we have
(4.4) |
with the equality only when . This is a paraphrase of the non-negativity of mutual information known in information theory. We can prove this as follows. Put , and . Then we have
by Jensen’s inequality, where the equality holds if and only if . For , and , a simple calculation shows
The consequence is the inequality (4.4).
Now for , let be the -canonical distributions of temperature . Then for any , we have
so that minimizes , which shows the claim for by the uniqueness of minimizer. On the other hand, if minimizes , then minimizes , so that we compute
Now the uniqueness of -canonical distribution completes the proof. ∎
4.2. Equilibrium and isothermality
Equilibrium and isothermality are fundamental notions in thermodynamics. To compare our argument, let us briefly recall how the notion of equilibrium is formalized in stochastic thermodynamics. In stochastic thermodynamics, a measure space (in a simple setup, a finite set with the counting measure) together with a function called Hamiltonian is interpreted as a thermodynamical system, which serves as the space of possible (deterministic) states. A probability measure on is interpreted as nonequilibrium state. Nonequilibrium entropy is defined by the relative entropy with respect to and its nonequilibrium internal energy is defined by . Equilibrium is described as entropy maximizer on a level set of internal energy.
By the method of Lagrange multiplier, we can describe equilibrium as critical point of for a proper choice of depending on energy level. The critical state can be written explicitly as
which is well known as canonical distribution. We can apply a similar argument to non-archimedean -entropy. We note maximizer of non-archimedean -entropy has no simple explicit description, which is different from stochastic thermodynamics.
4.2.1. Equilibrium
Let be a system. For , we consider the energy level set:
(4.5) |
We call a state equilibrium of internal -energy if it maximizes the entropy on the level set . For instance, the trivial state is equilibrium of internal -energy .
Our main interest is in the equilibrium of internal -energy in the interval
(4.6) |
for which we can show the existence. We should note even though we have proved the compactness of the sublevel set , we cannot conclude the compactness of the level set as is only lower semi-continuous, so the existence of equilibrium is not a direct consequence of our compactness result. We make use of the continuity of the family of -canonical distributions , which is a consequence of the uniqueness.
Theorem 4.2.
Let be a system. For , there exists a unique equilibrium of internal -energy .
Proof.
The uniqueness of equilibrium follows by the affinity of and the strict convexity of . This is a general fact for .
As for , equilibrium of this internal -energy is nothing but the -canonical distribution of temperature , so we already proved the existence.
Now we assume and show the existence of equilibrium of internal -energy . Recall we proved for the -canonical distribution is continuous on and its image is . It follows that there is satisfying , hence . For any other , we have
which shows is the equilibrium of internal -energy . ∎
We recall a system is called K-unstable if for some convex function . Thanks to Theorem 3.13, this is equivalent to say
(4.7) |
is nonempty. For a system and a K-unstable system , we have for the Minkowski sum, and when further is K-unstable.
Now let be a K-unstable system. By the monotonic continuity of , for each , the subset
(4.8) |
is a nonempty compact interval of , which has only one element except for at most countably many . By the above proof, the -canonical distribution of temperature is the equilibrium of internal -energy if and only if . Later we characterize this in terms of the equilibria rather than the -canonical distributions .
By Remark 3.14, for any with , we have
This allows us to interpret as “ability of destabilization” and equilibrium of internal -energy as “the most balanced state among all states of the same ability of destabilization”.
4.2.2. Isothermality
We firstly began our theory with the enigmatic parameter called temperature, studied the minimization of and finally reached the notion of equilibrium, which is a priori irrelevant to the parameter .
Now we can shift our perspective. According to the teaching of thermodynamics, we can rediscover the temperature at least in three essentially equivalent ways:
-
(1)
(Lagrange multiplier) .
-
(2)
(Mutual interaction) Mutual interaction with thermostat.
-
(3)
(Carnot theorem) The ratio of heats in Carnot cycle.
The aim of our observation is to understand the role of temperature and the free -energy in terms of mutual interaction.
Now consider equilibria of internal -energy on systems , respectively. We are interested in comparing the product state and the equilibrium of internal -energy of the composite system . Since in general we have the strict inclusion
(4.9) |
the equilibrium of the composite system may differ from the initial interaction-free state . In thermodynamics, this is known as thermalisation: the initial states are preserved under mutual interaction only when the systems are isothermal, otherwise, one subsystem is eventually warmed up and the other is cooled down.
Proposition 4.3.
Definition 4.4.
Let be equilibria of two K-unstable systems , respectively. We say and are isothermal if the product is the equilibrium of internal -energy on the composite system .
Let us observe the isothermality is reflexive.
Lemma 4.5.
Let be an equilibrium of a system . Then on is the equilibrium of internal -energy . In particular, and are isothermal.
Proof.
Put . Let be the equilibrium of internal -energy . We put
Suppose for some . Then for , we have
by the strict concavity of . This contradicts to the assumption that is the maximizer of on . Thus we have for every and , which shows the claim. ∎
We expect the isothermality is an equivalence relation for pairs , which in thermodynamics is assumed by the zeroth law. (Observe this is not true if we include K-stable systems. ) To see the transitivity, we need strict monotonicity of the canonical entropy on , but what we know at present is monotonicity in a weak sense.
We will prove the strict monotonicity under a slightly better regularity assumption on optimizer. Before discussing it, we observe the Lagrange multiplier interpretation of temperature.
4.2.3. Equilibrium temperature as Lagrange multiplier
Let be a system. For , we put
(4.10) |
We note on . The following superadditivity is clear from (4.9):
(4.11) |
For general , we say pairs and are isothermal if
This definition makes sense even when there is no equilibria of internal -energy . When , this definition is equivalent to the former definition: the equality holds if and only if the equilibria of internal -energy on , respectively, are isothermal. In particular, we have
by Lemma 4.5.
Proposition 4.6.
The functional is concave on whose maximum is achieved at . Moreover, it is strictly concave and strictly increasing on the interval .
Proof.
For with , we compute
which shows the concavity. This product trick is expressive of a thermodynamical intuition behind the concavity, though we present another proof in the following.
We can also see the concavity in a more direct way: for and , we have , so that
Since the last supremum is achieved for , the third inequality is strict for and by the strict concavity of . The strict monotonicity is a consequence of strict concavity. ∎
By the concavity, the right and left differentials exist. We put
(4.12) | ||||
(4.13) |
which are right and left continuous, respectively. We further put
(4.14) |
For , we have
by the strict concavity, where for two intervals we write (resp. ) if (resp. ). Now we show the following.
Proposition 4.7.
Let be K-unstable system. Then for , we have
Proof.
Since is K-unstable, for the -canonical distribution of temperature , we have : if , we have , so that is K-semistable. This implies
By the monotonicity of , we have
for . It follows that
(4.15) |
Indeed, if not, we have for , which is a contradiction.
We observe
For and , we have
by concavity. Thus for , we have
Now suppose is the equilibrium of internal -energy , then for another state , we compute
Thus is the -canonical distribution of temperature for , which shows .
As a general property on left and right derivative of convex function, we have
On the other hand, by and (4.15), we have
so that we get
It follows that
which shows . ∎
4.2.4. Thermalisation
Now we can characterize the isothermality by quantities for equilibrium.
Proposition 4.8.
Let be K-unstable systems and be the equilibria of internal -energy on , respectively. Then the equilibria and are isothermal if and only if
(4.16) |
In this case, we have
Proof.
Assume (4.16) and take . Since are the -canonical distribution of temperature , the product is also the -canonical distribution of temperature by Theorem 4.1. Thus we have and hence is the equilibrium of internal -energy . Therefore, and are isothermal.
Suppose conversely and are isothermal. Since is the equilibrium of internal -energy , it is the -canonical distribution of temperature . Again by Theorem 4.1, are also the -canonical distribution of temperature by the above proof. Thus we have . Since , we obtain the claim. ∎
Now we introduce the following notion.
Definition 4.9.
Let be a K-unstable system. We call mild if for every .
We speculate every K-unstable system is mild, but this is still a conjecture. In the next section, we will see the mildness of some systems including K-semistable systems as a consequence of slightly better regularity of equilibrium. We note if is a system and is a mild K-unstable system, then the composite system is also a mild K-unstable system.
In view of Proposition 4.8, the mildness assumption is essential for the transitivity of isothermaility. Here we conclude the following.
Corollary 4.10.
Let be K-unstable systems. Suppose is mild. If for , and , and are isothermal, respectively, then and are also isothermal.
When two systems are not isothermal, the composite system is thermalized to medium temperatue.
Proposition 4.11.
Let be K-unstable systems and be the composite system. For , we have either
Suppose , then we have
Proof.
Let be the equilibria associated to as in (4.1) and (4.2). Since and are isothermal, by the above proposition, we have
Now we have exactly three possibilities: , and . Suppose , then
Since , we get
by the same argument. It follows that .
Similarly, we obtain when , when . Since these are exclusive conditions, the claim is proved. ∎
We will see the medium temperature can be arbitrary close to when the system is sufficiently large.
4.3. Thermodynamics of non-archimedean -entropy
In the previous section, we discuss equilibrium and thermalisation. Here we rediscover the non-archimedean -entropy from a further exploration on thermalisation. This observation is well-known in thermodynamics. We present it in a mathematically rigorous way.
4.3.1. Temperature as variable
Let us firstly consider change of variables as usual in thermodynamics.
Let be a K-unstable system. Recall for , there exists unique such that . For a system and , we put
(4.17) |
This can be regarded as the inverse function of . It is continuous and increasing on . We further put
(4.18) |
which is also continuous and increasing.
Now for , we introduce
(4.19) |
This is a concave function on . Then we have the following.
Proposition 4.12.
For , we have
Moreover, is continuously differentiable on with
(4.20) |
Proof.
For the -canonical distribution of temperature , we have
by definiton.
The -canonical distribution of temperature has internal -energy , so that we have
Since for the equilibrium of internal -energy , we have
Therefore we get the first claim.
To see the second claim, note
for every . In particular for , we have
with the equality at thanks to the first claim. It follows that is a subdifferential of the convex function at . Since is continuous, is actually differentiable on with the derivative . ∎
4.3.2. Mild regularity hypothesis
The mildness of a K-unstable system is equivalent to the strict monotonicity of . Since is strictly increasing, it is also equivalent to the strict monotonicity of .
Proposition 4.13.
Let be a K-unstable system. Suppose for every there exists such that for the -canonical distribution of temperature . Then is strictly increasing, hence is mild.
Proof.
Take . It suffices to show for every . Suppose we could find a family in so that
-
•
-
•
are differentiable at and .
Since minimizes , we have . This implies that for every , there exists such that
for every . Then using , we compute
It follows by convexity that
which shows the strict monotonicity of .
Now we construct the family under the regularity assumption on . Take with so that . We put . For , we have for small by assumption. This implies the differentiability of around . Meanwhile, we have
by Cauchy–Schwarz inequality. This completes the proof. ∎
Corollary 4.14.
Let be the system associated to a polarized toric variety which is K-semistable for every . Then is either K-semistable or mild K-unstable system.
In view of crystal conjecture (Conjecture 3.16), the assumption of the above theorem, which we call the mild regularity hypothesis, is believed to be always true.
Remark 4.15.
As a weak evidence for the mild regularity hypothesis, we note
for -canonical distribution of any temperature. This could be viewed as a limit of .
For , we have pointwise monotone convergence by the convexity of exponential. Then by monotone convergence theorem, we compute
Since and are smooth for , we directly compute
where the inequality holds as is a maximizer of . Since , the integrals
are finite, thus is also finite by the inequality.
Remark 4.16.
Later we further consider the assumption that has strictly positive differential at some . In thermodynamical terminology, the quantity is called the heat capacity of .
If a polarized toric manifold admits -cscK metric for every , then we can show the family of optimal vectors are smooth. (Apply implicit function theorem to -cscK equation. Compare the proof of [22, Theorem 5.1]. ) It follows that for the associated system , is smooth. We note the positivity of differential implies the strict monotonicity, but the strict monotonocity does not necessarily imply the positivity of differential.
4.3.3. Illustration
Here we illustrate an explicit example with positive heat capacity.
Let be the one point-blowing up of and be the anti-canonical polarization . The associated polytope can be illustrated as follows.
It is shown in [22] that admits -cscK metric for every and the optimal vectors is of the form .
Similarly as [25, section 5.2], we can compute
Then we get the explicit expression
and
We can see for as illustrated in the following figure.

The optimal vector is the critical point of , so we have
If we put
this is equivalent to the condition .


For , we have and . It follows that
for .
4.3.4. Heat bath
It is well known in thermodynamics that the free energy can be derived as the entropy of a composition of the system of our interest and a sufficiently large system working as heat bath. We can realize large system as limit of infinitely many composition. This observation gives us a new interpretation of -canonical distribution: -canonical distribution is equilibrium of an infinite dimensional system.
Theorem 4.17 (Heat bath limit).
Let be a K-unstable system and be a mild K-unstable system. Fix and . For , consider the composite system
Let be the equilibrium of internal -energy . Then the following associated equilibrium on the subsystem
converges in -topology () to the -canonical distribution of temperature , which is independent of the choice of .
Proof.
We note . Since is mild, the composite system is also mild. Let be the element of the one point set
Since on is the -canonical distribution of temperature , on is also the -canonical distribution of temperature by Theorem 4.1.
Since
(4.21) |
we compute
Since is bounded, we get as . It follows that . By the continuity we already proved, we conclude converges to the -canonical distribution of temperature . ∎
Now we obtain a characterization of free -energy in terms of equilibrium of composite system.
Theorem 4.18 (Free -energy as composite entropy).
Let , and be the same as in the above theorem. Assume further the heat capacity of is positive. (See Remark 4.16. ) Namely we assume , is differentiable at and .
Let be the equilibrium of internal -energy on . Then for any of internal -energy , the difference of composite entropy
converges to
as , which is independent of the choice of the mild system .
Proof.
When , we have
so that . Then since is the -canonical distribution of temperature , we have . Then by , we can compute
for .
When , we have either
In either case, we have . Since is differentiable on by Proposition 4.12, is also differentiable at by our assumption. Then by , we have
Since , we have
We then compute
∎
Remark 4.19.
The above theorems are also valid for the case . We do not even need to assume K-instability of , but for the proofs we need separate arguments.
References
-
[1]
V. Apostolov, S. Jubert, A. Lahdili, Weighted K-stability and coercivity with applications to extremal Kähler and Sasaki metrics,
arXiv:2104.09709
. - [2] J. C. Baez, O. Lynch, J. Moeller, Compositional thermostatics, J. Math. Phys. 64 (2023), 2, 023304, 16pp.
-
[3]
R. J. Berman, Emergent Complex Geometry,
arXiv:2109.00307
. -
[4]
H. Blum, Y. Liu, C. Xu, Z. Zhuang, The existence of the Kähler–Ricci soliton degeneration,
arXiv:2103.15278
. - [5] S. Boucksom, M. Jonsson, Global pluripotential theory over a trivially valued field, Ann. Fac. Sci. Toulouse Math. (6) 31 (2022), 3, 647–836.
-
[6]
S. Boucksom, M. Jonsson, A non-archimedean approach to K-stability, I: Metric geometry of spaces of test configurations and valuations,
arXiv:2107.11221
. -
[7]
S. Boucksom, M. Jonsson, A non-archimedean approach to K-stability, II: divisorial stability and openness,
arXiv:2206.09492
. -
[8]
X. Chen, J. Cheng, On the constant scalar curvature Kähler metrics (I) – apriori estimates,
arXiv:1712.06697
. -
[9]
X. Chen, J. Cheng, On the constant scalar curvature Kähler metrics (II) – Existence results,
arXiv:1801.000656
. -
[10]
X. Chen, J. Cheng, On the constant scalar curvature Kähler metrics (II) – Existence results,
arXiv:1801.05907
. - [11] X. Chen, S. Sun, B. Wang, Kähler–Ricci flow, Kähler–Einstein metric, and K-stability, Geom. Topol. 22, 6 (2108), 3145–3173.
- [12] X. Chen, B. Wang, Space of Ricci flows (II) — Part B: weak compactness of the flows, J. Diff. Geom., 116 (2020), 1, 1–123.
- [13] D. A. Cox, J. B. Little, H. K. Schenck, Toric varieties, Graduate Studies in Mathematics, 124, American Mathematical Society, Providence, RI, 2011.
- [14] R. Dervan, G. Székelyhidi, The Kähler–Ricci flow and optimal degenerations, J. Diff. Geom. 116, 1 (2020), 187–203.
- [15] S. Donaldson, Scalar curvature and stability of toric varieties, J. Diff. Geom. 62, 2 (2002), 289–349.
- [16] S. Donaldson, Lower bounds on the Calabi functional, J. Diff. Geom. 70, 3 (2005), 453–472.
- [17] D. Gale, V. Klee, R. T. Rockafellar, Convex functions on convex polytopes, Proc. Amer. Math. Soc. 19 (1968), 867–873.
-
[18]
J. Han, C. Li, Algebraic uniqueness of Kähler–Ricci flow limits and optimal degenerations of Fano varieties,
arXiv:2009.01010
. -
[19]
T. Hisamoto, Stability and coercivity for toric polarizations,
arXiv:1610.07998
. -
[20]
T. Hisamoto, Geometric flow, multiplier ideal sheaves and optimal destabilizer for a Fano manifold,
arXiv:1901.08480
. - [21] E. Inoue, The moduli space of Fano manifolds with Kähler–Ricci solitons, Advances in Math. Volume 357 (2019), 106841.
- [22] E. Inoue, Constant -scalar curvature Kähler metric – formulation and foundational results, J. Geom. Anal. 32 (2022), Article number 145.
-
[23]
E. Inoue, Equivariant calculus on -character and K-stability of polarized schemes,
arXiv:2004.06393
. -
[24]
E. Inoue, Entropies in -framework of canonical metrics and K-stability, I – Archimedean aspect: Perelman’s -entropy and -cscK metrics,
arXiv:2101.11197
. -
[25]
E. Inoue, Entropies in -framework of canonical metrics and K-stability, II – Non-archimedean aspect: non-archimedean -entropy and K-semistability,
arXiv:2202.12168
. -
[26]
D. Joyce, Manifolds with analytic corners,
arXiv:1605.05913
. - [27] A. Lahdili, Kähler metrics with constant weighted scalar curvature and weighted K-stability, Proc. London Math. Soc. (3) 119 (2019), 1065–1114.
-
[28]
A. Lahdili, Convexity of the weighted Mabuchi functional and the uniqueness of weighted extremal metrics,
arXiv:2007.01345
. - [29] C. Li, Geodesic rays and stability in the cscK problem, Ann. Sci. Éc. Norm. Supér. (4) 55(2022), 6, 1529–1574.
-
[30]
C. Li, K-stability and Fujita approximation,
arXiv:2102.09457
. - [31] Y. Nakagwa, On generalized Kähler–Ricci solitons, Osaka J. Math. 48, 2 (2011), 497–513.
-
[32]
G. Perelman, The entropy formula for the Ricci flow and its geometric applications,
arXiv:0211159
. - [33] G. Szekelyhidi, Optimal test configurations for toric varieties, J. Diff. Geom. 80, 3 (2008), 501–523.
- [34] F. Wang, B. Zhou, Fano manifolds with weak almost Kähler–Ricci solitons, Int. Math. Res. Not. IMRN 2015, 9, 2437–2464.
- [35] M. Xia, On sharp lower bounds for Calabi type functionals and destabilizing properties of gradient flows, Anal. PDE 14, 6 (2021), 1951–1976.
- [36] Y. Yao, Mabuchi solitons and relative Ding stability of toric Fano varieties, Int. Math. Res. Not. IMRN 2022, 24, 19790–19853.