log-Coulomb gases in the projective line of a -field
Abstract
This article extends recent results on log-Coulomb gases in a -field (i.e., a nonarchimedean local field) to those in its projective line , where the latter is endowed with the -invariant Borel probability measure and spherical metric. Our first main result is an explicit combinatorial formula for the canonical partition function of log-Coulomb gases in with arbitrary charge values. Our second main result is called the “th Power Law”, which relates the grand canonical partition functions for one-component gases in (where all particles have charge 1) to those in the open and closed unit balls of in a simple way. The final result is a quadratic recurrence for the canonical partition functions for one-component gases in both unit balls of and in . In addition to efficient computation of the canonical partition functions, the recurrence provides their “” limits and “” functional equations.
Mathematics subject classification 2020: 05A18, 11S40, 12J25, 32A99, 82D05
1 Introduction
1.1 Canonical partition functions for log-Coulomb gases
Let be a topological space endowed with a metric and a finite positive Borel measure satisfying for every . A log-Coulomb gas in is a statistical model described as follows: Consider particles with fixed charge values and corresponding variable locations . Whether the charge values are distinct or not, we assume particles are distinguished by the labels , so that unique configurations of the system correspond to unique tuples . Each tuple is called a microstate and has an energy defined by
Note that has measure 0 in by our choice of , and that is identically zero if . We assume the system is in thermal equilibrium with a heat reservoir at some inverse temperature , so that its microstates form a canonical ensemble distributed according to the density . The canonical partition function is defined as the total mass of this density, namely
(1.1.1) |
Given and , the explicit formula for is of primary interest, as it yields fundamental relationships between the observable parameters of the system and its temperature. For instance, the system’s dimensionless free energy, mean energy, and energy fluctuation (variance) are respectively given by , , and , all of which are functions of (and hence of temperature). Below are two examples in which the formula for is known.
Example 1.1.
If with the standard metric , the standard Gaussian measure , and charge values , then Mehta’s integral formula [FW08] states that converges absolutely for all complex with , and in this case
At the special values , the probability density coincides with the joint density of the eigenvalues (counted with multiplicity) of the Gaussian orthogonal , unitary , and symplectic matrix ensembles.
Example 1.2.
If with , the Haar probability measure , and charges with , , and , then a general theorem in [Web21b] implies that converges absolutely for all complex with , and in this case
Note that always extends to a complex domain containing the line . To simultaneously treat all possible choices of , we extend further to a subset of as follows:
Definition 1.3.
For any and as above, we write for a complex tuple (the empty tuple if or ) and define and
Once the formula and domain for are known, then for any choice of , the formula and domain for follow by specializing to . Thus, given , the main problem is to determine the formula and domain of .
1.2 -fields, projective lines, and splitting chains
Of our two main goals in this paper, the first is to determine the explicit formula and domain of , where is a -field and its projective line is endowed with a natural metric and measure. To make this precise, we briefly recall well-known properties of -fields (see [Wei95], for instance) and establish some notation. Fix a -field , write for its canonical absolute value, write for the associated metric (i.e., ), and define
The closed unit ball is the maximal compact subring of , the open unit ball is the unique maximal ideal in , and the group of units is . The residue field is isomorphic to for some prime power , and there is a canonical isomorphism of the cyclic group onto the group of th roots of unity in . Fixing a primitive such root and sending extends the isomorphism to a bijection with inverse . Therefore is a complete set of representatives for the cosets of , i.e.,
(1.2.1) |
Fix a uniformizer (any element satisfying ) and let be the unique additive Haar measure on satisfying . The open balls in are precisely the sets of the form with and , and every such ball is compact with measure equal to its radius, i.e., . In particular, each of the cosets of in (1.2.1) is a compact open ball with measure (and radius) , and two elements satisfy if and only if they belong to different cosets. Henceforth, we reserve the symbols , , , , , , , , and for the items above, and we distinguish from the standard absolute value on by writing for the latter.
We now recall some useful facts from [FP15] in our present notation. The projective line of is the quotient space , where if and only if and for some . Thus we may understand concretely as the set of symbols with , subject to the relation for all and endowed with the topology induced by the quotient . The projective line is compact and metrizable by the spherical metric , which is defined via
(1.2.2) |
In particular, every open set in is a union of balls of the form
(1.2.3) |
with and , and every such ball is open and compact. The projective linear group is the quotient of by its center, namely . It is straightforward to verify that the rule
where and is any representative of , gives a well-defined transitive action of on .
Lemma 1.4 (-invariance [FP15]).
The spherical metric satisfies
for all and all . There is also a unique Borel probability measure on satisfying
for all and all Borel subsets . In particular, for each the relation defines a transitive action on the set of balls of radius , and thus depends only on .
It is routine to verify that and
for all , so we have suitable metrics and measures to define log-Coulomb gases in and . Definition 1.3 specializes to these as follows:
Definition 1.5.
With and as before, we have , and
for . The first integral is independent of , and equal to if or if .
The formulas and domains of absolute convergence for the integrals above can be stated neatly in terms of the following items from [Web21b]:
Definition 1.6 (Splitting chains).
A splitting chain of order and length is a tuple of partitions of satisfying
-
(a)
Each non-singleton part is called a branch of . We write for the set of all branches of , i.e.,
-
(b)
Each must appear in a final partition before refining into two or more parts in , so we define its depth and degree by
-
(c)
We say that is reduced if each satisfies .
Write for the set of reduced splitting chains of order and define
Proposition 3.15 and Theorem 2.6(c) in [Web21b] imply the following proposition, which shall be generalized slightly in order to prove the main results of this paper:
Proposition 1.7.
For , the integral converges absolutely if and only if , and in this case it can be written as the finite sum
Here stands for the degree falling factorial evaluated at the integer .
2 Statement of results
2.1 The projective analogue
Our first main result is the following analogue of Proposition 1.7:
Theorem 2.1.
For , the integral converges absolutely if and only if , and in this case it can be written as the finite sum
The summand for each is defined for all prime powers , as the denominator is cancelled by the factor inside the product over .
The evident similarities between Proposition 1.7 and Theorem 2.1 follow from explicit relationship between the metrics and measures on and those on . These relationships also play a role in the upcoming results, so they are worth recalling now. Note that if and only if , in which case is the unique element of satisfying . Therefore the rule defines a bijection and relates the metric structures of and in a simple way: Given , (1.2.2) implies and
(2.1.1) |
Using the definitions (1.2.3) and for , along with (2.1.1) and the strong triangle equality (i.e., whenever ), one easily verifies that
(2.1.2) |
whenever and . That is, sends the open ball of radius centered at onto the open ball of radius centered at , so is a homeomorphism that restricts to an isometry on and a contraction on .
The map also relates the measures on and in a simple way: Given and a complete set of representatives for the cosets of , applying (2.1.2) to the partition yields
Therefore -invariance of (Lemma 1.4) implies . On the other hand,
implies , which has measure 1. But , so the measure of must be , and therefore every ball with has measure . Combining this with (2.1.2), one concludes that the measure on pulls back along to an explicit measure on :
(2.1.3) |
Finally, (1.2.1) and (2.1.2) give a nice refinement of in terms of the th roots of unity in , which should be understood as the projective analogue of (1.2.1):
(2.1.4) |
Indeed, all of the parts in the partition are balls with measure and radius , and two points satisfy if and only if and belong to different parts. Note that sends onto the “equator” , i.e., the set of points in with -distance 1 from both the “south pole” and the “north pole” .
2.2 Relationships between grand canonical partition functions
So far we have only considered log-Coulomb gases with labeled (and hence distinguishable) particles. Our second main result concerns the situation in which all particles are identical with charge for all , in which case a microstate is regarded as unique only up to permutations of its entries. Since the energy and measure on are invariant under such permutations, each unlabeled microstate makes the contribution to the integral in (1.1.1) precisely times. Therefore the canonical partition function for the unlabeled microstates is given by . We further assume that the system exchanges particles with the heat reservoir with chemical potential and define the fugacity parameter . In this situation the particle number is treated as a random variable and the canonical partition function is replaced by the grand canonical partition function
(2.2.1) |
with the familiar convention . Many properties of the system can be deduced from the grand canonical partition function. For instance, if is fixed and is sub-exponential in , then is analytic in and the expected number of particles in the system is given by . The canonical partition function for each can also be recovered by evaluating the th derivative of with respect to at .
We are interested in the examples , , and , which turn out to share several common properties and simple relationships. By setting in Definition 1.5, one sees that , , and are bounded above by 1 for all and all , and hence , , and are analytic in when . Sinclair recently found an elegant relationship between the first two, which is closely related to the partition of in (1.2.1):
Proposition 2.2 (The th Power Law [Sin20]).
For we have
Roughly speaking, the th Power Law states that a log-Coulomb gas in exchanging energy and particles with a heat reservoir “factors” into identical sub-gases (one in each coset of ) that exchange energy and particles with the reservoir. For , note that the series equation is equivalent to the coefficient identities
(2.2.2) |
The case of (2.2.2) is given in [BGMR06], in which the positive number is recognized as the probability that a random monic polynomial in splits completely in . The more general case given in [Sin20] makes explicit use of the partition of into cosets of (as in (1.2.1)). In Section 3, we will use the analogous partition of into balls (recall (2.1.4)) to show that
(2.2.3) |
which immediately implies our second main result:
Theorem 2.3 (The th Power Law).
For all we have
Like the th Power Law, the th Power Law roughly states that a log-Coulomb gas in exchanging energy and particles with a heat reservoir “factors” into identical sub-gases in the balls , , , , , (each isometrically homeomorphic to ), with fugacity . The th Power Law allows the th Power Law to be written more crudely as
(2.2.4) |
which is to say that the gas in “factors” into two sub-gases: one in and one in (which are respectively isometrically homeomorphic to and ), both with fugacity .
2.3 Functional equations and a quadratic recurrence
Although Proposition 1.7 and Theorem 2.1 provide explicit formulas for and , they are not efficient for computation because they require a complete list of reduced splitting chains of order . For a practical alternative, we take advantage of both Power Laws and the following ideas from [BGMR06] and [Sin20]: Apply to the equation to get
then expand both sides as power series in to obtain the coefficient equations
(2.3.1) |
The identities follow easily from Definition 1.5 and eliminate all instances of in (2.3.1) while introducing powers of the form . For , a careful rearrangement of these powers, the factorials, and the terms in (2.3.1) yields the explicit recurrence
The expression at left is identically 1 if or , so induction confirms that it is polynomial in ratios of hyperbolic sines for all . In particular, its dependence on is carried only by the factor appearing inside the hyperbolic sines, which motivates the following lemma:
Lemma 2.4 (The Quadratic Recurrence).
Set for all and all . For , , and , define by the recurrence
-
(a)
For fixed and fixed , the function is holomorphic for .
-
(b)
For fixed and fixed , the function is defined, smooth, and even on .
Both parts of the Quadratic Recurrence are straightforward to verify by induction. An interesting and immediate consequence of The Quadratic Recurrence and the preceding discussion is the formula
It offers a computationally efficient alternative to the “univariate case” of Proposition 1.7 (when for all ) and extends to a smooth function of . Moreover, it transforms nicely under the involution :
The Quadratic Recurrence serves the projective analogue as well. Expanding (2.2.4) into powers of yields the coefficient equations
(2.3.2) |
and the identities and allow the th summand to be rewritten as
Thus, adding two copies of the sum in (2.3.2) together, pairing the th term of the first copy with the th term of the second copy, and dividing by gives
which is valid for . Through this formula, clearly extends to a smooth function of and is invariant under the involution . We conclude this section by summarizing these observations:
Theorem 2.5 (Efficient Formulas and Functional Equations).
Suppose and , and define as in Lemma 2.4. The th canonical partition functions are given by the formulas
which extend and to smooth functions of satisfying
3 Proofs of the main results
This section will establish the proofs of Theorems 2.1 and 2.3. The common step in both is a decomposition of into cells that are isometrically isomorphic to , which combines with the metric and measure properties in Section 1.2 to create the key relationship between the canonical partition functions for and . We will prove this relationship first, then conclude the proofs of Theorems 2.1 and 2.3 in their own subsections.
3.1 Decomposing the integral over
We begin with an integer that shall remain fixed for the rest of this section, reserve the symbol for a complex tuple , and fix the following notation to better organize the forthcoming arguments:
Notation 3.1.
Let be a subset of .
-
•
For any set we write for the product and assume has the product topology if is a topological space.
-
•
We write for the product Haar measure on satisfying , and we make this consistent for by giving the singleton space measure 1. We also write for the product measure on , with the same convention for .
-
•
For a measurable subset we set and
Note that is constant with respect to the entry if or , and it is equal to if .
-
•
Using the constant , we write for an ordered partition of into at most parts. That is, means are disjoint ordered subsets of with union equal to , where some may be empty.
In addition to the above, it will be useful to consider -analogues of splitting chains:
Definition 3.2.
Suppose . An -splitting chain of length is a tuple of partitions of satisfying
If , we define , , and just as in Definition 1.6. Otherwise will be treated as the empty set and there is no need to define or . Finally, we call an -splitting chain reduced if each satisfies , write for the set of reduced -splitting chains, and define
Note that because has no partitions, because is an intersection of subsets of taken over an empty index set, and for a similar reason. For each singleton , the set is comprised of a single splitting chain of length zero, for the same reason as the case, and similarly . At the other extreme, taking in Definition 3.2 recovers Definition 1.6 and gives .
Proposition 3.3.
For any and any nonempty subset , the integral converges absolutely if and only if , and in this case
In particular, we recover Proposition 1.7 by taking and .
Proof.
First suppose is a singleton, so that the product inside the integral is empty and hence
This integral is constant, and hence absolutely convergent, for all . On the other hand, consists of a single -splitting chain, namely the one-tuple . Then and imply
as well, so the claim holds for any singleton subset . Now suppose is not a singleton. By relabeling we may assume where . By Proposition 3.15 and Lemma 3.16(c) in [Web21b], the integral
converges absolutely if and only if belongs to the intersection
(3.1.1) |
and for such we have
Changing variables in the integral by the homothety gives
and the first equality implies that the domain of absolute convergence for is also the intersection appearing in (3.1.1). But every subset with appears as a branch in at least one reduced splitting chain of order , so the intersection in (3.1.1) is precisely . Therefore the claim holds for , and hence for any non-singleton subset .
∎
The case of Proposition 3.3 has an important relationship with the main result of this section, which is the following theorem.
Theorem 3.4.
For each , the integral converges absolutely if and only if , and in this case
Proof.
The partition of in (2.1.4) can be rewritten in the form
where is the element represented by if , if , or if . This leads to a partition of the -fold product,
where each part is a “cell” of the form
Accordingly, the integral breaks into a sum of integrals of the form
(3.1.2) |
summed over all . Since every cell has positive measure, the integral converges absolutely if and only if the integral in (3.1.2) converges absolutely for all . Fix one for the moment. By (2.1.4) and the definition of the ’s above, note that the entries of each tuple satisfy if and only if and belong to different parts of . Therefore the integrand in (3.1.2) factors as
and the measure on factors in a similar way, namely where is the product measure on . Now Fubini’s Theorem for positive functions and -invariance give
so the integral in (3.1.2) converges absolutely if and only if all of the integrals of the form
(3.1.3) |
converge absolutely. The change of variables given by in each coordinate, along with (2.1.1), (2.1.2), and (2.1.3), allows the integral in (3.1.3) to be rewritten as , and thus Proposition 3.3 implies that it converges absolutely if and only if . Therefore the integral over in (3.1.2) converges absolutely if and only if , and in this case Fubini’s Theorem for absolutely integrable functions, -invariance, and the change of variables above allow it to be rewritten as
Finally, since is the sum of these integrals over all , it converges absolutely if and only if
The last equality of intersections holds because each subset with appears as a part in at least one of the ordered partitions , and because none of the parts with affect the intersection (because for such ). The intersection of over all with is clearly equal to by Definition 3.2, so the proof is complete.
∎
3.2 Finishing the proof of Theorem 2.1
Theorem 3.4 established that the integral converges absolutely if and only if , and for such it gave
(3.2.1) |
It remains to show that the righthand sum can be converted into the sum over proposed in Theorem 2.1.
Proof of Theorem 2.1.
We begin by breaking the terms of the sum in (3.2.1) into two main groups. The simpler group is indexed by those with for some and for all , in which case and for all . Therefore each of the group’s terms (one for each ) contributes the quantity to the sum in (3.2.1) for a total contribution of
(3.2.2) |
by the and case of Proposition 3.3. The other group of terms is indexed by the ordered partitions satisfying . To deal with them carefully, we fix one such for the moment, and note that the number of nonempty parts must be at least 2. Thus we have indices with , and for every we have and hence . For the nonempty sets , Proposition 3.3 expands as a sum over (whose elements shall be denoted instead of ) and hence
We now make use of a simple correspondence between the tuples and the reduced splitting chains satisfying . To establish it, note that each corresponds uniquely to its branch set (Lemma 2.5(b) of [Web21b]), which generalizes in an obvious way to reduced -splitting chains (for any nonempty ). Now if satisfies , the corresponding branch set decomposes as
Each of the sets is the branch set for a unique , so in this sense “breaks” into a unique tuple . On the other hand, any tuple can be “assembled” as follows. Since is a partition of , taking the union of the branch sets and the singleton forms the branch set for a unique . It is clear that “breaking” and “assembling” are inverses, giving a correspondence under which each identification amounts to a branch set equation, i.e.,
In particular, each is contained in exactly one , and by Definition 1.6 in this case. These facts allow the sum over above to be rewritten as a sum over all with , and each product over inside it is simply a product over . We conclude that an ordered partition with contributes the quantity
(3.2.3) |
to the sum in (3.2.1), where is the (unordered) subset of nonempty parts in that particular ordered partition. We must now total the contribution in (3.2.3) over all possible with . Given a partition with , note that there are precisely ordered partitions such that . Therefore summing (3.2.3) over all with gives
Given a partition , those splitting chains with all have by Definition 1.6. Moreover, no is missed or repeated in the sum of sums above, so it can be rewritten as
Note that the summand for each is still defined for any prime power since the denominators and (which vanish when ) are cancelled by the numerator appearing in the product over . Finally, we evaluate the righthand side of (3.2.1) by combining the sum directly above with that in (3.2.2) and multiplying through by . This yields the desired formula for :
∎
3.3 Finishing the proof of Theorem 2.3
Our final task is to prove the th Power Law, which we noted in Section 2.2 is equivalent to the equations in (2.2.3). That is, it remains to prove
Proof.
Fix and , and fix via for all , so that and for any subset . The formula in Theorem 3.4 relates these functions of via
For each choice of ordered integers satisfying , there are precisely
ordered partitions satisfying . Finally, grouping ordered partitions according to all possible ordered integer choices establishes the desired equation:
∎
Acknowledgements: I would like to thank Clay Petsche for confirming several details about the measure and metric on , and I would like to thank Chris Sinclair for many useful suggestions regarding the Power Laws and the Quadratic Recurrence.
References
- [BGMR06] Joe Buhler, Daniel Goldstein, David Moews, and Joel Rosenberg, The probability that a random monic -adic polynomial splits, Experiment. Math. 15 (2006), no. 1, 21–32. MR 2229382
- [DM91] Jan Denef and Diane Meuser, A functional equation of Igusa’s local zeta function, Amer. J. Math. 113 (1991), no. 6, 1135–1152. MR 1137535
- [FP15] Paul Fili and Clayton Petsche, Energy integrals over local fields and global height bounds, Int. Math. Res. Not. IMRN (2015), no. 5, 1278–1294. MR 3340356
- [FW08] Peter J. Forrester and S. Ole Warnaar, The importance of the Selberg integral, Bull. Amer. Math. Soc. (N.S.) 45 (2008), no. 4, 489–534. MR 2434345
- [Sin20] Christopher D. Sinclair, Non-Archimedean Electrostatics, arXiv e-prints (2020), arXiv:2002.07121.
- [Vol10] Christopher Voll, Functional equations for zeta functions of groups and rings, Ann. of Math. (2) 172 (2010), no. 2, 1181–1218. MR 2680489
- [Web21a] J. Webster, The combinatorics of log-Coulomb gases in -fields, Ph.D. thesis, Eugene, OR, 2021.
- [Web21b] Joe Webster, log-Coulomb gas with norm-density in -fields, P-Adic Num Ultrametr Anal Appl 13 (2021), no. 1, 1–43.
- [Wei95] André Weil, Basic number theory, Classics in Mathematics, Springer-Verlag, Berlin, 1995, Reprint of the second (1973) edition. MR 1344916
Joe Webster
Department of Mathematics, University of Virginia, Charlottesville, VA 22903
email: [email protected]