1. Introduction
This paper is concerned with the geometric formulation of thermodynamic systems. While the geometric formulation of mechanical systems has given rise to an extensive theory, commonly called geometric mechanics, the geometric formulation of thermodynamics has remained more elusive and restricted.
Starting from Gibbs’ fundamental relation, contact geometry has been recognized since the 1970s as an appropriate framework for the geometric formulation of thermodynamics; see in particular [
1,
2,
3,
4,
5,
6,
7,
8]. More recently, the interest in contact-geometric descriptions has been growing, from different points of view and with different motivations; see, e.g., [
9,
10,
11,
12,
13,
14,
15,
16,
17,
18,
19,
20].
Despite this increasing interest, the current geometric theory of thermodynamics still poses major challenges. First, most of the work is on the geometric formulation of the equations of state, through the use of Legendre submanifolds [
1,
2,
3,
5,
8], while less attention has been paid to the geometric definition and analysis of non-equilibrium dynamics. Secondly, thermodynamic system models commonly appear both in energy and in entropy representation, while in principle, this corresponds to contactomorphic, but different contact manifolds. This is already demonstrated by rewriting Gibbs’ equation in energy representation
, with intensive variables
, into the entropy representation
, with intensive variables
. Thirdly, for reasons of analysis and control of composite thermodynamic systems, a geometric description of the interconnection of thermodynamic systems is desirable, but currently largely lacking.
A new viewpoint on the geometric formulation of thermodynamic systems was provided in [
21], by exploiting the well-known result in geometry that odd-dimensional contact manifolds can be naturally symplectized to even-dimensional symplectic manifolds with an additional structure of homogeneity; see [
22,
23] for textbook expositions. While the classical applications of symplectization are largely confined to time-dependent Hamiltonian mechanics [
23] and partial differential equations [
22], the paper [
21] argued convincingly that symplectization provides an insightful angle to the geometric modeling of thermodynamic systems as well. In particular, it yields a clear way to bring together energy and entropy representations, by viewing the choice of different intensive variables as the selection of different homogeneous coordinates.
In the present paper, we aim at expanding this symplectization point of view towards thermodynamics, amplifying our initial work [
24,
25]. In particular, we show how the symplectization point of view not only unifies the energy and entropy representation, but is also very helpful in describing the dynamics of thermodynamic processes, inspired by the notion of the contact control system developed in [
11,
12,
13,
17,
18,
19]; see also [
16]. Furthermore, it yields a direct and global definition of a metric on the submanifold describing the state properties, encompassing the locally-defined metrics of Weinhold [
26] and Ruppeiner [
27], and providing a new angle to the equivalence results obtained in [
3,
5,
7,
10]. Finally, it is shown how symplectization naturally leads to a definition of interconnection ports; thus extending the compositional geometric port-Hamiltonian theory of interconnected multi-physics systems (see, e.g., [
28,
29,
30]) to the thermodynamic realm. All this will be illustrated by a number of simple, but instructive, examples, primarily serving to elucidate the developed framework and its potential.
2. Thermodynamic Phase Space and Geometric Formulation of the Equations of State
The starting point for the geometric formulation of thermodynamic systems throughout this paper is an
-dimensional manifold
, with
, whose coordinates comprise the extensive variables, such as volume and mole numbers of chemical species, as well as entropy and energy [
31]. Emphasis in this paper will be on simple thermodynamic systems, with a single entropy and energy variable. Furthermore, for notational simplicity, and without much loss of generality, we will assume:
with
the entropy variable,
the energy variable, and
Q the
-dimensional manifold of remaining extensive variables (such as volume and mole numbers).
In composite (i.e., compartmental) systems, we may need to consider multiple entropies or energies; namely for each of the components. In this case,
is replaced by
, with
denoting the number of entropies and
the number of energies; see Example 3 for such a situation. This also naturally arises in the interconnection of thermodynamic systems, as will be discussed in
Section 5.
Coordinates for
throughout will be denoted by
, with
q coordinates for
Q (the manifold of remaining extensive variables). Furthermore, we denote by
the
-dimensional cotangent bundle
without its zero-section. Given local coordinates
for
, the corresponding natural cotangent bundle coordinates for
and
are denoted by:
where the co-tangent vector
will be called the vector of co-extensive variables.
Following [
21], the thermodynamic phase space
is defined as the projectivization of
, i.e., as the fiber bundle over
with fiber at any point
given by the projective space
. (Recall that elements of
are identified with rays in
, i.e., non-zero multiples of a non-zero cotangent vector.) The corresponding projection will be denoted by
.
It is well known [
22,
23] that
is a contact manifold of dimension
. Indeed, recall [
22,
23] that a contact manifold is an
-dimensional manifold
N equipped with a maximally non-integrable field of hyperplanes
. This means that
for a, possibly only locally-defined, one-form
on
N satisfying
. By Darboux’s theorem [
22,
23], there exist local coordinates (called Darboux coordinates)
for
N such that, locally:
Then, in order to show that
for any
-dimensional manifold
M is a contact manifold, consider the Liouville one-form
on the cotangent bundle
, expressed in natural cotangent bundle coordinates for
as
. Consider a neighborhood where
, and define the homogeneous coordinates:
which, together with
, serve as local coordinates for
. This results in the locally-defined contact form
as in (
3) (with
). The same holds on any neighborhood where one of the other coordinates
is different from zero, in which case division by the non-zero
results in other homogeneous coordinates. This shows that
is indeed a contact manifold. Furthermore [
22,
23],
is the canonical contact manifold in the sense that every contact manifold
N is locally contactomorphic to
for some manifold
M.
Taking
, it follows that coordinates for the thermodynamical phase space
are obtained by replacing the coordinates
for the fibers
by homogeneous coordinates for the projective space
. In particular, assuming
, we obtain the homogeneous coordinates:
defining the intensive variables of the energy representation. Alternatively, assuming
, we obtain the homogeneous coordinates (see [
21] for a discussion of
, or
, as a gauge variable):
defining the intensive variables of the entropy representation.
Example 1. Consider a mono-phase, single constituent, gas in a closed compartment, with volume , entropy S, and internal energy E, satisfying Gibbs’ relation . In the energy representation, the intensive variable γ is given by the pressure , and is the temperature T. In the entropy representation, the intensive variable is equal to , while equals the reciprocal temperature .
In order to provide the geometric formulation of the equations of state on the thermodynamic phase space
, we need the following definitions. First, recall that a submanifold
of
is called a Lagrangian submanifold [
22,
23] if the symplectic form
is zero restricted to
and the dimension of
is equal to the dimension of
(the maximal dimension of a submanifold restricted to which
can be zero).
Definition 1. A homogeneous Lagrangian submanifold is a Lagrangian submanifold with the additional property that: In the
Appendix A, cf. Proposition A2, homogeneous Lagrangian submanifolds are geometrically characterized as submanifolds
of dimension equal to
, on which not only the symplectic form
, but also the Liouville one-form
is zero.
Importantly, homogeneous Lagrangian submanifolds of
are in one-to-one correspondence with Legendre submanifolds of
. Recall that a submanifold
L of a
-dimensional contact manifold
N is a Legendre submanifold [
22,
23] if the locally-defined contact form
is zero restricted to
L and the dimension of
L is equal to
n (the maximal dimension of a submanifold restricted to which
can be zero).
Proposition 1 ([
23], Proposition 10.16)
. Consider the projection . Then, is a Legendre submanifold if and only if is a homogeneous Lagrangian submanifold. Conversely, any homogeneous Lagrangian submanifold is of the form for some Legendre submanifold L. In the contact geometry formulation of thermodynamic systems [
1,
2,
3,
5], the equations of state are formalized as Legendre submanifolds. In view of the correspondence with homogeneous Lagrangian submanifolds, we arrive at the following.
Definition 2. Consider and the thermodynamical phase space . The state properties of the thermodynamic system are defined by a homogeneous Lagrangian submanifold and its corresponding Legendre submanifold .
The correspondence between Legendre and homogeneous Lagrangian submanifolds also implies the following characterization of generating functions for any homogeneous Lagrangian submanifold
. This is based on the fact [
22,
23] that any Legendre submanifold
in Darboux coordinates
for
N can be locally represented as:
for some partitioning
and some function
(called a generating function for
L), while conversely, any submanifold
L as given in (
8), for any partitioning
and function
, is a Legendre submanifold.
Given such a generating function
for the Legendre submanifold
L, we now define, assuming
and substituting
,
Then a direct computation shows that:
implying, in view of (
8), that:
In its turn, this implies that
G as defined in (
9) is a generating function for the homogeneous Lagrangian submanifold
. If instead of
, another coordinate
is different from zero, then by dividing by this
, we obtain a similar generating function. This is summarized in the following proposition.
Proposition 2. Any Legendre submanifold L can be locally represented as in (
8)
, possibly after renumbering the index set , for some partitioning and generating function , and conversely, for any such , the submanifold L defined by (
8)
is a Legendre submanifold. Any homogeneous Lagrangian submanifold can be locally represented as in (
11)
with generating function G of the form (
9)
, and conversely, for any such G, the submanifold (
11)
is a homogeneous Lagrangian submanifold. Note that the generating functions
G as in (
9) are homogeneous of degree one in the variables
; see the
Appendix A for further information regarding homogeneity.
The simplest instance of a generating function for a Legendre submanifold
L and its homogeneous Lagrangian counterpart
occurs when the generating
F as in (
8) only depends on
. In this case, the generating function
G is given by:
with the corresponding homogeneous Lagrangian submanifold
locally given as:
A particular feature of this case is the fact that exactly one of the extensive variables, in the above
, is expressed as a function of all the others, i.e.,
. At the same time,
is unconstrained, while the other co-extensive variables
are determined by
. For a general generating function
G as in (
9), this is not necessarily the case. For example, if
, corresponding to a generating function
, then
are all expressed as a function of the unconstrained variables
.
Remark 1. In the present paper, crucial use is made of homogeneity in the co-extensive variables , which is different from homogeneity with respect to the extensive variables , as occurring, e.g., in the Gibbs–Duhem relations [31]. The two most important representations of a homogeneous Lagrangian submanifold
, and its Legendre counterpart
, are the energy representation and the entropy representation. In the first case,
is represented, as in (
12), by a generating function of the form:
yielding the representation:
In the second case (the entropy representation),
is represented by a generating function of the form:
yielding the representation:
Note that in the energy representation, the independent extensive variables are taken to be q and the entropy S, while the energy variable E is expressed as a function of them. On the other hand, in the entropy representation, the independent extensive variables are q and the energy E, with S expressed as a function of them. Furthermore, in the energy representation, the co-extensive variable is “free”, while instead in the entropy representation, the co-extensive variable is free. In principle, also other representations could be chosen, although we will not pursue this. For instance, in Example 1, one could consider a generating function where the extensive variable V is expressed as function of the other two extensive variables .
As already discussed in [
1,
2], an important advantage of describing the state properties by a Legendre submanifold
L, instead of by writing out the equations of state, is in providing a global and coordinate-free point of view, allowing for an easy transition between different thermodynamic potentials. Furthermore, if singularities occur in the equations of state,
L is typically still a smooth submanifold. As seen before [
21], the description by a homogeneous Lagrangian submanifold
has the additional advantage of yielding a simple way for switching between the energy and the entropy representation.
Remark 2. Although the terminology “thermodynamic phase space” for may suggest that all points in are feasible for the thermodynamic system, this is actually not the case. The state properties of the thermodynamic system are specified by the Legendre submanifold , and thus, the actual “state space” of the thermodynamic system at hand is this submanifold L; not the whole of .
A proper analogy with the Hamiltonian formulation of mechanical systems would be as follows. Consider the phase space of a mechanical system with configuration manifold Q. Then, the Hamiltonian defines a Lagrangian submanifold of given by the graph of the gradient of H. The homogeneous Lagrangian submanifold is analogous to , while the symplectized thermodynamic phase space is analogous to .
3. The Metric Determined by the Equations of State
In a series of papers starting with [
26], Weinhold investigated the Riemannian metric that is locally defined by the Hessian matrix of the energy expressed as a (convex) function of the entropy and the other extensive variables. (The importance of this Hessian matrix, also called the stiffness matrix, was already recognized in [
31,
32].) Similarly, Ruppeiner [
27], starting from the theory of fluctuations, explored the locally-defined Riemannian metric given by minus the Hessian of the entropy expressed as a (concave) function of the energy and the other extensive variables. Subsequently, Mrugała [
3] reformulated both metrics as living on the Legendre submanifold
L of the thermodynamic phase space and showed that actually, these two metrics are locally equivalent (by a conformal transformation); see also [
9]. Furthermore, based on statistical mechanics arguments, [
7] globally defined an indefinite metric on the thermodynamical phase space, which, when restricted to the Legendre submanifold, reduces to the Weinhold and Ruppeiner metrics; thus showing global conformal equivalence. This point of view was recently further extended in a number of directions in [
10].
In this section, crucially exploiting the symplectization point of view, we provide a novel global geometric definition of a degenerate pseudo-Riemannian metric on the homogeneous Lagrangian submanifold
defining the equations of state, for any given torsion-free connection on the space
of extensive variables. In a coordinate system in which the connection is trivial (i.e., its Christoffel symbols are all zero), this metric will be shown to reduce to Ruppeiner’s locally-defined metric once we use homogeneous coordinates corresponding to the entropy representation, and to Weinhold’s locally-defined metric by using homogeneous coordinates corresponding to the energy representation. Hence, parallel to the contact geometry equivalence established in [
3,
7,
10], we show that the metrics of Weinhold and Ruppeiner are just two different local representations of this same globally-defined degenerate pseudo-Riemannian metric on the homogeneous Lagrangian submanifold of the symplectized thermodynamic phase space.
Recall [
33] that a (affine) connection ∇ on an
-dimensional manifold
M is defined as an assignment:
for any two vector fields
, which is
-bilinear and satisfies
and
, for any function
f on
M. This implies that
only depends on
and the value of
Y along a curve, which is tangent to
X at
q. In local coordinates
q for
M, the connection is determined by its Christoffel symbols
, defined by:
The connection is called torsion-free if:
for any two vector fields
, or equivalently if its Christoffel symbols satisfy the symmetry property
. We call a connection trivial in a given set of coordinates
if its Christoffel symbols in these coordinates are all zero.
As detailed in [
34], given a torsion-free connection on
M, there exists a natural pseudo-Riemannian (“pseudo” since the metric is indefinite) metric on the cotangent-bundle
, in cotangent bundle coordinates
for
given as:
Let us now consider for
M the manifold of extensive variables
with coordinates
as before, where we assume the existence of a torsion-free connection, which is trivial in the coordinates
, i.e., the Christoffel symbols are all zero. Then, the pseudo-Riemannian metric
on
takes the form:
Denote by
the pseudo-Riemannian metric
restricted to the homogeneous Lagrangian submanifold
describing the state properties. Consider the energy representation (
15) of
, with generating function
. It follows that
equals (in shorthand notation):
where:
is recognized as Weinhold’s metric [
26]; the (positive-definite) Hessian of
E expressed as a (strongly convex) function of
q and
S.
On the other hand, in the entropy representation (
17) of
, with generating function
, an analogous computation shows that
is given as
, with:
the Ruppeiner metric [
27]; minus the Hessian of
S expressed as a (strongly concave) function of
q and
E. Hence, we conclude that:
implying
, with
T the temperature. This is basically the conformal equivalence between
and
found in [
3]; see also [
7,
10]. Summarizing, we have found the following.
Theorem 1. Consider a torsion-free connection on , with coordinates , in which the Christoffel symbols of the connection are all zero. Then, by restricting the pseudo-Riemannian metric to , we obtain a degenerate pseudo-Riemannian metric on , which in local energy-representation (
15)
for is given by , with the Weinhold metric (
24)
, and in a local entropy representation (
17)
by , with the Ruppeiner metric (
25)
. We emphasize that the degenerate pseudo-Riemannian metric
is globally defined on
, in contrast to the locally-defined Weinhold and Ruppeiner metrics
and
; see also the discussion in [
3,
5,
7,
9,
10]. We refer to
as degenerate, since its rank is at most
n instead of
. Note furthermore that
is homogeneous of degree one in
and hence does not project to the Legendre submanifold
L.
While the assumption of the existence of a trivial connection appears natural in most cases (see also the information geometry point of view as exposed in [
35]), all this can be directly extended to any non-trivial torsion-free connection ∇ on
. For example, consider the following situation.
For the ease of notation, denote
, and correspondingly denote
. Take any torsion-free connection on
given by symmetric Christoffel symbols
, with indices
, satisfying
whenever one of the indices
is equal to the index
E. Then, the indefinite metric
on
is given by (again in shorthand notation):
It follows that the resulting metric
on
is given by the matrix:
Here, the
-matrix at the right-hand side of
is the globally defined geometric Hessian matrix (see e.g., [
36]) with respect to the connection on
corresponding to the Christoffel symbols
,
.
5. Interconnections of Port-Thermodynamic Systems
In this section, we study the geometric formulation of interconnection of port-thermodynamic systems through their ports, in the spirit of the compositional theory of port-Hamiltonian systems [
28,
29,
30,
43]. We will concentrate on the case of power-port interconnections of port-thermodynamic systems, corresponding to power flow exchange (with total power conserved). This is the standard situation in (port-based) physical network modeling of interconnected systems. At the end of this section, we will make some remarks about other types of interconnection; in particular, interconnection by exchange of the rate of entropy.
Consider two port-thermodynamic systems with extensive and co-extensive variables:
and Liouville one-forms
. With the homogeneity assumption in mind, impose the following constraint on the co-extensive variables:
This leads to the summation of the one-forms
and
given by:
on the composed space defined as:
Leaving out the zero-section , this space will be denoted by and will serve as the space of extensive and co-extensive variables for the interconnected system. Furthermore, it defines the projectivization , which serves as the composition (through ) of the two projectivizations .
Let the state properties of the two systems be defined by homogeneous Lagrangian submanifolds:
with generating functions
. Then, the state properties of the composed system are defined by the composition:
with generating function
.
Furthermore, consider the dynamics on defined by the Hamiltonians . Assume that does not depend on the energy variable . Then, the sum is well-defined on for all . This defines a composite port-thermodynamic system, with entropy variables , total energy variable E, inputs , and state properties defined by .
Next, consider the power-conjugate outputs
; in the sequel, simply denoted by
. Imposing on the power-port variables
interconnection constraints that are satisfying the power-preservation property:
yields an interconnected dynamics on
, which is energy conserving (the
-term in the expression for
is zero by (
76)). This is summarized in the following proposition.
Proposition 6. Consider two port-thermodynamic systems with spaces of extensive variables , . Assume that does not depend on , . Then, , with given in (
75)
, defines a composite port-thermodynamic system with inputs and outputs . By imposing interconnection constraints on satisfying (
76)
, an autonomous (no inputs) port-thermodynamic system is obtained. Remark 8. The interconnection procedure can be extended to the case of an additional open power-port with input vector u and output row vector y, by replacing (
76)
by power-preserving interconnection constraints on satisfying: Proposition 6 is illustrated by the following examples.
Example 7 (Mass-spring-damper system). We will show how the thermodynamic formulation of the system as detailed in Example 4 also results from the interconnection of the three subsystems: mass, spring, and damper.
I. Mass subsystem (leaving out irrelevant entropy). The state properties are given by:with energy κ (kinetic energy) and dynamics generated by the Hamiltonian:corresponding to . II. Spring subsystem (again leaving out irrelevant entropy). The state properties are given by:with energy P (spring potential energy) and dynamics generated by the Hamiltonian:corresponding to . III. Damper subsystem. The state properties are given by:involving the entropy S and an internal energy . The dynamics of the damper subsystem is generated by the Hamiltonian:with d the damping constant and power-conjugate output:equal to the damping force. Finally, interconnect, in a power-preserving way, the three subsystems to each other via their power-ports as: This results (after setting ) in the interconnected port-thermodynamic system with total Hamiltonian given as:which is equal to the Hamiltonian for as obtained before in Example 4, Equation (
63)
. Example 8 (Heat exchanger)
. Consider two heat compartments as in Example 2, with state properties:The dynamics is given by the Hamiltonians:with the incoming heat flows and power-conjugate outputs , which both are equal to one. Consider the power-conserving interconnection:with λ the Fourier heat conduction coefficient. Then, the Hamiltonian of the interconnected port-thermodynamical system is given by:which equals the Hamiltonian (
59)
as obtained in Example 3. Apart from power-port interconnections as above, we may also define other types of interconnection, not corresponding to the exchange of rate of energy (power), but instead to the exchange of rate of other extensive variables. In particular, an interesting option is to consider interconnection via the rate of entropy exchange. This can be done in a similar way, by considering, instead of the variables as above, the variables . Imposing alternatively the constraint yields a similar composed space of extensive and co-extensive variables, as well as a similar composition of the state properties. By assuming in this case that the Hamiltonians do not depend on the entropies and by imposing interconnection constraints on and the “rate of entropy” conjugate outputs leads again to an interconnected port-thermodynamic system. Note however that while it is natural to assume conservation of total energy for the interconnection of two systems via their power-ports, in the alternative case of interconnecting through the rate of entropy ports, the total entropy may not be conserved, but actually increasing.
Example 9. As an alternative to the previous Example 8, where the heat exchanger was modeled as the interconnection of two heat compartments via power-ports, consider the same situation, but now with outputs being the “rate of entropy conjugate” to , i.e., equal (cf. the end of Example 2) to the reciprocal temperatures with , . This results in interconnecting the two heat compartments as, equivalently to (
89)
, This interconnection is not total entropy conserving, but instead satisfies , corresponding to the increase of total entropy.