Appendix A. Mathematical Derivation of the Thermosyphon Model
For the sake of completeness, we summarize succinctly the method described in Ref. [
24].
- (i)
Starting from Equations (
12), (
13), (
15) and (
16), we project them along a unitary vector tangent of the pipe (hereafter, we will refer to the coordinate along that tangent simply as
x).
- (ii)
Given the sufficiently small size of the thermosyphon’s cross-sectional area, denoted as A, we can safely disregard variations in the transverse direction of the pipe, ensuring that the process of averaging does not result in any appreciable loss of precision.
- (iii)
Finally, since the fluid is incompressible, the average velocity remains constant along the entire length of the pipe.
To emphasize the role of elasticity, we differentiate Equation (
12) to eliminate the stress component
(after multiplying by
). Thus,
Based on Refs. [
24,
27], it is assumed that the non-linear terms and viscous term,
, can be expressed through the wall law force,
, which is dependent on the average velocity (see details below).
According to the incompressibility hypothesis stated in Equation (
13), the average velocity
v remains constant throughout the thermosyphon and is solely dependent on time,
t, rather than the spatial coordinate along the pipe,
x. Additionally, due to the projection and averaging procedures, temperature,
T, and solute concentration,
S, are only functions of time,
t, and position along the loop,
x. By projecting in the direction of the pipe, we obtain Equation (
17).
In Equations (
12)–(
14), the velocity is linked with temperature and diffusion through the body force
, where the density
depends on temperature (hotter fluids have lower density) and solute concentration (higher solute concentrations lead to higher densities). Mathematically, it is possible to assume a linear correlation between density and temperature. Similarly, in this work, we also assume a linear relationship between density and solute concentration. Thus, in the end,
with
and
, being the so-called dilatation and thermophoretic coefficients, respectively.
In this work, we use the Boussinesq approximation [
24,
27], which assumes that the density, which is proportional to the dynamical terms
and
, remains approximately constant and equal to
. It also assumes that the effects of temperature and solute concentration on inertia can be neglected. Therefore, by integrating Equation (
17) over the length of the pipe
L, resulting in zero integrated pressure gradients and a constant gravitational term
, we obtain the following equation:
Here, the function
denotes a mathematical expression that relates
to the coordinate
x. Following Rodriguez-Bernal et al.’s study in 1995, we make the assumption that the local dependence of the wall law projection on velocity
v can be expressed as
. The function
exhibits specific general characteristics that are elucidated in the subsequent discussion.
To complete the closure of Equations (
A3) and (
15), it is necessary to supplement them with appropriate “constitutive equations” governing the behavior of temperature and solute concentrations. The primary mechanisms considered for temperature evolution are as follows:
Convection accounted for in the second term on the left-hand side of Equation (
15).
Thermal diffusion (conduction), expressed as
. Thus, from Equation (
15),
.
Similarly, for solute concentration, the following physical mechanisms are taken into consideration:
Convection, incorporated in the second term on the left-hand side of Equation (16).
Solute diffusion, described by .
Onsager coupling (thermodiffusion or the Soret effect), represented by .
Hence, from Equation (16), .
Finally, we proceed to non-dimensionalize the resulting equations, whereby lengths are scaled by the loop length (with the entire loop represented by the interval
), temperatures are scaled by
, and solute concentration by
, to arrive at the main focus of our work, summarized in the system of Equation
18.
Appendix A.1. Well-Posedness and Boundedness: Global Attractor
For completeness, we include the details of well-posedness and boundedness for the present model as it requires specific work, although the techniques used are similar to those in Refs. [
28,
29].
In this context, we enforce the condition that the thermal diffusivity
maintains physical consistency. The system of Equation (
18) represents an ODE/PDE system governing the evolution of the observables
v,
T, and
S. The solute concentration
adheres to a solute diffusivity
and a positive Soret coefficient
b, similar to previous models for binary fluids, like in [
6,
15,
30].
In subsequent discussions, we utilize the notation
for integration along a closed circuit path, which can be associated with integration over periodic functions with a period of 1. The function
f in the equation characterizes the loop’s geometry and gravitational forces [
24,
27], where
, since the integral of
over a closed loop yields zero.
The parameter
in Equation (
18) represents the non-dimensional form of
, with time units. Roughly,
determines the (non-dimensional) timescale for the material to transition from an elastic to a fluid-like behavior.
We assume that the friction law
at the inner wall of the loop is positive and remains bounded away from zero. Previous works have considered
as a constant for linear friction [
27] (Stokes flow) or
for quadratic friction [
25,
26]. A more general form of
is given by
, where
is the Reynolds number, and
is a function of
. Here, we consider a general function of the velocity, assumed to be large [
31,
32]. The functions
G,
f, and
l incorporate relevant physical constants of the model, such as the cross-sectional area
D, the length of the loop
L, and the Prandtl, Rayleigh, or Reynolds numbers. We require
G and
l to be continuous functions satisfying
and
, where
and
are positive constants. Additionally, we impose further constraints on
G in Equation (
A23), mandating
G to be a nonlinear friction function, such that there exists a constant
, satisfying the following:
We will introduce some function spaces that will be used to study the existence of solutions of (
18). Let
and consider the spaces
where
, and
(or
) if for every open set
one has
(or
, respectively). Finally, we consider functions with zero average, and we denote by the following:
Appendix A.2. Existence and Uniqueness of Solutions
In this section, we demonstrate the presence and singularity of solutions for the thermosyphon model (
18), where
f and
h belong to
,
is in
, and
is in
. Here,
and
are defined by (
A5). It should be noted that the term dot indicates functions that have zero averages and not the time derivatives of the functions. To begin with, we choose this framework by observing that for
, integrating the temperature equation along the loop, while considering the periodicity of
T, we have the following:
and
Next, if
, we have
. Therefore, the temperature is unbounded, as
, unless
. However, taking
and
reduces to the case
, since
would satisfy
Additionally, if we integrate the equation for the solute concentration along the loop and take into account the periodicity of S, we obtain the conditions and Since is constant, it means that the solute for all t.
Let
(not to be confused with the stress component
). Then, from the third equation of system (
18),
satisfies the equation:
Finally, since
, we have
, and the equation for
v is
Therefore,
satisfies system (
18) with
replacing
, respectively, and
and
for all
From now on, we consider all the functions of system (
18) to have zero average.
Furthermore, since
, the operators
and
, along with periodic boundary conditions, are unbounded, self-adjoint operators with a compact resolvent in
, which are positive when restricted to the space of zero average functions in
. Moreover, the equations for the temperature
T and the solute concentration
S in (
18) are parabolic types for
We express system (
18) as the following evolution system for acceleration, velocity, temperature, and solute concentrations:
That is:
with
and initial data
.
The operator
is a sectorial operator in
with domain
and has a compact resolvent, see Equation (
A5).
Using the results and techniques in the sectorial operator of [
33] to prove the existence of solutions of the system, we have Theorem A1.
Theorem A1. We assume that is locally Lipschitz, , . Then, given , there exists a unique solution of (18), satisfyingfor every In particular, (A7) defines a nonlinear semigroup, in with . Proof. In order to prove this, we cover several steps:
Step (i). We prove the local existence and regularity. This follows easily from the variation of the constants formula as presented in [
33]. In order to prove this, we write the system as (
A8), and we have:
where the operator
B is a sectorial operator in
with domain
, and has a compact resolvent. In this context, operator
must be understood in the variational sense, i.e., for every
,
and
coincides with the fractional space of exponent
, as in [
33]. We denote
as the dual space and
as the norm on the space
. If we prove that the nonlinearity
is well defined, Lipschitz, and bounded on bounded sets, we obtain the local existence for the initial data in
.
Using
as a locally Lipschitz function, with
, we can follow the approach in [
28], and we will prove the nonlinear terms,
in Equations (
A9)–(
A12), satisfying
,
and
; that is,
is well defined, Lipschitz, and bounded on bounded sets.
For
, in this case, we have that
In this way, is locally Lipschitz and bounded on bounded sets.
Applying the techniques of the variation of constants formula from [
33], we obtain the unique local solution
(with a suitable
) of (
A7), which is given by
with
.
where
and using the results from [
33], (the smoothing effect of the equations with the bootstrapping method), we obtain the regularity of solutions.
Step (ii). To prove global existence, we must show that the solutions are bounded in the norm on finite time intervals, and using the nonlinearity of maps bounded on bounded sets, we can conclude.
Part I: We study the norm of temperature, T.
To prove that the norm of
T is bounded on finite time, we multiply the equation for the temperature by
T in
Then, integrating by parts, we obtain:
since
.
By using Cauchy–Schwartz and Young inequalities, as well as the Poincaré inequality for functions of zero average, since
, and as
is the first nonzero eigenvalue of
in
, we obtain
for every
. Now, taking
, we have
and conclude that the norm of
T in
remains bounded in finite time.
Now, we will prove that the norm
remains bounded in finite time intervals. For this, multiply the second equation of (
A7) by
in
. By integrating by parts, applying the Young inequality, and taking into account that
since
is periodic, we obtain
for every
and
. Thus, taking
, and applying the Poincaré inequality for functions with zero average, we obtain
which proves that the norm of
T in
remains bounded in finite time.
Part II: We study the norm of solute concentration S.
In this step, we show that the norm of
S in
does not blow up in finite time. Multiplying the fourth equation of (
A7) by
S, integrating by parts, applying the Young inequality and again taking into account that
, since
S is also periodic, we have
for every
. Thus, taking
, and using (
A19) with the Poincaré inequality for functions with zero average, we obtain
with
. Therefore,
remains bounded in finite time.
Finally, since and are bounded in finite time, this implies that and remain bounded in finite time. Hence, we obtain a global solution in the nonlinear semigroup in □
Appendix A.3. Boundedness of the Solutions: Global Attractor
In this section, we adapt the results and techniques in Refs. [
15,
28] with Refs. [
6,
14] for a fluid with one component, to prove the existence of the global attractor for a binary fluid for the semigroup defined in the space
, in this case, when we consider this transfer law with temperature diffusion.
To obtain the asymptotic bounds on the solutions, as
, we consider the prescribed flux
, satisfying that there exists
, such that
Moreover, we consider the friction function
G, as in [
6,
14,
28], satisfying the hypotheses of the previous section; there exits a constant
, such that
Using the l’Hopital’s lemma proved in [
32], we have the following lemma proved in [
28].
Lemma A1. If we assume that and satisfy the hypotheses of Theorem A1, with (A23), then:with being a positive constant, such that if , and and . Remark A1. We note that the conditions (A23) are satisfied for all friction functions G considered in the previous works, i.e., the thermosyphon models where G is constant, or linear or quadratic laws. Moreover, the conditions (A23) are true for . Theorem A2. Under the above notation and hypotheses of Theorem A1, if we assume that G satisfies (A24) for a constant , and satisfies (A22), then - (i)
- (ii)
- (iii)
- (iv)
Moreover, if , then we also have - (v)
- (vi)
Here, when v satisfies (A29). In particular, we have a global compact and connected attractor in .
Proof. - (i)
From (
A17), using Gronwall’s lemma, for every
, we have
where
is given by (
A22), i.e.,
and we have (
A25).
- (ii)
From (
A19), and working as above, for every
,
then we obtain (
A26).
- (iii)
Using Gronwall’s lemma, from (
A21) with (
A26), we have (
A27).
- (iv)
From system (
A7), if
,
satisfies
this is the same kind of equation for
T. Therefore, working as in (i), we have
and we conclude (
A28).
- (v)
Working as ([
28]), from (
A7), we have
and
satisfies
where
and
. We rewrite (
A35) as
with
For any
, there exits
, such that
for any
, and integrating (
A36) with
, we obtain
Using L’Hopital’s lemma proved in [
32], we obtain the following two results:
for any
and
with
and from (
A38) with (
A24), we conclude for any
Thus, using
with the above results, we obtain (
A29).
(vi) From (
A34) with Gronwall’s lemma, we have
where
. Consequently, for any
, there exits
, such that for any
this is
for any
, and using the above results, we have (
A30).
Finally, since the sectorial operator
B, as defined in Theorem A1, has a compact resolvent, the rest follows from [
34] [Theorems 4.2.2 and 3.4.8]. □
Appendix A.4. Asymptotic Behavior: Reduction to Finite-Dimensional Systems
Finally, in this section, we aim to analyze the long-term behavior of system (
A7) by examining its asymptotic properties. To achieve this, we utilize the Fourier expansions of the functions involved in the system, namely the temperature and solute concentrations. Our focus lies on the Fourier coefficients of the loop’s geometric function, denoted as
, and the prescribed flux across the loop wall, denoted as
. These coefficients play a crucial role in studying the system’s asymptotic behavior. We establish the system’s asymptotic behavior (
A7) through the appropriate Fourier coefficients associated with
f and
h.
To investigate the system’s asymptotic behavior, we consider the coefficients
for temperature and
for the solute concentration, where
represent the
relevant modes. Thus, we obtain a finite system of differential equations:
Furthermore, it is worth noting that the Fourier expansion of any function
g belonging to
, where
, can be expressed as
, with
. Moreover, we have the following expression for the norm of
g in
:
Assume that
are given by the following Fourier series expansions:
with the initial data
are given by
and
is given by
Since all functions involved are real and periodic, we have
,
,
and
.
Proposition A1. Under the above notation and hypotheses of Theorem A1, we consider given by (A46) and the initial data given by and given by Let be the solution of system (18) given by Theorem A1, we then have the following: - (i)
Coefficients and in (A47) satisfy the following equations: - (ii)
The equation for velocity is
Proof. By using the following Fourier expansion for a generic function
of
, the Fourier coefficients
satisfy the following relations:
Consider the model (
18) and the Fourier series expansions of all functions, depending on the spatial variable
x, i.e.,
,
With
given by (
A47),
and
given by (
A46), with the expansions of initial data for temperature
and for solute
, we can easily find that the coefficients for temperature
and solute concentration
are the solutions of (
A48).
Moreover, it is sufficient to note that
since all the functions involved are real and periodic, we have for all
; this allows us to conclude that (
18) is equivalent to infinite systems of ODEs consisting of (
A48) coupled with
□
Remark A2. It is worth mentioning that system (18) is synonymous with system (A7) concerning acceleration, velocity, temperature, and solute concentration. Moreover, based on the proposition above, it can be equated to the following infinite system of ordinary differential Equation (A53): The system of Equation (A53) captures two significant characteristics: (i) the interaction between the modes occurs through the velocity, while diffusion acts as a linear damping term, and (ii) the evolution of temperature impacts the evolution of solute concentration. In the subsequent analysis, we will utilize this explicit equation for the Fourier modes of temperature and solute concentrations to examine the asymptotic behavior of the system and derive explicit low-dimensional models. Similar explicit constructions were presented in various previous studies, such as [
28], and by Bloch and Titi in [
35] for a nonlinear beam equation, where nonlinearity arises solely from the appearance of the
norm of the unknown. Stuart also provided a related construction in [
36] for a nonlocal reaction–diffusion equation.
In the following section, we will establish the boundedness of these coefficients, which enhances, to some extent, the boundedness of temperature and solute concentrations. This, in turn, proves the existence of the inertial manifold for system (
A7), employing similar techniques as in Refs. [
6,
14,
15,
28,
29].
Appendix A.5. Inertial Manifold
We consider the general case
with the heat transfer law across the loop wall given by a prescribed function
, and use inertial manifold techniques, in the spirit of the non-diffusion case of [
37]. In this case, the existence of an inertial manifold does not rely on the existence of significant gaps in the spectrum of the elliptic operator but on the invariance of particular sets of Fourier modes.
Proposition A2. Under the above notation and hypotheses of Theorem A2, with initial conditions , for every solution of system (A7), and for every , and recalling the expansionswith the initial data given by and given by , we have - (i)
and is a positive constant, such that - (ii)
In particular, if , we have the global compact and connected attractor , where are the upper bounds for acceleration and velocity, as given in (A61) and (A60), respectively, and where and is a compact set in . Proof. (i) From (
A48), we have
and taking into account that
we obtain:
and we have (
A54), i.e.,
.
From (
A48), we have
therefore, from (
A65), we have
Finally, taking into account that (
A66) with (
A54), we obtain
and we have (
A55), i.e.,
.
From (iii) in Theorem A2, with
and using (
A54) and (
A55), we have
where
. Hence, from (
A40), we obtain (
A56), namely
and using (
A43), we obtain (
A57), i.e.,
(ii) Using Theorem A2 and taking into account Equation (
A45) with
we find that for any solution of (
A7), in conjunction with (
A58), (
A59), (
A60), and (
A61), given that the sectorial operator
B defined above (in Section 2.1.1) has a compact resolvent from [Theorem 4.2.2 and 3.4.8] in Ref. [
30], the system has a global, compact, and connected attractor,
, in
.
Next, we show that , where , with .
From Equations (
A54) and (
A55), for any
, we have
and
; therefore,
and
, i.e., if
then
, and we have
; that is, we have
.
Finally, we show that is compact in .
Certainly, given any sequence
in
, we can derive a subsequence denoted as
, which exhibits weak convergence to a function
R. Moreover, for any
, the Fourier coefficients satisfy
as
, where
represents the kth Fourier coefficient of
R. Consequently, we have
for every integer
.
since there exists a positive constant
C, such that
where
denotes the norm in
. Hence, the first term goes to zero, such as
, and the second can be made arbitrarily small, such as
, since
with
. Consequently,
and
in
, and the result is proved. □
Now, we will prove that there exists an inertial manifold
(see a definition in Ref. [
38]) for the semigroup
in the phase space
i.e., a submanifold of
, such that
- (i)
for every
- (ii)
there exists
, satisfying that for every bounded set,
there exists
, such that
; see, for example, [
38,
39].
Assume that
with
with
for every
with
since
We denote by
and
the closure of the subspaces of
and
, respectively, generated by
Theorem A3. Assume that and Then, the set is an inertial manifold for the flow of in the space . Moreover, if K is a finite set, the dimension of is where is the number of elements in K.
Proof. Step (i). First, we show that
is invariant. Note that if
then
; therefore, if
from (
A62), we have that
for every
t, i.e.,
, and if
using
from (
A65), we have
for every
t, i.e.,
Therefore, if
then
for every
i.e.,
is invariant.
Step (ii). From previous assertions,
, the flow on
is given by
Now, we consider the following decomposition in , where is the projection of T on and is the projection of T on the subspace generated by i.e., and
Analogously, we consider the decomposition
in
, where
is the projection of
S on
i.e.,
and
. Then, given
, we decompose
, and
and we consider
and
From (
A64), and taking into account that
for
we derive
; moreover, with
for every
with (
A45), it implies that
i.e.,
in
if
.
Moreover, we have
; therefore,
Since
for
from (
A67), we have
Thus,
Therefore,
and
as
with the exponential decay rate
where
. Thus,
attracts
with the exponential rate
in
. □
Remark A3. If from and taking into account of (A45), we have ; Furthermore, we note that, by working as above, we have:and the invariant attracts the solutions inwith exponential rate . Appendix A.6. The Explicit Reduced Subsystem
Under the hypotheses and notation of Theorem A3, we suppose that
with
for every
and
if
. On the inertial manifold
Hence, the behaviors of velocity v and acceleration w are influenced by the Fourier coefficients of T and S within the set . To solve the complete system, we first determine the coefficients of T and S belonging to , and then proceed to solve the equations for the coefficients of T and S, where (i.e., ).
It is worth noting that . Since and , the set contains an even number of elements, denoted as . Therefore, the number of positive elements in , denoted as , is .
Corollary A1. Under the assumptions and notation of Theorem A3, if we assume that the set is finite with , then the asymptotic behavior of system (A7) can be described by a system of coupled equations in . These equations determine the variables for , along with a family of linear non-autonomous equations. Proof. On the inertial manifold:
Consequently, the system’s dynamics rely on the coefficients within . Additionally, the equations for and are conjugates of the equations for and , respectively. This implies that and .
From this, taking real and imaginary parts of
and
in (
A53) with
we conclude. □
Remark A4. Taking the real and imaginary parts of the coefficients of the temperature, , the heat flux at the wall of the loop, , the geometry of the circuit, , and the solute concentration, asthe asymptotic behavior of system (A7) is given by a reduced explicit system of ODEs in with given by Thus, we reduced the asymptotic behavior of the initial system (
A7) to the dynamics of the reduced explicit system (
A72). It is worth noting that, from the above analysis, it is possible to design the geometry of the circuit and/or the external heating source, by properly choosing the functions
f and/or the ambient temperature,
h, so that the resulting system has an arbitrary number of equations of the form
.
Note that K and J may be infinite sets, but their intersection is finite. For instance, for a circular circuit, we have , i.e., , and then is either or the empty set.
If we take, for the sake of simplicity, a circular geometry, then
and
Also, if we take
and omit the equation for
, the conjugate of
k, we have the following transformed set of equations:
where the variables of interest are
, representing the fluid acceleration,
, the fluid velocity,
, the Fourier mode of temperature, and the Fourier mode of solute concentration,
. To simplify the notation, we can omit the subscripts as we only have a single coefficient.
To reduce the number of independent parameters, we introduce a change of variables:
and
. Additionally, we express the equations in terms of their real and imaginary components using the following approach:
with
,
,
,
,
A,
. Finally, we integrate the system to arrive at Equation (
19) in
Section 3.