5.1. Semi-Classical Ideal Gas
Thermal fluctuations and other finite-size effects are often assumed to negligibly alter the average properties of large systems [
65,
66,
67]. However, we now show that finite-size effects may be necessary to find the true thermal equilibrium in systems of any size. First focus on a large volume (
V~1 m
3) containing on the order of Avogadro’s number of atoms (
N~
NA = 6.022 × 10
23 atoms/mole). Assume monatomic atoms at temperature
T with negligible interactions (ideal gas), so that the average internal energy comes only from their kinetic energy,
= 3
N(½
kT). Gibbs’ paradox [
18,
19,
20,
21] is often used to argue that the entropy of such thermodynamic systems must be additive and extensive. Nanothermodynamics is based on assuming standard thermodynamics in the limit of large systems, while treating non-extensive contributions to thermal properties of small systems in a self-consistent manner. Here, we review and reinterpret several results given in chapters 10 and 15 of Hill’s Thermodynamics of Small Systems [
13]. We emphasize that sub-additive entropy, a fundamental property of quantum-mechanics [
23,
68], often favors subdividing a large system into an ensemble of nanoscale regions, increasing the total entropy and requiring nanothermodynamics for a full analysis.
Table 1 gives the partition function, fundamental thermodynamic function (entropy, free energy, or subdivision potential) and other thermal quantities for an ideal gas of mass
m in the four ensembles of
Figure 1, similar to the tables in [
26]. (Subscripts on the entropy and subdivision potential denote the ensemble.) Other symbols used in
Table 1 include the thermal de Broglie wavelength
(where
is Planck’s constant), and the absolute activity
.
Table 1 elucidates several aspects of nanothermodynamics of the ideal gas in various ensembles. The microcanonical partition function comes from the multiplicity of microscopic states that have energy
E. Partition functions in other ensembles come from one or more Legendre transforms to yield other sets of environmental variables. If the transform involves a continuous variable, it should be done using an integral over the variable. However, if the variable is discrete (e.g.,
N), in nanothermodynamics it is especially important to use a discrete summation, thereby maintaining accuracy down to individual atoms, which also often simplifies the math and removes Stirling’s formula for the factorials. Similarly, note that the chemical potential in the canonical ensemble is calculated using a difference equation, not a derivative, so that again Stirling’s formula can be avoided. Another general feature to be emphasized is that the variables shown in the “Ensemble” column are fixed by the environment (e.g., types of walls surrounding a subsystem); hence they do not fluctuate. In contrast, each conjugate variable fluctuates due to contact with the environment, so that these conjugate variables are shown as averages. Thus, as expected for small systems [
14], it is essential to use the correct ensemble for determining which variables fluctuate, and by how much.
Now focus on the entropy. Recall that the Sackur–Tetrode formula for the entropy of an ideal gas is
. Note that to make this entropy extensive, the partition function is divided by
N!, which assumes that all atoms in the system are indistinguishable, usually attributed to quantum symmetry across the entire system. However, the need to use macroscopic quantum mechanics for the semi-classical ideal gas remains a topic of debate [
18,
19,
20,
21].
Table 1 shows that in nanothermodynamics, entropy is non-extensive due to contributions from subtracting the subdivision potential
(see Equation (4)). For example, in the canonical ensemble
, which comes from Stirling’s formula for
N!. Because the Legendre transformation from
N to
μ is done by a discrete sum over all
N, Stirling’s formula is eliminated from the grand-canonical and nanocanonical ensembles. Instead, a novel non-extensive contribution to entropy arises in the nanocanonical ensemble from
. Because this negative subdivision potential is subtracted from
S/
k, the entropy per particle increases when the system subdivides into smaller regions. This entropy increase appears only in the nanocanonical ensemble, where the sizes of the regions are unconstrained, a feature that is unique to nanothermodynamics.
Figure 2 is a cartoon sketch of how net entropy may change if a single system subdivides into subsystems: decreasing if subsystems are constrained to have fixed
V and
N (canonical ensemble), but increasing if subsystems have variable
V and
N (nanocanonical ensemble). As expected, total entropy increases if most atoms can be distinguished by their nanoscale region, even if they may soon travel to other regions to become indistinguishable with other atoms. In fact, for the semi-classical ideal gas, the fundamental requirement of sub-additive quantum entropy [
23,
68] is found only in the nanocanonical ensemble.
The subdivision potentials from
Table 1 can be used to obtain the non-extensive corrections to entropy of specific atoms in various ensembles. As an example, consider one mole (
N = 6.022 × 10
23 atoms) of argon gas (mass
m = 6.636 × 10
−26 kg) at a temperature of 0 C (
T = 273.15 K), yielding the thermal de Broglie wavelength Λ =
≈ 16.7 pm. At atmospheric pressure (101.325 kPa), the number density of one amagat (
N/
V = 2.687 × 10
25 atoms/m
3) gives an average distance between atoms of
3.34 nm, and a mean-free path of
59.3 nm (using a kinetic diameter of
d = 0.376 nm for argon). Under these conditions the Sackur–Tetrode formula predicts a dimensionless entropy per atom of
18.39 (equal to 152.9 J/mole-K). In the canonical ensemble the subdivision potential is positive,
, so that when subtracted from the Sackur–Tetrode formula the entropy is reduced. Although the magnitude of this entropy reduction per atom is microscopic,
= 4.70 × 10
−23, even such a small reduction is used to justify the standard thermodynamic hypothesis of a single homogeneous system. However, the hypothesis breaks down if subsystems are not explicitly constrained to have a fixed size. Indeed,
regions in the nanocanonical ensemble have a sub-additive entropy that increases upon subdivision. Specifically,
is negative when
, confirming that any system of ideal gas atoms favors subdividing into an ensemble of regions whenever the size of each small region is not externally constrained. Thermal equilibrium in the nanocanonical ensemble is usually found by setting
[
57], yielding
and an increase in entropy per atom of:
= 1, about 5.4% of the Sackur–Tetrode component. However, the Sackur–Tetrode formula has been found to agree with measured absolute entropies of four monatomic gases, with discrepancies (0.07–1.4%) that are always within two standard deviations of the measured values [
69]. Thus, the experiments indicate that
>> 1 in real gases, presumably due to quantum symmetry on length scales of greater than 10 nm. For example, if quantum symmetry (indistinguishability) occurs for atoms over an average distance of the mean-free path (
ℓ = 58.3 nm), then
5600 atoms. Now the subdivision potential per atom yields
, well within experimental uncertainty. In any case, nature should always favor maximum total entropy, no matter how small the gain, so that the statistics of indistinguishable particles may apply to semi-classical ideal gases across nanometer-sized regions, but not across macroscopic volumes.
Having for a semi-classical ideal gas implies that many atoms can be distinguished by their local region within the large system. Thus, as expected, a large system of indistinguishable atoms can increase its entropy by making many atoms distinguishable by their location. Furthermore, because the nanocanonical ensemble allows fluctuations around , local regions may adapt their size and shape to encompass atoms that are close enough to collide, or at least to have wavefunctions that may overlap, which is the usual criterion for the onset of quantum behavior.
Figure 3 is a cartoon sketch depicting two ways of mixing gases from two boxes, with the color of each box representing the particle density of each type of gas. The upper-left sketch shows two boxes containing different gases, but with the same volume and particle density, that combine irreversibly with a large increase in entropy due to mixing, whereas the lower-left sketch shows two identical boxes containing the same type of gas that combine reversibly, with negligible change in total entropy. First consider the upper-left picture showing boxes with the same volume
V, but different types of gases. Let one box contain
N1 particles of ideal gas 1, and the other box
N2 =
N1 particles of ideal gas 2, so that when combined, both specific densities are halved, e.g.,
. The Sackur–Tetrode formula yields an increased entropy from mixing:
. This entropy of mixing dominates all ensembles. In fact, because the subdivision potentials in
Table 1 depend on the number of particles in the system, but not on the volume, finite-size effects in the entropy are unchanged by mixing two types of gases. Specifically, for the canonical ensemble:
. Similarly, for the nanocanonical ensemble:
.
Next consider the lower-left picture in
Figure 3 showing identical boxes, each of volume
V with
N1 particles of ideal gas 1. When the boxes are combined, the particle density does not change
. From the Sackur–Tetrode formula for effectively infinite systems of indistinguishable particles, the total entropy also does not change:
= 0. Adding finite-size effects to the canonical ensemble (row 2 in
Table 1), combining identical systems increases the total entropy:
Quantitatively, if each box initially contains one mole of particles, then
N1 = 6.022 × 10
23 yields
27.95. Although the entropy increase per particle is extremely small, any increase in entropy inhibits heterogeneity in bulk systems, supporting the standard thermodynamic assumption of large homogeneous systems. However, this entropy increase applies only to ensembles having subsystems of fixed size. In contrast, combining boxes in the nanocanonical ensemble decreases the total entropy. Specifically, in thermal equilibrium at constant density, both
and
remain constant so that
, yielding a decrease in total entropy when boxes are combined:
. The per-particle entropy change is again extremely small for large boxes, but the inverse process of subdividing into small internal regions should proceed until the increase in per-particle entropy reaches its maximum:
. As previously discussed (following
Figure 2), the fact that real gases do not show such deviations from the Sackur–Tetrode formula [
69] implies
; but any increase in entropy is favored by the second law of thermodynamics, and required by a fundamental property of quantum mechanics for sub-additive entropy [
23,
68]. Moreover, similarly uncorrelated small regions are found to dominate the primary response measured in liquids and solids [
36,
37,
38,
39,
40,
41,
42,
43,
44,
45,
46,
47,
48,
49].
To summarize this subsection, all ensembles yield primary contributions to entropy that match the Sackur–Tetrode formula for combining ideal gases. However, nanothermodynamics allows an ideal gas to maximize its entropy and mimic measured changes in entropy, without resorting to macroscopic quantum behavior for semi-classical ideal gas particles that may be meters apart, and therefore distinguishable by their location. Furthermore, because the nanocanonical ensemble allows the number of particles in a particular region to fluctuate, the number of indistinguishable particles in a specific region may be N >> 1, due to particles that are close enough to collide, or to have coherent wave functions. In any case, nature favors maximizing the total entropy whenever possible using any allowed mechanism. Hence, a novel solution to Gibbs’ paradox comes from including finite-size effects in the entropy of ideal gases, without requiring quantum symmetry for macroscopic systems. This fundamental result stresses the importance of treating energy, entropy, and symmetry across multiple size scales, which requires nanothermodynamics for a fully-accurate analysis.
5.2. Finite Chain of Ising Spins
Simple models of magnetic spins provide a basic scenario for studying finite-size thermal effects between interacting particles. The fundamental equation of nanothermodynamics for reversible processes in magnetic systems is given in
Figure 4. As in Equation (2), the equation in
Figure 4 gives changes in total internal energy of a macroscopic system from changes in total quantities, plus finite-size effects from the subdivision potential,
.
Figure 4 also shows a set of cartoon sketches of energy-level diagrams indicating how various contributions change the internal energy. Each sketch shows three energy levels, with dots depicting the relative occupation of each level. The occupation of these levels for an initial internal energy is shown by the left-most energy-level diagram. The next three energy-level diagrams, from left-to-right, respectively, show that when done reversibly: adding heat (
) alters the relative occupation of the levels, doing magnetic work (
) changes the energy of the levels, while adding spins (
) increases the occupation of all levels. The right-most energy-level diagram represents novel contributions to energy from the subdivision potential (
dη). Inside a system of fixed total size (
Nt), when the number of subsystems increases (
dη > 0), the average subsystem size (
N) decreases, the energy levels may broaden from finite-size effects due to surface states, interfaces, thermal fluctuations, etc. The subdivision potential in nanothermodynamics uniquely allows systematic treatment of these finite-size effects, thereby ensuring that energy is strictly conserved, even on the scale of nanometers.
The Ising model for uniaxial spins (binary degrees of freedom) demonstrates the power and utility of nanothermodynamics for finding the thermal equilibrium of finite-sized systems. Exact results can be obtained analytically in 1-D in zero magnetic field,
B = 0, but first consider
B > 0. Assume
N Ising spins, each having magnetic moment
that can align in the +
B or −
B direction, with interactions only between nearest-neighbor spins. Let the spins favor ferromagnetic alignment, so that the energy of interaction (exchange energy) is −
J if the two neighboring spins are aligned, and +
J if they are anti-aligned. The usual solution to the 1-D Ising model includes contributions to energy from
B and from the exchange interaction, yielding the partition function [
2,
3]
If
B = 0, the resulting free energy becomes
The approximations in Equations (12) and (13) come from assuming large systems with negligible end effects, or equivalently spins in a ring. However, most real spin systems do not form rings, so that these equations are valid only in the limit of large systems, . We now address finite-size effects explicitly.
Consider a finite linear chain of
N + 1 spins, yielding a total of
N interactions (“bonds”) between nearest-neighbor spins [
70]. It is convenient to write the energy in terms of the binary states of each bond,
bi = ±1. Using +
J for the energy of anti-aligned neighboring spins, and −
J between aligned neighbors. The Hamiltonian is
Assuming
x high-energy bonds (
bi = −1), with (
N −
x) low-energy bonds (
bi = +1), the internal energy is
. The multiplicity of ways for this energy to occur is given by the binomial coefficient
The factor of 2 in Equation (15) is needed to accommodate both alignments of neighboring spins for each type of bond. The thermal properties of this finite-chain Ising model in various ensembles are given in
Table 2. Note that although the summation for the nanocanonical ensemble starts at
, because the number of spins is
every region contains at least one spin, as required for spontaneous changes in the number of subsystems [
57]. Additionally, note that due to end effects, the Helmholtz free energy from
Table 2 for
bonds (
N spins) is
, approaching Equation (13) only when
. Thus, if an unbroken chain is forced to have a macroscopic number of spins, all ensembles yield similar results. However, if the length of the chain can change by adding or removing spins at either end, thermal equilibrium requires the nanocanonical ensemble. As with the ideal gas, this nanocanonical ensemble is the only ensemble that does not externally constrain the sizes of the regions, so that the system itself can find its equilibrium average and distribution of sizes. From
Table 2, setting the subdivision potential to zero yields an average number of spins in each chain of:
. Thus, at high temperatures the average chain contains two spins connected by one bond, whereas when
the average chain length diverges.
As expected,
Table 2 shows that the entropy of Ising spins increases with decreasing constraints, so that again (as in
Table 1) the nanocanonical ensemble has the highest total entropy. Specifically, the entropy per bond in the nanocanonical ensemble exceeds that in the canonical ensemble by the difference
. At high
T where
,
. At low
T where
,
. Numerical solution yields a maximum entropy difference of nearly 6% (
= 0.0596601…) at
kT/
J = 0.687297… where
= 2.25889… Hence, Ising spins in the nanocanonical ensemble always have higher entropy than if they were constrained to be in the canonical ensemble, but the excess is small at both low, and high
T. Nevertheless, if a mechanism exists to change the length of the system, an infinite chain will shrink until there is on average
spins in each region, thereby maximizing the entropy of system plus its environment with no external constraints on the internal heterogeneity. In fact, because it can be difficult to fix the size of internal regions, their size should vary without external constraints, limiting the usefulness of the canonical ensemble for describing finite-size effects inside most real systems.
A key feature of the nanocanonical ensemble is that thermal equilibrium is found by setting the subdivision potential to zero [
57]. Indeed,
ensures that the system finds its own equilibrium distribution of regions, without external constraint, similar to how
μ = 0 in standard statistical mechanics yields the equilibrium distribution of phonons and photons, without external constraint. Specifically, because
is the change in the total energy with respect to the number of subsystems, spontaneous changes in
η occur unless
. However,
requires
without any normalization, so that all factors must be carefully included in the partition function. For example, suppose that the factor of 2 in the numerator of Ω (Equation (15)) is ignored from neglecting the degeneracy of each sequence of spins and its inversion. Because averages in the canonical ensemble (e.g.,
) are normalized by the partition function, they do not change, but the average number of bonds in the nanocanonical ensemble becomes
, twice the value of
from
Table 2.
5.3. The Subdivided Ising Model: Ising-Like Spins with a Distribution of Neutral Bonds
Results similar to those for the finite-size Ising model in the nanocanonical ensemble (
Section 5.2) can be obtained in the canonical ensemble by modifying the Ising model to include “neutral bonds,” from nearest-neighbor spins that do not interact. (Our model differs from dilute Ising models [
71] that assume empty lattice sites at fixed locations.) Physically, neutral bonds may come from neighboring spins having negligible quantum exchange (which suppresses their interaction), or from neighboring spins with uncorrelated fluctuations so that their interaction is time averaged to zero. Again, start with the standard Ising model having
spins (
N bonds), but now let there be
η’ neutral bonds (yielding
η =
η’ + 1 subsystems). In addition, let there be
x high-energy bonds between anti-aligned spins, leaving
N −
η’ + 1 −
x low-energy bonds between aligned spins.
Figure 5 shows a specific configuration of 11 spins (
N = 10 bonds) with
x = 2 high-energy bonds (
X),
η’ = 3 neutral bonds (
O), and
N −
η’ −
x = 5 low-energy bonds (●). It is again convenient to write the energy in terms of the bonds, which may now have three distinct states, yielding the Hamiltonian
The internal energy of the system is
. The canonical ensemble involves two sums. The first sum is over
x for fixed
η’, with a multiplicity given by the trinomial coefficient for the number of ways that the high- and low-energy bonds can be arranged among
N −
η’ interacting bonds. An extra factor of 2
η’ arises because each neutral bond has two possible states for its neighboring spin. This first sum yields a type of canonical ensemble for the system with fixed
η’. A second sum is over all values of
η’. The multiplicity is given by the binomial for the number of ways that the neutral bonds can be distributed, which arises from the trinomial after the first summation. The behavior of this model is summarized in
Table 3.
We now compare the results in
Table 3 for the subdivided Ising model with those from
Table 2 for the finite-size Ising model. In the canonical ensemble, the average energy of the subdivided system is higher (not as negative) as that of the finite system, as expected when neutral bonds replace an equilibrium mixture of predominantly low-energy bonds. However, the average energy per interacting bond (
from
Table 3, divided by
) is
, the same as
from
Table 2. Another similarity comes from using the average number of subsystems,
, to obtain the average number of spins in each region,
Hence, in the limit of
, Equation (17) gives
for the average number of bonds in the subdivided Ising model in the canonical ensemble, approaching
from
Table 2 for the finite-size Ising model in the nanocanonical ensemble. In other words, if the initial system is large enough, both approaches to nanothermodynamics are equivalent: a large ensemble of small systems (
Table 2) and a large system that is repeatedly subdivided into independent subsystems (
Table 3). Furthermore, models with distinct Hamiltonians—Equation (16) here for a system of three-state bonds and Equation (14) for a system of two-state bonds—may yield equivalent results. However, equivalence requires that the correct ensemble is used for each system, canonical ensemble here for the subdivided Ising model, and nanocanonical in
Section 5.2. Thus, the choice of ensemble is crucial for obtaining fully-accurate behavior, even for systems that are in thermal equilibrium and in the thermodynamic limit.
5.5. Entropy and Heat in an Ideal 1-D Polymer
Although adiabatic demagnetization provides a well-known connection between entropy and heat in spin systems [
4,
41,
72], this connection is often more familiar in the context of ideal polymers. Furthermore, the basic behavior of the polymer can be experienced at home using a rubber band [
73]. Here, for a simplified analysis related to the 1-D Ising model, we treat an ideal polymer comprised of freely jointed monomers (units) in 1-D [
6]. Consider a polymer of
N units, each of length
a. Let one end of the 1st unit be freely jointed (free to invert) about the origin (
X = 0), with the far end of the
Nth unit unconnected. All other units have both ends freely jointed to neighboring units. For simplicity assume all units are uniaxial (1-D), with
x segments pointing in the –
X direction and
N −
x segments in the +
X direction. The free end of the polymer is at position
X(
x) = (
N − 2
x)
a. The multiplicity matches that of Equation (18a) for the standard Ising model, and
Table 2 gives the resulting microcanonical entropy
. The elastic restoring force from the entropy [
1] is:
F = −
T ΔS/
ΔX Note that this model involves differences (not differentials) because it is comprised of discrete polymer units in 1-D. Such discrete differences circumvent Stirling’s formula, improving the accuracy, especially for small systems. For an incremental shortening of the polymer
ΔX = −2
a, using half integers to best represent the average values at each integer, the change in configurational entropy of the polymer is
. Solving for the average number of negatively aligned units as a function of
F gives
, yielding the equilibrium endpoint of the polymer
. At high-temperatures
X(
) ≈
, similar to the standard expression for the ideal 1-D polymer if
N >> 1.
This model shows a common characteristic of polymers under tension: their average length varies inversely proportional to temperature. This decrease in
X with increasing
T, opposite to the behavior of most other solids, arises from the dominance of configurational entropy in polymers. Such length contraction can be observed by heating a rubber band that holds a hanging mass, demonstrating a simple conversion of heat into work. The converse conversion of work into heat can be experienced by the increased temperature of a rubber band that is rapidly stretched while in contact with your lips. Cyclic heat-to-work conversion is shown by the heat engine made from a wheel with rubber-band spokes, where an incandescent lamp causes the wheel to rotate continuously [
73]. The heat-to-work mechanism comes from increased thermal agitation around the polymer, increasing its entropy and coiling it more tightly. Similarly, the work-to-heat conversion involves energy added to the heat bath when entropy is reduced as the polymer is stretched.
The purpose of this brief digression is to emphasize how changing the configurational entropy of a polymer by changing its length alters the energy of the heat bath, thus altering the entropy of the bath. We assume that an analogous change in the entropy of the polymer during a fluctuation causes a similar exchange of entropy with the bath. In other words, we assume that the heat bath cannot discern whether changes in entropy of a polymer are due to external forces, or internal fluctuations. Next, we assume that equilibrium (reversible) fluctuations occur with no net loss in entropy, so that the second law of thermodynamics is strictly obeyed. Specifically, we make the ansatz that during equilibrium fluctuations the entropy of the polymer plus the entropy of its local heat bath (
SL(
X)) never deviate from a maximum value:
We now expand on the concept of local heat baths in nanothermodynamics [
35,
74] to show how Equation (19) facilitates reversible fluctuations. In standard thermodynamics, reversible processes must proceed at infinitesimal rates, allowing the system to couple uniformly to the effectively infinite heat reservoir. However, many thermal fluctuations are fast and heterogeneous. Nanothermodynamics is based on independent small systems, often with energy and entropy isolated from neighboring systems, consistent with the energy localization and local
T deduced from experiments [
36,
37,
38,
39,
40,
41,
42] and simulations [
50]. For the specific model presented here, imagine a large sample containing a polymer melt. Let independent polymers (or their independent monomers [
49,
58]) and a local heat bath occupy a nanoscale volume inside the sample. During fast fluctuations, each volume is effectively isolated from neighboring volumes, conserving local energy and local entropy (Equation (19)), characteristic of the microcanonical ensemble (upper-left boxes in
Figure 1). Note, however, that for fluctuations about equilibrium, the microcanonical boxes would have a distribution of sizes and shapes, basically a frozen snapshot of the lower-right regions in
Figure 1. During sufficiently slow fluctuations, energy and particles can transfer freely between variable volumes to maximize the total entropy and maintain a thermal equilibrium distribution of regions in the nanocanonical ensemble. Thus, accurate evaluation of thermal fluctuations often involves two ensembles, the fully-closed ensemble for fast fluctuations, and the fully-open ensemble for slow fluctuations. Partially-open ensembles (e.g., canonical and grand-canonical), which restrict the exchange of some quantities but not others, are often artificially constrained. Specifically, because excess energy is persistently localized during the primary response in liquids, glasses, polymers, and crystals [
36,
37,
38,
39,
40,
41,
42], sometimes for seconds or longer, particle exchange will usually accompany these slow changes in energy. Therefore, the relatively fast transfer of energy needed for a well-defined local temperature, without also changing size and particle number, is unlikely for the primary fluctuations inside most realistic systems.
We now evaluate equilibrium fluctuations that include energy from configurational entropy, which is often ignored in standard fluctuation theory. Let there be no external force on the polymer, so that the average position of its endpoint is
, corresponding to its maximum configurational entropy. From Equation (19), as
S(
X) decreases when
,
SL(
X) must increase, and vice versa. Quantitatively, using the microcanonical entropy for the polymer (adapted from
Table 2), allowing fast fluctuations that are localized and reversible, the entropy of the local bath is:
We assume that Boltzmann’s factor, commonly used to weight large-reservoir states, also weights the local-bath states
. Thus, when thermally averaged, every length of the polymer is equally likely,
. In other words, maintaining maximum entropy during equilibrium fluctuations removes degeneracies from systems of classical particles that have the same macrostate (e.g., same
X,
N, or
E), mimicking the statistics of indistinguishable particles. In previous work it has been shown that removing the alignment degeneracy from systems with the same
X yields 1/
f-like noise, with several features that match the low-frequency fluctuations measured in metal films and tunnel junctions [
17,
74,
75,
76]. Here, we describe how removing the energy degeneracy from systems with the same
E yields similar 1/
f-like noise at lower frequencies, combined with Johnson-Nyquist-like (white) noise at higher frequencies.
5.6. Simulations of Finite Chains of Effectively Indistinguishable Ising-Like Spins
We explore consequences of including contributions from configurational entropy in the total energy during equilibrium fluctuations. The manner in which we add this entropy reduces the degeneracy of most energy states, mimicking the statistics of indistinguishable particles, Equation (18b). We study the 1-D Ising model using Monte Carlo (MC) simulations for the dynamics. The 1-D Ising model is used for simplicity, having its multiplicity of energy states given exactly by the binomial coefficient in Equation (18a). Although MC simulations are too simplistic for microscopic dynamics, they can accurately simulate slow thermal processes around equilibrium [
77,
78]. A novel ingredient in our simulations is to introduce a type of orthogonal dynamics, where changes in energy are independent of changes in alignment. Specifically, each MC step conserves energy, or alignment, with no step allowing both to change simultaneously. Such constraints on the dynamics can be justified by the fact that energy and alignment contribute to distinct thermodynamic variables, and each is governed by a separate conservation law. Analogous decoupling of degrees of freedom has been found in supercooled fluids [
79,
80]. This orthogonal Ising model yields a combination of 1/
f-like noise at low frequencies, and white noise at higher frequencies, similar to behavior often found in nature.
We start with a finite chain of Ising spins (
Section 5.2), with ferromagnetic interaction
J between nearest-neighbor spins. The Hamiltonian is given by Equation (14). Consider a state containing
x high-energy bonds and
N −
x low-energy bonds. The interaction energy of this state is
, and its multiplicity is the binomial coefficient
. Equilibrium behavior of this model in various thermodynamic ensembles is given in
Table 2, but these results restrict Boltzmann’s factor to include only the internal energy from interactions. We now explore how adding the energy from configurational entropy alters the behavior.
The Metropolis algorithm is often used to efficiently yield the Boltzmann distribution of energy states in MC simulations. A standard MC simulation of the Ising model involves choosing a spin at random, then inverting the spin if its change in interaction energy (Δ
E) meets the Metropolis criterion
where [0,1) is a random number uniformly distributed between 0 and 1. This criterion comes from energy transfer with an ideal (effectively infinite and homogeneous) heat reservoir due to changes in interaction energy; but crucial sources of energy from configurational entropy and the local thermal bath are neglected. We therefore add a second criterion that must also be met if configuration changes
where Δ
S(
X) =
SL(
X). Note that Equation (22) favors high entropy in the local bath, just as Equation (21) favors high energy (and hence high entropy) in a large reservoir. Additionally, note that Equation (19) gives
SL(
X) =
Smax −
S(
X), the offset between the maximum configurational entropy and its current value, not just the change in entropy between initial and final states. Justification comes from the assumption that fast fluctuations involve local properties that do not have time to couple to the large reservoir, and even localized thermal process must not diminish the total entropy. Furthermore,
S(
X) for finite systems involves nonlinear terms that cannot be reduced to linear differentials.
Symbols in
Figure 6 show histograms of the energy from the Ising model at four temperatures, using the standard Metropolis algorithm (open), and when contributions from configurational entropy are added (solid). Solid curves come from the integrand in the numerator of Equation (18a), showing that using Equation (21) alone yields the statistics of distinguishable states, whereas dashed lines come from the integrand in the numerator of Equation (18b), showing that adding Equation (22) yields behavior characteristic of the statistics of indistinguishable states.
We now describe a dynamical sequence for simulating the Ising model which separates energy-changing steps from alignment-changing steps, thereby separating the fundamental laws of conservation of energy and conservation of angular momentum. For this orthogonal Ising model, alignment is conserved using Kawasaki dynamics [
81], where neighboring spins exchange their alignments, or equivalently they exchange their locations without changing their alignments. This exchange always conserves the net alignment, but net energy often changes. Alternatively, alignment is changed without changing energy by inverting only spins that have oppositely oriented neighbors. The simulation proceeds by first choosing a spin at random from the chain, then randomly choosing whether to attempt a spin exchange or a spin flip. If spin flip is chosen, and only if the spin’s neighbors are oppositely aligned, then the spin is inverted, changing the net alignment without changing the interaction energy. If instead spin exchange is chosen, and if the configurational-entropy criterion is met (Equation (22)), then one of its nearest-neighbor spins is chosen at random. If exchanging the alignments of these two spins also meets the interaction-energy criterion (Equation (21)) then spin exchange occurs, always preserving net alignment, but often changing the interaction energy.
Solid lines in
Figure 7 show frequency-dependent power spectral densities (
S(
f)) from simulations of the orthogonal Ising model (solid lines) and from measured flux noise in a qubit (symbols) [
82]. Simulated
S(
f) comes from the magnitude squared of the Fourier transform of time-dependent fluctuations in alignment. Note the general feature that large chains exhibit a low-frequency 1/
f-like regime that crosses over to a white noise regime at higher frequencies. Thus, this model has a single thermodynamic variable that exhibits both types of thermal noise that are usually found together in nature. The basic mechanism involves slowly-fluctuating energy (with 1/
f-like noise due to the entropy-change constraint), which slowly modulates an envelope that limits the maximum amplitude for the fast-fluctuating alignment. The orthogonal dynamics is crucial to prevent all other intermixing between distinct thermodynamic variables.
Broken lines in
Figure 7 show linear fits to the 1000-bond chain for the 1/
f-like (dashed) and white (dotted) noise regimes. The intersection of these lines (marked by an arrow) yields the crossover frequency,
fc(1000). The inset in
Figure 7 shows the chain-size dependence of this
fc. Three distinct features shown by the simulations in
Figure 7 mimic measured noise in quantum bits [
82,
83]: 1/
f-like noise with a slope of magnitude 0.92 ± 0.02;
S(
f) in smaller chains with discrete Lorentzian spectra; and white noise at higher frequencies.
Figure 7 also shows that there are two ways to reduce low-frequency noise in the orthogonal Ising model. Specifically, at log(
f/f0) = 4 maximum noise occurs when
N ≈ 50. Noise decreases for larger
N as
fc shifts to lower frequencies, and decreases for smaller
N as 1/
f-like noise saturates at low frequencies in small systems, avoiding the divergence as
f → 0 [
76]. Thus,
Figure 7 shows that the orthogonal Ising model has fluctuations in alignment that yield measured frequency exponents for 1
/f-like noise, a crossover to white noise at higher
f, and discrete Lorentzian responses; three distinct features that mimic measured spectra.