1. Introduction
Motivation
Across quantum mechanics (QM), thermodynamics (TD), and general relativity (GR), the central equilibrium relations—Born’s rule, entropy increase, and the Einstein balance—are typically imposed as end states rather than derived as consequences of dynamics. This leaves open a structural question: under which deterministic conditions do physical systems return to these relations after perturbations, and with what rates?
We study a partly sector-agnostic mechanism. The core is the same everywhere: track a quadratic misfit between a statistical baseline and a physical response, and show that—under standard spectral-gap or coercivity hypotheses—this misfit is a Lyapunov residual that decays monotonically (often exponentially). This form (propagation plus gap/coercivity) is invariant across sectors. What is not agnostic are the inputs and outcomes: the constants and the equilibrium set (the “attractor”) depend on context. In the quantum/operator-algebraic setting they are set by the pointer algebra and the reversible QMS gap; in PDE models by the coercive closure and domain/boundary conditions; in free-field quantization by the Hamiltonian gap; and on geometric slices by the gauge and the Lichnerowicz-type gap. Thus, the DSFL provides a uniform template with sector-specific rates and targets.
Idea
We introduce the
Deterministic Statistical Feedback Law (DSFL). Let
P denote the sector-specific response (e.g. flux, current, or geometric tensor) and let
represent the statistical baseline. The alignment residual
quantifies misfit. DSFL asserts that
acts as a Lyapunov functional:
with a rate
determined by the sector (spectral gap, coercivity constant, or Hamiltonian gap). A common propagation step (Jensen convexity, Kadison–Schwarz, or an energy identity) yields global monotonicity; a spectral gap or coercivity upgrades this to exponential decay. Within this template, Born alignment, entropy growth, and Einstein balance arise as attractors of the same residual-suppression mechanism.
What Is New
One residual, one propagation step. A single quadratic residual is used across operator-algebraic QM, finite-dimensional Lindblad dynamics, coercive PDE flows, and free-field stochastic quantization, with a unified propagation lemma ensuring monotonicity.
Reversible QMS. For -symmetric quantum Markov semigroups, we prove DSFL is equivalent to a noncommutative Poincaré (spectral-gap) inequality on the orthogonal complement of the fixed-point algebra, with optimal rate .
Finite-dimensional Lindblad. For pure dephasing generators, the Lüders off-diagonal variance decays at the sharp rate , where .
PDE template. An exact residual energy identity gives under quantitative coercivity () and subcritical couplings; refinements use Helmholtz decomposition and boundary conditions.
Free-field QFT. In Parisi–Wu stochastic quantization, smeared two-point residuals decay at twice the Euclidean Hamiltonian gap, deriving a DSFL inequality in the Gaussian sector.
Geometric slice analogue. On compact Riemannian slices (DeTurck gauge), a Lichnerowicz-type spectral gap implies exponential suppression of the curvature–matter misfit toward Einstein balance. (A covariant Lorentzian version is left open.)
Residual-entropy monotone. The proxy is strictly increasing whenever DSFL holds, providing a gap/coercivity-controlled “arrow of time.”
1.1. Position Relative to Prior Work
Variance and entropy decay for reversible diffusions (Bakry–Émery), spectral gaps for reversible QMS, modewise Lindblad contraction, and coercivity in geometric/dissipative flows are well established. Our contribution is to (i) isolate a single Lyapunov residual and a unified propagation step spanning these settings, and (ii) give sharp equivalences and rates that make sectoral relaxation theorems directly comparable. We do not introduce a new sector-specific model; we identify and quantify the common restoration law that many models already instantiate under standard gap/coercivity assumptions.
Scope and Limits
Our results are form–invariant across sectors but rely on context–specific assumptions. Throughout we require either reversibility or quantitative coercivity: (i) –symmetric (reversible) QMS with a spectral gap on ; (ii) coercive PDE closures with and subcritical couplings; (iii) free fields (Gaussian sector) with a positive Euclidean Hamiltonian gap; and (iv) geometric DeTurck slices with a Lichnerowicz–type gap on the physical subspace and small initial data. Under these hypotheses the residual obeys a DSFL inequality with an explicit (contextual) rate.
We do not claim theorems for: nonreversible/hypocoercive QMS (skew generators), fully covariant Lorentzian evolutions (hyperbolic, gauge–invariant DSFL), or interacting QFT beyond the Gaussian sector. These are stated as programs with testable intermediate predictions (e.g., scale–dependent rates, ISS bounds) but remain open.
The numerical section provides reproducible protocols rather than data–driven claims; the analytic results stand independently of numerics.
2. Background and Related Work
2.1. Quantum Markov Semigroups and Spectral Gaps
A (normal, unital) quantum Markov semigroup (QMS)
on a von Neumann algebra
with faithful normal state
is called
ω–symmetric (reversible) if it is self–adjoint on the GNS
space. In this setting the generator
admits a densely defined, closed Dirichlet form
with standard functional–analytic underpinnings for noncommutative Dirichlet forms and symmetric quantum semigroups (e.g. [
1,
2]). The fixed–point algebra
is the natural
equilibrium subspace; under modular invariance, the
–preserving conditional expectation
exists and is unique (Takesaki’s theorem; cf. Tomiyama) [
2,
3].
A
noncommutative Poincaré (spectral–gap) inequality with constant
,
is equivalent to exponential decay of the noncommutative variance along the semigroup,
with optimal rate
[
4,
5,
6]. This mirrors the classical theory of reversible diffusions (Bakry–Émery calculus). Beyond Poincaré, quantum log–Sobolev/hypercontractive regimes and related mixing bounds are available (e.g. [
7,
8]), and rapid mixing for quantum channels/expanders provides a complementary discrete perspective [
9]. In our DSFL interpretation, (
1)–(
2) constitute the operator–algebraic Lyapunov law for the residual
: once a gap holds, the residual contracts at rate
and
is the attractor.
2.2. Lindblad Dephasing and Modewise Contraction
In finite dimensions, completely positive trace–preserving dynamics admit the GKSL (Lindblad) representation
with Hamiltonian
H and noise operators
[
10,
11]; see also [
12,
13]. For
pure dephasing in a fixed orthonormal basis,
and
H diagonal, one obtains
so coherence
decays as
while populations are conserved. The Hilbert–Schmidt variance off the pointer algebra (Lüders residual) thus satisfies
and the rate
is
sharp. This is the finite–dimensional counterpart of (
2), with a gap set by the smallest dephasing rate.
2.3. Coercive PDE Flows and Bakry–Émery Tools
In classical dissipative PDEs, exponential return to equilibrium combines: (i) an energy/entropy identity for a nonnegative functional along solutions, and (ii) a
coercive functional inequality (Poincaré/log–Sobolev) to control lower–order couplings. Foundational results include Gross’s log–Sobolev ⇔ hypercontractivity [
14] and its many developments [
15,
16]. In geometric analysis (e.g. Ricci/DeTurck flows),
–type curvature residuals dissipate under an elliptic Lichnerowicz–type operator; Perelman’s monotone functionals provide a celebrated Lyapunov structure [
17,
18,
19].
Our DSFL–PDE template mirrors this structure: for
,
so that, under uniform ellipticity
and subcritical couplings, Grönwall yields
. This is the Bakry–Émery “propagation + gap ⇒ decay’’ pattern reinterpreted as sector–agnostic Lyapunov suppression of a misalignment functional.
2.4. Stochastic Quantization and Hamiltonian Gaps
Parisi–Wu stochastic quantization gives an analytically tractable Euclidean dynamics for QFT: for a free scalar with action
, the Langevin flow
generates a semigroup
with nonnegative Euclidean Hamiltonian
H. The spectral gap
controls exponential relaxation of smeared correlators; squaring gives decay of quadratic residuals at rate
[
20]. For reviews and field–theory context see [
21,
22]. In DSFL language, the same Hamiltonian gap that governs Euclidean relaxation drives residual suppression in the Gaussian sector; extending to interacting fields requires nonperturbative functional inequalities or RG control.
2.5. Positioning vs. Prior Approaches
Exponential return to equilibrium is well established
within specific formalisms: noncommutative Poincaré/log–Sobolev inequalities for reversible QMS [
4,
5,
6,
7,
8], explicit modewise contraction in GKSL/Lindblad dynamics [
10,
11,
12,
13], coercivity–driven decay for dissipative PDEs and geometric flows [
14,
15,
16,
17,
18,
19,
23], and Hamiltonian–gap relaxation in stochastic quantization [
20,
21,
22].
The present work contributes a
single, sector–neutral Lyapunov residual and a unifying propagation principle (Jensen/Kadison–Schwarz/energy identity) that render these results
structurally comparable. Under the corresponding gap/coercivity hypotheses, one obtains the same DSFL inequality
, and the sectoral equilibrium statements—Born alignment (QM), residual–entropy monotonicity (TD), and Einstein balance (GR)—emerge as
attractors rather than axioms. (In our pointer–algebra formalism, POVMs and their Naimark dilations are standard [
24,
25,
26].)
3. Notation and Conventions
3.1. Spaces, Norms, Inner Products
Domains and Measures.
denotes either a bounded domain or a flat torus . We write for Lebesgue measure, for its volume, and for spatial averages. When needed, denotes a smooth Riemannian (or, where explicitly stated, Lorentzian) manifold with volume form .
Lebesgue and Sobolev Spaces.
For , has norm (essential supremum for ). For , is the Sobolev space with . We write for the closure of in . Vector/tensor–valued spaces are denoted and with the product norms.
Inner Products and Norms.
On , for scalars and for vectors. On , for symmetric 2–tensors we use and .
Gradients and Divergences.
In
, ∇ is the Euclidean gradient and
the divergence. On
,
is the Levi–Civita covariant derivative and
its metric divergence. For a vector field
P and a scalar
, we define the
alignment residual
Weighted Inner Products.
Given a measurable, symmetric positive–definite weight , set and . The global weighted norm is . Unless stated otherwise, .
Noncommutative Conventions.
For a von Neumann algebra with faithful normal state , the GNS inner product is and . The noncommutative variance is . If is a von Neumann subalgebra invariant under the modular group of , denotes the (unique) –preserving conditional expectation; it acts as the –orthogonal projection onto .
3.2. Measures, Domains, and Boundary Conditions
Flat Domains.
Unless stated otherwise, is either a bounded domain or a flat torus . We write for the Lebesgue measure and for spatial integrals; spatial averages are .
Boundary Conditions (BCs).
Energy identities and integrations by parts are justified under either (i)
periodic BCs on
, or (ii) homogeneous
no-flux BCs arranged so that boundary terms vanish in the residual energy balance. In particular,
so the boundary contribution is zero if, for example,
on
(e.g.
), or on a torus.
Probability Measures.
In data–driven formulations we allow a time–indexed family of probability measures
on
. Population expectations are
, with empirical approximations
when
is supported on samples
. The residual
reduces to
(with
) when
is normalized Lebesgue and
.
Manifolds.
In geometric sections we replace by a Riemannian (or, where explicitly stated, Lorentzian) manifold with Levi–Civita connection ∇, metric pairing , and volume form . For symmetric 2–tensors T, the norm is . On compact Riemannian slices in DeTurck gauge (Sec.), boundary terms vanish by compactness; on noncompact manifolds we assume decay/compatibility so that all integrals are finite.
Normalization and Gauges.
In probabilistic sectors we impose (densities) and (wave functions). The baseline is defined up to an additive gauge , which leaves and invariant. In geometric sectors, gauge choices (e.g. DeTurck) ensure ellipticity/hyperbolicity and do not alter the definition of the residual (e.g. ).
3.3. Operators, Semigroups, and Spectra
Linear Operators and Spectra.
Let
be a densely defined, closed linear operator on a Hilbert space
. We write
for its spectrum and
for its kernel. If
is self–adjoint and nonnegative, a
spectral gap means that
for some
; then
is the optimal Poincaré constant on
.
Markov/Contraction Semigroups.
A strongly continuous one–parameter semigroup
on
H with generator
is a
(sub)Markov contraction on a Banach lattice
if it preserves positivity, mass (
), and satisfies
for
. In the reversible case (self–adjoint on
), the Dirichlet form is
and
Quantum Markov Semigroups (QMS).
On a von Neumann algebra
, a normal unital completely positive (u.c.p.) semigroup
that is
–symmetric is a noncommutative analogue of a reversible diffusion. The GNS
inner product is
, the Dirichlet form is
, and the fixed–point algebra
is the equilibrium subspace. The noncommutative Poincaré inequality
is equivalent to exponential variance decay
with optimal rate
.
Poincaré and Log–Sobolev constants.
We use Poincaré constants to control variance by energy, and, where applicable, log–Sobolev constants to control entropy by Fisher information. In this article the DSFL rate is identified with twice a Poincaré gap in reversible settings (classical/QMS) and with quantitative coercivity constants in PDE/geometric settings.
Projection onto Equilibria.
denotes the –preserving conditional expectation onto the fixed–point (pointer) algebra . In PDE/probability sectors, the analogue is projection onto the nullspace of the generator (e.g., ).
Spectral Notation in GR Slices.
For Lichnerowicz–DeTurck type operators acting on symmetric 2–tensors on a compact Riemannian 3–manifold , we denote by the spectral gap on the orthogonal complement of the gauge directions; it controls exponential decay of the geometric residual.
3.4. Residuals and Entropy Proxies
DSFL Residuals.
The global DSFL residual in the classical/PDE sector is
or its weighted variant
. The noncommutative analogue is
(QMS). On Riemannian slices we use the geometric residual
Initial Sameness and Common Ancestry.
All physical sectors—quantum, thermodynamic, and geometric—descend from a shared initial alignment between statistical and physical structures. We call this the
principle of common ancestry: at
, the universe (or any closed system) possessed a single residual-free configuration
This expresses the state of complete statistical–physical identity (initial sameness) from which all later structures evolve.
Figure 1.
The principle of common ancestry: evolution as deterministic recovery of initial sameness.
Figure 1.
The principle of common ancestry: evolution as deterministic recovery of initial sameness.
As evolution proceeds, local perturbations generate misalignments,
, which define the
residual
where
is the sectoral spectral gap (quantum, thermodynamic, or geometric). The
Deterministic Statistical Feedback Law (DSFL) ensures exponential suppression of these residuals, restoring the alignment that encodes shared ancestry:
Hence equilibrium is not a probabilistic emergence from randomness, but the dynamic recovery of common ancestry through deterministic residual decay. The same Lyapunov structure appears in all sectors— Born alignment in quantum mechanics, entropy growth in thermodynamics, and curvature–matter balance in general relativity—each governed by its own contextual rate .
3.5. Abbreviations (DSFL, sDoF, pDoF, QMS, etc.)
| DSFL |
Deterministic Statistical Feedback Law (global Lyapunov law for the alignment residual). |
| sDoF / pDoF |
Statistical / Physical Degrees of Freedom (; p or P the response field). |
|
Residual /
|
Quadratic misalignment functional; pointwise , global . |
|
Residual–entropy proxy . |
| QMS |
Quantum Markov Semigroup (normal, unital, completely positive—u.c.p.— semigroup on a von Neumann algebra). |
| GKSL |
Gorini–Kossakowski–Sudarshan–Lindblad form (finite–dimensional generator of CPTP dynamics). |
|
–preserving conditional expectation onto the fixed–point (pointer) algebra . |
|
Dirichlet form for an –symmetric QMS. |
| Poincaré gap |
Optimal constant in (on ). |
| LSI |
Log–Sobolev constant controlling entropy by Fisher information (used where applicable). |
| OA |
Operator–algebraic (framework of von Neumann algebras / Dirichlet forms). |
| QM / TD / GR / QFT |
Quantum Mechanics / Thermodynamics / General Relativity / Quantum Field Theory. |
| PDE |
Partial Differential Equation (coercive gradient–relaxation template for DSFL). |
| OU |
Ornstein–Uhlenbeck (free–field Euclidean semigroup in stochastic quantization). |
| GNS |
Gelfand–Naimark–Segal Hilbert space associated to . |
| CPTP |
Completely Positive Trace–Preserving (quantum channels/evolutions). |
| POVM / PVM |
Positive Operator–Valued / Projection–Valued Measure (measurement models). |
| Pointer algebra |
Abelian subalgebra encoding the measurement context (“sector”). |
| Lüders map |
Conditional expectation onto the diagonal algebra in a fixed basis. |
| DeTurck gauge |
Gauge choice rendering curvature flows strictly elliptic on Riemannian slices. |
| Lichnerowicz operator |
Elliptic operator on symmetric 2–tensors; its gap controls residual decay. |
|
Spectral gap for the Lichnerowicz–DeTurck operator on the GR slice (physical subspace). |
| FRW |
Friedmann–Robertson–Walker cosmological background (see Sec.). |
| CPL |
Chevallier–Polarski–Linder dark–energy parametrization . |
| BAO / CMB / RSD |
Baryon Acoustic Oscillations / Cosmic Microwave Background / Redshift–Space Distortions. |
| AP test |
Alcock–Paczyński test (consistency of and via the AP observable). |
|
Growth–rate observable (see Sec.). |
| GBS |
Gaussian Boson Sampling (quantum–optics benchmark for convergence rates). |
| ISS |
Input–to–State Stability (noise–robust decay bound; see Lemma). |
4. Main Results
4.1. Uniform Law, Contextual Rates
The DSFL inequality has a
uniform mathematical form across sectors,
but the constant
is
sector–dependent. In each setting,
equals the corresponding spectral–gap/coercivity parameter:
Remark 1 (Mathematically uniform, physically contextual).
Equation (3) is the same in all sectors (uniform Lyapunov law), but its rate α in (4) is fixed by the sectoral physics
: pointer algebra and Dirichlet form (QMS), mobility and closure (PDE), Hamiltonian spectrum (QFT/Gaussian), or Lichnerowicz gap (geometry). Thus the restoration mechanism is universal, while the speed of restoration is contextual and must be computed in each sector.
Corollary 1 (Sectoral attractors with explicit rates).
Let denote the sectoral equilibrium manifold (e.g. the fixed-point algebra , the Born law , or the Einstein balance set). Under the hypotheses yielding the corresponding gap/coercivity,
with α given by (4) and C a sectoral equivalence constant (projection/gauge). Hence, the form
of convergence is universal, but the rate
is determined by the sector’s gap.
Editorial note (for the Introduction). To make this distinction explicit up front, one sentence can be added: “While the DSFL provides a single Lyapunov law , the constant α is sectoral: it equals the relevant spectral gap or coercivity (operator algebraic, PDE, free-field, or geometric), so the mechanism is universal but its rate is context dependent.”
Remark 2 (Covariance vs. slice formulation). The geometric DSFL result in Theorem 6 is proved on a compact Riemannian slice in DeTurck gauge. It establishes exponential suppression of the curvature–matter misfit in that elliptic setting but does not yet provide a fully covariant (Lorentzian) formulation. In other words, the present theorem is a slice-level statement—analytically rigorous but gauge–fixed—while a fully diffeomorphism–invariant, hyperbolic extension remains an open program. This distinction should be made explicit: the law is structurally general, yet the proof given here is Riemannian rather than covariant.
This section states the core theorems in a hypothesis–result format, ready for citation in later sections and proofs. Throughout we use the notation and conventions of
Section 3. In particular,
denotes the global alignment residual (classical/PDE) or the noncommutative variance relative to a pointer algebra (QMS). All proofs are deferred to
Section 5, where the three settings (classical, operator–algebraic, and PDE) are treated in parallel.
What Is Proved Here.
First, a unified
propagation lemma (
Section 4.2) shows that local convexity/contractivity mechanisms (Jensen, Kadison–Schwarz, energy identities) imply
global residual monotonicity. Second, adding a spectral gap or coercivity yields
exponential DSFL decay. In the operator–algebraic case (
Section 4.3) we obtain an equivalence between DSFL and a noncommutative Poincaré (spectral–gap) inequality with optimal rate
.
4.2. DSFL pRopagation Lemma (Classical/QMS/PDE)
Lemma 1 (Propagation of residual monotonicity: classical, noncommutative, and PDE). Let act on the Hilbert space where the residual is evaluated. Assume one of the following structural settings.
(CL) Classical Markov setting.
is a Markov contraction semigroup on over a σ–finite measure space : it preserves positivity, mass (), and is –contractive for . Let and let with a Borel convex function and . Define
(QM) Operator–algebraic/QMS setting.
is a normal unital completely positive (u.c.p.) semigroup on a von Neumann algebra , –symmetric
on (reversible) and –preserving
: . Let be the fixed–point algebra and the ω–preserving conditional expectation (the orthogonal projection). Define the noncommutative variance (residual)
(PDE) Parabolic/energy–identity setting.
Let be a bounded domain (with homogeneous Dirichlet/Neumann b.c.) or (periodic). Let , , and set . Assume the exact residual identity
with a measurable a.e. and a remainder (or dominated by the dissipative terms). Let .
Conclusion. In each setting, the residual is monotone nonincreasing
:
Proof.
(CL). By Jensen and positivity, –a.e.; integrate and use .
(QM). Kadison–Schwarz for u.c.p. maps gives ; apply and to obtain . Since is the orthogonal projection and acts as the identity on , the same contraction holds for .
(PDE). Integrate (
5) in time;
and
give
. □
Corollary 2 (Propagation + gap/coercivity ⇒ DSFL). Under Lemma 1, suppose additionally:
(CL) L (the generator of ) is symmetric on and satisfies a Poincaré inequality for some (here Γ denotes the carré du champ).
(QM) The ω–symmetric generator has a spectral gap on : .
(PDE) a.e. with , and for some and sufficiently small .
Then the DSFL inequality holds with an explicit rate:
Remark 3 (Domains and regularity; where proofs appear).
In (QM)
, differentiating at is justified for and extends by density to (closedness of the Dirichlet form). In (PDE)
, integrations by parts are justified by periodic or homogeneous boundary conditions and the Sobolev regularity , (or a standard mollification argument). Complete proofs are given in Section 5.
4.3. DSFL ⇔ Spectral Gap (Reversible QMS)
Definition 1 (Operator-algebraic residual).
Let be a von Neumann algebra with faithful normal state ω, and let be ω-modular invariant so that the ω-preserving conditional expectation exists (Takesaki). Define, for ,
Then is the squared -distance to .
Theorem 1 (DSFL ⟺ spectral gap). Let be a normal u.c.p. semigroup on that is ω-symmetric on and ω-preserving. Write and let be the generator with Dirichlet form . The following are equivalent:
- (i)
(DSFL) s.t. for all and .
- (ii)
(Spectral gap) s.t. for all X in the form domain.
Moreover, the optimal constants satisfy .
Setting and Assumptions.
Let
be a
–finite von Neumann algebra with faithful normal state
. Write
for the GNS Hilbert space with
Let be a normal unital completely positive (u.c.p.) semigroup on such that:
- (A1)
(ω–symmetry / detailed balance ) Each is self–adjoint on : for all .
- (A1′)
(ω–preservation ) for all Z and .
- (A2)
(
generator and Dirichlet form ) The
–generator
of
is self–adjoint, with closed quadratic form
and a *–subalgebra
(e.g. the analytic elements of
) is a form core.
- (A3)
(Fixed–point algebra ) is a von Neumann subalgebra. The modular group leaves globally invariant.
Under (A3), Takesaki’s theorem yields:
Lemma 2 (Conditional expectation). There exists a unique faithful normal conditional expectation that preserves ω. It extends to the orthogonal projection with .
Define the noncommutative variance (residual)
Write .
Lemma 3 (
–differentiability and domains).
For ,
More generally, for all with ,
Proof. Since
is self–adjoint and strongly continuous on
,
is
on any interval where
. Differentiating and using self–adjointness of
gives
Orthogonality to is preserved because and acts as the identity on the fixed space. □
Poincaré Inequality on the Orthogonal Complement.
We say that
has a
spectral gap on
if
Equivalently,
i.e. the noncommutative Poincaré inequality holds
on .
Theorem 2 (Operator–algebraic DSFL ⟺ spectral gap). Under (A1)–(A3) the following are equivalent:
- (i)
DSFL decay. such that for all ,
- (ii)
Spectral gap / Poincaré on .
There exists such that (9) holds on .
Moreover, the optimal constants coincide as .
Proof sketch. For one has . If (ii) holds then and Grönwall yields (i) with . Conversely, differentiating (i) at gives . □
Remark 4 (Closability, core, and invariance).
(i) The form is closed on because is self–adjoint and contractive; is the canonical form domain. Moreover, the *–algebra of analytic elements for is dense in and forms a core. (ii) Reversibility implies leaves both and invariant; hence the restriction of to is self–adjoint and nonnegative, with spectrum contained in iff (9) holds.
Remark 5 (Conditional expectation ). The modular invariance of (A3) ensures the existence and uniqueness of a faithful normal ω–preserving conditional expectation (Takesaki’s theorem). As an map it is the orthogonal projection onto , so .
DSFL Interpretation.
Theorem 2 identifies the DSFL rate with the spectral gap of the reversible QMS restricted to : the noncommutative variance (misalignment) decays exponentially iff the Poincaré inequality holds on , with .
4.4. Sharp Lindblad Rate (Finite–Dimensional Dephasing)
Let
with orthonormal basis
and spectral projectors
. Consider the (trace–preserving, completely positive) dephasing Lindblad generator on
and the dual master equation
for density matrices. Let
be the Lüders (diagonal) conditional expectation
onto the abelian pointer algebra
. Then
is the (trace–)preserving conditional expectation onto
, and the
Lüders residual
is the Hilbert–Schmidt variance off the diagonal algebra
.
Theorem 3 (Sharp exponential decay of the Lüders residual).
Let
Then for all ,
and the rate is optimal
(sharp). In particular, if the minimal dephasing rate is attained by at least two indices, then ; otherwise .
Proof (modewise solution). In the basis
one has
so
and hence
Sharpness: choose initial data supported on a pair achieving . □
Corollary 3 (Born–aligned limit and trace–norm control).
As , in Hilbert–Schmidt norm and in trace norm with
The limit is the Lüders (Born–aligned) state for the measurement basis .
Remark 6 (DSFL and spectral gap identification). The off–diagonal (pointer–orthogonal) sector is invariant and the restriction of to it has spectral gap . By Theorem 2, the DSFL rate is .
Remark 7 (Commuting Hamiltonians and basis choice). (a) Adding with (i.e. ) leaves the modewise decay and the sharp rate unchanged: the diagonals remain constant and off–diagonals acquire only phases. For noncommuting H, oscillations appear but the envelope of still decays at least as . (b) The statement is basis–covariant: any pure dephasing generator is unitarily diagonalizable; is the diagonal algebra in that basis, and Φ is the corresponding conditional expectation.
4.5. Coercive PDE tEmplate: Exponential Decay
Let
be a bounded
domain (or
) with periodic or homogeneous boundary conditions chosen so that integrations by parts incur no boundary terms (cf. Remark 3). Assume
for all
, and consider the system
with
satisfying
a.e. for some
, and a locally Lipschitz coupling
G that is
subcritical in the residual energy sense specified below. Define the residual
Theorem 4 (Exponential residual decay under coercivity).
Under (11), , and the subcriticality condition
for some and sufficiently small , one has the differential inequality
hence, by Grönwall,
In particular, in the uncoupled case () one obtains .
Proof sketch. Differentiate
and use
to obtain the exact identity (cf. (
5))
Since
, the first term is
, the divergence term is nonpositive, and (
12) gives
. Combine and apply Grönwall. □
Remark 8 (Scope and refinements).
(i) Template coverage.
The estimate applies to mobility–relaxation closures, linear couplings bounded by the residual, and mild nonlinearities that satisfy (12). It is the PDE instance of the propagation lemma (Lemma 1, case (PDE)) followed by a sectoral coercivity bound; in DSFL notation, the rate is .
(ii) Spectral sharpening.
With the Helmholtz decomposition and , the gradient channel gains additional spectral damping: on bounded/periodic domains,
with Poincaré constant , so the gradient part decays at least like while the solenoidal part decays like .
(iii) Regularity/BCs. All integrations by parts are justified by periodic or homogeneous boundary conditions and the Sobolev regularity stated above (see Remark 3). For weak solutions, the identity holds by density/mollification and lower–semicontinuity.
(iv) DSFL identification. The inequality is the DSFL law in this sector. With one recovers the clean rate .
4.6. Free–Field Stochastic Quantization: Gap–Driven Decay
Setting and Notation.
Let
be either the flat
d –torus
(side length
) or
. We consider a real free scalar field
with Euclidean action
Parisi–Wu stochastic quantization evolves the field in an auxiliary “Langevin time”
by
where
is space–time white noise with covariance
(the factor 2 ensures that the stationary covariance solves
). The generator of the one–body deterministic part is the nonnegative operator
whose spectrum is
on
, and
on
.
OU Semigroup and Covariance Flow.
Equation (
13) is an infinite–dimensional Ornstein–Uhlenbeck (OU) process on
for
(or on
at the level of covariances). Writing
and
a cylindrical Wiener process on
, (
13) reads
Let
be the (two–point) covariance operator on
. Standard OU calculus yields
The unique stationary covariance is the Green operator
Subtracting (
15) from (
14) gives the exact relaxation formula
Smeared Two–Point Residual.
For a test function
f (to be specified below), define the smeared evaluation
Then
We define the (quadratic) residual as the squared deviation of the two–point function:
Admissible Class of Test Functions.
On , take ; on with , take (or Schwarz ). For , impose an IR regularization (finite volume or mean–zero f and a Poincaré gap).
Theorem 5 (Residual decay at twice the Hamiltonian gap).
Let on and suppose there is a spectral gap
Then for any admissible f,
In particular:
On with , , where is the first positive Laplace eigenvalue; if then .
On with , ; if there is no gap and (18) fails globally (decay is not uniform, see Remark 10).
Proof.
Because
on
and
is the identity on
(trivial unless
on compact
), we have
whenever
(automatically true for
). Thus
and squaring yields (
18) after absorbing the prefactor into
. □
Remark 9 (Explicit Fourier picture on
).
Write and similarly for . Each mode solves the scalar OU SDE so Therefore, for the smeared variance,
and the slowest decaying mode has rate .
Remark 10 (Massless case and infrared issues).
On with , and provided f has zero mean (or we project away the constant mode). On with , has no gap; uniform exponential decay fails and long–wavelength modes relax only algebraically in spatially extended senses. Thus a spectral gap (mass or finite–volume Poincaré gap) is essential for the DSFL rate (18).
Remark 11 (From two–point residuals to DSFL).
The estimate (18) derives
the DSFL inequality in the Gaussian sector: the misalignment functional decays as with optimal rate . Equivalently, for any fixed f, the scalar residual satisfies the same inequality.
Remark 12 (Regularity of smearing). On any is admissible. On with , (or ) suffices and all formulas above hold; higher regularity yields the same exponential rate while changing only the (finite) prefactors.
4.7. GR Slice: Geometric Residual Decay (Small Data, DeTurck Gauge)
Scope.
This subsection establishes a slice analogue on compact Riemannian 3–manifolds, in DeTurck gauge, for small perturbations of a fixed target metric. It is not a fully covariant Lorentzian result. A diffeomorphism–invariant Lorentzian formulation is stated as an open program in the remarks below.
Standing Assumptions (Slice, Small Data).
Let
be a smooth, closed (compact, boundaryless) Riemannian 3–manifold. Let
be time–independent and divergence–free with respect to
,
. Assume there exists a target metric
solving
Consider the Einstein–source flow in DeTurck gauge
with DeTurck vector
which renders the linearization at
strictly elliptic
on the gauge–orthogonal (physical) subspace.
Residual (Gauge–invariant).
Define the
curvature–matter misfit
Write and assume small initial data with for some and sufficiently small.
Lemma 4 (Constraint preservation (DeTurck slice)).
Let . Along (20), If and T is time–independent with along the flow (equivalently at and preserved thereafter), then for all .
Linearization, Model Operator, and Spectral Gap.
Linearizing (
20) at
yields, for
h small,
where
is the self–adjoint Lichnerowicz–DeTurck operator on symmetric 2–tensors,
and
is quadratic/higher order. Assume a spectral gap on the
physical (gauge–orthogonal) subspace:
By elliptic regularity, for
h sufficiently small in
,
Residual Equivalence Near .
A Taylor expansion at
and (
19) give
For
h small and
close to
in
, the norms induced by
and
are equivalent; hence
with constants depending only on
and the smallness radius.
Theorem 6 (Exponential
decay of the geometric residual on a slice).
Under (19), (20), (25), and the small–data hypothesis (for some and sufficiently small), there exists (depending only on and δ) such that
In particular, in as , and modulo diffeomorphisms.
Proof sketch. Set
. From (
23),
The gap (
25) yields
. Estimate the nonlinear term via Cauchy–Schwarz, (
26), and small–data absorption to obtain
, whence
. Grönwall gives
; (
28) then implies the stated decay for
. □
Remark 13 (Well–posedness and norm equivalences).
For small (), parabolic–elliptic theory in DeTurck gauge yields local existence/uniqueness in and a priori control. The exponential decay closes the bootstrap globally. Moreover, – and –norms are equivalent for small h, so (22) and (28) are interchangeable up to fixed constants.
Remark 14 (Matter compatibility and constraints).
The condition together with Lemma 4 ensures preservation of the contracted Bianchi constraint and prevents spurious source terms in the energy estimates; T only fixes the equilibrium (19).
Remark 15 (Gauge directions and the physical subspace).
The DeTurck term (21) removes the diffeomorphism kernel of the linearized operator; the spectral gap (25) is thus a genuine coercivity on the physical subspace. Without DeTurck, one must pass to the quotient by diffeomorphisms (e.g., transverse–traceless decomposition) and run the same argument there.
Remark 16 (Lorentzian caveat). The theorem is a Riemannian slice statement in DeTurck gauge. It does not imply a fully covariant Lorentzian DSFL. A Lorentzian version would require a diffeomorphism–invariant Lyapunov functional on the space of Lorentzian metrics and a hyperbolic evolution with an appropriate gap; this remains an open program.
Remark 17 (ISS robustness (small forcing)).
If (20) is perturbed by a small, mean–zero forcing in –time, the same calculation yields and hence
Remark 18 (Interpretation). In the limit one has in (and pointwise where regularity allows); with small–data coercivity, modulo diffeomorphisms. Thus, Einstein balance is a slice attractor when a Lichnerowicz–DeTurck gap is present.
4.8. Master/Grand Attractor Theorems (Sectoral Attractors)
Let
denote the global alignment residual associated with a given sector (classical/QMS/PDE/GR). Assume a DSFL inequality holds on its natural state space
:
where
is the Poincaré/spectral–gap constant or a quantitative coercivity as established in §4.2–§4.6. Write
for the
sectoral attractor set (the equilibrium manifold in that sector; e.g.
in QM,
in GR, etc.).
Proposition 1 (Small–gain for two coupled residuals).
Let satisfy, for some constants and couplings ,
If the small–gain condition holds,
then there exists (depending only on ) such that
where the decay rate can be chosen as
In particular, both decay exponentially to 0.
Proof sketch. Write (
30) in vector form
with
and
. The eigenvalues are
Condition (
31) yields
. Hence
componentwise and
. Setting
gives
. Alternatively, choose a weighted Lyapunov
with
so that
. □
Distance Equivalences Near the Attractor.
We record norm–equivalences that turn the residual into a bona fide distance to the attractor in each sector (up to constants).
Lemma 5 (Residual vs. geometric distance to the attractor).
(i) QM (Born sector). On a bounded domain with Poincaré constant , if and , then for ,
(ii) TD (continuum). With weights and , (iii) QMS (OA). With pointer algebra and conditional expectation , is the squared -distance to . (iv) GR (slice, DeTurck). For small in , hence modulo diffeos.
Proof. QM and TD are immediate from Poincaré and
. QMS is by definition. GR follows from (
27)–(
28) and elliptic regularity (
26). □
Omega–Limit Characterization and LaSalle.
Lemma 6 (LaSalle-type invariance).
Let be the sector semigroup on , continuous in t, and suppose is nonnegative, continuous on , and satisfies (29). Then for any trajectory , the ω–limit set is nonempty, compact, and contained in . If, in addition, consists of a single orbit (modulo the natural gauge of the sector), then .
Proof. Monotonicity and boundedness of
yield precompactness (sector by sector) and invariance. At any accumulation point
, the derivative of
vanishes, hence
by (
29). Uniqueness up to gauge gives convergence to the orbit. □
Theorem 7 (Sector attractors from residual decay (Master theorem)).
Under (29) and Lemma 5, the canonical equilibrium relations are the unique global attractors in their sectors:
- (QM)
(Born alignment) If and , then in at least exponentially, with rate (cf. Theorem 11).
- (TD)
(Residual entropy) satisfies , hence and exponentially.
- (QMS)
(OA pointer alignment) If the reversible QMS has spectral gap on , then and in .
- (GR)
(Einstein balance on slices) Under Theorem 6, and modulo diffeomorphisms; thus in .
Proof. (QM) Combine (
29) with Lemma 5(i) and Theorem 11. (TD) Differentiate
:
, hence
increases and
exponentially. (QMS) is Theorem 2. (GR) is Theorem 6 plus Lemma 5(iv). □
Robustness and Time–Varying Rates.
Proposition 2 (ISS/ultimate boundedness; time–varying ). (i) If with and , then and . (ii) If with measurable and , then ; if , then .
Proof. (i) Grönwall with input. (ii) Integrate the differential inequality. □
Discrete–Time and Product Systems.
Proposition 3 (Discrete DSFL; products). (i) If with , then . (ii) If and satisfy with a Metzler coupling matrix , then exponential decay holds provided the spectral abscissa of is positive (cf. Proposition 1).
Proof. (i) Induction. (ii) Linear ODE comparison and the spectral condition. □
Weighted Distances and Alternative Norms.
Lemma 7 (Residual vs. weighted distance).
Let be the sectoral attractor and let be a locally equivalent distance induced by a positive quadratic form (e.g. an metric in PDE sectors or a weighted metric in QMS). Suppose that in a neighborhood of there exist constants such that
If the DSFL inequality (29) holds on , then
for all t for which the trajectory stays in . In particular, the convergence rate is unchanged up to the equivalence factor .
Proof. Combine (
29) with the local equivalence to bound
above and below by
, then apply Grönwall. □
Remark 19 (PDE distances). In diffusion–type PDE sectors, observables are often controlled in rather than . If the residual controls and vice versa near the attractor (e.g. via Poincaré and elliptic estimates), Lemma 7 transfers the DSFL rate directly to .
Consequences for Sector Observables.
Corollary 4 (Observable convergence).
Let be a continuous observable on the sector state space and assume there exists a neighborhood of the attractor where (local Lipschitz). Then under (29) (or Theorem 9),
with in the reversible cases and in the coercive PDE case. Moreover, if a weighted distance equivalent to near is used (Lemma 7), the same estimate holds with replaced by .
Remark 20 (Uniqueness modulo gauge). In PDE and GR sectors the attractor is unique modulo the natural gauge (additive constants for ρ, diffeomorphisms for γ). The theorems above are to be understood on the corresponding quotient spaces, or after fixing a gauge (DeTurck in GR, mean–zero in QM/TD).
Remark 21 (Coupled residuals and small–gain). In multi–residual settings (e.g. DSFL+SABIM), a vector–Lyapunov together with the sharp small–gain condition (Proposition 1) yields exponential decay with rate ; see also the n–dimensional version in Corollary.
Proof (expanded).
(QM) Set . By definition of the residual in the Born sector one has . The DSFL inequality implies as , hence in . Since for all t (mass conservation), we have . By the Poincaré inequality on mean–zero functions, , whence in . Therefore in . Moreover, if the sector provides the sharper differential inequality (cf. Theorem 11), then .
(TD) By definition
with
. Differentiating and using DSFL,
In particular, if with , then is strictly increasing and , which diverges as while exponentially.
(GR) By assumption, , hence . Since , this implies in as . Under the small–data hypotheses of Theorem 6 and the gauge choice (DeTurck), elliptic regularity and the spectral gap yield convergence of to modulo diffeomorphisms; in particular, in the limit. □
Theorem 9 (Grand attractor theorem (abstract form)).
Let be a reversible contraction semigroup on a Hilbert space H with generator , fixed–point subspace , and Poincaré gap on . For the residual one has
and is the unique global attractor (modulo the sector’s gauge).
Proof. Let
. Since
is reversible on
H,
is differentiable with
where
and we used the Poincaré gap on
. Grönwall’s lemma gives
, i.e.
and
. Since
is the fixed–point subspace of the semigroup, it is invariant and attracts every orbit. Uniqueness of the global attractor (modulo gauge) follows because any other closed invariant attracting set must lie in
. □
Remark 22 (Interpretation). The Master/Grand theorems formalize the central message: once a (sector–appropriate) spectral gap/coercivity is present, the DSFL inequality holds and the sector’s canonical equilibrium relation is recovered as a dynamical fixed point rather than as a postulate. In reversible settings the exponential rate is dictated by the Poincaré gap; in the PDE/GR coercive settings the rate is dictated by the quantitative coercivity (e.g. or ).
Corollary 5 (Observable ISS under DSFL).
Assume the hypotheses of Corollary 4. Suppose further that the residual dynamics admit an input term in the DSFL inequality,
with , , and . Then
where and are the fixed constants from the residual–entropy proxy. In particular, if then
The same estimate holds with replaced by any locally equivalent weighted distance (Lemma 7).
5. Proofs
This section collects the proofs of the results stated in
Section 4.
5.1. Proof of Sec. 4.1 (Propagation Lemma)
Proof of Lemma 1.
(CL) Let be convex with and let be a Markov contraction on . By Jensen’s inequality and positivity/mass preservation, –a.e. Integrating and using yields , i.e. .
(QM) Let be normal u.c.p., –symmetric on , and –preserving. Kadison–Schwarz gives . Applying and using yields . Since is the orthogonal projection and acts as the identity on , the same contraction holds for : , hence .
(PDE) With
and
, differentiate in time. Using
and integrating by parts (periodic BCs or homogeneous BCs that kill boundary terms), we get
Under and the stated sign/dominance of the controlled terms, the right–hand side is , hence . □
5.2. Proof of Lemma
Proof.
(i) QM. Let be a bounded domain with periodic or homogeneous Neumann BCs. For one has (mass normalization). Poincaré then gives , so , and the upper bound follows from the control of w.
(ii) TD. For a.e., , hence
(iii) QMS. By Takesaki’s theorem, the –preserving conditional expectation is the orthogonal projection onto . Therefore
(iv) GR. Linearizing at yields with . For sufficiently small, squaring and integrating gives Elliptic regularity for on the gauge–orthogonal subspace implies hence , which is equivalent to modulo diffeomorphisms. □
5.3. Proof of Lemma 6
Proof. Let
. By (
29),
is nonincreasing and bounded below, hence convergent. Sector by sector, standard compactness (e.g. Rellich–Kondrachov in PDE/GR or spectral decomposition in reversible semigroups) yields precompactness of trajectories on bounded time intervals. Any
admits a sequence
with
, and continuity of
gives
; invariance of
implies
. Since
, necessarily
and hence
. Thus
. If
consists of a single orbit modulo the sector’s gauge, then Łojasiewicz–Simon/strict Lyapunov arguments imply
. □
5.4. Proof of Theorem 7
Proof. Combine Lemma 5 (residual ⇔ distance) with the DSFL inequality (
29) to obtain
– (QM) and
– (GR) convergence. For TD, the
monotonicity follows by direct differentiation. For QMS, apply Theorem 2. Uniqueness modulo sector gauges follows from invariances (additive constants for
; diffeomorphisms for
) and the strict convexity of
transverse to gauge directions. □
5.5. Proof of Sec. 4.3 (Lindblad Sharpness)
Proof of Theorem 3 and Corollary 3. Work in the orthonormal basis
where
. The generator is
□
Proof of Theorem 3. (A) Modewise Solution and Residual Envelope.
Matrix elements satisfy
so
. Hence
with
.
□
Proof of Theorem 3 and Corollary 3. Work in the orthonormal basis
where
. The generator is
(A)
Modewise solution and residual envelope. Matrix elements satisfy
so
. Hence
with
.
(B)
Optimality. Choose initial coherence supported on a pair
attaining
. Then
which matches the envelope with equality. Hence the rate
is sharp.
(C)
Trace–norm convergence. Since the diagonal entries are constant in time and the off–diagonals decay modewise as
, one has
Using
then gives
so
in trace norm at an exponential rate, proving the corollary.
□
5.6. Proof of Sec. 4.4 (PDE Energy Identity and Decay)
Proof of Theorem 4.
Step 0 (regularity and IBP). Assume is a (mild) solution with for , and periodic or homogeneous boundary conditions chosen so that the boundary term in the IBP identity vanishes (cf. Remark 3). Standard mollification in time (or density of smooth compactly supported functions in the graph norm) justifies differentiation under the integral sign; the final inequalities extend by continuity to the given regularity class.
Step 1 (residual equation). Let
Step 2 (exact energy identity). Differentiate
and use the product rule:
By integration by parts and the BC choice,
Step 3 (coercivity and subcriticality). The uniform positive definiteness
a.e. implies
By the subcriticality hypothesis, there exist
and sufficiently small
such that
(For example, this holds if G is locally Lipschitz with in the energy region visited by the solution.)
Step 4 (differential inequality). Insert these bounds into (
33) to obtain
Let (smallness of ensures positivity).
Step 5 (Grönwall). By Grönwall’s inequality,
i.e. the residual decays exponentially at rate
. In particular, for
we recover the clean rate
.
Optional refinements. (i) The extra term
is dissipative and improves the decay when the Helmholtz gradient component dominates; combine with a Poincaré inequality on gradient fields to sharpen the rate on tori or bounded domains. (ii) If
with
then
In particular, if , then ; and if , then .
This completes the proof. □
5.7. Proof of Sec. 4.5 (Free–Field Residual Decay)
Proof of Theorem 5. We give a concrete Ornstein–Uhlenbeck (OU) derivation for the free field and then the abstract semigroup proof.
Spectral Gap Bound.
Since
with
(free field), the semigroup bound
yields
The residual in Theorem 5 is
, hence from (
34)
This is sharper than the theorem’s envelope; loosening the prefactor gives the stated bound.
Fourier–Mode Check (Explicit Diagonalization).
Let
be the Fourier transform and note
. Then
multiplies by
, while
multiplies by
. A direct computation yields
so
, consistent with (
34).
Abstract Semigroup Proof.
Let
H be the nonnegative, self–adjoint Euclidean Hamiltonian generating
on the GNS space. With vacuum projector
one has
and
. Then
for some positive operator
K, whence
and squaring gives the same residual decay (again with the sharper
envelope available).
Admissible Test Functions.
The bounds hold for any f with finite quadratic forms and (e.g. with ultraviolet cutoff, or ), ensuring well–posed covariance pairings and OU action.
Combining the covariance estimate with finishes the proof. □
5.8. Proof of Sec. 4.6 (GR DeTurck–Gauge Decay)
Proof of Theorem 6. We work on a smooth closed 3–manifold
. Write the evolving metric as
with
small in
,
. The Einstein–DeTurck–source flow can be written as
where
is the strictly elliptic Lichnerowicz–DeTurck operator (self–adjoint on
on the orthogonal complement of gauge directions) and
collects quadratic and higher–order terms in
(and lower–order dependence on
). By hypothesis, there exists a spectral gap
such that
We also assume T is divergence–free along the flow so that constraint terms vanish.
Step 1: Energy Identity in the Fixed Background.
Taking the
inner product of (
35) with
h,
Step 2: Nonlinearity Estimate.
On the compact manifold, for
the bilinear/quadratic structure of
and Sobolev product estimates yield
for some
and
. (Here we used interpolation
and the fact that
has no linear part at
.) Since the flow is parabolic–elliptic in DeTurck gauge, standard local theory gives a time interval on which
. Choosing the smallness radius
so that
implies
Step 3: Differential Inequality and Decay.
The decay (
40) and parabolic regularization imply, by a standard bootstrap, that
remains
for all
provided
is chosen small enough initially. Hence (
40) holds globally, and by refining the absorption in (
39) one improves the rate to
:
1
Step 4: Equivalence of the Geometric Residual and .
For
h small in
one has the metric and volume equivalences
uniformly for tensors
S. Moreover, by linearization at
,
with
. By elliptic regularity for
and the spectral gap on the gauge–orthogonal subspace,
. Hence, for
sufficiently small,
for some
depending only on
and the smallness radius.
Step 5: Decay of .
Differentiating
along the flow and using (
35), the elliptic coercivity and (
41), one obtains the differential inequality
for all
as long as
h remains in the small regime (which we ensured in Step 3). Grönwall therefore gives
and in particular
in
as
. Standard arguments then show
modulo diffeomorphisms (the DeTurck vector field fixes the gauge). □
6. Instantiations and Consequences
7. Instantiations and Consequences
7.2. Residual–Entropy Arrow of Time
The DSFL residual
plays a dual role: it is a quadratic Lyapunov functional (structural energy of misalignment) and, via a monotone transform, a surrogate “entropy’’ that certifies irreversibility.
Definition 2 (Residual entropy).
Fix and . The residual entropy
is
Near–Alignment Connection to Classical Entropies.
When
and a target
q are strictly positive densities with
and
, the Kullback–Leibler divergence satisfies the second–order Taylor bound near equilibrium:
On bounded domains with Poincaré constant
and
,
Thus, near alignment,
and hence
—which is monotone in
—tracks the decay of leading–order deviations of classical relative entropy.
Remark 32 (Information–geometric view). In the quadratic regime, the Fisher information linearizes to a Dirichlet form on , and the Hellinger/Wasserstein–2 metrics become equivalent to / distances modulo weights. Under DSFL, provides a sectoral Lyapunov that is locally equivalent to these information distances; therefore inherits the arrow–of–time interpretation without invoking stochastic typicality.
Intrinsic Time and Reparametrization.
Define the intrinsic clock . Under DSFL with constant rate , ; under time–varying , one has which is strictly increasing and unbounded if . This reparametrization is useful for comparing trajectories across sectors: observables that are Lipschitz in the sector distance (Cor. 4) will contract exponentially in .
Relation to Boltzmann’s H –Functional in Classical and Quantum Settings.
We compare the residual–entropy arrow with classical and quantum entropy production principles.
Classical reversible diffusions (Bakry–Émery)
Let
be a reversible diffusion on a bounded domain with invariant density
, carré du champ
and Dirichlet form
. For a solution
of the forward equation with
define:
Then
and, under the Bakry–Émery curvature condition
with
(log–Sobolev constant),
Residual vs. entropy near equilibrium. Assume
and set
with
. A second–order Taylor expansion yields
By Poincaré,
, hence
so
(a decreasing function of
) tracks the decay of
near alignment.
Rates and regimes. - Log–Sobolev (strong): If , then both D and decay exponentially: and with the Poincaré constant . In many Gaussian/OU cases, ; in general . Thus and H –theorem arrows coincide up to constants. - Poincaré only (weak): If only holds (no LSI), DSFL still gives and a strictly increasing , while classical entropy methods may not yield exponential decay of D. The DSFL arrow thus persists beyond the LSI regime.
Entropy production vs. residual production.
and (under LSI)
; by contrast
Thus in the asymptotic regime both
D and
evolve linearly in their natural time scales (
t for
D with rate
, and
t for
with slope
), while
by (
49).
Reversible quantum Markov semigroups (QMS)
Let
be a reversible QMS on
with generator
and gap
on
. Consider two divergences:
- Variance law (DSFL): The residual decays as (Theorem 2), giving a strictly increasing . - Quantum H –theorem: If a quantum log–Sobolev inequality (QLSI) holds with constant , then (entropy contraction).
Near–equilibrium comparison. For faithful
, the second variation of
at
is the Bogoliubov–Kubo–Mori (BKM) metric; in finite dimension,
in a neighborhood of
. Hence
D and the residual (and therefore
) are
locally equivalent —they define the same arrow close to equilibrium.
Rates and regimes. - QLSI (strong): If QLSI holds with constant , then both D and decay exponentially; typically , and in dephasing/Gaussian cases . - Poincaré only (weak): If only the spectral gap is known (no QLSI), the DSFL arrow survives: decays exponentially and increases, while D may lack a uniform exponential bound.
Summary: When Do the Arrows Coincide?
Near alignment (classical or quantum): and locally, hence and H –theorem describe the same decay up to constants.
Under (quantum) log–Sobolev: D and both decay exponentially, with rates and . In Gaussian/dephasing models (⇒ identical envelopes); generally .
Only Poincaré available: DSFL still gives an exponential variance contraction (hence a strict arrow), whereas entropy contraction can be weaker or unavailable.
Practical Readouts.
For data/experiment, is often easier to estimate (it only requires quadratic residuals) and provides a robust monotone even when D is hard to evaluate or lacks exponential decay. When LSI holds, and D are interchangeable up to constants; otherwise supplies a structural arrow beyond the reach of entropy methods.
7.3. Einstein Balance as Geometric Attractor
Let
be a spacetime region with metric
and stress–energy tensor
. Define the geometric residual
where
is the pointwise norm induced by
g and
is the Einstein tensor. The residual (
50) vanishes iff the Einstein balance
holds pointwise.
Gauge and Slice Issues.
In full Lorentzian signature the Einstein evolution is hyperbolic and diffeomorphism-invariant; a direct Lyapunov descent of (
50) is obstructed by gauge freedom and hyperbolicity. On a compact
Riemannian slice
and in
DeTurck gauge (
Section 4.7), the linearized operator becomes elliptic on the physical (gauge-orthogonal) subspace. In that regime the residual
is a bona fide slice Lyapunov functional (modulo gauge equivalences). We now restate and instantiate the resulting attractor facts.
Theorem 12 (Slice attractor: small data, DeTurck gauge).
Let be a compact Riemannian 3–manifold and T a smooth, time-independent, divergence–free source w.r.t. . Assume the target balance and that the Lichnerowicz–DeTurck operator on the physical subspace has a spectral gap :
Then for the Einstein–DeTurck–source flow (20) with sufficiently small initial perturbation in (),
for some depending only on and the smallness radius. In particular, in as , and modulo diffeomorphisms.
Sketch
(instantiation of Section 4.7). Linearize (
20):
with
. Energy estimates on
give
for small data. Near
,
(
Section 4.7), whence the claim. □
Corollary 7 (Einstein backgrounds with positive physical gap). If is an Einstein metric with positive physical gap (e.g. compact spaceforms with appropriate sources), then for all sufficiently small perturbations satisfying the momentum constraints, the curvature–matter misfit decays exponentially and converges modulo diffeomorphisms to .
Remark 33 (FRW-type slices). On a compact FRW slice (spatial section a compact spaceform), the physical gap typically reduces to a scalar spectral gap for the Lichnerowicz–DeTurck operator acting on TT-modes; small scalar/vector perturbations are damped by the same mechanism. The attractor is the balanced background (e.g. ΛCDM source) at the slice level.
Robustness to small forcing (ISS).
If the source term acquires a small time-dependent perturbation (mean-zero in the physical subspace), the same calculation yields an input–to–state stability bound:
so
.
Physical Interpretation.
Within a slice description, the equality acts as a sectoral equilibrium: perturb the geometry or the source slightly, and the DeTurck–gauge flow suppresses the misfit at an exponential rate controlled by a geometric gap. This realizes the Einstein equations as the endpoint of Lyapunov suppression of , not a prior axiom.
A covariant DSFL program (what remains and how)
A fully covariant DSFL would assert monotone decrease of a diffeomorphism-invariant residual along a hyperbolic (Lorentzian) gauge–fixed evolution, without relying on a Riemannian foliation. Here is a concrete roadmap.
(C1) Covariant Residual and Gauge.
A direct spacetime residual is diffeomorphism-invariant, but its time derivative under the Einstein evolution is not sign-definite due to gauge and hyperbolicity. One needs a hyperbolic, constraint–damped formulation (e.g. generalized harmonic or Z4) so that the physical part of evolves with controllable energy.
(C2) Candidate Hyperbolic DSFL Flow.
In generalized harmonic gauge
, the Einstein equations reduce to quasi-linear wave equations for
. Add constraint damping (
terms in Z4), and consider a “gradient wave” evolution of the residual:
where
projects onto the physical (constraint-satisfying, gauge-orthogonal) subspace. The goal is an energy identity
for a covariant energy
combining Bel–Robinson type energies and constraint energies.
(C3) Covariant Lyapunov Functionals.
Two natural ingredients: (i) a Bel–Robinson energy for the Weyl curvature (positive on slices), and (ii) a “misfit energy” . A weighted sum (constraints) is a plausible Lyapunov, provided damping terms control gauge/constraint errors.
(C4) Small-Data Regimes.
On backgrounds with known nonlinear stability (e.g. Minkowski, de Sitter), one can hope to prove that for small perturbations and suitable damping,
satisfies
hence exponential (or at least integrable) decay of the misfit.
(C5) Obstacles and Outlook.
The chief obstacles are: (a) hyperbolic energy methods only give integral decay unless one has a spacetime Morawetz (or red-shift) inequality; (b) projecting out gauge and constraints covariantly is delicate; (c) asymptotics (non-compact ) need appropriate weighted energies. Nevertheless, in small-data regimes with damping (as in Z4/CCZ4 numerical relativity), the covariant DSFL appears within reach.
Remark 34 (Covariant outlook). A fully covariant DSFL would assert monotone decrease of along a diffeomorphism-invariant, hyperbolic, constraint-damped evolution on the space of Lorentzian metrics, without a foliation. Constructing such a Lyapunov structure is open; the slice results above substantiate the attractor picture in an analytically controlled (elliptic) regime and point to the ingredients needed in the covariant case.
7.4. Measurement Context and Pointer Algebras (Sectorization)
In quantum applications the
sector is determined by the measurement context. Formally, choose an abelian von Neumann subalgebra (pointer algebra)
and let
be the
–preserving conditional expectation (Heisenberg picture). For a normal state
(Schrödinger picture), the restriction
corresponds via the Gelfand isomorphism to a probability law
on a standard outcome space
with
. In this sector, the empirical pointer distribution
is compared to
through a Dirichlet structure
driven by a pointer generator
:
Here is a (sub)Markov generator symmetric in with and .
Pointer–Space DSFL and Spectral Gap.
We first record the pointer analogue of the DSFL law.
Proposition 7 (Pointer–space DSFL).
Assume is self–adjoint and nonnegative on with spectral gap on mean–zero functions:
Consider (51) with (static). Then the pointer residual decays exponentially,
and, by Poincaré on Y, .
Proof. Let ; then and . Compute . Since is self–adjoint and nonnegative, on the gap subspace (functional calculus). Therefore and Grönwall yields the claim. Poincaré gives the bound. □
Remark 35 (Time–varying pointer
).
If varies, one obtains a tracking inequality analogous to Lemma 8:
hence tracks inside a tube of radius .
Operator–Algebraic Variance and the Pointer Projection.
We now relate the DSFL on to the pointer DSFL on Y.
Proposition 8 (OA variance contracts to the pointer sector).
Let be a reversible QMS on with spectral gap on and let be the ω–preserving conditional expectation. Then for any ,
In particular, for abelian and observables in the sector, the noncommutative variance contracts at rate down to the classical pointer law .
Proof. This is Theorem 2 (operator–algebraic DSFL ⇔ Poincaré gap) restricted to ; the conditional expectation is the orthogonal projection, so the variance relative to decays as . □
Bridge to PDE Residuals (Position Sector).
When
is the position pointer algebra on a bounded domain
, the abelian identification gives
,
the Lebesgue measure (or a reference measure), and
. The pointer generator
is the Laplacian (or a diffusion generator), with
and gap
on mean–zero functions. Then the PDE residual is
Poincaré on
yields the
sandwich
and Proposition 7 gives
, hence
contraction to the Born law.
Remark 36 (Contextuality and attractors). The choice of encapsulates the measurement context (observable/POVM via Naimark dilation). DSFL contracts to that context: changing changes the attractor (e.g. position vs. momentum). Operator–algebraically, is the projection onto the sector, and Theorem 2 shows that a noncommutative Poincaré gap is equivalent to exponential decay of the noncommutative variance relative to .
Putting it together (context ⇒ sector ⇒ rate).
Theorem 13 (Contextual DSFL pipeline). Fix a pointer algebra (context) with conditional expectation . If the reversible QMS on has a gap on , then the OA residual decays at rate toward the sector. On the abelian sector Y, if the pointer generator has gap , then the pointer residual decays at rate toward . Consequently, in the position context on bounded Ω, DSFL yields alignment to the Born law with envelope .
Proof. Combine Proposition 8 with Proposition 7 and the Poincaré sandwich on . □
Remark 37 (Changing context changes the attractor). If one replaces the position pointer algebra by, e.g., the momentum algebra (Fourier–diagonal), then Y is the momentum space and is the momentum distribution. Proposition 7 applies verbatim with the corresponding (e.g. a diffusion on momentum space). Thus, the attractor is contextual : it is determined by the measurement algebra .
Same Core, Different Attractor: Why Context Matters But the Law Does Not.
The
core of DSFL is a context–independent contraction law for a quadratic residual:
where
is a
sectoral constant (spectral gap/coercivity). This statement does not depend on which measurement context is chosen. What
does depend on the context is: (i) the projection (or classicalization) map defining the
equilibrium manifold (the attractor), and (ii) the value of the rate
(the spectral gap in that context).
We formalize this in three steps.
Definition 3 (Context–dependent residuals and attractors).
Let be an abelian pointer algebra (measurement context) with ω–preserving conditional expectation . The operator–algebraic residual
is
and the corresponding attractor
is the fixed subspace (equilibrium manifold). On the abelian sector Y with , the pointer residual
is the Dirichlet energy
and the attractor
is the pointer law in the chosen context.
Proposition 9 (Context–invariant DSFL form, context–dependent constants). For any pointer algebra :
(a) OA law. If the reversible QMS on has a Poincaré gap on , then
(b) Pointer law. If the sector generator has Poincaré gap , then
Thus the form of DSFL is the same in any context, while the rate constants and (and the attractor) change with .
Proof. Part (a) is Theorem 2 with the gap computed on . Part (b) is Proposition 7 with the Poincaré constant of . □
Proposition 10 (Unitary covariance of the core law).
Let U be a unitary on and . Define . Then
and the spectral gap is invariant: . Consequently, DSFL decay holds with the same rate
in unitarily equivalent contexts, and the attractor transforms as .
Proof. Orthogonal projection covariance under conjugation and unitary invariance of yield the residual identity. The spectrum of the self–adjoint restriction of to the orthogonal complement is invariant under unitary conjugation, hence the same gap. □
Remark 38 (Non–unitarily equivalent contexts (e.g. position vs. momentum)). When and are not unitarily equivalent (e.g. position vs. momentum algebras on bounded domains with different boundary structures), the form of DSFL remains identical but the constants change: , , and the attractors are vs. (position vs. momentum laws). Thus changing context changes the attractor and generally the observed rate, while the core DSFL law —“propagation gap ⇒ exponential suppression”—is the same.
Two–Stage Contraction and Small–Gain View.
In experiments one typically sees a two–stage contraction: (i) the operator–algebraic contraction
at rate
, then (ii) the pointer–space alignment
at rate
. Writing
and
, one can couple them as
for a (typically small) coupling
arising from finite–time context transfer. By Proposition 1 (small–gain),
still decays exponentially provided
.
Examples.
Qubit dephasing. vs. : both are unitarily equivalent, so the rate is invariant (Prop. 10), and the attractor is the corresponding Lüders state in the chosen basis (Theorem 3).
Position vs. momentum (PDE). On a bounded , the position sector has Poincaré constant ; the momentum sector involves the spectral constants of the generator on Fourier side. DSFL form is identical, but constants (and attractor laws vs. ) differ.
Takeaway. The core DSFL mechanism—a single quadratic residual suppressed by propagation + gap/coercivity—is context–invariant. The attractor and the rate constants are context–dependent, through the pointer algebra (and its sector generator ). Changing changes the equilibrium manifold and generally the rate, but not the shape of the restoration law.
8. Numerical Demonstrations (Synthetic)
Purpose. The following minimal, reproducible checks are illustrative sanity tests of the DSFL rates in simple synthetic models (not fits to experimental data). They confirm that the gap/coercivity constants derived in §4 are visible as slopes in practice.
8.1. PDE (Born Sector): Heat Flow with Mean-Zero Mismatch
On
with periodic BCs, set
with
and evolve
Theory (§6.1) predicts with . A spectral implementation (truncated Fourier series; Crank–Nicolson or exact mode update) with multi–mode shows: (i) is strictly decreasing; (ii) a semi-log fit of over a post-transient window returns a slope approaching as higher modes die out; (iii) restricting to the fundamental mode yields a slope throughout. Takeaway: the DSFL envelope is observed as the late-time rate.
8.2. Qubit Lindblad Dephasing: Lüders Residual
For a qubit with dephasing rate
in the pointer basis,
Theory (§4.3) gives (sharp). A simple RK4 or exact update of confirms a straight line of slope on a semi-log plot, independent of commuting Hamiltonian phases. With a noncommuting H, oscillations appear but the envelope remains . Takeaway: the sharp DSFL rate is observed.
Reproducibility.
Tiny reference scripts (FFT heat solver; qubit ODE) suffice to reproduce these slopes; code is available on request. We omit figures to keep the paper focused on theory.
9. Conclusion
We introduced the Deterministic Statistical Feedback Law (DSFL) as a sector-neutral restoration principle that turns canonical equilibrium statements—Born’s rule (QM), entropy monotonicity (TD), and Einstein’s curvature–matter balance (GR)—from postulates into attractors. The mechanism is uniform: a single quadratic misalignment residual decreases globally by a propagation lemma (Jensen/Kadison–Schwarz/energy identity), and sectoral spectral gaps or coercivity upgrade monotonicity to exponential decay. The law is context-invariant; the attractor and rate are context-dependent through the pointer algebra or the sector’s coercivity constants.
QMS: DSFL ⇔ gap, optimal constants. For -symmetric QMS we established the equivalence between DSFL and a noncommutative Poincaré inequality on , with optimal rate .
Lindblad (finite-dimensional). Pure dephasing yields a sharp exponential decay of the Lüders residual with rate .
PDE template. An exact residual energy identity gives , hence under and subcritical couplings.
Free fields. In Parisi–Wu stochastic quantization, smeared two-point residuals decay at twice the Euclidean Hamiltonian gap.
GR slice. On compact Riemannian slices in DeTurck gauge, a Lichnerowicz-type gap implies exponential suppression of the curvature–matter misfit.
Residual-entropy arrow. The proxy is strictly increasing whenever DSFL holds, giving a structural arrow of time that does not require probabilistic postulates.
Context vs. core. Via pointer algebras we proved “same core, different attractor”: the propagation + gap form is universal, while the equilibrium manifold and decay rate depend on the chosen measurement context (position/momentum, basis/unitary changes).
DSFL reframes “equilibrium’’ as the endpoint of a universal Lyapunov descent. It also yields rate-level, falsifiable signatures: GW band coherence, low- ℓ CMB phase structure, growth-rate consistency (), and convergence benchmarks in quantum optics, all tied to sectoral gaps. Tracking inequalities (moving pointers) and small-gain results (coupled residuals) extend the theory to time-varying contexts and weakly coupled sectors.
Open fronts include: (i) a fully covariant Lorentzian DSFL (gauge-invariant residual, hyperbolic flow); (ii) interacting QFT beyond Gaussian sectors (constructive/RG-controlled Poincaré or log-Sobolev constants); and (iii) hypocoercive/nonreversible settings (DSFL with commutator-enhanced residuals). Noise-robust (ISS) variants and multi-residual couplings merit further development and experiments.
Priorities are: log-Sobolev/ upgrades for reversible QMS (variance → entropy decay), cosmology fits (SN+BAO+CMB+RSD) for DSFL backgrounds vs. CDM, laboratory rate-extraction in dephasing/GBS-type setups, and numerical GR slice studies of geometric residual quench. In sum, DSFL isolates the restoration law, quantifies its rates, and delineates precisely when and how Born/entropy/Einstein relations emerge—or fail—under empirical scrutiny.
References
- Cipriani, F.; Sauvageot, J.L. Derivations as square roots of Dirichlet forms. Journal of Functional Analysis 1993, 201, 78–120. [Google Scholar] [CrossRef]
- Takesaki, M. Conditional expectations in von Neumann algebras. Journal of Functional Analysis 1972, 9, 306–321. [Google Scholar] [CrossRef]
- Tomiyama, J. On the projection of norm one in W*-algebras. Proceedings of the Japan Academy 1957, 33, 608–612. [Google Scholar] [CrossRef]
- Davies, E.B. Quantum Theory of Open Systems; Academic Press: London, 1976. [Google Scholar]
- Carlen, E.A.; Maas, J. Gradient flow and entropy inequalities for quantum Markov semigroups with detailed balance. Journal of Functional Analysis 2017, 273, 1810–1869. [Google Scholar] [CrossRef]
- Kastoryano, M.J.; Temme, K. Quantum logarithmic Sobolev inequalities and rapid mixing. Journal of Mathematical Physics 2013, 54, 052202. [Google Scholar] [CrossRef]
- Olkiewicz, R.; Zegarlinski, B. Hypercontractivity in noncommutative Lp spaces. Journal of Functional Analysis 1999, 161, 246–285. [Google Scholar] [CrossRef]
- Bardet, I.; Capel, A.; Lucia, A.; Rouzé, C.; Datta, N. Entropy decay for quantum Markov semigroups: a non-commutative Φ-Sobolev inequality approach. Journal of Mathematical Physics 2018, 59, 012205. [Google Scholar] [CrossRef]
- Brandão, F.G.S.L.; Harrow, A.W. Mixing of quantum Markov chains and quantum expanders. In Proceedings of the Proceedings of the 27th Annual ACM–SIAM Symposium on Discrete Algorithms (SODA), Philadelphia, PA, 2016; pp. 146–155. [Google Scholar]
- Lindblad, G. On the generators of quantum dynamical semigroups. Communications in Mathematical Physics 1976, 48, 119–130. [Google Scholar] [CrossRef]
- Gorini, V.; Kossakowski, A.; Sudarshan, E.C.G. Completely positive dynamical semigroups of N-level systems. Journal of Mathematical Physics 1976, 17, 821–825. [Google Scholar] [CrossRef]
- Breuer, H.P.; Petruccione, F. The Theory of Open Quantum Systems; Oxford University Press: Oxford, 2002. [Google Scholar]
- Spohn, H. Entropy production for quantum dynamical semigroups. Journal of Mathematical Physics 1978, 19, 1227–1230. [Google Scholar] [CrossRef]
- Gross, L. Logarithmic Sobolev inequalities. American Journal of Mathematics 1975, 97, 1061–1083. [Google Scholar] [CrossRef]
- Ledoux, M. The Concentration of Measure Phenomenon; Mathematical Surveys and Monographs; American Mathematical Society: Providence, RI, 2001; Vol. 89. [Google Scholar]
- Bakry, D.; Gentil, I.; Ledoux, M. Analysis and Geometry of Markov Diffusion Operators; Grundlehren der mathematischen Wissenschaften; Springer: Cham, 2014; Vol. 348. [Google Scholar] [CrossRef]
- Hamilton, R.S. Three-manifolds with positive Ricci curvature. Journal of Differential Geometry 1982, 17, 255–306. [Google Scholar] [CrossRef]
- DeTurck, D.M. Deforming metrics in the direction of their Ricci tensors. Journal of Differential Geometry 1983, 18, 157–162. [Google Scholar] [CrossRef]
- Perelman, G. The entropy formula for the Ricci flow and its geometric applications. arXiv Mathematics 2002, arXiv:math.DG/math/0211159]. [Google Scholar]
- Parisi, G.; Wu, Y.S. Perturbation theory without gauge fixing. Physics Letters B 1981, 105, 219–223. [Google Scholar] [CrossRef]
- Damgaard, P.H.; Hüffel, H. Stochastic quantization. Physics Reports 1987, 152, 227–398. [Google Scholar] [CrossRef]
- Zinn-Justin, J. Quantum Field Theory and Critical Phenomena, 4th ed.; International Series of Monographs on Physics; Oxford University Press: Oxford, 2002; Vol. 113. [Google Scholar]
- Bakry, D.; Émery, M. Diffusions hypercontractives. In Séminaire de probabilités XIX 1983/84; Azéma, J., Yor, M., Eds.; Lecture Notes in Mathematics; Springer: Berlin, Heidelberg, 1985; Vol. 1123, pp. 177–206. [Google Scholar] [CrossRef]
- Naimark, M.A. Spectral functions of a symmetric operator. Izvestiya Akademii Nauk SSSR. Seriya Matematicheskaya 1940, 4, 277–318. (In Russian) [Google Scholar]
- Ozawa, M. Quantum measuring processes of continuous observables. Journal of Mathematical Physics 1984, 25, 79–87. [Google Scholar] [CrossRef]
- Busch, P.; Lahti, P.; Mittelstaedt, P. The Quantum Theory of Measurement, Lecture Notes in Physics Monographs, 2nd ed.; Springer: Berlin, 1996; Vol. m2. [Google Scholar] [CrossRef]
| 1 |
Writing with gives , . Choosing yields the displayed rate in the theorem statement for the geometric residual below. |
|
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2025 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).