1. Introduction
Diffusive models of belief adoption, protest mobilization, and organizational change share the following core intuition:
influence radiates outward through social ties but attenuates where saturation, resistance, or repression occur. The present paper formalizes this intuition in a degenerate anisotropic logistic–diffusion framework (introduced by us in [
1]) and analyses three geometric mechanisms: one-dimensional pinning, two-dimensional curvature–induced quenching, and curvature–targeted control. Before outlining the mathematical results, we motivate the study by surveying empirical domains in which these mechanisms manifest with striking regularity.
Protest diffusion and containment: Recent cross-national panel analyses covering more than thirty European states (2000–2015) reveal robust spatial contagion of protest: unrest in one country raises both domestic and neighboring protest frequency in subsequent periods, producing traveling waves of contention. Yet such waves rarely expand indefinitely. Data from the 2009–2010 Iranian Green Movement, the 2019 Hong Kong demonstrations, and the 2020 Belarus protests show abrupt “arrests” of expansion after highly localized crackdowns, corridor blockades, or internet throttling, tactics that leave low-curvature arterial routes open but pin high-curvature flanks. The pattern echoes our one-dimensional analysis, that is, once a single depth layer of a protest hierarchy is saturated by repression, upward diffusion of mobilization stalls.
Diffusion and suppression of misinformation: At the level of online discourse, large-scale Twitter studies demonstrate that false news diffuses more rapidly and broadly than factual information, often through a few high-degree hub accounts. Platform interventions increasingly adopt a geometry-aware approach, focusing resources on hub removal, flagging, or rate-limiting where the network curvature, informally, the concentration of potential diffusion paths, is highest. Experiments show that deleting or down-ranking a malicious post within thirty minutes can cut cumulative downstream engagement by over 90%, whereas uniform throttling of all content achieves far less per unit effort. These findings parallel our two-dimensional quenching criterion, that is, targeting regions of maximal projected curvature offers disproportionate leverage.
Organizational inertia and targeted dissent suppression: Inside bureaucratic hierarchies, information can likewise become “pinned”. Morrison and Milliken’s organizational–silence framework, replicated across corporate and governmental settings, shows that employees suppress negative feedback when punitive norms saturate a particular managerial layer, effectively halting the upward flow of corrective signals. Recent fieldwork in United Nations peacekeeping departments reveals how “anticipatory obedience” concentrates along specific reporting corridors, quenching reform initiatives transversely while allowing day-to-day operations along the main vertical chain to proceed unchecked.
Historical precedents of strategic control: Eastern Germany’s two-stage containment strategy (rapid arrest of focal organizers followed by selective media blackouts prior to construction of the Berlin Wall) illustrates curvature–targeted repression avant la lettre. Similarly, Soviet disaster archives on the Chernobyl accident document top-layer narrative pinning (“all systems normal”) that froze internal dissent until the curvature of external scrutiny overwhelmed institutional inertia, precipitating a preference cascade.
These examples demonstrate that pinning, quenching, and curvature–sensitive control are not abstract artifacts but recurring features of social and political dynamics. Our goal is therefore twofold.
First, we develop a minimal anisotropic logistic–diffusion model that captures the emergence and arrest of propagating fronts under nonlinear saturation and geometrically localized resistance.
Second, we derive and analyze a variational principle for curvature–targeted control, identifying how limited repressive or corrective resources can be deployed to maximize suppression efficiency.
The subsequent sections are structured as follows:
Section 2 introduces the probabilistic foundations of the underlying network–influence model;
Section 3 derives its continuum limit and formulates the associated degenerate parabolic logistic equation;
Section 4 establishes well-posedness and long-time behavior in the one-dimensional case;
Section 5 investigates front selection and pinning phenomena;
Section 6 extends the analysis to two dimensions, detailing curvature–induced quenching under anisotropic diffusion;
Section 7 presents the optimal control formalism and numerical illustrations;
Section 8 interprets these mathematical results in light of empirical evidence from protest dynamics, organizational sociology, and digital media research; and
Section 9 concludes with broader implications and directions for future work.
The central question addressed in this work is how spatially heterogeneous belief propagation, governed by network-derived anisotropic logistic–diffusion dynamics, can be systematically described and optimally suppressed. We investigate how anisotropy and interface curvature interact to pin or quench belief fronts, and derive a control framework that prescribes the spatial allocation of limited repressive resources so as to maximize retardation of propagation.
2. Probabilistic Foundations of the Network–Influence Model
Let a finite population of N agents be indexed by i = 1,…,N, each of whom is exposed to a particular informational proposition, social norm, innovation, or ideological stance, denoted collectively as a designated statement A. To formalize the adoption of A, we associate with each agent i a random variable Xi ∈ {0,1}, where the event signifies that agent i has accepted or internalized the statement. All probabilistic quantities introduced below are rigorously defined with respect to the canonical product probability space generated by the family , where is the minimal -algebra capturing joint distributions over all binary adoption states in the system. The prior probability that agent i independently adopts A prior to any social interaction or informational exchange is denoted as , where is the agent-specific information field summarizing personal history, preferences, heuristics, and possibly exposure to external media or previous encounters.
To account for socially mediated influence, we introduce a row-stochastic matrix
, where each entry
denotes the conditional probability that agent
j takes agent
i as an epistemic reference point when updating her belief. Structurally,
defines a weighted influence graph over the agent population, whereas probabilistically, it serves as a transition kernel for a Markov chain on the social network, reflecting the directional flow of informational trust. Each agent is further endowed with an individual-level parameter
, representing
obstinacy, i.e., the cognitive inertia or resistance to social influence. The posterior probability that agent
j adopts
A after a single equilibrium round of belief pooling is then given by the convex combination
which may be interpreted as the conditional probability
after integrating private priors with network-structured second-order beliefs, where
is the
-algebra generated by agent
j’s direct observations of the prior beliefs held by her reference group as encoded in
. This expression extends classical DeGroot-type models [
2] by allowing each agent to partially retain individual priors with strength
, a feature empirically validated by studies in political psychology and communication science that associate increased obstinacy with ideological commitment, identity-protective cognition, and motivated reasoning [
3,
4].
In matrix notation, denoting by
the vector of prior adoption probabilities and by
the diagonal obstinacy matrix, we write the posterior update rule in fixed-point form as
which resolves uniquely to the closed-form solution
The resolvent operator S encapsulates the cumulative outcome of all higher-order influence pathways over the social graph, incorporating iterated consensus-forming interactions under the constraint of individual obstinacy. Each entry may, thus, be interpreted as the marginal contribution of agent j’s prior belief to agent i’s stabilized posterior, and operator S plays a central role in the subsequent dynamical model by fixing the weights of influence propagation in both discrete- and continuous-time regimes.
Within the probabilistic framework described above, we distinguish two fundamentally different learning regimes based on the nature of exposure accumulation over time.
In the
fully correlated (memory-preserving) regime, agents integrate all influence signals received from the onset of observation (
) without any decay or reset, leading to temporally entangled adoption dynamics. Let
denote the cumulative probability of adoption for agent
i after
discrete rounds of exposure, each governed by the same influence operator
S. Then, the adoption probability satisfies the exact formula
as derived in [
1], where the integral spans all possible configurations of prior belief probabilities
. This formulation captures the Bayesian updating of adoption likelihoods under perfect memory retention and statistical independence of priors across agents. The factor
represents the cumulative probability that agent
i has failed to adopt in
t successive exposures, while the integrand weights this failure by the total social signal
received at each round. As
, the decay of
becomes sub-exponential, resulting in heavy tails in the distribution of adoption times and the divergence of the expected learning time. Consequently, convergence in this regime is slow and strongly influenced by the global distribution of initial signals.
By contrast, the
memory-less or Bernoulli–hazard regime assumes that each exposure event is statistically independent of prior ones, modeling probabilistic adoption as a continuous-time process governed by instantaneous hazard rates induced by socially aggregated influence. Letting
for
, we obtain the network-coupled nonlinear differential equation
as derived in [
1]. Here,
denotes the instantaneous probability that agent
i remains unconverted at time
t, while the aggregated influence
represents the social pressure from already-adopting peers. The structure of the influence matrix
S encodes both the topology and weighting of inter-agent effects.
Solutions
to Equation (
5) exhibit strictly monotonic and sigmoidal dynamics under general assumptions, namely, that
S is nonnegative and irreducible, and
. Indeed, monotonicity follows since for all
, both multiplicative factors on the right-hand side are positive, which implies that
is strictly increasing. The upper bound
is maintained at finite time due to the vanishing of
as
, and saturation occurs in the long-time limit:
. These properties confirm that
for all
.
To rigorously establish the sigmoidal profile of the adoption trajectory
, consider its second derivative:
Introducing the auxiliary quantities
and
we obtain the simplified form
Since
is increasing and bounded,
initially grows faster than
but eventually lags behind as
flattens. Hence, there exists a critical time
, such that
, with
for
and
for
. This confirms that
has a single inflection point and, hence, is sigmoidal. The logistic structure of (
5), therefore, guarantees that all solution trajectories are smooth, strictly increasing, and exhibit a single peak in their growth rate. Consequently, the probabilistic interpretation of
as a cumulative hazard function remains valid, and the definition of the
mean learning time (MLT) given by the first moment of the activation rate,
is mathematically well-defined and directly applicable to the matrix–logistic dynamics. The quantity
in (
5) thus serves as a valid probability density function on
for each agent’s stochastic adoption time. The heterogeneous structure of
S implies agent-dependent learning curves, which reflect both local influence topology and initial configuration.
The global behavior of belief propagation in the memory-less regime is ultimately governed by the spectral properties of the base conformity matrix
, which indirectly shapes the influence matrix
S via (
3). As a row-stochastic matrix,
admits a spectral decomposition
where
corresponds to the trivial consensus eigenvalue, and the associated eigenvector is the invariant distribution
satisfying
,
. The spectral gap
controls the rate at which the discrete-time dynamics
converge to the rank-one projection
. In the symmetric case (i.e., when
is reversible), this convergence is diffusive and uniform; in the asymmetric case, non-normality can lead to directional biases, transient amplification, or localization of influence. Obstinacy modulates this convergence by transforming the effective influence kernel to
, scaling all nontrivial eigenvalues by factors bounded above by
. Consequently, obstinate agents resist rapid shifts in belief, leading to local smoothing but also global delays. This introduces a structural trade-off—while high obstinacy dampens local volatility and slows saturation, it can impede large-scale propagation, especially in sparse or weakly connected regions. This interplay between spectral structure, individual inertia, and global coordination speed will be quantitatively analyzed in subsequent sections, particularly through the lens of front velocity, saturation depth, and mean learning time.
3. Continuum Limit of the Matrix–Logistic Dynamics
We now derive the continuum limit of the matrix–logistic Equation (
5), which governs probabilistic adoption dynamics over a structured influence network. Consider a large-scale population embedded in a spatially regular lattice graph
, either one- or two-dimensional, with
nodes and uniform spacing
such that
defines the spatial position of node
i. For a smooth adoption profile
interpolating the discrete probabilities
, we assume that
in the limit
. The influence matrix
S is specified through the network structure and obstinacy parameters via Equation (
3), and we assume that its entries correspond to a discrete diffusive kernel with locally varying coefficients, for instance,
or more generally, a smoothed convolution operator over local neighborhoods.
Under these assumptions, the aggregated influence term in (
5) admits the diffusive approximation:
where
is a spatially dependent effective diffusion coefficient emerging from the composition of the local conformity weights
and obstinacy profile
, and
is the Laplacian operator in the continuous spatial domain. Substituting into (
5) and passing to the continuum limit yields the nonlinear degenerate diffusion equation:
This characterizes adoption under locally modulated persuasion and saturating nonlinearity. If the aggregated influence includes both diffusive dispersion and a mass amplification term, we assume
where
reflects autocatalytic amplification of beliefs in ideologically resonant regions. The resulting continuum model becomes a nonlinear logistic–diffusion PDE:
To interpret the coefficients
and
, we expand the Neumann series for the influence matrix:
In the regime of weak obstinacy
, this yields
indicating a diffusive process governed by
. In the opposite limit
, we obtain
and the influence structure collapses to identity, halting propagation. Consequently, the coefficients are approximately
where
denotes a measure of the local spectral spread or diffusion strength encoded in
. To formalize this, let
denote the
i-th row of the conformity matrix, interpreted as a probability distribution over neighboring nodes. The local spectral spread at node
i is then defined as the second spatial moment:
which quantifies the variance of the influence distribution around agent
i. In regular lattices with uniform spacing
, this yields
, so the diffusion coefficient becomes
, in agreement with classical finite-difference approximations of diffusive transport. For general graphs with spatial embeddings,
characterizes the effective width of the influence kernel and reflects how broadly social influence disperses from each node.
In the continuum limit, belief dynamics synthesize two structurally distinct mechanisms: logistic self-reinforcement and saturation-limited diffusion. Both mechanisms admit closed-form traveling wave solutions, but only in the one-dimensional setting, where spatial ordering allows reduction to ordinary differential equations [
5]. In higher-dimensional or networked domains, such explicit solutions generally fail, and front propagation becomes sensitive to geometry, anisotropy, and boundary effects. The analysis in [
5] applies specifically to depth-ordered hierarchical structures, where a one-dimensional continuum approximation is valid.
The self-reinforcing component is governed by the classical Fisher–Kolmogorov–Petrovsky–Piskunov (FKPP) equation [
6,
7]:
which supports traveling wave solutions
of sigmoidal form,
with propagation speed
In contrast, the nonlinear diffusion mechanism is described by
which degenerates as
, reflecting the inhibition of influence propagation in saturated regions. This equation admits an exact traveling wave solution [
5] of the form
, where
, given by
with
denoting the principal branch of the Lambert function. Here,
represents a global phase shift determined by initial conditions, specifying the spatiotemporal location of the front, while
denotes the internal coordinate at which the front reaches full saturation
. The difference
encodes the effective delay of front initiation and governs the onset position of activation in space-time. The normalization condition
leads to the linear front velocity
in contrast to the square-root scaling of FKPP fronts (
20). The resulting profile
remains bounded in
and rises monotonically for
, describing a purely diffusive sigmoidal front with finite support and delayed activation.
5. Pinning and Front Selection in One Dimension
In this section, we develop a rigorous theory of front propagation in the one-dimensional degenerate logistic–diffusion model (
24), focusing on the following two dynamically distinct mechanisms: (i) the emergence and spatial localization of pinned plateaus near saturation; and (ii) the nonlinear selection of front velocity via matched asymptotics between the weak-signal and degenerate regions. The mobility coefficient
vanishes near saturation and encodes the mechanism of spontaneous arrest. As shown below, any saturated point
initiates a region of finite-speed propagation governed by the traveling-wave reduction (
49), whose phase-plane structure reveals a critical half-width below which no admissible front can connect the saturated and unsaturated states [
5].
We define a
pinned plateau as a nontrivial interval on which the solution remains saturated irreversibly. More precisely, for some
and interval
,
since the flux
vanishes identically at saturation [
19]. In the socio-physical interpretation of (
24), this corresponds to a
frozen-conviction zone: a saturated agent cluster becomes dynamically inert, and external influence cannot penetrate it. The domain splits into an active sector
, where diffusion operates, and a pinned core whose survival depends on initial overshoot and the ratio
.
A refined maximum principle explains the irreversibility of such plateaus. Let
u be a bounded weak solution, and suppose
is achieved in the interior. The degeneracy annihilates diffusion, and evaluating the PDE yields
so that all inequalities are saturated. The strong maximum principle [
8,
19] then forces
in a neighborhood, and unique continuation implies persistence. A sharp threshold follows from matched asymptotics—equating the exponential tail
, where
, to the interior curvature scale yields the minimal half-width
so that initial plateaus of width less than
vanish, while wider ones persist.
Further insight comes from the phase portrait of the traveling-wave system. Setting
, Equation (
49) becomes
with degenerate saddle at
. Linearization reveals one zero and one diverging eigenvalue, placing the origin on a non-hyperbolic center manifold. Center-manifold theory [
31] shows that orbits launched at
remain pinned, while those with
escape immediately. The entire segment
is invariant, structurally enforcing pinning.
Numerical integration (see
Figure 1) confirms this: given a saturated initial patch, the edge recedes at finite velocity
, and the interior remains flat. Rescaling
collapses interface profiles to a similarity shape, consistent with the law
. Motion halts precisely when the profile shrinks to size
, after which the segment remains permanently pinned.
Finally, we examine front selection numerically. The critical velocity
arises from linearization at the leading edge and marks the pulled/pushed transition. Subcritical waves violate monotonicity and decay conditions, while supercritical ones connect
to
smoothly.
Figure 2 confirms this dichotomy. Thus, despite the absence of spectral bifurcations, the system exhibits sharp thresholds, as follows:
for plateau survival, and
for admissible propagation.
This completes the analytic picture—pinned regions emerge and persist due to degeneracy, propagation occurs when saturation is sufficient, and coherent fronts exist only when their speed exceeds the linear threshold. No instability or bifurcation was observed; hence, the degenerate logistic–diffusion model in one dimension admits only robust, threshold-governed dynamical regimes.
6. Two-Dimensional Geometry, Anisotropy, and Curvature–Induced Quenching in Degenerate Parabolic Logistic Models
We now turn to the analysis of degenerate logistic belief propagation in spatial domains of dimension , focusing on geometric and tensorial extensions of the one-dimensional dynamics studied previously. In higher dimensions, the propagation of conviction fronts becomes sensitive not only to the degeneracy structure , which encodes saturation-induced diffusion arrest, but also to anisotropy arising from the directional asymmetries of the influence network.
The motivating discrete structure is depicted in
Figure 3, and consists of a two-dimensional directed grid of agents arranged on a square lattice augmented by diagonal links, forming a full 8-neighbor Moore-type connectivity pattern. Each node
influences its immediate neighbors in the horizontal, vertical, and diagonal directions, with direction-dependent conformity weights as indicated. The central node at the spatial index
, marked in red, acts as a fully convinced teacher—its belief state is fixed at unity and its obstinacy coefficient normalized as
. All other agents share homogeneous low obstinacy
, and evolve under the matrix logistic dynamics. This configuration defines a canonical anisotropic two-dimensional influence network whose macroscopic limit inherits both degeneracy and directional asymmetry.
We define the influence matrix
S via the posterior amplification structure (
3), with homogeneous obstinacy
for all
, and a full Moore-type neighborhood weighted by direction-dependent conformity coefficients. For each node
, the matrix element
is nonzero only if
, where
denotes the ordered list of horizontal, vertical, and diagonal predecessors of
i, each linked by weight
according to their geometric orientation. The resulting belief dynamics on the lattice take the form
with fixed boundary condition
. The structure of
S encodes both local amplification and anisotropic influence geometry, inducing direction-sensitive nonlinear propagation and degenerate self-reinforcement at saturation.
Rescaling lattice indices by
and Taylor-expanding the Moore–sum to
converts the matrix–logistic rule to the degenerate reaction–diffusion form
where the anisotropic tensor
and the point source
inherit the directional weights of the lattice. Anisotropy enters through the off-diagonal component
. To eliminate shear while retaining axis heterogeneity, we set the two diagonal weights equal,
, yielding the diagonal model
with
Equation (
60) captures the minimal two-dimensional geometry—axis-weighted elliptic spreading governed solely by the ratio
and quenched as
, without the additional tilt that would arise from unequal diagonal links. Throughout the numerical experiments, we fix
and choose the conformity weights
so that the effective diffusivities satisfy
and
.
Passing to the sharp–interface limit of (
60), we separate an outer region, where
u is essentially binary (
or
), from an inner layer of width
centered on the isoconviction curve
.
In stretched normal coordinates
, a matched–asymptotic expansion à la Fife–McLeod [
32] shows that the leading inner profile
travels with speed
c and effective diffusivity
, where
is the unit normal to
. The Euclidean curvature of the interface is defined by the following:
Flux matching gives the normal velocity law
with
, denoting the mid-layer value of
U. Equation (
63) describes anisotropic mean–curvature motion attenuated by saturation; if
or
becomes too small, the interface stalls, a phenomenon we term
curvature–induced quenching. For an initially circular patch of radius
, one obtains the self-similar ellipse
verified numerically and exhibiting pinning along the slow axis when
.
With a diagonal tensor
the law (
63) reduces to
so that only
modulate curvature. A radially symmetric contour evolves via
, giving the critical stopping radius
where
is the one-dimensional wave speed extracted from
. If
, the level set elongates with
; hence, the fast axis expands while the slow one can hit
first, arresting motion transversally, a behavior reproduced in
Figure 4.
A rigorous quenching criterion emerges by minimizing the mobility in (
65). Setting
yields the maximal admissible curvature
so the interface stops whenever
. Analytically, this inequality defines an anisotropic Wulff shape bounding all feasible fronts; linear stability of a planar wave gives the neutral wavenumber
. Numerically, level-set or phase-field discretizations of (
60) corroborate these predictions; an initially elliptic interface grows according to (
64) until the minor semi-axis meets the curvature threshold
, after which motion ceases along the slow direction while persisting along the fast one, reproducing the pinned contours in
Figure 4. Thus, mapping
and validating its geometric realization constitute the next step in both analytical refinement and high-resolution computation.
Numerical confirmation of this criterion is provided in
Figure 5, where we visualize the evolution of an initially circular front under strongly anisotropic diffusion. The profile
forms an elongated lobe along the fast axis, while transversal expansion halts due to curvature pinning.
7. Curvature–Targeted Control of Belief Propagation in Two Dimensions
We now address the problem of curvature–targeted suppression of front propagation under resource constraints [
33]. In many realistic settings, ranging from protest diffusion and epidemic spread to adversarial information campaigns, blanket intervention across the domain is either infeasible or counterproductive.
Instead, inhibition must be applied selectively and strategically, focusing on geometrically sensitive regions where the front is most susceptible to arrest. To this end, we developed a sharp–interface control formulation that quantifies how a limited repressive rate , concentrated on a thin neighborhood of the isoconviction manifold , perturbs the matched–asymptotic balance governing front motion and prescribes an optimal spatial allocation strategy that maximizes inhibitory efficiency under a finite enforcement budget.
Augmenting the anisotropic logistic–diffusion field with a nonnegative
repressive rate concentrated in a thin tubular neighborhood of the isoconviction manifold
, perturbs the matched–asymptotic balance that governs front motion. Expressed in arc length–time coordinates
, the uncontrolled normal speed reads
, where
projects the mobility tensor onto the unit normal
, and the Euclidean curvature is
. Introducing
q removes growth rather than adds it, so the sharp–interface velocity becomes
During an elementary interval
, the occupied area therefore varies by
where the sign convention is such that
denotes outward motion. If an instantaneous enforcement budget
is prescribed, the optimization problem becomes the allocation of
q along
so as to maximize the negative first term in (
69) while respecting (
70); equivalently, one seeks to deploy scarce repression where curvature is high and projected diffusivity is low, thereby exerting the greatest leverage on the advance of the front without resorting to blanket suppression.
Maximizing the retardation of front propagation under the constraint of a fixed enforcement budget leads to a variational optimization problem formulated over the interface
. The control density
is determined by maximizing the inhibitory effect per unit resource, giving rise to the Lagrangian functional
where
enforces the linear constraint
. Taking the first variation of
and imposing non-negativity of
q leads to a complementary–slackness condition, that ism optimal control vanishes on segments where curvature is below the quenching threshold
, and is positive only where the local geometry is sufficiently convex to allow effective intervention. On these segments, the optimal profile is given by
where
denotes the positive part and
is the curvature threshold derived in Equation (
67). The function
acts as a normalizing factor, ensuring that the total enforcement matches the budget
. The resulting distribution of repression is thus sharply focused: it concentrates entirely on high-curvature flanks where the interface exhibits reduced projected mobility and is therefore most susceptible to quenching. In such regions, even modest local inhibition suffices to arrest propagation, enabling targeted suppression without recourse to full-domain intervention. The addition of a bounded, spatially localized repressive term
preserves the degenerate parabolic structure of Equation (
60). The existence of weak solutions for such semilinear degenerate parabolic PDEs with measure-valued sources was established in, e.g., Pierre and Schmitt (1985) [
34] and Quittner and Souplet (2007, Chapter 4) [
35]. This ensures that the control-augmented problem remains mathematically well-posed under the proposed feedback law.
To demonstrate the efficacy of curvature–targeted suppression, we integrate Equation (
60) numerically in a rectangular domain with strongly anisotropic diffusion coefficients
. The initial profile consists of a compactly supported teacher source centered within the domain, and the control law
prescribed by Equation (
72) is activated within a dynamically evolving elliptic shell that closely tracks the support of the uncontrolled interface
.
The resulting simulations, summarized in
Figure 6a, reveal a pronounced asymmetry in the suppression effect—the active repressive shell, visible as a dark-blue trench encircling the peak of the belief field, induces substantial retardation along the transverse (
y) direction while exerting negligible influence along the fast (
x) axis. Quantitatively, the radial extent in the
y-direction is reduced by approximately 40% at time
, despite the total resource budget
remaining fixed and the feedback being confined to a thin boundary layer. This outcome aligns precisely with the theoretical prediction based on the projected mobility
, which is minimal along the short axis of the elliptic front and thereby maximizes the leverage of localized repression. To further highlight the spatial localization of the repressive feedback,
Figure 6b presents a top-view projection of the same simulation with arrows marking the regions of highest curvature. These flanks correspond to the loci of maximal
intensity and exhibit clear suppression effects.
This example substantiates the central hypothesis that finely tuned curvature–selective feedback can yield nontrivial spatial inhibition with minimal energetic expenditure. Operationally, the prescription (
72) admits efficient implementation; once the evolving front
is approximated (e.g., as a level set of the solution field), the curvature
and projected diffusivity
can be computed locally, and the control density
is obtained by thresholding against the analytically derived critical curvature
and allocating the available budget proportionally. The conceptual and computational tractability of this scheme offers a robust baseline for more sophisticated control frameworks, such as those incorporating spatial heterogeneity in diffusivity, temporally adaptive budgets, or second-order penalization. These generalizations present natural extensions for future work, particularly in domains where spatial anisotropy and limited intervention capacity dominate system dynamics.
Although the control law (
72) depends on local curvature estimation, our numerical tests indicate that the integral budget constraint and the positive-part thresholding both help regularize noise effects. In practice, smoothing of the level-set field or local averaging of curvature estimates can further mitigate discretization artifacts. The resulting allocation remains sharply localized but stable, suggesting practical robustness even on moderate-resolution grids.
The numerical experiments were conducted on uniform Cartesian grids with typical resolution , and time steps chosen to respect the diffusive CFL condition. Discretization employed standard second-order central differences in space and forward Euler in time. Boundary conditions were implemented as homogeneous Neumann via ghost cells. The control law was computed at each time step using level–set–extracted contours to define , with curvature estimated via finite differences and smoothed over 3–5 grid cells to mitigate noise. Convergence was verified by halving the grid spacing and time step, confirming consistent front profiles and suppression effects within variation. Typical runs over used fewer than 200 steps and completed in under a minute on a standard desktop machine. While explicit methods suffice for this regime, future work will explore adaptive meshing and semi-implicit solvers for large-scale or long-time simulations.
Energetic and thermal interpretation.
Equation (
68) admits a natural gradient–flow structure in the sense of Otto’s calculus [
36,
37]; there exists a Lyapunov (free-energy) functional
such that the deterministic dynamics can be rewritten as
that is, as a zero-temperature steepest descent in
with mobility
. The quenching threshold defined in (
72) acts as an
energy barrier; that is, front advance ceases when the local energetic cost
exceeds the driving force
, rendering the interface metastable. The optimal control law (
72) can be reinterpreted as a minimal-action principle: the density
elevates the landscape
, and the Lagrange multiplier
distributes the finite budget
where
is maximal, namely, along high-curvature/low-mobility arcs.
Introducing environmental volatility amounts to supplementing (
68) with a Dean-type stochastic flux [
38]
where
is an
effective temperature quantifying exogenous information noise and
is vectorial space–time white noise. Finite
T allows thermally activated barrier crossing: Kramers’ theory [
39] predicts a depinning rate
so that the sharp threshold in (
72) is blurred and long-time spreading resumes once
becomes comparable to the energetic gap
. In microscopic terms, the same
appears as the variance of adoption–noise at the agent level, reducing effective obstinacy and enlarging
D. Numerically, grid–scale truncation error already supplies a vanishing background temperature; explicit simulations of (
75) would require Itô time-stepping and ensemble averaging, yet our tests show that the geometry of
remains robust provided
. We will explore the finite-temperature regime in future work.
9. Conclusions
This study has advanced a unified geometric-probabilistic framework for the spatiotemporal evolution of collective conviction, protest mobilization, and organizational change.
Starting from a network–influence model with agent-level obstinacy, we derived a continuum-limit equation of degenerate anisotropic logistic–diffusion, established well-posedness in one dimension, and characterized front selection and pinning via spectral and entropy methods.
Extending the analysis to two spatial dimensions revealed a curvature-induced quenching mechanism: that is, propagation arrests once the product of projected mobility and interface curvature falls below a critical threshold.
We further developed a variational law of curvature–targeted control, showing that scarce repressive (or corrective) resources should be concentrated along high-curvature, low-mobility flanks to maximize retardation. Numerical experiments confirm the analytic predictions, while an extensive discussion maps the following three principal phenomena: one-dimensional pinning, two-dimensional quenching, and optimal control onto data-rich cases from protest science, misinformation studies, and organizational sociology.
Broader implications: The degenerate logistic-diffusion paradigm provides a minimal yet expressive template for modeling opinion fronts, behavioral contagion, and institutional inertia in heterogeneous media. The curvature law captures the intuitive idea that diffusion slows in saturated regions and at sharply curved protrusions, which is insightful for risk assessment in domains as diverse as crowd management, algorithmic content moderation, and policy rollout. The control result offers a quantitative rationale for selective, geometry-aware interventions, guiding law-enforcement tactics, fact-checking priorities, or organizational feedback channels, while cautioning against blanket suppression strategies that dissipate resources without the proportional effect.
Future work: Three natural extensions are immediate. (i) Stochastic heterogeneity: Allowing diffusion coefficients, obstinacy, or amplification rates to vary randomly in space and time would bridge to empirical settings with dynamic network topologies and shifting media ecologies. Rigorous quantification of front speed distributions and quenching probabilities under noise remains largely open. (ii) Fully tensorial mobility: Relaxing the diagonal assumption and analyzing shear-induced oblique pinning could uncover new geometric regimes, including non-convex Wulff envelopes and orientation locking. (iii) Adaptive or adversarial control: Embedding the curvature-targeted feedback into a game-theoretic setting, where propagators and suppressors co-evolve strategies, would link the present theory to real-time protest policing, cyber-warfare over narratives, and organizational learning under dissent. Empirically, high-resolution protest trajectories, platform-level misinformation logs, and intra-organizational communication audits constitute fertile datasets for calibrating and validating the model.
By integrating geometric PDE analysis with probabilistic network foundations and empirically grounded control concepts, the work lays a mathematical cornerstone for the quantitative social science of collective dynamics under constraint.