Spatially-invariant opinion dynamics on the circle

Giovanna Amorim1, Alessio Franci2, and Naomi Ehrich Leonard1 This work was supported in part by ONR grant N00014-19-1-2556.1 G. Amorim, N.E. Leonard are with the Department of Mechanical and Aerospace Engineering, Princeton University, Princeton, NJ 08544 USA, {giamorim, naomi}@princeton.edu2 A. Franci is with the Department of Electrical Engineering and Computer Science, University of Liege, 10 Allée de la Découverte, 4000 Liège, Belgium, and with the WEL Research Institute, Wavre, Belgium, afranci@uliege.be
Abstract

We propose a nonlinear opinion dynamics model for an agent making decisions about a continuous distribution of options in the presence of distributed input. Inspired by perceptual decision-making, we develop the theory for options distributed on the circle, representing, e.g., the set of possible heading directions in planar robotic navigation problems. Interactions among options are encoded through a spatially invariant kernel. We design the kernel to ensure that decision-making is robust, in the sense that only a finite subset of options can be favored over the continuum. We prove the spatial invariance of the model linearization and use this result to prove an opinion-forming bifurcation in the model with zero input. We then use space and time frequency domain analysis of the model linearization to infer the ultra-sensitive input-output behavior of the nonlinear dynamics with input. We illustrate our model’s versatility with a robotic navigation example in crowded spaces.

I INTRODUCTION

In perceptual decision-making, animals use sensory information to respond to their environment. Both the sensory information and the perceptual response are spatially distributed. For example, visual information, like object positions, is distributed both in the physical world and in its neural representation in the visual cortex [1]. Similarly, in robotics, an agent can use a distributed representation of its visual field and the objects in it to drive decisions, e.g., to navigate crowded spaces or track targets as in  [2, 3, 4, 5].

Motivated by perceptual decision-making, we propose a distributed nonlinear opinion dynamics model for an agent making decisions about a continuous distribution of options in the presence of continuously distributed input. We study option and input spaces defined by the unit circle, representing both a one-dimensional visual field and the set of possible heading directions. Our model generalizes the nonlinear opinion dynamics (NOD) of [6, 7] by extending it from a finite set to a continuous distribution of options. NOD has been used in a variety of applications, such as robotics and automated driving [8, 9, 10, 11, 12, 13]. These works leverage the fast-and-flexible decision-making of NOD resulting from a bifurcation in the dynamics [7].

The distributed model we present here inherits the same features as a result of a bifurcation in its infinite-dimensional dynamics. The proposed dynamics also allow for increased scalability and use in more complex applications. For instance, in an obstacle avoidance scenario as the one in [8], the representation of objects in the distributed NOD does not require prior knowledge of the number of objects to be avoided. In the finite-dimensional NOD this prior information is, in principle, fundamental to the formulation.

Our model and the bifurcation behavior we prove is closely related to the spatiotemporal neural field models of mathematical neuroscience [14, 15, 1, 16], which consider spatially-invariant integro-differential equations with a convolution kernel that captures excitatory and inhibitory interactions between different regions of the neural field. Response to input is considered in [16, 17, 18] but only for specific classes of inputs. Our approach differs from those works by using a spatiotemporal transfer function approach, which allows us to predict the response of the nonlinear model to arbitrary inputs from the response of its linearization.

For illustration, we consider a robot navigation scenario where the visual input corresponds to the locations of (possibly many) obstacle-free “gaps”. This scenario is inspired by classical empirical approaches exploring the use of neural fields for robotic navigation [2, 3, 4, 5]. Our model provides a theoretical foundation to such approaches that can be used for analysis, design, and control in more general scenarios.

Our contributions are as follows. First, we propose a nonlinear opinion dynamics model for an agent making decisions about a continuous distribution of options and in the presence of continuously distributed inputs. Second, we prove the spatial invariance of the model linearization. Third, we use spatial-invariance of the linearized dynamics to prove the existence of an opinion forming bifurcation for the model with zero input. Fourth, we use space and time frequency domain analysis of the model linearization to infer the input-output behavior of the nonlinear dynamics to arbitrary inputs. Fifth, we propose a framework for designing kernels for an application of our model to robotic navigation.

Notation and mathematical background are in Section II. We introduce the nonlinear opinion dynamics for a continuous distribution of options in Section III. We prove the spatial invariance of the model linearization in Section IV. In Section V we leverage spatial-invariance of the linearized dynamics to prove an opinion forming bifurcation in the model with zero input. In Section VI, we use spatial and time frequency analysis of the model linearization to infer the model’s input-output behavior. We propose a frequency-domain kernel design approach and illustrate the benefits of our approach for robotic navigation in Section VII. A discussion and final remarks are provided in Section VIII.

II MATHEMATICAL PRELIMINARIES

We denote the set of integer values as {\mathbb{Z}}blackboard_Z, the set of non-negative integer values as {\mathbb{N}}blackboard_N, the set of real numbers as {\mathbb{R}}blackboard_R, and the set of complex numbers as {\mathbb{C}}blackboard_C. The unit circle is denoted by 𝕊1superscript𝕊1\mathbb{S}^{1}blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, i.e., 𝕊1=/superscript𝕊1\mathbb{S}^{1}={\mathbb{R}}/{\mathbb{Z}}blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT = blackboard_R / blackboard_Z. For a,b𝑎𝑏a,b\in{\mathbb{R}}italic_a , italic_b ∈ blackboard_R, the notation ab𝑎𝑏a\nearrow bitalic_a ↗ italic_b indicates the limit ab𝑎𝑏a\to bitalic_a → italic_b with a<b𝑎𝑏a<bitalic_a < italic_b. For a complex number s=σ+iω𝑠𝜎𝑖𝜔s=\sigma+i\omegaitalic_s = italic_σ + italic_i italic_ω, the real and imaginary parts are denoted as (s)𝑠\Re(s)roman_ℜ ( italic_s ) and (s)𝑠\Im(s)roman_ℑ ( italic_s ), respectively. We represent the complex conjugate as s¯=σiω¯𝑠𝜎𝑖𝜔\bar{s}=\sigma-i\omegaover¯ start_ARG italic_s end_ARG = italic_σ - italic_i italic_ω, the modulus as |s|=ss¯𝑠𝑠¯𝑠|s|=\sqrt{s\bar{s}}| italic_s | = square-root start_ARG italic_s over¯ start_ARG italic_s end_ARG end_ARG and the argument as arg(s)=limnn(s/|s|n)𝑠subscript𝑛𝑛𝑛𝑠𝑠\arg(s)=\lim_{n\to\infty}n\Im(\sqrt[n]{s/|s|})roman_arg ( italic_s ) = roman_lim start_POSTSUBSCRIPT italic_n → ∞ end_POSTSUBSCRIPT italic_n roman_ℑ ( nth-root start_ARG italic_n end_ARG start_ARG italic_s / | italic_s | end_ARG ) for n{0}𝑛0n\in{\mathbb{N}}-\{0\}italic_n ∈ blackboard_N - { 0 }.

The Hilbert space of square-integrable real functions on 𝕊1superscript𝕊1\mathbb{S}^{1}blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT is denoted by L2(𝕊1)subscript𝐿2superscript𝕊1L_{2}(\mathbb{S}^{1})italic_L start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ). Given v,wL2(𝕊1)𝑣𝑤subscript𝐿2superscript𝕊1v,w\in L_{2}(\mathbb{S}^{1})italic_v , italic_w ∈ italic_L start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ), their inner product is v,w=𝕊1v(θ)w(θ)𝑑θ𝑣𝑤subscriptsuperscript𝕊1𝑣𝜃𝑤𝜃differential-d𝜃\langle v,w\rangle=\int_{\mathbb{S}^{1}}v(\theta)w(\theta)d\theta⟨ italic_v , italic_w ⟩ = ∫ start_POSTSUBSCRIPT blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_v ( italic_θ ) italic_w ( italic_θ ) italic_d italic_θ. The induced norm is v=v,v1/2norm𝑣superscript𝑣𝑣12||v||=\langle v,v\rangle^{1/2}| | italic_v | | = ⟨ italic_v , italic_v ⟩ start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT.

We denote operators with capital roman letters. Let A:L2(𝕊1)L2(𝕊1):𝐴subscript𝐿2superscript𝕊1subscript𝐿2superscript𝕊1A:L_{2}(\mathbb{S}^{1})\to L_{2}(\mathbb{S}^{1})italic_A : italic_L start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) → italic_L start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) be a linear operator. We let the set Sp(A)={λk}Sp𝐴subscript𝜆𝑘\text{Sp}(A)=\{\lambda_{k}\}Sp ( italic_A ) = { italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT } denote the point spectrum of A𝐴Aitalic_A, if it is not empty. Each eigenvalue λkSp(A)subscript𝜆𝑘Sp𝐴\lambda_{k}\in\text{Sp}(A)italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∈ Sp ( italic_A ) satisfies Avk(θ)=λkvk(θ)𝐴subscript𝑣𝑘𝜃subscript𝜆𝑘subscript𝑣𝑘𝜃Av_{k}(\theta)=\lambda_{k}v_{k}(\theta)italic_A italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_θ ) = italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_θ ), where vkL2(𝕊1)subscript𝑣𝑘subscript𝐿2superscript𝕊1v_{k}\in L_{2}(\mathbb{S}^{1})italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∈ italic_L start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) denotes the eigenfunction corresponding to λksubscript𝜆𝑘\lambda_{k}italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT. We denote λmax=argmax{(λk)}subscript𝜆subscript𝜆𝑘\lambda_{\max}=\arg\!\max\{\Re(\lambda_{k})\}italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = roman_arg roman_max { roman_ℜ ( italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) } as the leading eigenvalue of A𝐴Aitalic_A, and its corresponding eigenfunction, as vmaxL2(𝕊1)subscript𝑣subscript𝐿2superscript𝕊1v_{\max}\in L_{2}(\mathbb{S}^{1})italic_v start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ∈ italic_L start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ), the leading mode.

Definition 1 (Differential operator):

Let F:L2(𝕊1)L2(𝕊1):𝐹subscript𝐿2superscript𝕊1subscript𝐿2superscript𝕊1F:L_{2}(\mathbb{S}^{1})\to L_{2}(\mathbb{S}^{1})italic_F : italic_L start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) → italic_L start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) be a nonlinear operator. The differential of F𝐹Fitalic_F in the direction of z𝑧zitalic_z at a point zsuperscript𝑧z^{*}italic_z start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, is

AF=DzF(z):=limϵ01ϵ(F(ϵz+z)F(z)),subscript𝐴𝐹subscript𝐷𝑧𝐹superscript𝑧assignsubscriptitalic-ϵ01italic-ϵ𝐹italic-ϵ𝑧superscript𝑧𝐹superscript𝑧\textstyle A_{F}=D_{z}F(z^{*}):=\lim_{\epsilon\to 0}\frac{1}{\epsilon}\big{(}F% (\epsilon z+z^{*})-F(z^{*})\big{)},italic_A start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = italic_D start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_F ( italic_z start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) := roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG italic_ϵ end_ARG ( italic_F ( italic_ϵ italic_z + italic_z start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) - italic_F ( italic_z start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ) ,

provided that the limit exists. If this limit exists, then F𝐹Fitalic_F is said to be differentiable in the direction z𝑧zitalic_z at point zsuperscript𝑧{z}^{*}italic_z start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT.

Definition 2 (Multiplication Operator):

A multiplication operator M𝑀Mitalic_M is defined by [Mh](x):=M(x)h(x)assigndelimited-[]𝑀𝑥𝑀𝑥𝑥[Mh](x):=M(x)h(x)[ italic_M italic_h ] ( italic_x ) := italic_M ( italic_x ) italic_h ( italic_x ), where hhitalic_h is in the domain of M𝑀Mitalic_M. Multiplication operators are the infinite-dimensional equivalent of diagonal matrices.

Definition 3 (Spatial shift operator [19], [20]):

The spatial shift operator denoted by Tψ:L2(𝕊1)L2(𝕊1):subscript𝑇𝜓superscript𝐿2superscript𝕊1superscript𝐿2superscript𝕊1T_{\psi}:L^{2}(\mathbb{S}^{1})\to L^{2}(\mathbb{S}^{1})italic_T start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT : italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) → italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) is defined as h(θ)[Tψh](θ):=h(θψ)maps-to𝜃delimited-[]subscript𝑇𝜓𝜃assign𝜃𝜓h(\theta)\mapsto[T_{\psi}h](\theta):=h(\theta-\psi)italic_h ( italic_θ ) ↦ [ italic_T start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT italic_h ] ( italic_θ ) := italic_h ( italic_θ - italic_ψ ) for ψ𝕊1𝜓superscript𝕊1\psi\in\mathbb{S}^{1}italic_ψ ∈ blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT and hL2(𝕊1)superscript𝐿2superscript𝕊1h\in L^{2}(\mathbb{S}^{1})italic_h ∈ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ).

Definition 4 (Spatially invariant operator [19], [20]):

An operator F𝐹Fitalic_F is spatially invariant if TψF=FTψsubscript𝑇𝜓𝐹𝐹subscript𝑇𝜓T_{\psi}F=FT_{\psi}italic_T start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT italic_F = italic_F italic_T start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT.

We mainly work with a special class of spatially invariant linear operators, namely, spatial convolution operators

[Az](θ):=𝕊1W(θϕ)z(ϕ)𝑑ϕ,assigndelimited-[]𝐴𝑧𝜃subscriptsuperscript𝕊1𝑊𝜃italic-ϕ𝑧italic-ϕdifferential-ditalic-ϕ\textstyle[Az](\theta):=\int_{\mathbb{S}^{1}}W(\theta-\phi)z(\phi)d\phi,[ italic_A italic_z ] ( italic_θ ) := ∫ start_POSTSUBSCRIPT blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_W ( italic_θ - italic_ϕ ) italic_z ( italic_ϕ ) italic_d italic_ϕ , (1)

where the convolution kernel W:𝕊1:𝑊superscript𝕊1W:\mathbb{S}^{1}\to{\mathbb{R}}italic_W : blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT → blackboard_R.

Definition 5 (Spatially Invariant Linear System [19], [20]):

Consider a spatiotemporal input-output linear system. Let u(,t),z(,t)L2(𝕊1)𝑢𝑡𝑧𝑡subscript𝐿2superscript𝕊1u(\cdot,t),z(\cdot,t)\in L_{2}(\mathbb{S}^{1})italic_u ( ⋅ , italic_t ) , italic_z ( ⋅ , italic_t ) ∈ italic_L start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) be the scalar-valued input and output functions at time t0𝑡subscriptabsent0t\in\mathbb{R}_{\geq 0}italic_t ∈ blackboard_R start_POSTSUBSCRIPT ≥ 0 end_POSTSUBSCRIPT, respectively. Let θ𝕊1𝜃superscript𝕊1\theta\in\mathbb{S}^{1}italic_θ ∈ blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT be the spatial coordinate. A linear system of the form

zt(θ,t)=[Az](θ,t)+[Bu](θ,t),𝑧𝑡𝜃𝑡delimited-[]𝐴𝑧𝜃𝑡delimited-[]𝐵𝑢𝜃𝑡\textstyle\frac{\partial z}{\partial t}(\theta,t)=[Az](\theta,t)+[Bu](\theta,t),divide start_ARG ∂ italic_z end_ARG start_ARG ∂ italic_t end_ARG ( italic_θ , italic_t ) = [ italic_A italic_z ] ( italic_θ , italic_t ) + [ italic_B italic_u ] ( italic_θ , italic_t ) , (2)

is spatially invariant if the linear operators A𝐴Aitalic_A, B𝐵Bitalic_B are spatially invariant.

Definition 6 (Spatial Fourier transform [19], [20]):

Let f,g:𝕊1×0:𝑓𝑔superscript𝕊1subscriptabsent0f,g:\mathbb{S}^{1}\times{\mathbb{R}}_{\geq 0}italic_f , italic_g : blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT × blackboard_R start_POSTSUBSCRIPT ≥ 0 end_POSTSUBSCRIPT be spatiotemporal fields with spatial and time coordinates θ𝕊1𝜃superscript𝕊1\theta\in\mathbb{S}^{1}italic_θ ∈ blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT and t0𝑡subscriptabsent0t\in{\mathbb{R}}_{\geq 0}italic_t ∈ blackboard_R start_POSTSUBSCRIPT ≥ 0 end_POSTSUBSCRIPT. Suppose f(,t),g(,t)L2(𝕊1)𝑓𝑡𝑔𝑡superscript𝐿2superscript𝕊1f(\cdot,t),g(\cdot,t)\in L^{2}(\mathbb{S}^{1})italic_f ( ⋅ , italic_t ) , italic_g ( ⋅ , italic_t ) ∈ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) for all t0𝑡subscriptabsent0t\in{\mathbb{R}}_{\geq 0}italic_t ∈ blackboard_R start_POSTSUBSCRIPT ≥ 0 end_POSTSUBSCRIPT. The spatial Fourier transform maps f(θ,t)𝑓𝜃𝑡f(\theta,t)italic_f ( italic_θ , italic_t ) into its spatial Fourier coefficients

f^(k,t):=𝕊1f(θ,t)ei2πkθ𝑑θ,assign^𝑓𝑘𝑡subscriptsuperscript𝕊1𝑓𝜃𝑡superscript𝑒𝑖2𝜋𝑘𝜃differential-d𝜃\textstyle\hat{f}(k,t):=\int_{\mathbb{S}^{1}}f(\theta,t)e^{-i2\pi k\theta}d\theta,over^ start_ARG italic_f end_ARG ( italic_k , italic_t ) := ∫ start_POSTSUBSCRIPT blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_f ( italic_θ , italic_t ) italic_e start_POSTSUPERSCRIPT - italic_i 2 italic_π italic_k italic_θ end_POSTSUPERSCRIPT italic_d italic_θ , (3)

where k𝑘k\in{\mathbb{Z}}italic_k ∈ blackboard_Z is the spatial frequency.

The spatial Fourier transform is a coordinate transformation that expresses f(θ,t)𝑓𝜃𝑡f(\theta,t)italic_f ( italic_θ , italic_t ) in terms of the spatial Fourier modes ηk(θ)=ei2πkθsubscript𝜂𝑘𝜃superscript𝑒𝑖2𝜋𝑘𝜃\eta_{k}(\theta)=e^{i2\pi k\theta}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_θ ) = italic_e start_POSTSUPERSCRIPT italic_i 2 italic_π italic_k italic_θ end_POSTSUPERSCRIPT, i.e., the Fourier basis on 𝕊1superscript𝕊1\mathbb{S}^{1}blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, and the Fourier coefficients f^(k,t)^𝑓𝑘𝑡\hat{f}(k,t)over^ start_ARG italic_f end_ARG ( italic_k , italic_t ). The inverse spatial Fourier transform can be used to recover f(θ,t)𝑓𝜃𝑡f(\theta,t)italic_f ( italic_θ , italic_t ) from its Fourier coefficients f^(k,t)^𝑓𝑘𝑡\hat{f}(k,t)over^ start_ARG italic_f end_ARG ( italic_k , italic_t ) :

f(θ,t)=kf^(k,t)ei2πkθ.𝑓𝜃𝑡subscript𝑘^𝑓𝑘𝑡superscript𝑒𝑖2𝜋𝑘𝜃\textstyle f(\theta,t)=\sum_{k\in\mathbb{Z}}\hat{f}(k,t)e^{i2\pi k\theta}.italic_f ( italic_θ , italic_t ) = ∑ start_POSTSUBSCRIPT italic_k ∈ blackboard_Z end_POSTSUBSCRIPT over^ start_ARG italic_f end_ARG ( italic_k , italic_t ) italic_e start_POSTSUPERSCRIPT italic_i 2 italic_π italic_k italic_θ end_POSTSUPERSCRIPT . (4)

Parseval’s Identity [21] ensures that f^,g^=f,g^𝑓^𝑔𝑓𝑔\langle\hat{f},\hat{g}\rangle=\langle f,g\rangle⟨ over^ start_ARG italic_f end_ARG , over^ start_ARG italic_g end_ARG ⟩ = ⟨ italic_f , italic_g ⟩. The spatial Fourier transform operator is denoted by ()\mathcal{F}(\cdot)caligraphic_F ( ⋅ ), and the inverse spatial Fourier transform operator by 1()superscript1\mathcal{F}^{-1}(\cdot)caligraphic_F start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( ⋅ ).

The spatial Fourier transform (3) diagonalizes convolution operators [19], i.e., if A𝐴Aitalic_A is a convolution operator (1), then [Ah]^(k):=W^(k)h^(k)assign^delimited-[]𝐴𝑘^𝑊𝑘^𝑘\widehat{[Ah]}(k):=\hat{W}(k)\hat{h}(k)over^ start_ARG [ italic_A italic_h ] end_ARG ( italic_k ) := over^ start_ARG italic_W end_ARG ( italic_k ) over^ start_ARG italic_h end_ARG ( italic_k ), where W^^𝑊\hat{W}over^ start_ARG italic_W end_ARG is the Fourier transform of the kernal of A𝐴Aitalic_A. Thus, A𝐴Aitalic_A is mapped by \mathcal{F}caligraphic_F into a multiplication operator over the spatial frequency k𝑘k\in{\mathbb{Z}}italic_k ∈ blackboard_Z. For linear systems of the form (2), if A𝐴Aitalic_A and B𝐵Bitalic_B are convolution operators (1),

z^(k,t)t=W^A(k)z^(k,t)+W^B(k)u^(k,t),^𝑧𝑘𝑡𝑡subscript^𝑊𝐴𝑘^𝑧𝑘𝑡subscript^𝑊𝐵𝑘^𝑢𝑘𝑡\frac{\partial\hat{z}(k,t)}{\partial t}=\hat{W}_{A}(k)\hat{z}(k,t)+\hat{W}_{B}% (k)\hat{u}(k,t),divide start_ARG ∂ over^ start_ARG italic_z end_ARG ( italic_k , italic_t ) end_ARG start_ARG ∂ italic_t end_ARG = over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_k ) over^ start_ARG italic_z end_ARG ( italic_k , italic_t ) + over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_k ) over^ start_ARG italic_u end_ARG ( italic_k , italic_t ) , (5)

where W^Asubscript^𝑊𝐴\hat{W}_{A}over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT and W^Bsubscript^𝑊𝐵\hat{W}_{B}over^ start_ARG italic_W end_ARG start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT are the Fourier transforms of the kernels of A𝐴Aitalic_A and B𝐵Bitalic_B, respectively. Following [19], we refer to (5) as the diagonalization of (2).

Definition 7 (Temporal Laplace Transform):

Let z:𝕊1×0:𝑧superscript𝕊1subscriptabsent0z:\mathbb{S}^{1}\times{\mathbb{R}}_{\geq 0}italic_z : blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT × blackboard_R start_POSTSUBSCRIPT ≥ 0 end_POSTSUBSCRIPT be a spatiotemporal field with spatial and time coordinates θ𝕊1𝜃superscript𝕊1\theta\in\mathbb{S}^{1}italic_θ ∈ blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT and t0𝑡subscriptabsent0t\in{\mathbb{R}}_{\geq 0}italic_t ∈ blackboard_R start_POSTSUBSCRIPT ≥ 0 end_POSTSUBSCRIPT, respectively. Then, the temporal Laplace transform maps z(θ,t)𝑧𝜃𝑡z(\theta,t)italic_z ( italic_θ , italic_t ) into

[z](θ,s)=0z(θ,t)est𝑑t,delimited-[]𝑧𝜃𝑠superscriptsubscript0𝑧𝜃𝑡superscript𝑒𝑠𝑡differential-d𝑡[\mathcal{L}z](\theta,s)=\int_{0}^{\infty}z(\theta,t)e^{-st}dt,[ caligraphic_L italic_z ] ( italic_θ , italic_s ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_z ( italic_θ , italic_t ) italic_e start_POSTSUPERSCRIPT - italic_s italic_t end_POSTSUPERSCRIPT italic_d italic_t , (6)

where s𝑠s\in{\mathbb{C}}italic_s ∈ blackboard_C, whenever the integral exists.

III OPINION DYNAMICS ON THE CIRCLE

We propose a nonlinear opinion dynamics model for an agent making decisions about a continuous distribution of options on the circle. For every option θ𝕊1𝜃superscript𝕊1\theta\in\mathbb{S}^{1}italic_θ ∈ blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, z(θ,t)𝑧𝜃𝑡z(\theta,t)\in{\mathbb{R}}italic_z ( italic_θ , italic_t ) ∈ blackboard_R is the opinion of the agent for option θ𝜃\thetaitalic_θ at time t𝑡titalic_t, where the more positive (negative) z(θ,t)𝑧𝜃𝑡z(\theta,t)italic_z ( italic_θ , italic_t ) is the more the agent favors (disfavors) option θ𝜃\thetaitalic_θ. When z(θ,t)=0𝑧𝜃𝑡0z(\theta,t)=0italic_z ( italic_θ , italic_t ) = 0, the agent is neutral about option θ𝜃\thetaitalic_θ. The relationship between each option is encoded by the Lipschitz continuous kernel W:𝕊1:𝑊superscript𝕊1W:\mathbb{S}^{1}\to{\mathbb{R}}italic_W : blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT → blackboard_R. A positive (negative) value of W(θϕ)𝑊𝜃italic-ϕW(\theta-\phi)italic_W ( italic_θ - italic_ϕ ) corresponds to excitatory (inhibitory) interactions between the options θ𝜃\thetaitalic_θ and ϕitalic-ϕ\phiitalic_ϕ. The opinion z(θ,t)𝑧𝜃𝑡z(\theta,t)italic_z ( italic_θ , italic_t ) evolves according to

τzt(θ,t)𝜏𝑧𝑡𝜃𝑡\displaystyle\textstyle\tau\frac{{\partial}z}{\partial t}\!(\theta,t)\!italic_τ divide start_ARG ∂ italic_z end_ARG start_ARG ∂ italic_t end_ARG ( italic_θ , italic_t ) =\displaystyle== z(θ,t)+α𝕊1W(θϕ)S(z(ϕ,t))𝑑ϕ+u(θ,t)𝑧𝜃𝑡𝛼subscriptsuperscript𝕊1𝑊𝜃italic-ϕ𝑆𝑧italic-ϕ𝑡differential-ditalic-ϕ𝑢𝜃𝑡\displaystyle\textstyle\!-z(\theta,t)\!+\!\alpha\!\!\int\limits_{\mathbb{S}^{1% }}\!W\!(\theta\!-\!\phi)S(z(\phi,t))d\phi\!+\!u(\theta,t)- italic_z ( italic_θ , italic_t ) + italic_α ∫ start_POSTSUBSCRIPT blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_W ( italic_θ - italic_ϕ ) italic_S ( italic_z ( italic_ϕ , italic_t ) ) italic_d italic_ϕ + italic_u ( italic_θ , italic_t ) (7)
=\displaystyle== [Gz](θ,t)+u(θ,t),delimited-[]𝐺𝑧𝜃𝑡𝑢𝜃𝑡\displaystyle[Gz](\theta,t)+u(\theta,t),[ italic_G italic_z ] ( italic_θ , italic_t ) + italic_u ( italic_θ , italic_t ) ,

where u(θ,t)𝑢𝜃𝑡u(\theta,t)\in{\mathbb{R}}italic_u ( italic_θ , italic_t ) ∈ blackboard_R is the input and τ>0𝜏subscriptabsent0\tau\in{\mathbb{R}}_{>0}italic_τ ∈ blackboard_R start_POSTSUBSCRIPT > 0 end_POSTSUBSCRIPT is the characteristic timescale. The parameter α>0𝛼subscriptabsent0\alpha\in{\mathbb{R}}_{>0}italic_α ∈ blackboard_R start_POSTSUBSCRIPT > 0 end_POSTSUBSCRIPT is the attention paid to option interactions, i.e., α𝛼\alphaitalic_α models the agent commitment to forming strong opinions. The nonlinear nature of (7) comes from S::𝑆S:{\mathbb{R}}\to{\mathbb{R}}italic_S : blackboard_R → blackboard_R, a saturating function with S(0)=0𝑆00S(0)=0italic_S ( 0 ) = 0, S(0)=1superscript𝑆01S^{{}^{\prime}}(0)=1italic_S start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT ( 0 ) = 1.

IV SPECTRAL ANALYSIS OF LINEARIZATION

We study the spectrum of the linearization of (7) at the neutral equilibrium z(θ,t)=0𝑧𝜃𝑡0z(\theta,t)=0italic_z ( italic_θ , italic_t ) = 0, θ𝕊1for-all𝜃superscript𝕊1\forall\theta\in\mathbb{S}^{1}∀ italic_θ ∈ blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT. We first prove spatial invariance, a property that enables the linearized system to be diagonalized. Using the diagonalization, we then compute the eigenvalues and eigenfunctions of the linearized system and prove their relationship with the Fourier coefficients of the kernel and the spatial Fourier modes.

Lemma 1 (Spatial invariance of the model linearization):

Define the nonlinear operator in (7) as

[Fz](θ,t)=𝕊1W(θϕ)S(z(ϕ,t)).delimited-[]𝐹𝑧𝜃𝑡subscriptsuperscript𝕊1𝑊𝜃italic-ϕ𝑆𝑧italic-ϕ𝑡\textstyle[Fz](\theta,t)=\int_{\mathbb{S}^{1}}W(\theta-\phi)S(z(\phi,t)).[ italic_F italic_z ] ( italic_θ , italic_t ) = ∫ start_POSTSUBSCRIPT blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_W ( italic_θ - italic_ϕ ) italic_S ( italic_z ( italic_ϕ , italic_t ) ) . (8)

The differential of F𝐹Fitalic_F in the direction z𝑧zitalic_z at z(θ,t)=0𝑧𝜃𝑡0z(\theta,t)=0italic_z ( italic_θ , italic_t ) = 0 is

[AFz](θ,t)=[DzF(0)](θ,t)=𝕊1W(θϕ)z(ϕ,t)𝑑ϕ.delimited-[]subscript𝐴𝐹𝑧𝜃𝑡delimited-[]subscript𝐷𝑧𝐹0𝜃𝑡subscriptsuperscript𝕊1𝑊𝜃italic-ϕ𝑧italic-ϕ𝑡differential-ditalic-ϕ\textstyle[A_{F}z](\theta,t)\!=\![D_{z}F(0)](\theta,t)\!=\!\int_{\mathbb{S}^{1% }}W(\theta\!-\!\phi)z(\phi,t)d\phi.[ italic_A start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_z ] ( italic_θ , italic_t ) = [ italic_D start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_F ( 0 ) ] ( italic_θ , italic_t ) = ∫ start_POSTSUBSCRIPT blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_W ( italic_θ - italic_ϕ ) italic_z ( italic_ϕ , italic_t ) italic_d italic_ϕ . (9)

The linearization of (7) at the neutral equilibrium z(θ,t)=0𝑧𝜃𝑡0z(\theta,t)=0italic_z ( italic_θ , italic_t ) = 0,

zt(θ,t)=1τ(z(θ,t)+α[AFz](θ,t)+u(θ,t))=1τ([AGz](θ,t)+u(θ,t)),𝑧𝑡𝜃𝑡1𝜏𝑧𝜃𝑡𝛼delimited-[]subscript𝐴𝐹𝑧𝜃𝑡𝑢𝜃𝑡1𝜏delimited-[]subscript𝐴𝐺𝑧𝜃𝑡𝑢𝜃𝑡\begin{split}\textstyle\frac{{\partial}z}{\partial t}(\theta,t)\!&=\!% \textstyle\frac{1}{\tau}\Big{(}\!-\!z(\theta,t)+\!\alpha[A_{F}z](\theta,t)\!+% \!u(\theta,t)\!\Big{)}\!\\ &\textstyle=\!\frac{1}{\tau}([A_{G}z](\theta,t)+u(\theta,t)),\end{split}start_ROW start_CELL divide start_ARG ∂ italic_z end_ARG start_ARG ∂ italic_t end_ARG ( italic_θ , italic_t ) end_CELL start_CELL = divide start_ARG 1 end_ARG start_ARG italic_τ end_ARG ( - italic_z ( italic_θ , italic_t ) + italic_α [ italic_A start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_z ] ( italic_θ , italic_t ) + italic_u ( italic_θ , italic_t ) ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = divide start_ARG 1 end_ARG start_ARG italic_τ end_ARG ( [ italic_A start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT italic_z ] ( italic_θ , italic_t ) + italic_u ( italic_θ , italic_t ) ) , end_CELL end_ROW (10)

is a spatially invariant system in the sense of Definition 2.

  •       Proof:

    Consider the expansion S(ϵz)=n=01n!(ϵz)nS(n)(0)𝑆italic-ϵ𝑧superscriptsubscript𝑛01𝑛superscriptitalic-ϵ𝑧𝑛superscript𝑆𝑛0S(\epsilon z)=\sum_{n=0}^{\infty}\frac{1}{n!}(\epsilon z)^{n}S^{(n)}(0)italic_S ( italic_ϵ italic_z ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_n ! end_ARG ( italic_ϵ italic_z ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_S start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( 0 ). Then, we can express DzF(0)=limϵ0ϵ𝕊1W(θϕ)S(0)z(ϕ)𝑑ϕ+𝒪(ϵ2)ϵsubscript𝐷𝑧𝐹0subscriptitalic-ϵ0italic-ϵsubscriptsuperscript𝕊1𝑊𝜃italic-ϕsuperscript𝑆0𝑧italic-ϕdifferential-ditalic-ϕ𝒪superscriptitalic-ϵ2italic-ϵD_{z}F(0)=\lim_{\epsilon\to 0}\frac{\epsilon\int_{\mathbb{S}^{1}}W(\theta-\phi% )S^{\prime}(0)z(\phi)d\phi+\mathcal{O}(\epsilon^{2})}{\epsilon}italic_D start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_F ( 0 ) = roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT divide start_ARG italic_ϵ ∫ start_POSTSUBSCRIPT blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_W ( italic_θ - italic_ϕ ) italic_S start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 0 ) italic_z ( italic_ϕ ) italic_d italic_ϕ + caligraphic_O ( italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_ϵ end_ARG, where 𝒪(ϵ2)𝒪superscriptitalic-ϵ2\mathcal{O}(\epsilon^{2})caligraphic_O ( italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) denotes higher order terms in ϵitalic-ϵ\epsilonitalic_ϵ. As ϵ0italic-ϵ0\epsilon\to 0italic_ϵ → 0, the higher order terms vanish and we are left with (9). Note that AFsubscript𝐴𝐹A_{F}italic_A start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT is a spatial convolution operator, which is spatially invariant. Then, by linearity so is AGsubscript𝐴𝐺A_{G}italic_A start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT. Therefore, by Definition 2, (10) is a spatially invariant system. ∎

As a consequence of Lemma 1, we can diagonalize the model linearization (10). Since (9) is a convolution operator, we use (5) to get

z^t(k,t)=1τ(1+αW^(k))z^(k,t)+1τu^(k,t).^𝑧𝑡𝑘𝑡1𝜏1𝛼^𝑊𝑘^𝑧𝑘𝑡1𝜏^𝑢𝑘𝑡\textstyle\frac{{\partial}\hat{z}}{\partial t}(k,t)=\frac{1}{\tau}\big{(}-1+% \alpha\hat{W}(k)\big{)}\hat{z}(k,t)+\frac{1}{\tau}\hat{u}(k,t).divide start_ARG ∂ over^ start_ARG italic_z end_ARG end_ARG start_ARG ∂ italic_t end_ARG ( italic_k , italic_t ) = divide start_ARG 1 end_ARG start_ARG italic_τ end_ARG ( - 1 + italic_α over^ start_ARG italic_W end_ARG ( italic_k ) ) over^ start_ARG italic_z end_ARG ( italic_k , italic_t ) + divide start_ARG 1 end_ARG start_ARG italic_τ end_ARG over^ start_ARG italic_u end_ARG ( italic_k , italic_t ) . (11)
Lemma 2 (Eigenvalues and eigenfunctions of the linearized system):

The eigenvalues λkSp(AG)subscript𝜆𝑘Spsubscript𝐴𝐺\lambda_{k}\in\text{Sp}(A_{G})italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∈ Sp ( italic_A start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ) of the linearized system (10) can be computed as

λk=1τ(1+αW^(k)).subscript𝜆𝑘1𝜏1𝛼^𝑊𝑘\textstyle\lambda_{k}=\frac{1}{\tau}(-1+\alpha\hat{W}(k)).italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_τ end_ARG ( - 1 + italic_α over^ start_ARG italic_W end_ARG ( italic_k ) ) . (12)

for k𝑘k\in{\mathbb{Z}}italic_k ∈ blackboard_Z. The corresponding eigenfunctions are the spatial Fourier modes ηk(θ)=ei2πkθsubscript𝜂𝑘𝜃superscript𝑒𝑖2𝜋𝑘𝜃\eta_{k}(\theta)=e^{i2\pi k\theta}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_θ ) = italic_e start_POSTSUPERSCRIPT italic_i 2 italic_π italic_k italic_θ end_POSTSUPERSCRIPT.

  •       Proof:

    The form of the eigenvalues follows directly from the diagonalization (11) of the linearized dynamics (10). The eigenfunctions are the spatial Fourier modes because they form the basis of the Fourier transformation that is used to diagonalize the system. ∎

Lemma 12 reaffirms that, because of spatial invariance, the spatial Fourier modes are the eigenfunctions of the model linearization for any kernel design, provided it is spatially-invariant. Since the Fourier coefficients of the kernel determine the eigenvalues associated with each mode, they dictate which modes dominate. More precisely, if all modes are stable, i.e., (λk)<0subscript𝜆𝑘0\Re(\lambda_{k})<0roman_ℜ ( italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) < 0 for all k𝑘k\in{\mathbb{Z}}italic_k ∈ blackboard_Z, spatiotemporal inputs u(θ,t)𝑢𝜃𝑡u(\theta,t)italic_u ( italic_θ , italic_t ) will be predominantly amplified along the Fourier modes with largest (λk)subscript𝜆𝑘\Re(\lambda_{k})roman_ℜ ( italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) as detailed in Section VI.

When the leading modes becomes unstable, that is, when the real part of their eigenvalues change from negative to positive through, e.g., an increase of the attention parameter α𝛼\alphaitalic_α, the nonlinear model (7) undergoes a bifurcation that enables robust opinion formation even in the absence of inputs. The leading Fourier modes determine the number of maxima of the stable steady-state opinion patterns emerging at the bifurcation, as detailed in the next section.

V OPINION-FORMING BIFURCATIONS

We revisit the results presented in [15], [16] for (7) with zero input. We prove that (7) undergoes a bifurcation and compute the bifurcation point. A local bifurcation occurs when the number and/or stability of the equilibrium solutions changes due to one or more eigenvalues of the model linearization crossing the imaginary axis as a parameter is varied. The state and parameter value at which this occurs is the bifurcation point. We study how opinion patterns emerge and the role of kernel W𝑊Witalic_W and show a bistability that enables rapid formation of strong opinions. We make the following assumption to ensure the eigenvalues of (10) are real.

Assumption 1 (Symmetric kernels):

The kernel W𝑊Witalic_W in (7)) is symmetric, i.e. W(ψ)=W(ψ)𝑊𝜓𝑊𝜓W(\psi)=W(-\psi)italic_W ( italic_ψ ) = italic_W ( - italic_ψ ). In particular, its Fourier coefficients W^^𝑊\hat{W}over^ start_ARG italic_W end_ARG are real and W^(k)=W^(k)^𝑊𝑘^𝑊𝑘\hat{W}(k)=\hat{W}(-k)over^ start_ARG italic_W end_ARG ( italic_k ) = over^ start_ARG italic_W end_ARG ( - italic_k ).

Lemma 3 (Existence of a bifurcation point at the neutral equilibrium):

Consider (7). Let Assumption 1 hold. Let ±kmax=argmaxkW^(k)plus-or-minussubscript𝑘subscript𝑘^𝑊𝑘\pm k_{\max}=\arg\max_{k\in{\mathbb{Z}}}\hat{W}(k)± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = roman_arg roman_max start_POSTSUBSCRIPT italic_k ∈ blackboard_Z end_POSTSUBSCRIPT over^ start_ARG italic_W end_ARG ( italic_k ), kmax>0subscript𝑘𝑚𝑎𝑥0k_{max}>0italic_k start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT > 0, denote the spatial frequency corresponding to the largest W^(k)^𝑊𝑘\hat{W}(k)over^ start_ARG italic_W end_ARG ( italic_k ). Then, system (7) undergoes a bifurcation at the neutral equilibrium z(θ,t)=0𝑧𝜃𝑡0z(\theta,t)=0italic_z ( italic_θ , italic_t ) = 0 and α=1W^(kmax)superscript𝛼1^𝑊subscript𝑘\alpha^{*}=\frac{1}{\hat{W}(k_{\max})}italic_α start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG over^ start_ARG italic_W end_ARG ( italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ) end_ARG. In particular, for 0<α<α0𝛼superscript𝛼0<\alpha<\alpha^{*}0 < italic_α < italic_α start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT the neutral equilibrium is locally asymptotically stable and for α>α𝛼superscript𝛼\alpha>\alpha^{*}italic_α > italic_α start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT the neutral equilibrium is unstable. The bifurcation branches emerging at bifurcation for α=α𝛼superscript𝛼\alpha=\alpha^{*}italic_α = italic_α start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT are tangent to the subspace of L2(𝕊1)superscript𝐿2superscript𝕊1L^{2}(\mathbb{S}^{1})italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) generated by cos(2πkmaxθ)2𝜋subscript𝑘𝜃\cos(2\pi k_{\max}\theta)roman_cos ( 2 italic_π italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT italic_θ ) and sin(2πkmaxθ)2𝜋subscript𝑘𝜃\sin(2\pi k_{\max}\theta)roman_sin ( 2 italic_π italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT italic_θ ). That is, steady-state opinion patterns along the bifurcation branches have the form z(θ,t)=Acos(2πkmaxθ)+Bsin(2πkmaxθ)𝑧𝜃𝑡𝐴2𝜋subscript𝑘𝜃𝐵2𝜋subscript𝑘𝜃z(\theta,t)=A\cos(2\pi k_{\max}\theta)+B\sin(2\pi k_{\max}\theta)italic_z ( italic_θ , italic_t ) = italic_A roman_cos ( 2 italic_π italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT italic_θ ) + italic_B roman_sin ( 2 italic_π italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT italic_θ ) for A,B𝐴𝐵A,B\in{\mathbb{R}}italic_A , italic_B ∈ blackboard_R. In particular, the number of maxima exhibited in the opinion patterns forming at bifurcation is fixed by kmaxsubscript𝑘k_{\max}italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT.

  •       Proof:

    From Lemma 12, λkSp(AG)subscript𝜆𝑘Spsubscript𝐴𝐺\lambda_{k}\in\text{Sp}(A_{G})italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∈ Sp ( italic_A start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ) are given by (12). We solve for λk=1τ(1+αW^(k))=0subscript𝜆𝑘1𝜏1𝛼^𝑊𝑘0\lambda_{k}=\frac{1}{\tau}(-1+\alpha\hat{W}(k))=0italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_τ end_ARG ( - 1 + italic_α over^ start_ARG italic_W end_ARG ( italic_k ) ) = 0. Then, the first eigenvalue crossing occurs at α=1W^(kmax)=1W^(kmax)superscript𝛼1^𝑊subscript𝑘1^𝑊subscript𝑘\alpha^{*}=\frac{1}{\hat{W}(k_{\max})}=\frac{1}{\hat{W}(-k_{\max})}italic_α start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG over^ start_ARG italic_W end_ARG ( italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ) end_ARG = divide start_ARG 1 end_ARG start_ARG over^ start_ARG italic_W end_ARG ( - italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ) end_ARG with two eigenvalues crossing at 00. For α<α𝛼superscript𝛼\alpha<\alpha^{*}italic_α < italic_α start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT we have λk<0subscript𝜆𝑘0\lambda_{k}<0italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT < 0 kfor-all𝑘\forall k∀ italic_k so the origin is stable. However, once α>α𝛼superscript𝛼\alpha>\alpha^{*}italic_α > italic_α start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, there exist at least two positive eigenvalues so the origin will be unstable. For the linearization, at the bifurcation, λk<0subscript𝜆𝑘0\lambda_{k}<0italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT < 0, k±kmaxfor-all𝑘plus-or-minussubscript𝑘\forall k\neq\pm k_{\max}∀ italic_k ≠ ± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, so the corresponding spatial Fourier modes belong to the stable manifold. The spatial Fourier modes ηkmax(θ)subscript𝜂subscript𝑘𝜃\eta_{k_{\max}}(\theta)italic_η start_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_θ ) and ηkmax(θ)subscript𝜂subscript𝑘𝜃\eta_{-k_{\max}}(\theta)italic_η start_POSTSUBSCRIPT - italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_θ ) form a basis for the center manifold; hence, they determine the dominating bifurcation direction and emerging pattern of the model linearization. ∎

Refer to caption
Figure 1: Bifurcation diagrams illustrating the effect of the shift value ξ𝜉\xiitalic_ξ on the dynamics (7) with shifted sigmoid (13). Stable (unstable) branches of equilibria are shown as solid (dashed) lines.

The bifurcation of system (7) with zero input is studied in [15], [16]. There a perturbation analysis is used to show that the spatial pattern that appears is determined by the leading modes. There are infinitely many branches of non-zero equilibria which exhibit the same pattern with kmaxsubscript𝑘k_{\max}italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT maxima up to spatial translation. As discussed in [15][Theorem 4.1], [16][Section 4.2.1], the stability of the bifurcating branches can be computed as functions of W^(kmax)^𝑊subscript𝑘\hat{W}(k_{\max})over^ start_ARG italic_W end_ARG ( italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ), S′′(0)2superscript𝑆′′superscript02S^{\prime\prime}(0)^{2}italic_S start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( 0 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and S′′′(0)superscript𝑆′′′0S^{\prime\prime\prime}(0)italic_S start_POSTSUPERSCRIPT ′ ′ ′ end_POSTSUPERSCRIPT ( 0 ). Generally, when S′′(0)=0superscript𝑆′′00S^{\prime\prime}(0)=0italic_S start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( 0 ) = 0 all of the non-zero branches of equilibria that bifurcate from the origin are stable. When S′′(0)0superscript𝑆′′00S^{\prime\prime}(0)\neq 0italic_S start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( 0 ) ≠ 0, all non-zero branches bifurcating at the origin are unstable; however, due to higher order terms, stable branches of non-zero equilibria exist farther away from the origin. Fig. 1 illustrates znorm𝑧||z||| | italic_z | | at the equilibria as a function of the bifurcation parameter α𝛼\alphaitalic_α with

S(z)=tanh(zξ)tanh(ξ)sech2(ξ),𝑆𝑧𝑧𝜉𝜉superscriptsech2𝜉\textstyle S(z)=\frac{\tanh(z-\xi)-\tanh(-\xi)}{\operatorname{sech}^{2}(\xi)},italic_S ( italic_z ) = divide start_ARG roman_tanh ( italic_z - italic_ξ ) - roman_tanh ( - italic_ξ ) end_ARG start_ARG roman_sech start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ξ ) end_ARG , (13)

a shifted hyperbolic tangent with ξ0𝜉0\xi\geq 0italic_ξ ≥ 0 shift. Note that S′′(0)=0superscript𝑆′′00S^{\prime\prime}(0)=0italic_S start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( 0 ) = 0 for ξ=0𝜉0\xi=0italic_ξ = 0 and S′′(0)0superscript𝑆′′00S^{\prime\prime}(0)\neq 0italic_S start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( 0 ) ≠ 0 ξ0for-all𝜉0\forall\xi\neq 0∀ italic_ξ ≠ 0.

We see that for ξ0𝜉0\xi\neq 0italic_ξ ≠ 0, there are regions below the bifurcation point for which the neutral and a non-neutral solution are both stable. This bistable region enables rapid formation of strong opinions in response to spatially distributed input, as discussed in Section VI. The patterns of opinion formation depend on the kernel, which can be designed. Fig. 2 shows how kmaxsubscript𝑘k_{\max}italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, the spatial frequency corresponding to the largest Fourier coefficient of the kernel, determines the number of maxima exhibited in the steady-state opinion pattern of the agent for (7) with zero-input. Spatial invariance ensures that for all initial conditions the solution converges to the same opinion pattern modulo a spatial translation (Fig. 2c).

VI DISTRIBUTED DECISION-MAKING ON THE CIRCLE WITH TUNABLE SENSITIVITY

We reintroduce distributed inputs to the model, and use its linearization, together with spatial and temporal frequency analysis, to infer the nonlinear input-output behavior.

We make the following assumptions.

Assumption 2 (Shifted sigmoid):

We assume S′′(0)0superscript𝑆′′00S^{\prime\prime}(0)\neq 0italic_S start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( 0 ) ≠ 0 to ensure that a bistable region exists (see Fig. 1 for ξ0𝜉0\xi\neq 0italic_ξ ≠ 0).

Assumption 3 (Input assumptions):

Inputs u(θ,t)L2(𝕊1)𝑢𝜃𝑡superscript𝐿2superscript𝕊1u(\theta,t)\in L^{2}(\mathbb{S}^{1})italic_u ( italic_θ , italic_t ) ∈ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ) for all t0𝑡0t\geq 0italic_t ≥ 0. Furthermore, for all θ𝕊1𝜃superscript𝕊1\theta\in\mathbb{S}^{1}italic_θ ∈ blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, u(θ,):0:𝑢𝜃subscriptabsent0u(\theta,\cdot):{\mathbb{R}}_{\geq 0}\to{\mathbb{R}}italic_u ( italic_θ , ⋅ ) : blackboard_R start_POSTSUBSCRIPT ≥ 0 end_POSTSUBSCRIPT → blackboard_R is slowly varying, that is, it is Lipschitz continuous with Lispschitz constant 0<lτ10𝑙much-less-thansuperscript𝜏10<l\ll\tau^{-1}0 < italic_l ≪ italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, for τ𝜏\tauitalic_τ in (7).

The condition lτ1much-less-than𝑙superscript𝜏1l\ll\tau^{-1}italic_l ≪ italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT implies that inputs vary much more slowly than the characteristic time constant of model (7). Hence, under Assumption 3, we can use the quasi-static input approximation and let u(θ,t)uh(θ)𝑢𝜃𝑡subscript𝑢𝜃u(\theta,t)\equiv u_{h}(\theta)italic_u ( italic_θ , italic_t ) ≡ italic_u start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_θ ).

For any diagonalized spatially distributed system of the form (5), if [u^](k,s)delimited-[]^𝑢𝑘𝑠[\mathcal{L}\hat{u}](k,s)[ caligraphic_L over^ start_ARG italic_u end_ARG ] ( italic_k , italic_s ), [z^](k,s)delimited-[]^𝑧𝑘𝑠[\mathcal{L}\hat{z}](k,s)[ caligraphic_L over^ start_ARG italic_z end_ARG ] ( italic_k , italic_s ) exist, then the transfer function H(k,s)𝐻𝑘𝑠H(k,s)italic_H ( italic_k , italic_s ) characterizes the input-output response in terms of the Laplace transforms of z^(k,t)^𝑧𝑘𝑡\hat{z}(k,t)over^ start_ARG italic_z end_ARG ( italic_k , italic_t ) and u^(k,t)^𝑢𝑘𝑡\hat{u}(k,t)over^ start_ARG italic_u end_ARG ( italic_k , italic_t ), i.e.,

[z^](k,s)=H(k,s)[u^](k,s).delimited-[]^𝑧𝑘𝑠𝐻𝑘𝑠delimited-[]^𝑢𝑘𝑠[\mathcal{L}\hat{z}](k,s)=H(k,s)[\mathcal{L}\hat{u}](k,s).[ caligraphic_L over^ start_ARG italic_z end_ARG ] ( italic_k , italic_s ) = italic_H ( italic_k , italic_s ) [ caligraphic_L over^ start_ARG italic_u end_ARG ] ( italic_k , italic_s ) .

By the Final Value Theorem, if the input is constant in time and  (5) is stable, then limtz^(k,t)=sH(k,0)(u^h(k)s)=H(k,0)u^h(k)subscript𝑡^𝑧𝑘𝑡𝑠𝐻𝑘0subscript^𝑢𝑘𝑠𝐻𝑘0subscript^𝑢𝑘\lim_{t\to\infty}\hat{z}(k,t)\!=\!sH(k,0)\Big{(}\frac{\hat{u}_{h}(k)}{s}\Big{)% }\!=\!H(k,0)\hat{u}_{h}(k)roman_lim start_POSTSUBSCRIPT italic_t → ∞ end_POSTSUBSCRIPT over^ start_ARG italic_z end_ARG ( italic_k , italic_t ) = italic_s italic_H ( italic_k , 0 ) ( divide start_ARG over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_k ) end_ARG start_ARG italic_s end_ARG ) = italic_H ( italic_k , 0 ) over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_k ). This leads us to the following definition.

Refer to caption
Figure 2: Influence of the kernel design on the steady-state opinion patterns of (7) with zero-input. (a) Two kernel designs. (b) Fourier coefficients of the kernel. Top: ±kmax=±1plus-or-minussubscript𝑘plus-or-minus1\pm k_{\max}=\pm 1± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = ± 1. Bottom: ±kmax=±3plus-or-minussubscript𝑘plus-or-minus3\pm k_{\max}=\pm 3± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = ± 3. (c) Steady-state opinion pattern z(θ,)𝑧𝜃z(\theta,\infty)italic_z ( italic_θ , ∞ ), of dynamics (7) for initial conditions z(θ,0)𝑧𝜃0z(\theta,0)italic_z ( italic_θ , 0 ). The number of maxima of z(θ,)𝑧𝜃z(\theta,\infty)italic_z ( italic_θ , ∞ ) equals kmaxsubscript𝑘k_{\max}italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT of the corresponding kernel. Parameters: d=1𝑑1d=1italic_d = 1, α=0.98𝛼0.98\alpha=0.98italic_α = 0.98, p=3𝑝3p=3italic_p = 3, ξ=0.7𝜉0.7\xi=0.7italic_ξ = 0.7.
Definition 8 (Spatial transfer function):

The spatial transfer function of (5) is

H~(k)=H(k,0).~𝐻𝑘𝐻𝑘0\textstyle\tilde{H}(k)=H(k,0).over~ start_ARG italic_H end_ARG ( italic_k ) = italic_H ( italic_k , 0 ) . (14)

Spatial transfer function H~(k)~𝐻𝑘\tilde{H}(k)over~ start_ARG italic_H end_ARG ( italic_k ) determines the steady-state output of (5) in response to input that is constant in time.

Theorem 4 (Spatial transfer function of (10)):

Let Assumptions 13 hold. Let ±kmax=argmaxkW^(k)plus-or-minussubscript𝑘subscript𝑘^𝑊𝑘\pm k_{\max}=\arg\max_{k\in{\mathbb{Z}}}\hat{W}(k)± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = roman_arg roman_max start_POSTSUBSCRIPT italic_k ∈ blackboard_Z end_POSTSUBSCRIPT over^ start_ARG italic_W end_ARG ( italic_k ) denote the spatial frequencies corresponding to the largest W^(k)^𝑊𝑘\hat{W}(k)over^ start_ARG italic_W end_ARG ( italic_k ). Spatial transfer function (14) of the linearized model (10) is

H~(k)=τ1αW^(k),k.formulae-sequence~𝐻𝑘𝜏1𝛼^𝑊𝑘𝑘\textstyle\tilde{H}(k)=\frac{\tau}{1-{\alpha}\hat{W}(k)},\quad k\in{\mathbb{Z}}.over~ start_ARG italic_H end_ARG ( italic_k ) = divide start_ARG italic_τ end_ARG start_ARG 1 - italic_α over^ start_ARG italic_W end_ARG ( italic_k ) end_ARG , italic_k ∈ blackboard_Z . (15)

In particular, for αα𝛼superscript𝛼{\alpha}\nearrow\alpha^{*}italic_α ↗ italic_α start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, H~(±kmax)~𝐻plus-or-minussubscript𝑘\tilde{H}(\pm k_{\max})\to\inftyover~ start_ARG italic_H end_ARG ( ± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ) → ∞.

  •       Proof:

    From Lemma 12 we know the eigenvalues of (11), the diagonalization of (10). So we can compute H(k,s)=1sλk𝐻𝑘𝑠1𝑠subscript𝜆𝑘{H}(k,s)=\frac{1}{s-\lambda_{k}}italic_H ( italic_k , italic_s ) = divide start_ARG 1 end_ARG start_ARG italic_s - italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG. Then H~(k)=τ1αW^(k).~𝐻𝑘𝜏1𝛼^𝑊𝑘\tilde{H}(k)=\frac{\tau}{1-{\alpha}\hat{W}(k)}.over~ start_ARG italic_H end_ARG ( italic_k ) = divide start_ARG italic_τ end_ARG start_ARG 1 - italic_α over^ start_ARG italic_W end_ARG ( italic_k ) end_ARG . As αα𝛼superscript𝛼{\alpha}\nearrow\alpha^{*}italic_α ↗ italic_α start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, H(±kmax,0)𝐻plus-or-minussubscript𝑘0H(\pm k_{\max},0)\to\inftyitalic_H ( ± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT , 0 ) → ∞. ∎

Theorem 15 implies that close to bifurcation, i.e., for αα𝛼superscript𝛼{\alpha}\nearrow\alpha^{*}italic_α ↗ italic_α start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, the input-output response of the linearized system is dominated by the leading spatial Fourier modes ηk±max(θ)subscript𝜂subscript𝑘plus-or-minus𝜃\eta_{k_{\pm\max}}(\theta)italic_η start_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT ± roman_max end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_θ ). This means that the alignment of u(θ)𝑢𝜃u(\theta)italic_u ( italic_θ ) with η±kmax(θ)subscript𝜂plus-or-minussubscript𝑘𝜃\eta_{\pm k_{\max}}(\theta)italic_η start_POSTSUBSCRIPT ± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_θ ) is the main determinant of the model response to inputs.

For the nonlinear system with αα𝛼superscript𝛼{\alpha}\nearrow\alpha^{*}italic_α ↗ italic_α start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, the input-output response takes place in the bistable region. If η±kmax(θ),uh(θ)=u^h(kmax)0subscript𝜂plus-or-minussubscript𝑘𝜃subscript𝑢𝜃subscript^𝑢subscript𝑘0\langle\eta_{\pm k_{\max}}(\theta),{u_{h}(\theta)}\rangle=\hat{u}_{h}(k_{\max}% )\neq 0⟨ italic_η start_POSTSUBSCRIPT ± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_θ ) , italic_u start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_θ ) ⟩ = over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ) ≠ 0, the input is aligned with the leading modes η±kmax(θ)subscript𝜂plus-or-minussubscript𝑘𝜃\eta_{\pm k_{\max}}(\theta)italic_η start_POSTSUBSCRIPT ± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_θ ). These modes filter input nonlinearity and amplify the input because, by Theorem 15, the direction of these modes are ultrasensitive to input. The result is a steady-state opinion pattern with kmaxsubscript𝑘k_{\max}italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT maxima and z(θ)uh(θ)much-greater-thannorm𝑧𝜃normsubscript𝑢𝜃||z(\theta)||\gg||u_{h}(\theta)||| | italic_z ( italic_θ ) | | ≫ | | italic_u start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_θ ) | |, as illustrated in the top row of Fig. 3. As shown in Fig. 3a (top), the Fourier coefficients of the input for ±kmax=±1plus-or-minussubscript𝑘plus-or-minus1\pm k_{\max}=\pm 1± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = ± 1 are nonzero meaning the input is aligned with η±kmax(θ)=η±1(θ)subscript𝜂plus-or-minussubscript𝑘𝜃subscript𝜂plus-or-minus1𝜃\eta_{\pm k_{\max}}(\theta)=\eta_{\pm 1}(\theta)italic_η start_POSTSUBSCRIPT ± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_θ ) = italic_η start_POSTSUBSCRIPT ± 1 end_POSTSUBSCRIPT ( italic_θ ). The input distribution in Fig. 3b (top) is small in magnitude (less than 0.01), while the resulting steady-state opinion pattern in Fig. 3c (top) has kmax=1subscript𝑘1k_{\max}=1italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 1 maximum that is greater than 1.5 in magnitude.

If η±kmax(θ),uh(θ)=0subscript𝜂plus-or-minussubscript𝑘𝜃subscript𝑢𝜃0\langle\eta_{\pm k_{\max}(\theta)},{u}_{h}(\theta)\rangle=0⟨ italic_η start_POSTSUBSCRIPT ± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ( italic_θ ) end_POSTSUBSCRIPT , italic_u start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_θ ) ⟩ = 0, the input is unaligned with the leading modes η±kmax(θ)subscript𝜂plus-or-minussubscript𝑘𝜃\eta_{\pm k_{\max}}(\theta)italic_η start_POSTSUBSCRIPT ± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_θ ) and the steady-state opinion pattern will not exhibit large maxima. I.e., because the input does not have a component along the ultrasensitive direction, by Theorem 15 it does not get amplified, as illustrated in the bottom row of Fig. 3. As shown in Fig. 3a (bottom), the Fourier coefficients of the input for ±kmax=±1plus-or-minussubscript𝑘plus-or-minus1\pm k_{\max}=\pm 1± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = ± 1 are zero meaning the input is unaligned with η±kmax(θ)=η±1(θ)subscript𝜂plus-or-minussubscript𝑘𝜃subscript𝜂plus-or-minus1𝜃\eta_{\pm k_{\max}}(\theta)=\eta_{\pm 1}(\theta)italic_η start_POSTSUBSCRIPT ± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_θ ) = italic_η start_POSTSUBSCRIPT ± 1 end_POSTSUBSCRIPT ( italic_θ ). The input distribution in Fig. 3b (bottom) is small in magnitude (less than 0.01) and the resulting steady-state opinion pattern in Fig. 3c (bottom) has no large maximum.

Our results show how even very small distributed inputs can trigger the formation of a strong opinion and whether or not this happens depends on the design of kernel W𝑊Witalic_W. Thus, W𝑊Witalic_W can be designed to tune response to be ultrasensitive to inputs that matter for function and robust to inputs that don’t.

Refer to caption
Figure 3: Input-output behavior of the dynamics (7) with input distributions aligned or unaligned with the Fourier mode corresponding to ±kmax=±1plus-or-minussubscript𝑘plus-or-minus1\pm k_{\max}\!=\!\pm 1± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = ± 1. Top row: Aligned. Bottom row: Unaligned. (a) Magnitude of the Fourier coefficients of the input. (b) Input distribution. (c) Steady-state opinion pattern z(0,)𝑧0z(0,\infty)italic_z ( 0 , ∞ ). Parameters: d=1𝑑1d\!=\!1italic_d = 1, α=0.98𝛼0.98\alpha\!=\!0.98italic_α = 0.98, p=3𝑝3p\!=\!3italic_p = 3, ξ=0.7𝜉0.7\xi\!=\!0.7italic_ξ = 0.7.

VII APPLICATION TO ROBOT NAVIGATION

We illustrate the benefits of our approach to perceptual decision-making with an application of the dynamics (7) to robot spatial navigation. We consider the case of a robot moving in a crowded space, such as an airport, where it must pass through gaps of different sizes (e.g., between people in a line) that may change over time. We assume the robot has a (visual) sensor so that it can perceive these gaps.

We specialize to a scenario where a robot finds itself trapped inside a circle of people and needs to choose and cross through a large enough gap between people. Choosing a gap is challenging as people may be distributed unevenly around the circle, resulting in multiple gap options, only a select few of which may be suitable for the robot to cross. Also, the size of the gaps may change over time due to people moving for their own purposes or in response to the robot, e.g., people may move to make space for the robot to cross.

In Section VII-A we present a framework for designing W𝑊Witalic_W from its Fourier coefficients to allow the robot to select a single gap. We discuss four scenarios that demonstrate how our model can be used for fast-and-flexible decision-making in this robotic navigation problem. In Section VII-B two scenarios demonstrate the robot’s ability to choose a single gap, while in Section VII-C the two other scenarios show how the robot can quickly adapt to changes in gap sizes.

We take 𝕊1superscript𝕊1\mathbb{S}^{1}blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT to represent the circular visual field for the robot. Then an option θ𝕊1𝜃superscript𝕊1\theta\in\mathbb{S}^{1}italic_θ ∈ blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT represents the angle associated to a point in the visual field. The input u(θ,t)𝑢𝜃𝑡u(\theta,t)italic_u ( italic_θ , italic_t ) is the visual observation (e.g., pixel) at angle θ𝜃\thetaitalic_θ at time t𝑡titalic_t. We let a point in the input distribution that reflects a gap be represented by u(θ,t)>0𝑢𝜃𝑡0u(\theta,t)>0italic_u ( italic_θ , italic_t ) > 0, in blue in Fig. 4 (bottom). We assume changes in gaps occur slowly enough that Assumption 3 holds. The opinion z(θ,t)𝑧𝜃𝑡z(\theta,t)italic_z ( italic_θ , italic_t ) as shown in Fig. 4 (top), captures the robot’s preference over time for one gap, where the preference corresponds to the strongest opinion (in yellow).

VII-A Fourier-Based Kernel Design

We leverage the results of Theorem 15 to design kernel W𝑊Witalic_W on 𝕊1superscript𝕊1\mathbb{S}^{1}blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT that imposes the desired opinion formation behavior in response to distributed input on 𝕊1superscript𝕊1\mathbb{S}^{1}blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT. Options (angles) that are close (far) to each other should have an excitatory (inhibitory) interaction. And the opinion pattern should have a single maximum, so that the robot selects a single gap. From the results summarized in Section VI, to achieve this we design W𝑊Witalic_W as follows and ensure that ±kmax=±1plus-or-minussubscript𝑘plus-or-minus1\pm k_{\max}=\pm 1± italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = ± 1:

W^(k)={0k=0,f^(k)k0^𝑊𝑘cases0𝑘0^𝑓𝑘for-all𝑘0\hat{W}(k)=\begin{cases}0\qquad&k=0,\\ \hat{f}(k)\qquad&\forall k\neq 0\end{cases}over^ start_ARG italic_W end_ARG ( italic_k ) = { start_ROW start_CELL 0 end_CELL start_CELL italic_k = 0 , end_CELL end_ROW start_ROW start_CELL over^ start_ARG italic_f end_ARG ( italic_k ) end_CELL start_CELL ∀ italic_k ≠ 0 end_CELL end_ROW (16)

where f^(k)::^𝑓𝑘\hat{f}(k):{\mathbb{Z}}\to{\mathbb{R}}over^ start_ARG italic_f end_ARG ( italic_k ) : blackboard_Z → blackboard_R is strictly decreasing, square-summable and symmetric. The strictly decreasing property ensures that kmax=±1subscript𝑘plus-or-minus1k_{\max}=\pm 1italic_k start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = ± 1 while the square-summability property ensures that the inverse Fourier transform of (16) exists and that W(θϕ)L2(𝕊1)𝑊𝜃italic-ϕsubscript𝐿2superscript𝕊1W(\theta-\phi)\in L_{2}(\mathbb{S}^{1})italic_W ( italic_θ - italic_ϕ ) ∈ italic_L start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ), by Parseval’s identity [21]. Symmetry is required to satisfy Assumption 1. We take W^(0)=0^𝑊00\hat{W}(0)=0over^ start_ARG italic_W end_ARG ( 0 ) = 0 so that the homogeneous state is never a solution. Examples of functions that meet this criteria include fk^=1/kp^subscript𝑓𝑘1superscript𝑘𝑝\hat{f_{k}}=1/k^{p}over^ start_ARG italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG = 1 / italic_k start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT with p>0.5𝑝0.5p>0.5italic_p > 0.5 and a Gaussian function of the form f^(k)=e(k1)2/p2^𝑓𝑘superscript𝑒superscript𝑘12superscript𝑝2\hat{f}(k)=e^{-(k-1)^{2}/p^{2}}over^ start_ARG italic_f end_ARG ( italic_k ) = italic_e start_POSTSUPERSCRIPT - ( italic_k - 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_p start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT where p𝑝pitalic_p adjusts the width of the Gaussian. For the following simulations, we take f^(k)=e(k1)2/p2^𝑓𝑘superscript𝑒superscript𝑘12superscript𝑝2\hat{f}(k)=e^{-(k-1)^{2}/p^{2}}over^ start_ARG italic_f end_ARG ( italic_k ) = italic_e start_POSTSUPERSCRIPT - ( italic_k - 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_p start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT.

VII-B Choosing the Best Gap and Avoiding Deadlock

We illustrate the model’s ability to pick the best among multiple gaps and to rapidly avoid deadlock when faced with two equally suitable gaps. We assume the people in the circle are not moving. In Fig. 4a (bottom) there are several gaps but one that is clearly wider than the others. In Fig. 4a (top), the robot forms a strong preference for the widest gap. In Fig. 4b (bottom), there are two equally wide gaps. In Fig. 4b (top), the robot forms strong opinions for one of the two widest gaps and avoids deadlock.

Refer to caption
Figure 4: Decision-making of a robot selecting a gap through which to cross through a circle of non-moving people. Bottom row: gap distribution over time where gaps are indicated by u(θ,t)>0𝑢𝜃𝑡0u(\theta,t)>0italic_u ( italic_θ , italic_t ) > 0 in blue. Top row: opinion pattern over time (strongest opinion in yellow). (a) One widest gap. (b) Two wide gaps of same size. Parameters: d=1𝑑1d=1italic_d = 1, u=0.96𝑢0.96u=0.96italic_u = 0.96, ξ=0.35𝜉0.35\xi=0.35italic_ξ = 0.35, p=3𝑝3p=3italic_p = 3.

VII-C Robustness and Responsiveness to Change

We illustrate the model’s robustness to unimportant change and responsiveness to important change in input. We assume the people in the circle are moving. In Fig. 5 (bottom), there is initially one very wide gap and one narrow gap. However, over time, the wide gap becomes narrower, and the narrower gap becomes wider. In Fig. 5a, the decrease in size of the initially wide gap is small enough that the robot can still fit through it and thus it does not change its choice. In Fig. 5b, the decrease in size of the initially wide gap is large enough that the robot changes its choice to the other emerging gap.

VIII DISCUSSION AND FINAL REMARKS

We presented a nonlinear opinion dynamics model for an agent making decisions about a continuous distribution of options in response to distributed input on the circle. We proved spatial invariance of the model linearization and a bifurcation of the model with zero input, which yields fast and flexible decision-making. We studied the input-output behavior of the model and design of the kernel. We demonstrated key advantages of the model in robot perceptual decision-making problem. In future work we aim to generalize the analysis to a continuous distribution of agents and extend the analysis to manifolds other than 𝕊1superscript𝕊1\mathbb{S}^{1}blackboard_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, e.g. Nsuperscript𝑁{\mathbb{R}}^{N}blackboard_R start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT. We also aim to implement this model for perceptual decision-making in robotics where the dynamics are in a closed loop with the physical dynamics of the agent.

ACKNOWLEDGMENTS

The authors thank Prof. Anastasia Bizyaeva (Cornell University) for important input and ideas provided during the initial phase of this project. They thank Dr. Juncal Arbelaiz (Princeton University) for insights and references on spatially-invariant systems.

Refer to caption
Figure 5: Decision-making of an agent selecting a gap through which to cross through a circle of moving people. Bottom row: gap distribution over time where gaps are indicated by u(θ,t)>0𝑢𝜃𝑡0u(\theta,t)>0italic_u ( italic_θ , italic_t ) > 0 in blue. Top row: opinion about where to cross the line over time (strongest opinion in yellow). (a) Small decrease over time in size of initially widest gap. (b) Large decrease over time in size of initially widest gap. Parameters: d=1𝑑1d=1italic_d = 1, u=0.96𝑢0.96u=0.96italic_u = 0.96, ξ=0.35𝜉0.35\xi=0.35italic_ξ = 0.35, p=3𝑝3p=3italic_p = 3.

References

  • [1] P. C. Bressloff, J. D. Cowan, M. Golubitsky, P. J. Thomas, and M. C. Wiener, “What geometric visual hallucinations tell us about the visual cortex,” Neural Comput., vol. 14, no. 3, pp. 473–491, 2002.
  • [2] M. Oubbati, M. Schanz, and P. Levi, “Neural fields for behavior-based control of mobile robots,” IFAC Proc. Vol., vol. 39, no. 15, pp. 61–66, 2006.
  • [3] W. Erlhagen and E. Bicho, “The dynamic neural field approach to cognitive robotics,” J. Neural Eng., vol. 3, no. 3, p. R36, jun 2006.
  • [4] P. Dahm, C. Bruckhoff, and F. Joublin, “A neural field approach for robot motion control,” in SMC’98 Conf. Proc. 1998 IEEE Int. Conf. Syst., Man, Cybern., vol. 4, 1998, pp. 3460–3465.
  • [5] G. Schöner, M. Dose, and C. Engels, “Dynamics of behavior: Theory and applications for autonomous robot architectures,” Robot. Auton. Syst., vol. 16, no. 2, pp. 213–245, 1995.
  • [6] A. Bizyaeva, A. Franci, and N. E. Leonard, “Nonlinear opinion dynamics with tunable sensitivity,” IEEE Trans. Autom. Control, vol. 68, no. 3, pp. 1415–1430, 2023.
  • [7] N. E. Leonard, A. Bizyaeva, and A. Franci, “Fast and flexible multiagent decision-making,” Annu. Rev. Control Robot. Auton. Syst., vol. 7, no. 1, 2024.
  • [8] C. Cathcart, M. Santos, S. Park, and N. E. Leonard, “Proactive opinion-driven robot navigation around human movers,” in IEEE/RSJ Int. Conf. Intell. Robots Syst. (IROS), 2023, pp. 4052–4058.
  • [9] A. Bizyaeva, G. Amorim, M. Santos, A. Franci, and N. E. Leonard, “Switching transformations for decentralized control of opinion patterns in signed networks: Application to dynamic task allocation,” IEEE Control Syst. Lett., vol. 6, pp. 3463–3468, 2022.
  • [10] G. Amorim, M. Santos, S. Park, A. Franci, and N. E. Leonard, “Threshold decision-making dynamics adaptive to physical constraints and changing environment,” in 2024 Eur. Control Conf. (ECC), 2024, pp. 1908–1913.
  • [11] S. Park, A. Bizyaeva, M. Kawakatsu, A. Franci, and N. E. Leonard, “Tuning cooperative behavior in games with nonlinear opinion dynamics,” IEEE Control Syst. Lett., vol. 6, pp. 2030–2035, 2022.
  • [12] H. Hu, K. Nakamura, K.-C. Hsu, N. E. Leonard, and J. F. Fisac, “Emergent coordination through game-induced nonlinear opinion dynamics,” in IEEE Conf. Decis. Control (CDC), 2023, pp. 8122–8129.
  • [13] H. Hu, J. DeCastro, D. Gopinath, G. Rosman, N. E. Leonard, and J. F. Fisac, “Think deep and fast: Learning neural nonlinear opinion dynamics from inverse dynamic games for split-second interactions,” 2024, arXiv: 2406.09810.
  • [14] S.-I. Amari, “Dynamics of pattern formation in lateral-inhibition type neural fields,” Biol. Cybern., vol. 27, no. 2, pp. 77–87, 1977.
  • [15] R. Curtu and B. Ermentrout, “Pattern formation in a network of excitatory and inhibitory cells with adaptation,” SIAM J. App. Dyn. Syst., vol. 3, no. 3, pp. 191–231, 2004.
  • [16] G. B. Ermentrout, S. E. Folias, and Z. P. Kilpatrick, “Spatiotemporal pattern formation in neural fields with linear adaptation,” in Neural Fields: Theory and Applications.   Springer Berlin Heidelberg, 2014, pp. 119–151.
  • [17] S. E. Folias and P. C. Bressloff, “Breathing pulses in an excitatory neural network,” SIAM J. App. Dyn. Syst., vol. 3, no. 3, pp. 378–407, 2004.
  • [18] S. E. Folias, “Nonlinear analysis of breathing pulses in a synaptically coupled neural network,” SIAM J. App. Dyn. Syst., vol. 10, no. 2, pp. 744–787, 2011.
  • [19] B. Bamieh, F. Paganini, and M. Dahleh, “Distributed control of spatially invariant systems,” IEEE Trans. Autom. Control, vol. 47, no. 7, pp. 1091–1107, 2002.
  • [20] J. Arbelaiz, B. Bamieh, A. E. Hosoi, and A. Jadbabaie, “Optimal estimation in spatially distributed systems: how far to share measurements from?” 2024, arXiv: 2406.14781.
  • [21] E. Titchmarsh, The Theory of Functions.   Oxford Univ. Press, 1939.