Fact-checked by Grok 2 weeks ago

Turing pattern

A Turing pattern is a self-organizing spatial structure that emerges in reaction-diffusion systems, where interacting chemical substances known as morphogens react and diffuse at different rates, leading to diffusion-driven instability and the spontaneous formation of periodic patterns such as spots, stripes, or waves from an initially homogeneous state. This phenomenon was first mathematically described by British mathematician in his seminal 1952 paper, "," which proposed that such mechanisms could explain biological during embryonic development without requiring pre-patterned structures. The core mechanism of Turing patterns relies on the interplay between local activation and long-range inhibition in a - framework. In these systems, an activator promotes its own production locally (self-enhancement) while an inhibitor diffuses farther and suppresses the activator's activity over larger distances, destabilizing uniform concentrations and amplifying small perturbations into stable patterns. This instability condition requires the inhibitor to diffuse faster than the activator, typically by a factor of at least 10, and is mathematically captured by partial differential equations describing concentrations over space and time. Early refinements, such as the Gierer-Meinhardt model, formalized this activator-inhibitor dynamic, demonstrating how it generates diverse morphologies like labyrinthine or spotted patterns depending on parameters like diffusion coefficients and . Turing patterns have been observed and experimentally validated across biological contexts, illustrating their role in morphogenesis and self-organization. In developmental biology, they contribute to pigment patterns on animal coats, such as the stripes of zebras or spots on leopards, where genetic factors regulate morphogen interactions to produce species-specific designs. They also underlie shell pigmentation in mollusks, like the spiral bands on cone snails, and wing scale arrangements in butterflies, where reaction-diffusion simulations match observed variations. Beyond multicellular organisms, Turing-like mechanisms appear in bacterial systems, such as Min protein oscillations in Escherichia coli that position the division site, and in embryonic processes like sea urchin skeleton formation via Nodal signaling. These patterns extend to non-biological applications, including chemical oscillators and material science for designing textured surfaces, underscoring Turing's enduring influence on understanding emergent complexity.

Introduction

Definition and Overview

Turing patterns are self-organizing spatial formations that emerge in chemical or biological systems through the interplay of and processes, resulting in regular structures such as spots, stripes, or . These patterns arise spontaneously from initially uniform states, driven by instabilities that amplify small perturbations into stable, periodic configurations without requiring external templates or pre-patterns. The concept was introduced by British mathematician in his seminal 1952 paper, "," where he proposed that such mechanisms could underlie the development of biological forms during embryogenesis. Central to the formation of Turing patterns is the dynamic interaction between morphogens—diffusible chemical substances—acting as activators that promote their own production and s that suppress it, with the typically diffusing more rapidly than the activator. This differential allows local activations to build up while broader inhibition prevents overgrowth, propagating simple local rules into complex global order across the system. Turing's framework highlights how , often seen as a homogenizing force, can instead foster heterogeneity when coupled with nonlinear reactions. The significance of Turing patterns lies in their ability to explain diverse natural phenomena, such as the pigmentation markings on animal coats—including stripes on zebras and spots on leopards—and the branching venation patterns in leaves, all emerging from decentralized chemical signaling rather than genetic blueprints alone. By providing a mathematical basis for in , these patterns bridge chemistry and biology, influencing fields from to and even .

Historical Development

The concept of Turing patterns originated with Alan Turing's seminal 1952 paper, "," published in Philosophical Transactions of the Royal Society, where he proposed that -driven instability in reaction- systems could generate spatial patterns from initially homogeneous states, providing a chemical mechanism for biological . Turing's mathematical framework demonstrated how interacting chemical substances, or morphogens, with differing rates could lead to spontaneous , such as spots or stripes, through analysis of reaction- equations. Despite its theoretical elegance, Turing's work encountered significant skepticism from the biological and chemical communities during the and , primarily due to its abstract mathematical density, which made it inaccessible to many experimentalists, and the absence of direct empirical validation. Early observations of oscillating reactions, such as the Belousov-Zhabotinsky (BZ) reaction discovered in the , were often dismissed as artifacts or violations of thermodynamic principles, further delaying acceptance. Experimental biologists largely overlooked the model, favoring gene-centric explanations for development, while chemists focused on equilibrium systems rather than nonequilibrium dynamics. The revival of interest in Turing patterns began in the 1970s with theoretical advancements, notably the 1972 activator-inhibitor model by Hans Meinhardt and Alfred Gierer, which provided a nonlinear approximation of Turing's ideas by incorporating short-range activation and long-range inhibition to explain biological patterning. This was bolstered in the 1980s and 1990s by computer simulations that visualized pattern emergence and experimental observations in chemical systems, including the reaction, where spatiotemporal oscillations and waves aligned with reaction-diffusion principles. A landmark confirmation came in 1990 with sustained stationary Turing patterns observed in the chlorite-iodide-malonic acid (CIMA) reaction, demonstrating diffusion-driven instability in a controlled nonequilibrium setting. Modern experimental validations in , such as studies on patterning in mice involving Wnt signaling, have further solidified Turing's legacy by showing how reaction-diffusion mechanisms govern periodic structures. Recent advances as of 2024 include the implementation of synthetic three-node Turing gene circuits in mammalian cells, demonstrating controlled in biological systems. These developments have established Turing patterns as a foundational bridge between , , and , inspiring interdisciplinary research into across scales from molecular networks to ecological systems.

Mathematical Foundations

Reaction-Diffusion Systems

Reaction-diffusion systems provide the mathematical foundation for understanding Turing patterns, modeling the spatiotemporal evolution of interacting chemical or biological species through coupled partial differential equations that incorporate both kinetics and processes. These systems describe how concentrations of substances, known as morphogens, change over time and space due to local chemical reactions and spatial transport via . In the seminal work by , such systems are formulated for at least two interacting morphogens whose nonlinear interactions and differential diffusivities can lead to the emergence of stable spatial patterns from an initially homogeneous state. The general form of a two-species reaction-diffusion system is given by \begin{align} \frac{\partial u}{\partial t} &= D_u \nabla^2 u + f(u,v), \\ \frac{\partial v}{\partial t} &= D_v \nabla^2 v + g(u,v), \end{align} where u(\mathbf{x},t) and v(\mathbf{x},t) represent the concentrations of the two morphogens at position \mathbf{x} and time t, D_u and D_v are the respective diffusion coefficients, \nabla^2 is the Laplacian operator accounting for spatial diffusion, and f(u,v) and g(u,v) are nonlinear reaction terms describing the local production and degradation rates of each species. This formulation assumes a continuous medium, such as a tissue or chemical solution, where the morphogens interact via autocatalytic or cross-catalytic reactions, with the reaction terms satisfying certain conditions for bounded growth. A key assumption is that the two species have distinct diffusion rates, typically with the inhibitor diffusing faster than the activator (D_v > D_u), which allows short-range activation and long-range inhibition to drive pattern formation. This differential diffusivity is central to the model's ability to generate spatial heterogeneity, as explored in early extensions like the activator-inhibitor framework. To analyze the potential for , analysis is applied to the system around a homogeneous , where concentrations are uniform and constant, satisfying f(h,k) = 0 and g(h,k) = 0 for steady-state values u = h and v = k. Small perturbations to this state are decomposed into modes, expressed as spatial waves with wave numbers corresponding to different length scales, such as u(\mathbf{x},t) = h + \sum \hat{u}_k e^{\lambda_k t + i \mathbf{k} \cdot \mathbf{x}} and similarly for v, where \lambda_k determines the growth rate of mode k. The eigenvalues \lambda_k are derived from the linearized reaction- , revealing how modifies the of the homogeneous state by introducing spatial dependence. Diffusion plays a crucial role in these systems by enabling the amplification of spatial inhomogeneities arising from inherent noise or random fluctuations in the initial conditions, transforming a stable uniform equilibrium into an unstable one for specific wavenumbers. Without diffusion, the homogeneous state remains stable under the kinetics alone; however, the interplay between and allows certain modes to grow exponentially, selecting preferred spatial periodicities that manifest as Turing patterns. This mechanism underpins the Turing instability, where -driven instability leads to .

Turing Instability Condition

The Turing instability arises in reaction-diffusion systems when a spatially homogeneous steady state, stable in the absence of diffusion, becomes unstable to spatial perturbations of certain wavelengths due to the interplay of reaction kinetics and differential diffusion rates. This mechanism, first proposed by Alan Turing, enables the emergence of periodic patterns from uniformity. To analyze this, consider small perturbations around the steady state in a two-component reaction-diffusion system, taking the form \delta u = \epsilon e^{\lambda t + i k x}, \delta v = \eta e^{\lambda t + i k x}, where \lambda is the growth rate and k is the wavenumber. The linearized system yields the characteristic equation for the growth rate: \lambda^2 - \operatorname{tr}(k) \lambda + \det(k) = 0, where \operatorname{tr}(k) = f_u + g_v - (D_u + D_v) k^2 and \det(k) = (f_u - D_u k^2)(g_v - D_v k^2) - f_v g_u, with f_u, f_v, g_u, g_v being the elements of the Jacobian matrix of the reaction terms at the steady state. The roots are \lambda = \frac{\operatorname{tr}(k) \pm \sqrt{\operatorname{tr}(k)^2 - 4 \det(k)}}{2}, and instability occurs if the real part of the dominant root is positive (\operatorname{Re}(\lambda) > 0) for some k \neq 0. This highlights how diffusion modifies the reaction-driven dynamics. For Turing instability to manifest, three key conditions must hold. First, the steady state must be stable without diffusion: the trace of the Jacobian \operatorname{tr}(J) = f_u + g_v < 0 and the determinant \det(J) = f_u g_v - f_v g_u > 0. Second, the reaction kinetics must exhibit activator-inhibitor characteristics: f_u > 0 (self-activation of the activator), g_v < 0 (self-inhibition of the inhibitor), and f_u g_v > f_v g_u (ensuring positive determinant with cross-inhibition). Third, there must be a disparity in diffusion rates, typically D_v > D_u, such that the inhibitor diffuses faster than the activator, allowing local activation to outpace global inhibition and destabilize the uniform state for finite k. These conditions ensure that \det(k) < 0 (while \operatorname{tr}(k) < 0) over a band of wavenumbers, leading to \operatorname{Re}(\lambda) > 0. The selected pattern wavelength emerges from the critical wavenumber k_c where \operatorname{Re}(\lambda(k)) achieves its maximum value, as this mode grows fastest during the initial linear phase. The corresponding wavelength is \lambda = 2\pi / k_c, which sets the spatial scale of the emerging and depends on the specific values satisfying the conditions. At the onset of instability (where \max \operatorname{Re}(\lambda(k)) = 0), k_c is determined by solving \frac{d \operatorname{Re}(\lambda)}{d k^2} = 0 at the marginal . Beyond the linear regime, the Turing instability leads to a where the uniform state gives way to finite- spatial patterns. This transition is typically a supercritical in the of the unstable , resulting in , nonlinear structures such as stripes or spots whose form is influenced by higher-order nonlinear terms in the reaction kinetics.

Mechanisms of Pattern Formation

Activator-Inhibitor Model

The activator-inhibitor model forms the core dynamical framework for Turing patterns, where two species interact nonlinearly while diffusing at different rates to drive from a homogeneous state. In this setup, the activator u stimulates its own synthesis and that of the inhibitor v, while the represses both the activator's production and its own, creating a feedback loop essential for instability. Specifically, the elements at the homogeneous satisfy \frac{\partial f}{\partial u} > 0 (activator self-enhancement), \frac{\partial f}{\partial v} < 0 (inhibition of the activator by the ), \frac{\partial g}{\partial u} > 0 (activation of the inhibitor by the activator), and \frac{\partial g}{\partial v} < 0 (inhibitor self-suppression). A widely adopted realization of this model is the Gierer-Meinhardt system, with kinetics f(u,v) = \rho \frac{u^2}{v} - \mu u for the activator (capturing self-enhancement via the quadratic term divided by inhibitor concentration) and g(u,v) = \rho u^2 - \nu v for the inhibitor (driven by activator production but subject to linear decay). These terms, combined with equations \frac{\partial u}{\partial t} = D_u \nabla^2 u + f(u,v) and \frac{\partial v}{\partial t} = D_v \nabla^2 v + g(u,v), where typically D_v \gg D_u, enable pattern emergence. The model's hallmark is short-range coupled with long-range : the slowly diffusing fosters local aggregation by amplifying concentrations in nascent peaks, whereas the rapidly diffusing spreads broadly to suppress activator growth elsewhere, enforcing regular spacing between pattern elements like spots or stripes. This dynamic ensures robust , as local competes with global . Pattern morphology—spots versus stripes—arises from specific parameter regimes, particularly the nonlinearity strength \rho and diffusion ratio D_v / D_u; increasing \rho enhances local amplification to favor compact spots, while moderate diffusion ratios select stripe-like structures through anisotropic growth in the nonlinear regime.

Spatial and Temporal Dynamics

The temporal evolution of Turing patterns typically initiates with the amplification of small random perturbations or initial noise through a linear . In this , modes corresponding to the most unstable spatial grow exponentially over time, driven by the diffusion-reaction dynamics that destabilize the homogeneous state. As the pattern reaches a where nonlinear terms become significant, the growth saturates, leading to the formation of stable, finite-amplitude steady states that maintain the spatial periodicity. This progression from linear growth to nonlinear equilibrium is a hallmark of Turing bifurcations in two-component reaction-diffusion systems. Spatial heterogeneity in Turing patterns arises prominently from the influence of geometry and size on pattern selection. In sufficiently large domains, where boundary effects are negligible, elongated patterns often dominate due to minimized energy costs associated with . Conversely, in finite or confined domains, spot-like patterns are preferentially selected, as the boundaries impose constraints that favor compact, localized structures to accommodate the overall scale. These domain-size dependencies highlight how global can override local instability preferences, altering the emergent without changing intrinsic kinetic parameters. Transitions between distinct Turing pattern types, such as from hexagonal arrays to labyrinthine networks or isolated spots, occur as system parameters—such as diffusion ratios or rates—are varied across the bifurcation landscape. Hexagonal patterns typically emerge near the onset of in isotropic settings, while labyrinthine structures form in regimes with stronger nonlinearities or , featuring interconnected, maze-like features. Spot patterns, in contrast, prevail when parameters favor discrete, non-connected elements. Boundaries further modulate these transitions by pinning pattern wavelengths or introducing topological defects, such as dislocations, that stabilize or irregular configurations against pure periodic states. Stability analysis of these patterns, particularly in the weakly nonlinear regime near the Turing bifurcation, relies on amplitude equations that capture the slow modulation of pattern envelopes. These reduced-order descriptions reveal how small deviations from the critical lead to Eckhaus instabilities, where patterns with wavelengths too far from the preferred value destabilize via perturbations. The Swift-Hohenberg model serves as a example for such analyses, providing a phenomenological framework to study the competition between stabilizing linear terms and destabilizing nonlinear interactions, thereby elucidating the robustness of steady patterns against spatiotemporal perturbations.

Applications in Biology

Morphogenesis and Development

Turing patterns contribute to by transforming broad, uniform gradients into finely patterned spatial domains that guide and formation during embryonic . In limb buds, for example, reaction-diffusion systems refine gradients of signaling molecules such as (BMP) into segmented regions that prefigure limb structures, enabling the emergence of digits and joints from an initially homogeneous field. This refinement occurs through activator-inhibitor interactions that amplify small fluctuations into stable patterns, as demonstrated in computational models validated against developmental data. Integration of Turing mechanisms with gene regulatory networks further specifies these patterns, particularly in formation among vertebrates. A BMP-Sox9-Wnt feedback loop acts as a Turing network, where serves as an activator promoting Sox9 expression for formation, while Wnt inhibits it distally; this system is modulated by opposing gradients of BMP and Wnt to orient and space primordia. additionally tune the wavelength of these Turing patterns by altering diffusion rates or reaction strengths in the network, thereby controlling number and identity—experimental perturbations of expression in limbs result in predictable shifts from five to fewer digits, confirming the mechanism's role. Experimental studies provide direct evidence for Turing-like dynamics in key developmental processes. In mouse somitogenesis, oscillatory in the presomitic generates traveling waves that resemble Turing patterns, synchronizing Wnt and signaling to form periodic somites; recent and models recreate this , showing how local interactions produce the rhythmic segmentation observed around embryonic day 8-12. In , melanin activators in melanophores drive Turing patterns during fin development, interacting with xanthophore inhibitors to form stripes and spots that regenerate post-injury, with cell ablation experiments revealing pattern recovery consistent with reaction-diffusion principles. These findings highlight Turing patterns' versatility in developmental timing and spatial control. Such patterns manifest across scales in early embryogenesis, from cellular resolutions—such as ~100 μm domains in pigment cell clusters or boundaries—to organismal levels in axial and limb outgrowth, ensuring robust structure emergence despite varying sizes. This multiscale underscores Turing mechanisms' efficiency in coordinating and without requiring precise positional cues.

Pigmentation and Tissue Patterns

Turing patterns manifest prominently in the pigmentation of animal coats, where reaction-diffusion mechanisms involving generate striking spatial arrangements such as stripes and spots. In zebras, the black-and-white stripes arise from a Turing instability driven by differential diffusion rates of morphogens that activate and inhibit production in skin cells. Similarly, spots result from short-range activation and long-range inhibition in epidermal distributions, as simulated in two-stage Turing models that first establish a pre-pattern of spots and then refine it during growth. A concrete biological validation comes from skin, where melanophores (black pigment cells) and xanthophores (yellow pigment cells) interact via short-range repulsive protrusions and long-range attractive signals, forming stripes that match Turing predictions; a 2014 identified these cell-cell contacts as key to pattern stability during . In , Turing-like patterns emerge in —the spiral arrangement of leaves and seeds—and leaf vein networks, primarily through acting as an activator in a reaction- framework. , a phytohormone, creates feedback loops where its efflux via PIN1 proteins polarizes cells, leading to convergent flows that self-organize vein precursors into hierarchical networks; mutants in PIN6 and PIN8 disrupt this, resulting in irregular venation that confirms the Turing mechanism's role in spatial periodicity. For , maxima at primordia sites drive inhibitory fields that space organs optimally, producing spirals observed in sunflowers and , with models integrating transport and diffusion to replicate these patterns without invoking mechanical forces alone. Human dermal patterns also exhibit Turing characteristics, particularly in fingerprint whorls and ridges, which form through a reaction-diffusion system involving WNT activation, BMP inhibition, and EDAR signaling to establish periodic epidermal thickenings around week 15 of gestation. Likewise, hair follicle spacing follows a Turing pre-pattern where WNT and FGF promote local aggregation of dermal cells, while BMP and DKK provide long-range inhibition, ensuring uniform distribution across the and body; disruptions in these pathways, as seen in models, lead to clustered or irregular follicles. Genetic variations can modulate these Turing patterns, altering outcomes from stripes to spots or blotches, as exemplified in tabby cats. Mutations in the Taqpep gene broaden epidermal pre-patterns, transforming mackerel tabby stripes into blotched forms by disrupting Wnt-inhibitor balance, while Dkk4 variants reduce spot size and increase density via enhanced Wnt signaling, demonstrating how allelic changes fine-tune and parameters in the underlying reaction-diffusion system.

Applications Beyond Biology

Chemical Reactions

Turing patterns have been experimentally realized in non-biological chemical systems, providing direct validation of Alan Turing's theoretical predictions for reaction- instabilities. These realizations typically involve far-from-equilibrium oscillatory reactions where rates differ sufficiently between species to destabilize uniform states and generate spatial heterogeneity. Key examples include the Belousov-Zhabotinsky (BZ) reaction and surface catalytic processes analogous to the Gray-Scott model, demonstrating transitions from temporal oscillations to stationary spatial structures under controlled conditions. The BZ reaction, an autocatalytic oxidation-reduction process involving bromate (BrO₃⁻), (CH₂(COOH)₂), and ions (Ce³⁺/Ce⁴⁺) in an acidic medium, exemplifies Turing pattern formation in solution. In bulk, it exhibits temporal oscillations, but spatial patterns emerge when is constrained, such as in thin layers or microemulsions, leading to traveling , spirals, and static spots or stripes. Early observations of complex wave patterns in the BZ reaction during the 1970s by Arthur Winfree confirmed aspects of Turing's morphogenesis theory, highlighting the system's potential for despite initial skepticism about non-biological pattern formation. Stationary Turing structures, including labyrinthine and spot-like domains, are achieved by immobilizing reactants in gels like , which suppresses and allows differential to dominate, stabilizing patterns near the oxidized state. The Gray-Scott model, originally a simplified mathematical of an autocatalytic between U and V with feed and decay terms, has been physically realized in heterogeneous catalytic systems, notably the oxidation of (CO) on (Pt) surfaces during the 1990s. In these experiments, CO and oxygen (O₂) adsorb on Pt(110) or Pt(100) catalysts under , forming standing waves, cellular structures, and Turing-like stripes due to and kinetics mimicking the model's . Pattern types in such systems evolve from temporal oscillations in well-mixed conditions to spatial Turing motifs when diffusion or boundary effects are introduced, as seen in photolithographically defined domains. Observing Turing patterns in chemical reactions presents significant challenges, primarily due to the acute sensitivity to initial concentrations, , and diffusion coefficients, which must differ by at least an between activator and to satisfy conditions—a rarity for small molecules with similar diffusivities around 10⁻⁵ cm²/s. often disrupts stationary patterns in unstirred solutions, necessitating techniques like gels or emulsions to isolate reaction-diffusion effects. These parameter constraints delayed experimental confirmation of Turing's predictions until the late , with ongoing refinements in microscale setups enabling reproducible observations.

Computational and Engineering Uses

Turing patterns are simulated computationally by solving the underlying reaction-diffusion partial differential equations (PDEs) using numerical methods such as schemes, which discretize and time to approximate solutions and capture pattern emergence in models like the Gray-Scott system. methods, including and Chebyshev approaches, offer higher accuracy for periodic patterns by expanding solutions in basis functions, enabling efficient computation of Turing instabilities in two- and three-dimensional domains. These techniques are implemented in software tools, such as libraries like rd-spiral, which employs pseudo- methods for 2D reaction-diffusion dynamics and Turing pattern visualization. MATLAB codes are also commonly used for custom simulations of Turing pattern formation in biological and chemical contexts, often incorporating approximations for parameter exploration. In , Turing-inspired diffusion processes have been harnessed for self-assembling materials, particularly in for creating high-resolution patterned coatings. For instance, thin-film solutions of can form Turing-like patterns through physical without chemical reactions, achieving sub-micrometer feature sizes suitable for optoelectronic devices. This approach, developed in the 2020s, leverages non-equilibrium solvent evaporation to spontaneously organize microstructures, demonstrating scalability for large-area coatings in . Biomedical modeling employs Turing frameworks to predict spatial patterns in tissue dynamics. Reaction-diffusion models incorporating Wnt signaling have been used to simulate emergent metabolic asymmetry in colorectal tumors, revealing how diffusion-driven instabilities contribute to heterogeneous growth patterns observed in clinical samples. Similarly, chemotaxis models with Turing instability analyze drug diffusion in biological tissues, where nonlinear diffusion and activator-inhibitor interactions predict that enhances targeted delivery by concentrating agents in tumor regions. These simulations aid in optimizing therapeutic strategies by forecasting how drug dispersal interacts with tissue architecture to form localized concentrations. Broader impacts include algorithmic design in , where Turing pattern models inspire in swarms. Algorithms adapting Turing reaction-diffusion principles enable active swarm robots to form and maintain spatial patterns using only local neighbor information, mimicking biological for tasks like collective exploration or structure assembly. Such designs promote robust, decentralized in large-scale robotic systems, with applications in search-and-rescue operations and adaptive manufacturing.

References

  1. [1]
    The chemical basis of morphogenesis - Journals
    A system of reactions and diffusion on a sphere is also considered. Such a system appears to account for gastrulation. Another reaction system in two dimensions ...
  2. [2]
    Turing's theory of morphogenesis of 1952 and the subsequent ... - NIH
    Feb 8, 2012 · In his pioneering work, Alan Turing showed that de novo pattern formation is possible if two substances interact that differ in their diffusion range.
  3. [3]
  4. [4]
    Turing pattern design principles and their robustness - PMC - NIH
    Nov 8, 2021 · (c) Patterns on snail and sea shells are one of the classical examples of Turing patterns in nature. ... In cell biology, we study, for example, ...
  5. [5]
    Pattern formation mechanisms of self-organizing reaction-diffusion ...
    Apr 1, 2020 · Turing's theory explains the complex self-organizing mechanisms underlying embryonic patterning using simple reactions of just two diffusible ...
  6. [6]
    Spots, stripes and more: Working out the logic of animal patterns
    May 23, 2024 · More than 70 years ago, mathematician Alan Turing proposed a mechanism that explained how patterns could emerge from bland uniformity.Missing: plant leaf veins
  7. [7]
    Patterning, From Conifers to Consciousness: Turing's Theory and ...
    This is a brief account of Turing's ideas on biological pattern and the events that led to their wider acceptance by biologists as a valid way to investigate ...
  8. [8]
    Turing patterns | Feature - Chemistry World
    May 30, 2012 · Turing showed how chemical reactions can create patterns. If you mix the ingredients he specified, they can separate into quasi-ordered patches of different ...Missing: original | Show results with:original
  9. [9]
    None
    Error: Could not load webpage.<|separator|>
  10. [10]
    A theory of biological pattern formation
    The theory is based on short range activation, long range inhibition, and a distinction between activator and inhibitor concentrations on one hand, and the ...Missing: Hans Alfred
  11. [11]
    Experimental evidence of a sustained standing Turing-type ...
    Jun 11, 1990 · We report the experimental observation of a sustained standing nonequilibrium chemical pattern in a single-phase open reactor.Missing: confirmation Zhabotinsky
  12. [12]
    Forging patterns and making waves from biology to geology - Journals
    Apr 19, 2015 · In Meinhardt and Gierer's model, stationary chemical patterns can result from two interacting ingredients—equivalent to Turing's morphogens—if ...<|separator|>
  13. [13]
  14. [14]
    Pattern formation outside of equilibrium | Rev. Mod. Phys.
    Jul 1, 1993 · Pattern formation outside of equilibrium. M. C. Cross and P. C. Hohenberg · M. C. Cross. Department of Physics, California Institute of ...
  15. [15]
    Stripes, spots, or reversed spots in two-dimensional Turing systems
    Striped patterns are produced when the equilibrium is equally distant from the upper and the lower limitations, but spotted patterns are produced when the ...
  16. [16]
    Cooperativity To Increase Turing Pattern Space for Synthetic Biology
    The growth of Turing patterns also depends on the size domain L, which, in any case, must be greater than the typical pattern size (i.e., L > π/kmax).
  17. [17]
    Characterizing topological transitions in a Turing-pattern-forming ...
    May 10, 2012 · In general, two-dimensional Turing patterns may spontaneously appear under three completely different spatial configurations, namely hexagonal ...
  18. [18]
    Digit patterning during limb development as a result of the BMP ...
    Dec 18, 2012 · We propose that the observed patterning during limb bud development may be the consequence of a Turing type patterning mechanism that arises ...Missing: segmentation | Show results with:segmentation
  19. [19]
    The Clock and Wavefront Self-Organizing model recreates the ...
    ABSTRACT. During mouse development, presomitic mesoderm cells synchronize Wnt and Notch oscillations, creating sequential phase waves that pattern somites.
  20. [20]
    Interactions between zebrafish pigment cells responsible for ... - PNAS
    May 26, 2009 · Studies using this model animal have made it possible to uncover the basic mechanism that generates the Turing pattern in biological systems.Missing: melanin | Show results with:melanin
  21. [21]
    Positional information and reaction-diffusion: two big ideas in ...
    Apr 1, 2015 · In the limb buds (ii), the positions of future digits are created as a Turing pattern driven by a feedback loop between Wnt and Bmp signalling ...
  22. [22]
    Of Turing and zebras: Turing diffusion inspires applications in nature ...
    May 15, 2023 · The Turing diffusion model emerges as an explanation for pattern formation in many species and across biological scales.Abstract · Zebra Stripes · A General Theory And...
  23. [23]
    (PDF) Two-stage Turing model for generating pigment patterns on ...
    Aug 6, 2025 · Starting from this spotted pattern, we successfully generate patterns of adult leopards and jaguars by tuning parameters of the model in the ...
  24. [24]
    Studies of Turing pattern formation in zebrafish skin - PMC
    Nov 8, 2021 · Turing's reaction–diffusion system [1] was first proposed as a principle to explain the autonomous nature of biological pattern formation. This ...
  25. [25]
    Patterning of Leaf Vein Networks by Convergent Auxin Transport ...
    Feb 21, 2013 · We show here that vein patterning in the Arabidopsis leaf is controlled by two distinct and convergent auxin-transport pathways.
  26. [26]
    Periodic pattern formation during embryonic development
    Turing RD systems have now been implicated in the formation of a diverse array of embryonic periodic patterns including the digits of the limb [3,17], hair ...<|control11|><|separator|>
  27. [27]
    Developmental genetics of color pattern establishment in cats - PMC
    Sep 16, 2021 · Early in development, we identify stripe-like alterations in epidermal thickness preceded by a gene expression pre-pattern. The secreted Wnt ...
  28. [28]
    Theory of Turing pattern formation and its experimental realization in ...
    Jul 30, 2024 · The first experimental realization of the Turing pattern was achieved in 1990 in a chlorite–iodide–malonic acid (CIMA) reaction system. Iodide ...
  29. [29]
    Patterns in the Belousov–Zhabotinsky reaction in water-in-oil ...
    We observe stationary Turing patterns at temperatures in the range 15–35 °C and bulk oscillations at T = 40–55 °C. When a temperature gradient ΔT is applied ...
  30. [30]
    Testing Turing's theory of morphogenesis in chemical cells - NIH
    We exploit an abiological experimental system of emulsion drops containing the Belousov–Zhabotinsky reactants ideally suited to test Turing's theory. Our ...
  31. [31]
    Introduction—A Bit of History
    In 1972, the cover of Science magazine featured Winfree's photo of spiral wave patterns in the BZ reaction (Winfree, 1972). Figure 1.7 shows such patterns.
  32. [32]
    (PDF) Turing patterns in confined gel and gel-free media
    Oct 26, 2025 · ... gel that is fed at the inner and outer rims. The reaction that we have studied is the Belousov–Zhabotinskii (BZ) reaction, which has become ...
  33. [33]
    “Black spots” in a surfactant-rich Belousov–Zhabotinsky reaction ...
    May 3, 2005 · Stationary spotlike and labyrinthine Turing patterns are found close to the fully oxidized state. ... Spatiotemporal Pattern Formation and Chaos ...
  34. [34]
    Effects of Boundaries on Pattern Formation: Catalytic Oxidation of ...
    The effect of boundaries on pattern formation was studied for the catalytic oxidation of carbon monoxide on platinum surfaces. Photolithography was used to ...
  35. [35]
    Pattern formation during CO oxidation in complex Pt domains
    Jul 1, 1995 · The exploration of pattern formation by reaction-diffusion systems in complex bounded domains has begun only recently.
  36. [36]
    Standing Wave Patterns in the CO Oxidation Reaction on a Pt(110 ...
    We present new experimental data that indicate the important role played by the formation of subsurface oxygen. The formation of these patterns is correlated ...
  37. [37]
    Testing Turing's theory of morphogenesis in chemical cells - PNAS
    Turing proposed that intercellular reaction-diffusion of molecules is responsible for morphogenesis. The impact of this paradigm has been profound.
  38. [38]
    [PDF] Notes on the Turing Instability and Chemical Instabilities
    Apr 25, 2006 · An additional difficulty in observing the Turing instability is that the diffusion constants of the some of the chemical participants must ...
  39. [39]
    Finite-Difference Schemes for Reaction–Diffusion Equations ...
    Feb 1, 2007 · We present two finite-difference algorithms for studying the dynamics of spatially extended predator–prey interactions with the Holling type II functional ...
  40. [40]
    Numerical simulations of Turing patterns in a reaction-diffusion ...
    Oct 4, 2018 · We apply Chebyshev spectral methods which proved to be numerical methods that can significantly speed up the computation of systems of reaction ...
  41. [41]
    Pattern formation and KPP equation - File Exchange - MathWorks
    This 15-line matlab program solves the nonlinear reaction diffusion equation, called Kolmogorov-Petrovskii-Piskunov (KPP) equation to generate patterns.Missing: toolbox | Show results with:toolbox
  42. [42]
    Turing patterns with high-resolution formed without chemical ...
    Dec 2, 2022 · Regular patterns can form spontaneously in chemical reaction-diffusion systems under non-equilibrium conditions as proposed by Alan Turing.
  43. [43]
    Mathematical modeling links Wnt signaling to emergent patterns of ...
    To understand this pattern, we developed a reaction–diffusion model that incorporates Wnt signaling, a pathway known to upregulate PDK1 and Warburg metabolism.
  44. [44]
    Chemotaxis Model for Drug Delivery Using Turing's Instability and ...
    This paper is devoted to the study of the chemotaxis model for drug delivery purposes. The pattern formation for a volume-filling with nonlinear diffusive ...Chemotaxis Model For Drug... · 2. Materials And Methods · Numerical Scheme And...<|separator|>
  45. [45]
    An algorithm applied the Turing pattern model to control active ...
    May 28, 2024 · This research focuses on such self-organizing robots or swarm robots. The proposed algorithm is a model that applies the Turing pattern, one of ...Missing: design | Show results with:design