Neurodevelopmental disorders modeling using isogeometric analysis, dynamic domain expansion and local refinement

Kuanren Qian Genesis Omana Suarez Toshi Nambara Takahisa Kanekiyo Ashlee S. Liao Victoria A. Webster-Wood Yongjie Jessica Zhang Department of Mechanical Engineering, Carnegie Mellon University, 5000 Forbes Ave, Pittsburgh, PA 15213, USA Department of Neuroscience, Mayo Clinic, 4500 San Pablo Road, Jacksonville, FL 32224, USA Department of Biomedical Engineering, Carnegie Mellon University, 5000 Forbes Ave, Pittsburgh, PA 15213, USA McGowan Institute for Regenerative Medicine, University of Pittsburgh, 450 Technology Drive, Pittsburgh, PA 15219, USA
Abstract

Neurodevelopmental disorders (NDDs) have arisen as one of the most prevailing chronic diseases within the US. Often associated with severe adverse impacts on the formation of vital central and peripheral nervous systems during the neurodevelopmental process, NDDs are comprised of a broad spectrum of disorders, such as autism spectrum disorder, attention deficit hyperactivity disorder, and epilepsy, characterized by progressive and pervasive detriments to cognitive, speech, memory, motor, and other neurological functions in patients. However, the heterogeneous nature of NDDs poses a significant roadblock to identifying the exact pathogenesis, impeding accurate diagnosis and the development of targeted treatment planning. A computational NDDs model holds immense potential in enhancing our understanding of the multifaceted factors involved and could assist in identifying the root causes to expedite treatment development. To tackle this challenge, we introduce optimal neurotrophin concentration to the driving force and degradation of neurotrophin to the synaptogenesis process of a 2D phase field neuron growth model using isogeometric analysis to simulate neurite retraction and atrophy. The optimal neurotrophin concentration effectively captures the inverse relationship between neurotrophin levels and neurite survival, while its degradation regulates concentration levels. Leveraging dynamic domain expansion, the model efficiently expands the domain based on outgrowth patterns to minimize degrees of freedom. Based on truncated T-splines, our model simulates the evolving process of complex neurite structures by applying local refinement adaptively to the cell/neurite boundary. Furthermore, a thorough parameter investigation is conducted with detailed comparisons against neuron cell cultures in experiments, enhancing our fundamental understanding of the mechanisms underlying NDDs.

keywords:
Neuron growth , Neurodevelopmental disorders , Phase field method , Isogeometric analysis , Dynamic domain expansion , Truncated T-splines , Local refinement
journal: Computer Methods in Applied Mechanics and Engineering

1 Introduction

Neurodevelopmental disorders (NDDs), often associated with impairments during the neuron developmental process, pose persistent complications that have a profound and long-lasting impact on patients [1, 2]. A wide range of potential factors behind the etiology of NDDs are gaining recognition recently, with emerging research highlighting their significant impact on the pathophysiology of these conditions [3, 4]. Investigations into the functional roles of these factors have revealed some potential influences on NDDs progression, indicating that they play crucial roles during the neurodevelopmental process [5, 6]. Moreover, emerging evidence suggests that certain factors involved during the neurodevelopmental process may play protective roles, potentially preventing the onset of NDDs [7, 8]. However, the lack of comprehensive disorder studies of the specific biological functions or biophysical processes highlights a significant gap in the current field. Furthermore, NDDs often manifest through complex neuron morphological transformations, such as beading, retraction, and atrophy. All of these behaviors add complexity to a thorough neurological study and pose challenges to therapeutic targeting and planning [9, 10]. Analyzing the complex morphological changes observed in NDDs and understanding the biological mechanisms that drive these conditions are crucial for develo** effective therapeutic strategies. Given the level of complexity associated with the sophisticated processes involved, unconventional approaches are necessary to enhance our understanding of the NDDs process. As such, there is a pressing need for a comprehensive computational model to study NDDs and unravel the intricate neurodevelopmental process. These models will be indispensable for shedding light on the fundamental causes and complexities of NDDs, consequently enabling the development of precise and potent treatments [11, 12].

In computational neuroscience, significant progress has been made in mathematically modeling neurodevelopmental processes, including initial neurite outgrowth [13], axon differentiation [14, 15] and axon guidance [16]. Phenomenological models have provided insights into a wide range of neurodevelopmental processes such as filopodia [17], external repulsive cues [18], stochastic mechanisms [19], generalized neurite characteristics for modeling morphology [20, 21], and incorporating interactions with surrounding substrates [22]. However, these models focus on phenomenological outcomes, often overlooking underlying biophysics [23, 24]. Integrating comprehensive biophysical mechanisms into neuron growth models comes with significant computational costs and numerical instabilities [25, 26], which are particularly aggravated in scenarios like intracellular material transport [27] and traffic jams in complex 3D structures [28]. Despite these complications, ongoing efforts are underway to enhance and validate computational models with biophysical phenomena from experimental observations. Among different approaches, some noteworthy work includes the utilization of phase field techniques for neuron growth modeling [29], exploration of neurotrophin interactions [30], and the development of biophysically coupled phase field neuron growth models [31]. There are also models that integrate neurite morphometric features and bridge the gap between theoretical and experimental neuroscience [32, 33]. Considering the complexity and large variety of the neural development process [34], robust computational approaches that tackle high-order equations on complex geometries are crucial for advancing our understanding of neurodevelopmental processes and paving the way toward effective targeted therapeutic interventions. This raises the need for isogeometric analysis (IGA) with high-fidelity spline modeling techniques that provide the necessary flexibility and precision [35, 36], particularly through domain expansion and localized refinements to accurately capture the dynamic and detailed evolution of neuron growth during development and degradation during apoptosis on response to damage or pathology.

Non-uniform rational B-splines are widely adopted [37, 38] and were initially chosen as the foundational basis for IGA [39, 40]. However, B-splines lack local refinement support, an important feature for our NDDs study. To address this limitation, T-splines have been developed to work with IGA and support the local refinements required for efficient and accurate analysis [41]. Local refinement in T-splines is achieved with T-junctions that are analogous to hanging nodes in conventional finite element methods (FEM), which break down global tensor product structures [42, 43]. This adaptability particularly benefits IGA because it lowers the degrees of freedom (DOFs) necessary while preserving the exact representations. In addition to local refinement, T-splines maintain essential properties of B-splines, including non-negativity and partition of unity. Thus, they extend the functionality of B-splines through local refinements while preserving the underlying mathematical properties. These properties make T-splines highly desirable in a wide range of problems, therefore leading to the development of T-splines into various forms such as analysis-suitable T-splines [44], LR-splines [45, 46], modified T-splines [47, 48], and weighted T-splines [49, 50]. Hierarchical approaches such as PHT-splines [51], hierarchical analysis-suitable T-splines [52], and truncated hierarchical Catmull–Clark subdivision [53, 54] have also been used for local refinement, with additional advancements in adaptive refinement techniques [55]. The capability of T-splines is further showcased in modeling heterogeneous solids [56], additive manufacturing analysis [57], and arbitrary-degree T-splines for IGA of Kirchhoff-Love shells [58]. In particular, the applications of T-splines in intracellular material transport modeling [28, 59, 60] further underline the versatility and potential of IGA and truncated T-splines in computational neuroscience.

To this end, we propose a novel IGA phase-field NDDs model coupling neuron growth with complex biophysics processes, built upon dynamic domain expansion, local refinement, and the Portable, Extensible Toolkit for Scientific Computation (PETSc) [61, 62] for Message Passing Interface (MPI) parallelization. With this model, we can investigate how neurons behave under NDDs and the underlying biophysics processes. The main contributions include:

  • 1.

    Development of a PETSc-based IGA phase field NDDs model. The NDDs model simulates complex neuron growth and disorder behaviors on truncated T-splines, leveraging PETSc for efficient parallel processing;

  • 2.

    Introduction of optimal neurotrophin concentration into the driving force and degradation of neurotrophin into the synaptogenesis process of the phase field model to simulate neurite retraction and atrophy. The optimal neurotrophin concentration effectively captures the inverse relationship between neurotrophin levels and neurite survival, and its degradation regulates concentration levels, providing critical insights into NDDs;

  • 3.

    Dynamic domain expansion that optimizes computational focus by expanding the domain based on neurite growth and interface-based local refinements that refine at evolving interface to preserve accuracy while minimizing computational load;

  • 4.

    Comprehensive NDDs study utilizing the computational model to investigate neurite retraction and atrophy, providing new insights into possible factors and biophysical mechanisms underlying NDDs; and

  • 5.

    Detailed validation through comparisons of the external-cue guided computational NDDs model with experimental healthy human induced pluripotent stem cell (iPSC)-derived neurons and rat hippocampus neurons undergoing excitotoxic apoptosis.

The rest of this paper is organized as follows. Section 2 outlines our NDDs model structure. Section 3 introduces our IGA-based phase field model for NDDs. Section 4 reviews the truncated T-splines and introduces our phase field variable interface-based local refinements. Section 5 walks through our dynamic domain expansion algorithm and interpolation process. Section 6 documents the experimental procedures for culturing human iPSCs-derived neurons and rat hippocampal neurons. Section 7 explains the parallelized computational model and showcases neuron growth and disorders simulation results with experimental comparisons. Section 8 concludes the NDDs model and discusses the potential future directions.

2 Algorithm overview

Refer to caption
Figure 1: Flow chart of the NDDs modeling pipeline. (A) The overall pipeline that conducts NDDs modeling. (B) Adaptive mesh refinement module that locally refines the mesh based on neuron outgrowth. (C) The IGA NDDs model simulates disorders using the phase field method, PETSc, and truncated T-splines. Parameters in red dashed boxes are selected to study their effects on NDDs. (D) Dynamic domain expansion module that directionally expands domain based on neurites near the domain boundary.

The flowchart illustrates our NDDs modeling workflow (Fig. 1). The model starts with predefined domain size and neuron growth parameters, which are used to generate the initial control mesh and variables (Fig. 1A). Then, the adaptive mesh refinement module (Fig. 1B) is called to generate truncated T-splines. This module checks for control mesh elements that require local refinements based on phase field variable ϕitalic-ϕ\phiitalic_ϕ value and selectively refines elements located on the ϕitalic-ϕ\phiitalic_ϕ interface, achieving sufficient resolution while kee** DOFs minimum. The relevant elements ID are selected and passed if refinement is needed to generate locally refined Bézier mesh [63]. Subsequently, mesh partitioning software METIS is called to partition the generated mesh for optimizing the computational load distribution across multiple processors [64]. The model then builds and solves all nonlinear and linear systems using PETSc (Fig. 1C) [61]. Afterward, the model checks whether the final iterations have been reached. If not, the model checks for neurites near domain boundaries and expands the domain as needed using the dynamic domain expansion module (Fig. 1D). If the mesh changes, the model regenerates the truncated T-splines to ensure that all Bézier elements have up-to-date control point information. The above steps are iterated until the final iteration is reached.

3 IGA-based phase field NDDs modeling

IGA and phase field methods are robust and powerful numerical methods for modeling complex engineering problems [65, 66]. IGA is a high-order numerical method that can capture the exact smooth representation of the geometry by eliminating the discretization needed by conventional FEM [39]. The phase field method, on the other hand, specializes in tackling evolving boundaries such as crack propagation and dendritic solidification [67]. Considering that NDDs are neuron growth processes with retracting cell boundaries, they are essentially an interface evolution problem. The convergence of IGA and phase field is highly effective in accurately simulating neurite morphological transformations. Utilizing these two techniques, an IGA-based phase field framework was introduced to depict the complex stages of healthy neuron growth by incorporating intracellular concentration [31, 32]. This model accounts for intracellular transport during the growth, enables the differentiation of the longest neurite into an axon, and models various growth dynamics across multiple stages to simulate the behavior of growth cones at neurite tips based on neurite morphometric features [68]. The proposed NDDs model extends upon it, consisting of five equations:

  1. 1.

    Phase field governing equation that captures the neuron morphological transformation with phase field variable;

  2. 2.

    Intracellular tubulin transport equation that models the effect of tubulin on neurite elongations;

  3. 3.

    Competitive tubulin consumption equation that captures the consumption of tubulin at neurite tips;

  4. 4.

    Synaptogenesis equation that captures the effect of neurotrophin particles supporting interface evolution. We include the effect of neurotrophin degradation that regulates its concentration level to model NDDs; and

  5. 5.

    Driving force equation that couples the effect of tubulin and neurotrophin back to the phase field equation. We introduce optimal neurotrophin concentration to model the inverse relationship between neurotrophin level and neurite survival. This inverse relationship adjusts the driving force magnitude and, therefore, drives the phase field interface evolution for NDDs modeling.

Phase field governing equation. To model NDDs in the context of phase field, we treat the neuron domain ΩΩ\Omegaroman_Ω as a binary phase field ϕitalic-ϕ\phiitalic_ϕ, where phase “1” indicates the neuron and phase “0” is the extracellular environment. The phase field interface evolution is achieved by constantly solving for energy minimization at interfaces. Starting with the simplified free energy functional for the phase field model as

Efree=Ω(Echem+Egrad+Edoub)𝑑Ω,subscript𝐸𝑓𝑟𝑒𝑒subscriptΩsubscript𝐸𝑐𝑒𝑚subscript𝐸𝑔𝑟𝑎𝑑subscript𝐸𝑑𝑜𝑢𝑏differential-dΩE_{free}=\int_{\Omega}\left(E_{chem}+E_{grad}+E_{doub}\right)d\Omega,italic_E start_POSTSUBSCRIPT italic_f italic_r italic_e italic_e end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT ( italic_E start_POSTSUBSCRIPT italic_c italic_h italic_e italic_m end_POSTSUBSCRIPT + italic_E start_POSTSUBSCRIPT italic_g italic_r italic_a italic_d end_POSTSUBSCRIPT + italic_E start_POSTSUBSCRIPT italic_d italic_o italic_u italic_b end_POSTSUBSCRIPT ) italic_d roman_Ω , (1)

where Echemsubscript𝐸𝑐𝑒𝑚E_{chem}italic_E start_POSTSUBSCRIPT italic_c italic_h italic_e italic_m end_POSTSUBSCRIPT is the chemical free energy density that expresses the bulk energy relationship, Egradsubscript𝐸𝑔𝑟𝑎𝑑E_{grad}italic_E start_POSTSUBSCRIPT italic_g italic_r italic_a italic_d end_POSTSUBSCRIPT is the gradient energy density [29], and Edoubsubscript𝐸𝑑𝑜𝑢𝑏E_{doub}italic_E start_POSTSUBSCRIPT italic_d italic_o italic_u italic_b end_POSTSUBSCRIPT is the double-well function that describes the barrier in free energy density as a function of the phase field parameter ϕitalic-ϕ\phiitalic_ϕ with two stable states representing the two phases of the system. Then, following the modified Allen–Cahn equation [67, 69], we have

ϕt=MϕδEfreeδϕ,italic-ϕ𝑡subscript𝑀italic-ϕ𝛿subscript𝐸𝑓𝑟𝑒𝑒𝛿italic-ϕ\frac{\partial\phi}{\partial t}=-M_{\phi}\frac{\delta E_{free}}{\delta\phi},\\ divide start_ARG ∂ italic_ϕ end_ARG start_ARG ∂ italic_t end_ARG = - italic_M start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT divide start_ARG italic_δ italic_E start_POSTSUBSCRIPT italic_f italic_r italic_e italic_e end_POSTSUBSCRIPT end_ARG start_ARG italic_δ italic_ϕ end_ARG , (2)

where the functional derivative δEfreeδϕ𝛿subscript𝐸𝑓𝑟𝑒𝑒𝛿italic-ϕ\frac{\delta E_{free}}{\delta\phi}divide start_ARG italic_δ italic_E start_POSTSUBSCRIPT italic_f italic_r italic_e italic_e end_POSTSUBSCRIPT end_ARG start_ARG italic_δ italic_ϕ end_ARG is derived based on Eqn. 1:

δEfreeδϕ=EchemϕEgrad(ϕ)+Edoubϕ,𝛿subscript𝐸𝑓𝑟𝑒𝑒𝛿italic-ϕsubscript𝐸𝑐𝑒𝑚italic-ϕsubscript𝐸𝑔𝑟𝑎𝑑italic-ϕsubscript𝐸𝑑𝑜𝑢𝑏italic-ϕ\frac{\delta E_{free}}{\delta\phi}=\frac{\partial E_{chem}}{\partial\phi}-% \nabla\cdot\frac{\partial E_{grad}}{\partial(\nabla\phi)}+\frac{\partial E_{% doub}}{\partial\phi},divide start_ARG italic_δ italic_E start_POSTSUBSCRIPT italic_f italic_r italic_e italic_e end_POSTSUBSCRIPT end_ARG start_ARG italic_δ italic_ϕ end_ARG = divide start_ARG ∂ italic_E start_POSTSUBSCRIPT italic_c italic_h italic_e italic_m end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_ϕ end_ARG - ∇ ⋅ divide start_ARG ∂ italic_E start_POSTSUBSCRIPT italic_g italic_r italic_a italic_d end_POSTSUBSCRIPT end_ARG start_ARG ∂ ( ∇ italic_ϕ ) end_ARG + divide start_ARG ∂ italic_E start_POSTSUBSCRIPT italic_d italic_o italic_u italic_b end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_ϕ end_ARG , (3)

with

Echemsubscript𝐸𝑐𝑒𝑚\displaystyle E_{chem}italic_E start_POSTSUBSCRIPT italic_c italic_h italic_e italic_m end_POSTSUBSCRIPT =p(ϕ)fs+(1p(ϕ))fl,absent𝑝italic-ϕsubscript𝑓𝑠1𝑝italic-ϕsubscript𝑓𝑙\displaystyle=p(\phi)f_{s}+(1-p(\phi))f_{l},= italic_p ( italic_ϕ ) italic_f start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + ( 1 - italic_p ( italic_ϕ ) ) italic_f start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT , (4)
Egradsubscript𝐸𝑔𝑟𝑎𝑑\displaystyle E_{grad}italic_E start_POSTSUBSCRIPT italic_g italic_r italic_a italic_d end_POSTSUBSCRIPT =a(Ψ)22|ϕ|2,absent𝑎superscriptΨ22superscriptitalic-ϕ2\displaystyle=\frac{a(\Psi)^{2}}{2}|\nabla\phi|^{2},= divide start_ARG italic_a ( roman_Ψ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG | ∇ italic_ϕ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (5)
Edoubsubscript𝐸𝑑𝑜𝑢𝑏\displaystyle E_{doub}italic_E start_POSTSUBSCRIPT italic_d italic_o italic_u italic_b end_POSTSUBSCRIPT =Wq(ϕ).absent𝑊𝑞italic-ϕ\displaystyle=Wq(\phi).= italic_W italic_q ( italic_ϕ ) . (6)

Following approach taken in existing literature [67, 70], Echemsubscript𝐸𝑐𝑒𝑚E_{chem}italic_E start_POSTSUBSCRIPT italic_c italic_h italic_e italic_m end_POSTSUBSCRIPT interpolates a monotomically increasing bulk free energy relationship between solid fSsubscript𝑓𝑆f_{S}italic_f start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT and liquid fLsubscript𝑓𝐿f_{L}italic_f start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT with p(ϕ)=ϕ3(1015ϕ+6ϕ2)𝑝italic-ϕsuperscriptitalic-ϕ31015italic-ϕ6superscriptitalic-ϕ2p(\phi)=\phi^{3}(10-15\phi+6\phi^{2})italic_p ( italic_ϕ ) = italic_ϕ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( 10 - 15 italic_ϕ + 6 italic_ϕ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). Egradsubscript𝐸𝑔𝑟𝑎𝑑E_{grad}italic_E start_POSTSUBSCRIPT italic_g italic_r italic_a italic_d end_POSTSUBSCRIPT is defined using the anisotropy gradient coefficient a(Ψ)𝑎Ψa(\Psi)italic_a ( roman_Ψ ) to account for thin neurite morphology, Edoubsubscript𝐸𝑑𝑜𝑢𝑏E_{doub}italic_E start_POSTSUBSCRIPT italic_d italic_o italic_u italic_b end_POSTSUBSCRIPT is defined using double well function q(ϕ)=ϕ2(1ϕ)2𝑞italic-ϕsuperscriptitalic-ϕ2superscript1italic-ϕ2q(\phi)=\phi^{2}(1-\phi)^{2}italic_q ( italic_ϕ ) = italic_ϕ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - italic_ϕ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. W𝑊Witalic_W is the energy barrier magnitude. Substituting Eqns. 3-6 into Eqn 2, we obtain

ϕt=Mϕ[(a(Ψ)2ϕ)+4Wϕ(1ϕ)(ϕ12+152Wϕ(1ϕ)(fLfS))].\frac{\partial\phi}{\partial t}=M_{\phi}\left[\bigtriangledown\cdot\left(a(% \Psi)^{2}\bigtriangledown\phi\right)+4W\phi(1-\phi)\left(\phi-\frac{1}{2}+% \frac{15}{2W}\phi(1-\phi)(f_{L}-f_{S})\right)\right].divide start_ARG ∂ italic_ϕ end_ARG start_ARG ∂ italic_t end_ARG = italic_M start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT [ ▽ ⋅ ( italic_a ( roman_Ψ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ▽ italic_ϕ ) + 4 italic_W italic_ϕ ( 1 - italic_ϕ ) ( italic_ϕ - divide start_ARG 1 end_ARG start_ARG 2 end_ARG + divide start_ARG 15 end_ARG start_ARG 2 italic_W end_ARG italic_ϕ ( 1 - italic_ϕ ) ( italic_f start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ) ) ] . (7)

The anisotropic gradient coefficient a(Ψ)𝑎Ψa(\Psi)italic_a ( roman_Ψ ) [29, 71] is defined as

a(Ψ)={a1(Ψ)=a¯1+ξ[1+ξcos(k(Ψθ))]for πi+θmΨθπ(i+1)θm,a2(Ψ)=a1(θm)cos(Ψθ)cosθmfor πiθm<Ψθ<πi+θm.𝑎Ψcasessubscript𝑎1Ψ¯𝑎1𝜉delimited-[]1𝜉𝑘Ψ𝜃for 𝜋𝑖subscript𝜃𝑚Ψ𝜃𝜋𝑖1subscript𝜃𝑚subscript𝑎2Ψsubscript𝑎1subscript𝜃𝑚Ψ𝜃subscript𝜃𝑚for 𝜋𝑖subscript𝜃𝑚Ψ𝜃𝜋𝑖subscript𝜃𝑚a(\Psi)=\begin{cases}a_{1}(\Psi)=\frac{\bar{a}}{1+\xi}\left[1+\xi\cos\left(k(% \Psi-\theta)\right)\right]&\text{for }\pi i+\theta_{m}\leq\Psi-\theta\leq\pi(i% +1)-\theta_{m},\\ a_{2}(\Psi)=\frac{a_{1}(\theta_{m})\cos(\Psi-\theta)}{\cos\theta_{m}}&\text{% for }\pi i-\theta_{m}<\Psi-\theta<\pi i+\theta_{m}.\end{cases}italic_a ( roman_Ψ ) = { start_ROW start_CELL italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( roman_Ψ ) = divide start_ARG over¯ start_ARG italic_a end_ARG end_ARG start_ARG 1 + italic_ξ end_ARG [ 1 + italic_ξ roman_cos ( italic_k ( roman_Ψ - italic_θ ) ) ] end_CELL start_CELL for italic_π italic_i + italic_θ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ≤ roman_Ψ - italic_θ ≤ italic_π ( italic_i + 1 ) - italic_θ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( roman_Ψ ) = divide start_ARG italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) roman_cos ( roman_Ψ - italic_θ ) end_ARG start_ARG roman_cos italic_θ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG end_CELL start_CELL for italic_π italic_i - italic_θ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT < roman_Ψ - italic_θ < italic_π italic_i + italic_θ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT . end_CELL end_ROW (8)

where a¯¯𝑎\bar{a}over¯ start_ARG italic_a end_ARG is a scaling constant, ξ𝜉\xiitalic_ξ the anisotrophy strength, k𝑘kitalic_k is the anisotrophy mode, i𝑖iitalic_i is 0 and 1, and θmsubscript𝜃𝑚\theta_{m}italic_θ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT is the missing orientation under high anisotrophy [71, 72]. θ𝜃\thetaitalic_θ is the orientation that denotes the variation in neurite elongation direction [29]. We also include the driving force term Fdrivsubscript𝐹𝑑𝑟𝑖𝑣F_{driv}italic_F start_POSTSUBSCRIPT italic_d italic_r italic_i italic_v end_POSTSUBSCRIPT to couple with the orientation field [31, 73]. Then, the phase field governing equation becomes

ϕt=Mϕ[(a(Ψ)2ϕ)x(a(Ψ)a(Ψ)Ψϕy)+y(a(Ψ)a(Ψ)Ψϕx)+ϕ(1ϕ)(ϕ12+Fdriv+6H|θ|)],\frac{\partial\phi}{\partial t}=M_{\phi}\left[\bigtriangledown\cdot\left(a(% \Psi)^{2}\bigtriangledown\phi\right)-\frac{\partial}{\partial{x}}\left(a(\Psi)% \frac{\partial a(\Psi)}{\partial\Psi}\frac{\partial\phi}{\partial y}\right)+% \frac{\partial}{\partial{y}}\left(a(\Psi)\frac{\partial a(\Psi)}{\partial\Psi}% \frac{\partial\phi}{\partial x}\right)+\phi(1-\phi)\left(\phi-\frac{1}{2}+F_{% driv}+6H|\bigtriangledown\theta|\right)\right],divide start_ARG ∂ italic_ϕ end_ARG start_ARG ∂ italic_t end_ARG = italic_M start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT [ ▽ ⋅ ( italic_a ( roman_Ψ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ▽ italic_ϕ ) - divide start_ARG ∂ end_ARG start_ARG ∂ italic_x end_ARG ( italic_a ( roman_Ψ ) divide start_ARG ∂ italic_a ( roman_Ψ ) end_ARG start_ARG ∂ roman_Ψ end_ARG divide start_ARG ∂ italic_ϕ end_ARG start_ARG ∂ italic_y end_ARG ) + divide start_ARG ∂ end_ARG start_ARG ∂ italic_y end_ARG ( italic_a ( roman_Ψ ) divide start_ARG ∂ italic_a ( roman_Ψ ) end_ARG start_ARG ∂ roman_Ψ end_ARG divide start_ARG ∂ italic_ϕ end_ARG start_ARG ∂ italic_x end_ARG ) + italic_ϕ ( 1 - italic_ϕ ) ( italic_ϕ - divide start_ARG 1 end_ARG start_ARG 2 end_ARG + italic_F start_POSTSUBSCRIPT italic_d italic_r italic_i italic_v end_POSTSUBSCRIPT + 6 italic_H | ▽ italic_θ | ) ] , (9)

where Mϕsubscript𝑀italic-ϕM_{\phi}italic_M start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT denotes the mobility coefficient of the phase field equation, Fdrivsubscript𝐹𝑑𝑟𝑖𝑣F_{driv}italic_F start_POSTSUBSCRIPT italic_d italic_r italic_i italic_v end_POSTSUBSCRIPT is the driving force for the evolution of the phase field variable. H𝐻Hitalic_H is a constant value, and the term 6H|θ|6𝐻𝜃6H|\nabla\theta|6 italic_H | ∇ italic_θ | is introduced to disrupt symmetry [73].

Intracellular tubulin transport equation. Tubulin is the building block of microtubules in cells, transported to neurite tips via active transport and diffusion and necessary to support neurite elongations [74, 75]. Develo** based on 1D tubulin model [76], we can model the effect of intracellular tubulin transport as

(ϕctubu)t=δt(ϕctubu)αt(ϕctubu)βt(ϕctubu)+ϵ0|(ϕ0)|2|(ϕ0)|2𝑑Ω,\frac{\partial(\phi\,c_{tubu})}{\partial t}=\delta_{t}\bigtriangledown\cdot\>(% \phi\,\bigtriangledown c_{tubu})-\mathbf{\alpha}_{t}\cdot\bigtriangledown(\phi% \,c_{tubu})-\beta_{t}(\phi\,c_{tubu})+\epsilon_{0}\frac{|\bigtriangledown(\phi% _{0})|^{2}}{\int{|\bigtriangledown(\phi_{0})|^{2}}\,d\Omega},divide start_ARG ∂ ( italic_ϕ italic_c start_POSTSUBSCRIPT italic_t italic_u italic_b italic_u end_POSTSUBSCRIPT ) end_ARG start_ARG ∂ italic_t end_ARG = italic_δ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ▽ ⋅ ( italic_ϕ ▽ italic_c start_POSTSUBSCRIPT italic_t italic_u italic_b italic_u end_POSTSUBSCRIPT ) - italic_α start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ⋅ ▽ ( italic_ϕ italic_c start_POSTSUBSCRIPT italic_t italic_u italic_b italic_u end_POSTSUBSCRIPT ) - italic_β start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_ϕ italic_c start_POSTSUBSCRIPT italic_t italic_u italic_b italic_u end_POSTSUBSCRIPT ) + italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT divide start_ARG | ▽ ( italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ∫ | ▽ ( italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d roman_Ω end_ARG , (10)

where δtsubscript𝛿𝑡\delta_{t}italic_δ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is the rate at which tubulin diffuses, αtsubscript𝛼𝑡\alpha_{t}italic_α start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is the active transport coefficient for tubulin, βtsubscript𝛽𝑡\beta_{t}italic_β start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is the decay coefficient for tubulin, and ϵ0|(ϕ0)|2|(ϕ0)|2𝑑Ωsubscriptitalic-ϵ0superscriptsubscriptitalic-ϕ02superscriptsubscriptitalic-ϕ02differential-dΩ\epsilon_{0}\frac{|\nabla(\phi_{0})|^{2}}{\int{|\nabla(\phi_{0})|^{2}}d\Omega}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT divide start_ARG | ∇ ( italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ∫ | ∇ ( italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d roman_Ω end_ARG represents constant tubulin production. ϕ0subscriptitalic-ϕ0\phi_{0}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the initial phase field, and ϵ0subscriptitalic-ϵ0\epsilon_{0}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is coefficient for tubulin production.

Competitive tubulin consumption equation. Tubulin concentration is then used to calculate competitive tubulin consumption at neurite tips, which is a crucial factor that determines neurite outgrowth:

dLdt=rgctubusg,𝑑𝐿𝑑𝑡subscript𝑟𝑔subscript𝑐𝑡𝑢𝑏𝑢subscript𝑠𝑔\frac{dL}{dt}=r_{g}\,c_{tubu}-s_{g},divide start_ARG italic_d italic_L end_ARG start_ARG italic_d italic_t end_ARG = italic_r start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_t italic_u italic_b italic_u end_POSTSUBSCRIPT - italic_s start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT , (11)

where rgsubscript𝑟𝑔r_{g}italic_r start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT and sgsubscript𝑠𝑔s_{g}italic_s start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT are the tubulin assembly and disassembly rate [75], and dLdt𝑑𝐿𝑑𝑡\frac{dL}{dt}divide start_ARG italic_d italic_L end_ARG start_ARG italic_d italic_t end_ARG reflects the effect of dynamic tubulin consumption balance that is crucial for neurite extension [74, 77].

Synaptogenesis equation. Drawing inspiration from previous studies on heat conduction in dendritic solidification [78], we adopt the concept to simulate the diffusion of neurotrophin concentration in neurons [30]. In order to model NDD behaviors, we incorporate the effect of neurotrophin into the model. Since neurotrophin diffusion is crucial for guiding growth cone and neurite pathways during synaptogenesis progress [79], we utilize the evolution of the ϕitalic-ϕ\phiitalic_ϕ interface, denoted by ϕtitalic-ϕ𝑡\frac{\partial\phi}{\partial t}divide start_ARG ∂ italic_ϕ end_ARG start_ARG ∂ italic_t end_ARG, as the neurotrophin source generation in the synaptogenesis equation. The survival or death of neurons is influenced by the differential binding of neurotrophins to receptors such as p75NTR𝑝75𝑁𝑇𝑅p75NTRitalic_p 75 italic_N italic_T italic_R receptors, a type of transmembrane proteins located near the neuron growth cone [80, 81]. Its balance is discovered to be a critical factor in neuron survival following injury [82, 83]. Since p75NTR𝑝75𝑁𝑇𝑅p75NTRitalic_p 75 italic_N italic_T italic_R receptors are proven to be fast-diffusive monomers and bind to neurotrophin concentration [84], we introduce a degradation kp75cneursubscript𝑘𝑝75subscript𝑐𝑛𝑒𝑢𝑟k_{p75}c_{neur}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT into the source generation in the synaptogenesis equation to model the associated degradation [30]. We obtain

cneurt=Dc2cneur+(Kkp75cneur)ϕtk2cneur,subscript𝑐𝑛𝑒𝑢𝑟𝑡superscript2subscript𝐷𝑐subscript𝑐𝑛𝑒𝑢𝑟𝐾subscript𝑘𝑝75subscript𝑐𝑛𝑒𝑢𝑟italic-ϕ𝑡subscript𝑘2subscript𝑐𝑛𝑒𝑢𝑟\frac{\partial c_{neur}}{\partial t}=D_{c}\bigtriangledown^{2}c_{neur}+(K-k_{p% 75}c_{neur})\frac{\partial\phi}{\partial t}-k_{2}c_{neur},divide start_ARG ∂ italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG = italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ▽ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT + ( italic_K - italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT ) divide start_ARG ∂ italic_ϕ end_ARG start_ARG ∂ italic_t end_ARG - italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT , (12)

where cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT is the neurotrophin concentration, Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is the diffusion coefficient, and K𝐾Kitalic_K is the latent neurotrophin. kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT is the degradation rate when binding to p75NTR𝑝75𝑁𝑇𝑅p75NTRitalic_p 75 italic_N italic_T italic_R receptors, and k2cneursubscript𝑘2subscript𝑐𝑛𝑒𝑢𝑟k_{2}c_{neur}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT is the sink term for the degradation [30, 85]. Because cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT is vital in neurite outgrowths and directly supports the ϕitalic-ϕ\phiitalic_ϕ interface balance and evolution, the degradation to cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT introduced by kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT and k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT can lead to abnormal neurite morphological transformations associated with NDDs.

Driving force equation. In neurons, neurotrophin concentrations are a determining factor in neuronal survival. An optimal amount of neurotrophin is necessary for the survival of neurons [86]. Lower levels typically support neuron survival, while higher concentrations can negate this effect. This inverse relationship between neurotrophin concentrations and cell survival is particularly notable at higher concentrations, where an observed survival-promoting effect at lower concentrations is reversed [87]. To model the aforementioned inverse relationship between neurotrophin concentrations and cell survival considering the intricate balance of neurotrophin concentration [88], we introduce coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT to the driving force equation and obtain

Fdriv=απtan1[Hϵ(dLdt)γ(copticneur)],subscript𝐹𝑑𝑟𝑖𝑣𝛼𝜋superscript1subscript𝐻italic-ϵ𝑑𝐿𝑑𝑡𝛾subscript𝑐𝑜𝑝𝑡𝑖subscript𝑐𝑛𝑒𝑢𝑟F_{driv}=\frac{\alpha}{\pi}\tan^{-1}\left[H_{\epsilon}\left(\frac{dL}{dt}% \right)\gamma(c_{opti}-c_{neur})\right],italic_F start_POSTSUBSCRIPT italic_d italic_r italic_i italic_v end_POSTSUBSCRIPT = divide start_ARG italic_α end_ARG start_ARG italic_π end_ARG roman_tan start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [ italic_H start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT ( divide start_ARG italic_d italic_L end_ARG start_ARG italic_d italic_t end_ARG ) italic_γ ( italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT - italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT ) ] , (13)

where απ𝛼𝜋\frac{\alpha}{\pi}divide start_ARG italic_α end_ARG start_ARG italic_π end_ARG is a scaling coefficient, Hϵsubscript𝐻italic-ϵH_{\epsilon}italic_H start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT is a Heaviside step function, and γ𝛾\gammaitalic_γ is the interfacial energy constant. By introducing coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT, Eqn. 13 can effectively model the inverse relationship between cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT and neurite survival. When the magnitude of cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT exceeds coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT, the effect of cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT is reversed, leading to the ϕitalic-ϕ\phiitalic_ϕ interface retractions.

In the implementation, we use IGA to solve the phase field governing equation (Eqn. 9) concurrently with the intracellular tubulin transport equation (Eqn. 10) and the synaptogenesis equation (Eqn. 12), coupled through the driving force equation (Eqn. 13) with competitive tubulin consumption equation (Eqn. 11).

4 Truncated T-splines and local refinement

Since the ϕitalic-ϕ\phiitalic_ϕ domain is mostly stable with values consistently at “0” and “1” except at the evolving interface, we utilize truncated T-splines to apply local refinements at ϕitalic-ϕ\phiitalic_ϕ interface within the IGA framework to accelerate the computation. This approach prioritizes computational efforts on the areas undergoing evolutions over iterations, maintaining high accuracy while avoiding extensive mesh refinement by substantially decreasing the number of DOFs. This leads to quicker solver iterations and increased computational efficiency.

4.1 Review of truncated T-spline

T-spline is developed to lift the limitations of the uniform control grid B-splines [42]. B-splines rely on a uniform, rectangular grid, which has limited capability when modeling intricate geometries and often requires more control points. A univariate B-spline of order p𝑝pitalic_p is defined on a non-decreasing sequence of real numbers, uisubscript𝑢𝑖u_{i}italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, which are used to build a knot vector U¯={u1,u2,un+p+1}¯𝑈subscript𝑢1subscript𝑢2subscript𝑢𝑛𝑝1\bar{U}=\{u_{1},u_{2},\cdots u_{n+p+1}\}over¯ start_ARG italic_U end_ARG = { italic_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_u start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , ⋯ italic_u start_POSTSUBSCRIPT italic_n + italic_p + 1 end_POSTSUBSCRIPT }, where n𝑛nitalic_n is the number of basis functions and p𝑝pitalic_p is the order of the B-spline [37]. The basis function Ni,p(u)subscript𝑁𝑖𝑝𝑢N_{i,p}(u)italic_N start_POSTSUBSCRIPT italic_i , italic_p end_POSTSUBSCRIPT ( italic_u ) is defined recursively as follows:

Ni,0(u)={1if uiu<ui+1,0otherwise,subscript𝑁𝑖0𝑢cases1if subscript𝑢𝑖𝑢subscript𝑢𝑖10otherwiseN_{i,0}(u)=\begin{cases}1&\text{if }u_{i}\leq u<u_{i+1},\\ 0&\text{otherwise},\end{cases}italic_N start_POSTSUBSCRIPT italic_i , 0 end_POSTSUBSCRIPT ( italic_u ) = { start_ROW start_CELL 1 end_CELL start_CELL if italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≤ italic_u < italic_u start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL otherwise , end_CELL end_ROW (14)
Ni,p(u)=uuiui+puiNi,p1(u)+ui+p+1uui+p+1ui+1Ni+1,p1(u),subscript𝑁𝑖𝑝𝑢𝑢subscript𝑢𝑖subscript𝑢𝑖𝑝subscript𝑢𝑖subscript𝑁𝑖𝑝1𝑢subscript𝑢𝑖𝑝1𝑢subscript𝑢𝑖𝑝1subscript𝑢𝑖1subscript𝑁𝑖1𝑝1𝑢N_{i,p}(u)=\frac{u-u_{i}}{u_{i+p}-u_{i}}N_{i,p-1}(u)+\frac{u_{i+p+1}-u}{u_{i+p% +1}-u_{i+1}}N_{i+1,p-1}(u),italic_N start_POSTSUBSCRIPT italic_i , italic_p end_POSTSUBSCRIPT ( italic_u ) = divide start_ARG italic_u - italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_u start_POSTSUBSCRIPT italic_i + italic_p end_POSTSUBSCRIPT - italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_N start_POSTSUBSCRIPT italic_i , italic_p - 1 end_POSTSUBSCRIPT ( italic_u ) + divide start_ARG italic_u start_POSTSUBSCRIPT italic_i + italic_p + 1 end_POSTSUBSCRIPT - italic_u end_ARG start_ARG italic_u start_POSTSUBSCRIPT italic_i + italic_p + 1 end_POSTSUBSCRIPT - italic_u start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT end_ARG italic_N start_POSTSUBSCRIPT italic_i + 1 , italic_p - 1 end_POSTSUBSCRIPT ( italic_u ) , (15)

where Ni,0(u)subscript𝑁𝑖0𝑢N_{i,0}(u)italic_N start_POSTSUBSCRIPT italic_i , 0 end_POSTSUBSCRIPT ( italic_u ) is a piece-wise linear basis function, Ni,p(u)subscript𝑁𝑖𝑝𝑢N_{i,p}(u)italic_N start_POSTSUBSCRIPT italic_i , italic_p end_POSTSUBSCRIPT ( italic_u ) is the B-spline basis function of degree p𝑝pitalic_p, recursively calculated based on combinations of two degree (p1)𝑝1(p-1)( italic_p - 1 ) basis functions over the knot vector U¯¯𝑈\bar{U}over¯ start_ARG italic_U end_ARG. Given a set of control points P={Pi}i=1n𝑃superscriptsubscriptsubscript𝑃𝑖𝑖1𝑛P={\{P_{i}\}}_{i=1}^{n}italic_P = { italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT, the B-spline curve C(u)𝐶𝑢C(u)italic_C ( italic_u ) is defined as:

C(u)=i=0nNi,p(u)Pi,0u1.formulae-sequence𝐶𝑢superscriptsubscript𝑖0𝑛subscript𝑁𝑖𝑝𝑢subscript𝑃𝑖0𝑢1C(u)=\sum_{i=0}^{n}N_{i,p}(u)P_{i},\quad 0\leq u\leq 1.italic_C ( italic_u ) = ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_i , italic_p end_POSTSUBSCRIPT ( italic_u ) italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , 0 ≤ italic_u ≤ 1 . (16)

We can obtain a T-spline control mesh by introducing T-junctions into a quadrilateral mesh in the physical domain. Each edge in the mesh is assigned a parametric value, known as the knot interval, that requires an equal cumulative sum across the opposite edge of the element to ensure continuity and uniform parameterization across the mesh. T-spline basis functions differ from B-splines in that they are defined using local knot vectors, determined by shooting rays across the T-mesh in the parametric directions [63]. A ray is extended to intersect with the mesh to determine the local knot vector for a particular point. The intersection coordinates are collected and arranged in ascending order to form the local knot vector. This allows the T-spline basis functions to be directly influenced by the local topology of the T-mesh, enhancing their flexibility and ability to accurately represent geometry with less restriction than the global knot vectors used in B-splines. T-splines enhance these capabilities by supporting local refinement via knot insertion [89]. Given a univariate T-spline basis function NU(u)subscript𝑁𝑈𝑢N_{U}(u)italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ( italic_u ) based on an initial knot vector U={u1,u2,u3,u4,u5}𝑈subscript𝑢1subscript𝑢2subscript𝑢3subscript𝑢4subscript𝑢5U=\{u_{1},u_{2},u_{3},u_{4},u_{5}\}italic_U = { italic_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_u start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_u start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_u start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , italic_u start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT }, we can extend U𝑈Uitalic_U by adding n𝑛nitalic_n knots, where n+𝑛superscriptn\in\mathbb{Z}^{+}italic_n ∈ blackboard_Z start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, to obtain an enlarged and refined knot vector U={u1=u1,u2,,un+5=u5}Usuperscript𝑈formulae-sequencesubscriptsuperscript𝑢1subscript𝑢1subscriptsuperscript𝑢2subscriptsuperscript𝑢𝑛5subscript𝑢5superset-of𝑈U^{\prime}=\{u^{\prime}_{1}=u_{1},u^{\prime}_{2},\ldots,u^{\prime}_{n+5}=u_{5}% \}\supset Uitalic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = { italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n + 5 end_POSTSUBSCRIPT = italic_u start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT } ⊃ italic_U. Here Ui={ui,ui+1,,ui+4}(i=1,,n+1)subscriptsuperscript𝑈𝑖subscriptsuperscript𝑢𝑖subscriptsuperscript𝑢𝑖1subscriptsuperscript𝑢𝑖4𝑖1𝑛1U^{\prime}_{i}=\{u^{\prime}_{i},u^{\prime}_{i+1},\ldots,u^{\prime}_{i+4}\}% \leavevmode\nobreak\ (i=1,\ldots,n+1)italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = { italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT , … , italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 4 end_POSTSUBSCRIPT } ( italic_i = 1 , … , italic_n + 1 ) are enriched local knot vectors. The locally refined basis function NU(u)subscript𝑁superscript𝑈𝑢N_{U^{\prime}}(u)italic_N start_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_u ) is then defined as:

NU(u)=i=1n+1ciNUi(u),subscript𝑁superscript𝑈𝑢superscriptsubscript𝑖1𝑛1subscript𝑐𝑖subscript𝑁subscriptsuperscript𝑈𝑖𝑢N_{U^{\prime}}(u)=\sum_{i=1}^{n+1}c_{i}N_{U^{\prime}_{i}}(u),italic_N start_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_u ) = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_u ) , (17)

where cisubscript𝑐𝑖c_{i}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT as the refinement coefficients from knot insertion [89], and NUisubscript𝑁subscriptsuperscript𝑈𝑖N_{U^{\prime}_{i}}italic_N start_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT are the active children of NU(u)subscript𝑁superscript𝑈𝑢N_{U^{\prime}}(u)italic_N start_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_u ). A bi-variate T-spline basis function, B(u,v)=NU(u)NV(v)𝐵𝑢𝑣subscript𝑁𝑈𝑢subscript𝑁𝑉𝑣B(u,v)=N_{U}(u)N_{V}(v)italic_B ( italic_u , italic_v ) = italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ( italic_u ) italic_N start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT ( italic_v ), is the tensor product of two uni-variate functions, and its refinement is achieved by refining NU(u)subscript𝑁𝑈𝑢N_{U}(u)italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ( italic_u ) and NV(v)subscript𝑁𝑉𝑣N_{V}(v)italic_N start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT ( italic_v ) [63].

While locally refined T-splines offer significant flexibility for geometric modeling, they can pose challenges for analysis due to difficulties in maintaining linear independence due to overlap** basis functions across different refinement levels. Analysis-suitable T-splines were introduced to address these problems with standard T-splines, such as ensuring linear independence and maintaining the partition of unity [44, 90] by limiting intersections except under specific conditions that ensure analysis suitability. Truncation is an approach that has been applied to T-splines to enhance their robustness and flexibility by allowing certain intersections, except for face-face types [63]. This approach, initially implemented in truncated hierarchical B-splines [91, 92], has been extended to truncated hierarchical Catmull-Clark subdivision [53, 54]. To preserve the partition of unity and geometric integrity, the truncation mechanism mitigates overlap** basis functions by selectively excluding redundant contributions from active children functions in the refinement hierarchy. For a T-mesh T𝑇Titalic_T and its refined counterpart Tsuperscript𝑇T^{\prime}italic_T start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT from truncated T-spline quadtree subdivision, a bi-variate partially refined basis function Bi(u,v)subscript𝐵𝑖𝑢𝑣B_{i}(u,v)italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_u , italic_v ) needs to exclude redundant children Bj(u,v)superscriptsubscript𝐵𝑗𝑢𝑣B_{j}^{\prime}(u,v)italic_B start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_u , italic_v ) to avoid overlap** influences of basis functions [63]. The truncated basis function, denoted as trunBi(u,v)𝑡𝑟𝑢𝑛subscript𝐵𝑖𝑢𝑣trunB_{i}(u,v)italic_t italic_r italic_u italic_n italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_u , italic_v ), is defined as

trunBi(u,v)=Bi(u,v)ji,BjBicijBj(u,v),𝑡𝑟𝑢𝑛subscript𝐵𝑖𝑢𝑣subscript𝐵𝑖𝑢𝑣subscriptformulae-sequence𝑗𝑖superscriptsubscript𝐵𝑗subscript𝐵𝑖subscript𝑐𝑖𝑗superscriptsubscript𝐵𝑗𝑢𝑣trunB_{i}(u,v)=B_{i}(u,v)-\sum_{j\neq i,B_{j}^{\prime}\in B_{i}}c_{ij}B_{j}^{% \prime}(u,v),italic_t italic_r italic_u italic_n italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_u , italic_v ) = italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_u , italic_v ) - ∑ start_POSTSUBSCRIPT italic_j ≠ italic_i , italic_B start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_B start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_u , italic_v ) , (18)

where cijsubscript𝑐𝑖𝑗c_{ij}italic_c start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT are coefficients determined in the knot insertion process, Bi(u,v)subscript𝐵𝑖𝑢𝑣B_{i}(u,v)italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_u , italic_v ) is the parent basis functions, and Bj(u,v)superscriptsubscript𝐵𝑗𝑢𝑣B_{j}^{\prime}(u,v)italic_B start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_u , italic_v ) represents the set of partially refined children basis functions being discarded. This ensures that trunBi(u,v)𝑡𝑟𝑢𝑛subscript𝐵𝑖𝑢𝑣trunB_{i}(u,v)italic_t italic_r italic_u italic_n italic_B start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_u , italic_v ) forms a partition of unity while preserving local refinement effects. The truncated T-spline surface, S(u,v)𝑆𝑢𝑣S(u,v)italic_S ( italic_u , italic_v ), is defined as

S(u,v)=i=0ntrunBij(u,v)Piji=0ntrunBij(u,v),𝑆𝑢𝑣superscriptsubscript𝑖0𝑛𝑡𝑟𝑢𝑛subscript𝐵𝑖𝑗𝑢𝑣subscript𝑃𝑖𝑗superscriptsubscript𝑖0𝑛𝑡𝑟𝑢𝑛subscript𝐵𝑖𝑗𝑢𝑣S(u,v)=\frac{\sum_{i=0}^{n}trunB_{ij}(u,v)P_{ij}}{\sum_{i=0}^{n}trunB_{ij}(u,v% )},italic_S ( italic_u , italic_v ) = divide start_ARG ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_t italic_r italic_u italic_n italic_B start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( italic_u , italic_v ) italic_P start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_t italic_r italic_u italic_n italic_B start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( italic_u , italic_v ) end_ARG , (19)

where Pijsubscript𝑃𝑖𝑗P_{ij}italic_P start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT are the control points and trunBij(u,v)𝑡𝑟𝑢𝑛subscript𝐵𝑖𝑗𝑢𝑣trunB_{ij}(u,v)italic_t italic_r italic_u italic_n italic_B start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( italic_u , italic_v ) are the truncated T-spline basis functions. Note that trunBij(u,v)𝑡𝑟𝑢𝑛subscript𝐵𝑖𝑗𝑢𝑣trunB_{ij}(u,v)italic_t italic_r italic_u italic_n italic_B start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( italic_u , italic_v ) at any particular point sums up to 1 to ensure the property of partition of unity [93], which makes truncated T-splines particularly effective for precise computational analysis with complex geometries. We refer readers to [63] for an in-depth discussion on truncated T-splines.

4.2 ϕitalic-ϕ\phiitalic_ϕ interface-based local refinements

Refer to caption
Figure 2: Local refinement using truncated T-splines. (A) Phase field variable ϕitalic-ϕ\phiitalic_ϕ initialization on a coarse mesh. (B) Identifying elements at the ϕitalic-ϕ\phiitalic_ϕ interface. (C) Interface elements are locally refined with subdivision. (D) Detecting face-face intersections created by face extension from subdivisions. (E) Applying 1-ring of bisections to resolve these face-face intersections and maintaining the integrity of the mesh for truncated T-splines. (F) Generating locally refined T-mesh for truncated T-splines with enhanced precision in regions of interest. (G) Interpolating ϕitalic-ϕ\phiitalic_ϕ initialization from coarse control mesh to the locally-refined mesh. (H) The locally refined ϕitalic-ϕ\phiitalic_ϕ initialization.

To solve the phase field model, ϕitalic-ϕ\phiitalic_ϕ initialization is needed. To apply local refinements, the model first needs to know ϕitalic-ϕ\phiitalic_ϕ initialization on a uniform coarse control mesh (Fig. 2A), refine elements at ϕitalic-ϕ\phiitalic_ϕ interface, and then obtain ϕitalic-ϕ\phiitalic_ϕ initialization on the locally refined mesh. Beginning with a uniform coarse control mesh, we can calculate ϕitalic-ϕ\phiitalic_ϕ initialization based on the initial cell center and cell radius r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Then, by selecting elements along ϕitalic-ϕ\phiitalic_ϕ interface (Fig. 2B), we can refine and enhance the effective mesh resolution at these critical locations of interest (Fig. 2C) while kee** the mesh coarse in areas where ϕitalic-ϕ\phiitalic_ϕ values are stable at either 0 or 1. This targeted refinement approach ensures efficient use of computational resources. Because a valid T-mesh in truncated T-splines must be strongly balanced and free of face-face intersections to satisfy linear independence [63, 90], we need to resolve these face-face intersections in the automated local refinement process (Fig. 2D). To address this issue, we apply bisections to 1-ring of elements surrounding the selected interface elements (Fig. 2E). This approach effectively resolves face-face intersections and allows the model to refine the mesh only where needed. Thus, we can obtain a locally-refined T-mesh for truncated T-splines tailored explicitly to the given ϕitalic-ϕ\phiitalic_ϕ value (Fig. 2F). Subsequently, ϕitalic-ϕ\phiitalic_ϕ values are interpolated and transferred onto this newly refined mesh using the KD tree-based interpolation method [94], providing accurate initialization for the computation of the evolving neuron outgrowth ϕitalic-ϕ\phiitalic_ϕ (Fig. 2G). Starting with an existing control point, a search using the KD tree data structure finds the nearest new control point. If a matching control point is found, the value at the input control point is directly used. If the new control point is situated among the existing points, the algorithm interpolates depending on their positions and connectivity in the control mesh (Fig. 2G). This conditional handling ensures values are correctly passed to the locally refined mesh. The KD tree approach effectively utilizes spatial indexing and positions, preserving the initialization accuracy while minimizing unnecessary computations.

5 Dynamic domain expansion

During neuron growth, neurites initiate along soma boundaries and then extend outwards toward the domain boundary. If using conventionally fixed domain size, this nature of the neuron growth process requires excessive DOFs in largely static areas where ϕitalic-ϕ\phiitalic_ϕ remains “0”, away from the center. To address this issue, the NDDs model incorporates dynamic domain expansion to accommodate the complex, ever-expanding neurite elongation process, followed by a KD tree-based interpolation to ensure fast and accurate variable pass-through [94]. This method reduces redundant DOFs associated with neurite extension, enabling the simulation of complex, high-resolution neuron outgrowth and enhancing both accuracy and computational efficiency.

The model incorporates a directional domain expansion algorithm (Algorithm 1) to dynamically adjust the control mesh in response to neurite approaching the domain boundary (double-sided arrows in Fig. 3A-D). The algorithm takes neuron growth ϕitalic-ϕ\phiitalic_ϕ and control mesh as the input and evaluates each element adjacent to the domain boundary, searching for any nonzero values of the ϕitalic-ϕ\phiitalic_ϕ. If nonzero ϕitalic-ϕ\phiitalic_ϕ is detected near the boundary, the algorithm flags the corresponding edge for expansion using expFlag𝑒𝑥𝑝𝐹𝑙𝑎𝑔expFlagitalic_e italic_x italic_p italic_F italic_l italic_a italic_g. Once boundaries are flagged, the algorithm modifies the domain size based on the predetermined expansion size (default 3×Δx3Δ𝑥3\times\Delta x3 × roman_Δ italic_x, where ΔxΔ𝑥\Delta xroman_Δ italic_x is coarse element size). It generates an enlarged control mesh (dashed zone in Figure 3A-D) and translates the entire control mesh according to the specified direction of expansion to maintain the position of the neuron ϕitalic-ϕ\phiitalic_ϕ. This directional domain expansion approach works in tandem with the truncated T-splines generation, discussed in Section 4, to ensure that the computational domain adapts to accommodate growing neurites with the least amount of elements, maintaining analysis accuracy and efficiency. The transition to the expanded mesh requires an interpolation operation that passes values from the original control points to the new control points (Fig. 2G). The interpolation process for a dynamically expanding mesh in the NDDs simulation leverages the KD tree algorithm to optimize spatial queries of control points cpts𝑐𝑝𝑡𝑠cptsitalic_c italic_p italic_t italic_s, significantly accelerating the search for the nearest cpts𝑐𝑝𝑡𝑠cptsitalic_c italic_p italic_t italic_s [94]. Starting with new control points cpts𝑐𝑝𝑡superscript𝑠cpts^{\prime}italic_c italic_p italic_t italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, each cpts𝑐𝑝𝑡superscript𝑠cpts^{\prime}italic_c italic_p italic_t italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT undergoes a rounding process to identify whether it is located on expanded elements. This procedure significantly speeds up the algorithm by skip** elements that are not expanded, and reduces the total number of DOFs while preserving the analysis accuracy.

Algorithm 1 Dynamic Domain Expansion (Figure 3)

Input: Control points cpts𝑐𝑝𝑡𝑠cptsitalic_c italic_p italic_t italic_s, control mesh Mesh𝑀𝑒𝑠Meshitalic_M italic_e italic_s italic_h, neuron outgrowth ϕitalic-ϕ\phiitalic_ϕ
Output: Expanded control points cpts𝑐𝑝𝑡superscript𝑠cpts^{\prime}italic_c italic_p italic_t italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, Expanded control mesh Mesh𝑀𝑒𝑠superscriptMesh^{\prime}italic_M italic_e italic_s italic_h start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT

1:Check for potential expansion direction based on ϕitalic-ϕ\phiitalic_ϕ value
2:Initialize expansion flag expFlag𝑒𝑥𝑝𝐹𝑙𝑎𝑔expFlagitalic_e italic_x italic_p italic_F italic_l italic_a italic_g for each element e𝑒eitalic_e of the domain.
3:for each element e𝑒eitalic_e in Mesh𝑀𝑒𝑠Meshitalic_M italic_e italic_s italic_h adjacent to the domain boundary do
4:     if ϕ(x,y)>0italic-ϕ𝑥𝑦0\phi(x,y)>0italic_ϕ ( italic_x , italic_y ) > 0 then
5:         expFlag[e]=true𝑒𝑥𝑝𝐹𝑙𝑎𝑔delimited-[]𝑒trueexpFlag[e]=\text{true}italic_e italic_x italic_p italic_F italic_l italic_a italic_g [ italic_e ] = true
6:         break \triangleright Terminate search if expansion is necessary
7:     end if
8:end for
9:Expanding domain based on expansion parameters
10:if any expFlag[e]𝑒𝑥𝑝𝐹𝑙𝑎𝑔delimited-[]𝑒expFlag[e]italic_e italic_x italic_p italic_F italic_l italic_a italic_g [ italic_e ] is true then \triangleright Expand mesh where elements are flaggeed
11:     Determine direxp𝑑𝑖subscript𝑟𝑒𝑥𝑝dir_{exp}italic_d italic_i italic_r start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT and szexp𝑠subscript𝑧𝑒𝑥𝑝sz_{exp}italic_s italic_z start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT based on flagged elements
12:     Create enlarged control mesh Mesh𝑀𝑒𝑠superscriptMesh^{\prime}italic_M italic_e italic_s italic_h start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT with new control points cpts𝑐𝑝𝑡superscript𝑠cpts^{\prime}italic_c italic_p italic_t italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT
13:     Translate cpts𝑐𝑝𝑡superscript𝑠cpts^{\prime}italic_c italic_p italic_t italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT based on direxp𝑑𝑖subscript𝑟𝑒𝑥𝑝dir_{exp}italic_d italic_i italic_r start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT \triangleright Offset mesh to keep ϕitalic-ϕ\phiitalic_ϕ position consistant
14:else
15:     Set cpts=cpts𝑐𝑝𝑡superscript𝑠𝑐𝑝𝑡𝑠cpts^{\prime}=cptsitalic_c italic_p italic_t italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_c italic_p italic_t italic_s and Mesh=Mesh𝑀𝑒𝑠superscript𝑀𝑒𝑠Mesh^{\prime}=Meshitalic_M italic_e italic_s italic_h start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_M italic_e italic_s italic_h \triangleright No expansion needed
16:end if
17:return cpts𝑐𝑝𝑡superscript𝑠cpts^{\prime}italic_c italic_p italic_t italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT and Mesh𝑀𝑒𝑠superscriptMesh^{\prime}italic_M italic_e italic_s italic_h start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT
Refer to caption
Figure 3: Sequential representation of directional domain expansion in neuron outgrowth simulation. (A) The initial soma ϕitalic-ϕ\phiitalic_ϕ and checking for neurites near the domain boundary. (B) Detected neurite near the bottom boundary and subsequent bottom directional domain expansion. (C) As growth continues, neurites approach the top and left boundaries, and the domain is expanded along these directions. (D) Neurites detected near the top, left, and right boundaries, and domain expansions follow. (E) The expanded computational domain for neurite development. The process showcasing the model can adapt to the evolving morphology of the neuron and minimize unnecessary computational costs.

6 Experimental human iPSC-derived and rat hippocampal neuron cultures

In this section, we outline the protocols and details for culturing human induced pluripotent stem cells (iPSCs) and rat hippocampal neuron cultures. Conducting experimental cultures to obtain neuron growth observations is necessary for computational modeling as it provides a reference to simulations and for comparison with realistic results. This approach allows us to better understand the intricacy of neurite outgrowth, improving the ability of the NDD model to reflect real-world biological behaviors in NDD studies. Human iPSC-derived neurons are selected to better correlate between the NDDs model and actual human neuron growth. Building upon our previous studies of healthy rat hippocampal neurons [32, 68], we now extend our focus to unhealthy rat hippocampal neurons to deepen our understanding of NDDs. In future work, we intend to include unhealthy human iPSC-derived neurons currently being developed at the Mayo Clinic to broaden our investigation of underlying disorder mechanisms.

Human iPSC-derived neuron culture. Human iPSCs, derived from a healthy 83-year-old female (MC0192), were cultured in TeSR-E7 complete medium on Matrigel-coated plates, with karyoty** conducted by the Mayo Clinic Genomics Core Facility [95]. The iPSCs were differentiated into neural precursor cells (NPCs) using the STEMdiff SMADi Neural Induction Kit as previously described [96]. After initial culturing in 24-well AggreWell800 plates, embryoid bodies were transferred to Matrigel-coated 6-well plates on day 8 to promote neural rosette formation for another 5 days. These NPCs were expanded, supplemented with ROCK inhibitor Y27632 on day 15, and eventually cryopreserved in Neural Progenitor Freezing Medium. NPCs were seeded on Poly-L-ornithine/Laminin-coated plates for neuronal differentiation in BrainPhys Neuronal Medium. Immunocytochemistry involved fixing the cells, staining them with primary antibodies against β𝛽\betaitalic_βIII-tubulin and secondary Alexa Fluor 488- or 568-conjugated secondary antibodies, and counterstaining nuclei with DAPI. Imaging was performed using a Keyence BZ-X800 fluorescence microscope. We refer readers to [95, 96] for detailed documentation of experimental protocols. The images of experimental human iPSC-derived neuron culture are used to compare with healthy neuron growth simulations to better analyze neurite outgrowth behaviors.

Rat hippocampal neuron cultures. For apoptotic rat hippocampal neuron culture results used in this paper, are drawn from failed cultures during protocol optimization for our previously published work creating a dataset and model of rat hippocampal neuron growth [32, 33] with slight modification. The neurons, taken from embryonic day 18 and cryopreserved (A36513, Gibco, USA), were grown in dishes coated with poly-D-lysine (P6407, Sigma-Aldrich, USA) following the manufacturer’s guidelines [97]. The seeding density was 10,000 cells per square centimeter in Neurobasal Plus medium (A3582901, Gibco, USA) supplemented with 2% B-27 Plus (A3582801, Gibco, USA). Cultures were maintained at 37 degrees Celsius with 5% CO2. It is common when establishing new neural cultures to optimize media formulation, media change frequency, and the fraction of media replaced during media changes (ADD REF). During this optimization process several cultures underwent apoptosis with neurite beading. Imaging was carried out using an Echo Revolve Microscope (Echo Revolve |||| R4, inverted, BICO, USA) with a 12-megapixel color camera at 20X and 40X magnifications and at intervals of 0.5, 1, 2, 3, 4, and 6 days post-inoculation. The images of experimental rat hippocampal neuron cultures undergoing persumed excitotoxic apoptosis are selected for qualitative comparison with neuron growth simulations affected by NDD parameters.

7 Numerical results and validation with expeirments

In this section, we first explain the implementation specifics of the proposed IGA phase field NDDs model. Then, we showcase simulation results of healthy and unhealthy neurons with several parameter studies designed to investigate NDDs. Finally, we compare our simulation outcomes with experimental data, focusing on cases with external guided-cue mechanisms. This comparative analysis aims to validate the effectiveness and accuracy of our NDDs model in simulating biomimetic neurite behaviors and disorders.

We implemented the IGA NDDs model in C++, leveraging the extensive capabilities and scalabilities of the PETSc library [61]. PETSc as the computational backbone offers a significant computational efficiency advantage over MATLAB-based implementations [31]. This is mainly because the PETSc library is written in compile language and tailored for high-performance computing environments, while MATLAB is a scripting language that faces limitations due to its inherently interpreted mode of execution, making it less efficient for tasks requiring intensive computation. Through the utilization of the MPI for parallel processing, our model facilitates efficient distribution of computations across multiple threads [62, 98], thereby cutting down execution times significantly when compared with MATLAB analysis. Although MPI leverages the immense computational power enabled by supercomputers, communication time among processors becomes a bottleneck. Mesh partitioning using METIS [64] ensures optimal load balancing for the computationally intensive simulations. This is crucial for the scalability and efficiency of the model, particularly when simulating complex neurite structures on locally refined truncated T-splines with a large number of DOFs because mesh alterations during execution often lead to unbalanced loads. Unbalanced loads across threads could lead to unnecessary slowdown and waiting during inter-threads/node communications. We ran simulations on the Bridges-2 supercomputer at Pittsburgh Supercomputing Center [99, 100] using 128-thread regular memory nodes.

Table 1: Parameters used in the NDDs model.
Parameter Description Value Parameter Description Value
coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT Optimal neurotrophin concentration [0, 1] H𝐻Hitalic_H Orientation constant coefficient 0.0070.0070.0070.007
Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT Neurotrophin diffusion coefficient [0, 6] δtsubscript𝛿𝑡\delta_{t}italic_δ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT Tubulin diffusion rate 4444 (μm2/h𝜇superscript𝑚2\mu m^{2}/hitalic_μ italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_h)
kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT Neurotrophin binding rate [0, 3] αtsubscript𝛼𝑡\alpha_{t}italic_α start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT Tubulin active transport rate 0.0010.0010.0010.001 (μm/h𝜇𝑚\mu m/hitalic_μ italic_m / italic_h)
k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT Neurotrophin degradation rate [0, 3] βtsubscript𝛽𝑡\beta_{t}italic_β start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT Tubulin decay coefficient 0.0010.0010.0010.001 (1/h11/h1 / italic_h)
γ𝛾\gammaitalic_γ Phase field interface driving force constant [0, 10] ϵ0subscriptitalic-ϵ0\epsilon_{0}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT Tubulin production coefficient 15151515
K𝐾Kitalic_K Dimensionless latent neurotrophin [0, 2] r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT Initial cell radius 15x15𝑥15\triangle x15 △ italic_x
Mϕsubscript𝑀italic-ϕM_{\phi}italic_M start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT Mobility coefficient 60606060 rgsubscript𝑟𝑔r_{g}italic_r start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT Tubulin assembly rate 5555
a¯¯𝑎\bar{a}over¯ start_ARG italic_a end_ARG Surface energy scaling constant 0.04 sgsubscript𝑠𝑔s_{g}italic_s start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT Tubulin disassembly rate 0.10.10.10.1
ξ𝜉\xiitalic_ξ Anisotropy strength 0.2 απ𝛼𝜋\frac{\alpha}{\pi}divide start_ARG italic_α end_ARG start_ARG italic_π end_ARG Scaling coefficient 0.28650.28650.28650.2865
k𝑘kitalic_k Anisotropy mode 6 θ𝜃\thetaitalic_θ Neurite growth orientation angle [0,1]01[0,1][ 0 , 1 ]
Note: Variables shaded in gray contribute to NDDs morphological transformation. For coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT, higher levels cause excessive branching
and lower values lead to retractions. For Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, high levels initiate neurite formation and low levels lead to atrophy. kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT increases neurite
thickness as values rise. k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT increases neurite thickness as values rise. γ𝛾\gammaitalic_γ maintains a constant effect on interface stability and scales the
effect of cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT. Higher K𝐾Kitalic_K values enhance neurite thickness and branching. Parameters require initializations are shown with their default
value. Dimensionless parameters are listed with default values without units. ΔxΔ𝑥\Delta xroman_Δ italic_x is the coarse element size

For clarity, we provide a detailed list of variables for NDDs and healthy neurons in Table 1. The domain ϕitalic-ϕ\phiitalic_ϕ is initialized with a central filled circle with radius r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, representing the cell. The initial values of θ𝜃\thetaitalic_θ are randomized to between [0,1]01[0,1][ 0 , 1 ] across the domain with T𝑇Titalic_T set to 0 at the beginning. Due to the significant variability in biophysics and growth behaviors across different types of neurons, the parameter values used in the phase field NDDs model were sourced from established literature [29, 30, 101] and fine-tuned empirically to capture realistic neuron growth. These parameters can be adjusted to better simulate biomimetic growth of specific neuron types  [32]. With these parameter settings, we first simulate healthy neuron growth and analyze single- and multiple-neuron growth scenarios (Section 7.1). Then, we conduct a simulation of abnormal neurite morphological transformation using the NDDs model with three case studies that each focuses on a specific parameter (Section 7.2). Finally, we compare simulation results with experimental observations to validate our NDDs model (Section 7.3).

7.1 Healthy neuron growth simulation

Refer to caption
Figure 4: Healthy neuron growth simulations. (A) Single neuron growth with many neurite morphologies. (B) Multiple-neuron growth simulations with neurite interactions. For multi-neuron cases, the initial soma placements are randomized in the domain.

Our NDDs model supports simulating healthy neuron growth. We set coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT to 1 to ensure sufficient concentration level to drive interface outwards expansion, Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT to 6 to ensure sufficient neurotrophin diffusion, and kp75,k2cneursubscript𝑘𝑝75subscript𝑘2subscript𝑐𝑛𝑒𝑢𝑟k_{p75},k_{2}c_{neur}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT , italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT to 0 to eliminate degradation effects. The rest parameters for healthy neuron growth are set following Table 1. Simulations are done with both single- and multiple-neuron configurations up to 350,000 iterations, and results near the final iteration that best illustrate the growth behaviors are shown (Fig. 4). The single-neuron scenarios aim to study individual neuron outgrowth behaviors, while the multiple-neuron setups focus on understanding neurite interactions that contribute to the formation of complex neurite networks.

In our results, the neuron, represented by ϕitalic-ϕ\phiitalic_ϕ, is shown in yellow, and the surrounding medium is shown in blue. For single neuron cases, the results (Fig. 4A) show that the model captures a broad spectrum of neurite outgrowth behaviors, from axon elongations to complex branching. Concurrently, our multiple-neuron simulations investigate interactions within complex neurite network formation, varying from 2-7 neurons (Fig. 4B). Starting from 2-neuron cases, we gradually increase complexity to configurations with up to 7 randomly placed neurons. These multi-neuron simulations showcase individual neurite growth patterns and the intricate neurite interactions with responses of other neurons in proximity. With the detailed dynamics of ctubusubscript𝑐𝑡𝑢𝑏𝑢c_{tubu}italic_c start_POSTSUBSCRIPT italic_t italic_u italic_b italic_u end_POSTSUBSCRIPT and cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT, these biomimetic single and multi-neuron simulation results showcase the potential of our model for in-depth neurodevelopmental research and provide a solid foundation for exploring NDDs.

7.2 Neurodevelopmental disorders

Extending from healthy neuron growth simulations, we conduct a comprehensive study (Figs. 5-7) to investigate the functional roles of different parameters and factors affecting the onset of NDDs, with a focus on critical parameters coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT, Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT, and k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT as detailed in Table 1. In this section, we categorize the parameters into three studies based on their effects on neurite morphological transformations:

  1. 1.

    Optimal neurotrophin concentration coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT. This parameter models the inverse relationship between neurotrophin concentration level and neurite survival and is analyzed for its impact mostly on neurite retraction;

  2. 2.

    Neurotrophin diffusion rate Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. This parameter is responsible for diffusing the concentration necessary for neurite outgrowths and predominantly triggers neurite atrophy; and

  3. 3.

    Degradation rates kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT and k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. These parameters are investigated for their effects on neurite thickness. Very thin neurites lead to atrophy.

For each study, the parameter being analyzed was adjusted based on the specified ranges listed in Table 1. All other parameters were set according to their default values listed in the same table. By adjusting these parameters in our NDD simulations, we can not only enhance our understanding of the pathophysiological mechanisms of NDDs but also examine the capability of our model to capture their intricate dynamics, which is critical for develo** targeted interventions and advancing neurodevelopmental research.

Refer to caption
Figure 5: Impact of coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT variations on neurite morphology and development. (A) Variations of neurite growth behaviors as influenced by coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT values ranging from 0.5 to 3.0 with zoomed-in views of the intricate neurite structures at around 40,000 iterations, illustrating the progressive neurite density and branching complexity changes. (B) Demonstration of the dynamic retraction behaviors in the simulated neuron growth process with an initial coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT value of 1, which is subsequently reduced to 0 at 150,000 iterations during the simulation to mimic the effect of increasing cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT magnitude and its inverse relationship on neuron survival. Atrophies are marked with cyan dashed circles, and retractions are traced with magenta dashed lines. The coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT reduction simulates neurite retraction, showcasing the potential impacts of decreasing neurotrophic support on neurite morphology and structural integrity over time.

Optimal neurotrophin concentration coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT. By adjusting the value of coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT while kee** other parameters consistent with Table 1, we can better understand its impact on neurite morphological transformations in NDDs. We first analyze neurite growth behaviors under different coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT values ranging from 0.50.50.50.5 to 3.03.03.03.0 at around 40,0004000040,00040 , 000 iterations (Fig. 5A). As coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT increases from 0.50.50.50.5 to 1.51.51.51.5, there is a noticeable improvement in neurite density and complexity, promoting more extended growth patterns. This process indicates a positive response of neurites to optimal neurotrophin levels, matching our previous expectations of enhanced growth behaviors. As coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT progresses from 1.51.51.51.5 to 3.03.03.03.0, there is a noticeable shift towards highly branched and extensively multidirectional neurite structures. This range of growth patterns, from minimal to excessive branching, shows a critical threshold level of neurotrophins necessary for neurodevelopmental processes. Notably, lower values of coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT lead to suppressed neurite outgrowth, validating that there is an inverse relationship between cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT and neurite survival [87], as well as a critical balance of cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT for maintaining healthy neurite development. We then explore the dynamic effects of varying coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT in our NDDs model to understand its impact on neurodevelopmental processes. For simulation cases in Fig. 5B, we initialize coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT as 1 to simulate healthy neurite outgrowth and then reduce coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT to 00 at 150,000150000150,000150 , 000 iterations across cases involving different numbers of neurons. The first two rows in Fig. 5B showcase the growth behavior of individual neurons, showing a straightforward process of neurite retraction as coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT decreases, highlighting the sensitivity of single neurons to cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT and coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT levels. The third and fourth rows extend this analysis to multi-neuron configurations with two and three neurons to investigate neurite retraction and interactions among multiple neurons. These simulation results highlight the intricate influence of coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT levels on the morphology transformation of neurites and the potential implications to neurite retractions and atrophy, vital for develo** effective therapeutic strategies targeting NDDs.

Refer to caption
Figure 6: Impact of neurotrophin diffusion rate Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT variations on neurite outgrowth. (A) Single- and (B) multiple-neuron NDDs simulations with different Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT values. It highlights how low values of Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, corresponding to inadequate diffusion, can lead to neurite atrophy. As the Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT value increases from 1.01.01.01.0 to 6.06.06.06.0, neurite continuity improves noticeably, showcasing the critical role of diffusion rates in supporting and maintaining neurite structures. Atrophies are marked in (A) with cyan dashed circles. Zoomed-in views in (B) provide a detailed view of atrophy behaviors.

Neurotrophin diffusion rate Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. In this study, we focus on the neurotrophin diffusion rate Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT that is responsible for controlling the diffusion of cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT to support ϕitalic-ϕ\phiitalic_ϕ interface balance. We analyze the effect of Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT with varying magnitude from 1.01.01.01.0 to 6.06.06.06.0 on single- and multiple-neuron cases, and results at around 100,000100000100,000100 , 000 iterations that best exhibit atrophy are shown in Fig. 6. The results show severe atrophy when the Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT value is below 2.02.02.02.0. As the Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT value approaches 3.03.03.03.0, atrophy becomes moderate. This atrophy behavior indicates that lower values of Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, particularly at 1.0, severely restrict the diffusion necessary for generating a neurotrophin gradient field for Eqn. 13 that stabilizes the phase field interface during neurite outgrowth. This insufficient gradient compromises neurite integrity and leads to the degeneration or atrophy of neurite structures due to inadequate neurotrophic support. As a result, as neurite tips continue growing out, there is not enough concentration to support and maintain the grown neurite structures, leading to neurite atrophy. On the other hand, as Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT increases, particularly beyond 5.05.05.05.0, a noticeable improvement in neurite continuity is observed in both single- (Fig. 6A) and multiple-neuron cases (Fig. 6B), markedly reducing the incidence of neurite atrophy. This highlights the necessity of maintaining a diffusion rate Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT to ensure a consistent concentration distribution surrounding the growing neurite, thereby supporting the uninterrupted neurodevelopmental process and potentially mitigating disorder progression in NDDs.

Refer to caption
Figure 7: Impact of variations in degradation rates kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT and k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT on neurite morphology. (A) Neurite growth results with varying kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT values. (B) Neurite growth results with varying k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT values. Both variables predominantly affect neurite thickness (marked with cyan arrows), subtly impacting the physical structure while maintaining the complexity of the neurite structures unchanged. Increasing values of these variables result in progressively thicker neurites but with diminishing effects.

Degradation rates kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT and k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. We initialize the simulation with kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT and k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT values ranging from 0.00.00.00.0 to 2.52.52.52.5, and results at around 30,0003000030,00030 , 000 iterations are shown in Fig. 7. The parameter kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT, when adjusted from 0.00.00.00.0 to 1.51.51.51.5, enhances the complexity of neurite branching, as well as from thin neurites to thicker neurites (Fig. 7A). This progression showcases the role of kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT in modulating neurotrophin generation and maintaining cneursubscript𝑐𝑛𝑒𝑢𝑟c_{neur}italic_c start_POSTSUBSCRIPT italic_n italic_e italic_u italic_r end_POSTSUBSCRIPT levels, which are vital for the healthy development of neurite structures. As increasing kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT value increases beyond 1.51.51.51.5 to 2.52.52.52.5, the results showcase a plateau in morphological complexity changes, indicating a point where further increases in kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT yield diminishing effects on morphological changes. Similarly, the impact of k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT on neurite morphology is demonstrated through its influence on increasing neurite thickness (Fig. 7B), enhancing the structures of neurites but without subtle changes to the overall branching patterns and the complexity of the neurite structures. As the k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT value rises, its effects also gradually diminish, highlighting diminishing effects similar to kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT. These results provide an interesting perspective on the degradation of neurotrophin in the pathophysiology of NDDs, which will contribute to the development of targeted therapeutic planning.

These detailed parameter studies utilizing our NDDs model demonstrate NDDs morphological transformations to specific simulation parameters. The impact of coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT is most dominantly on neurite retraction and branching, the influence of Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT leads to neurite atrophy, and the degradation rates kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT and k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT affect the thickness of neurite growth. By categorically investigating these parameters, we have identified crucial factors that predominantly influence the progression of NDDs in our computational model. Our analysis provides crucial insights into how variations in these key parameters can trigger or exacerbate the symptoms associated with NDDs. This knowledge is vital in guiding ongoing research toward understanding the pathophysiological mechanisms underlying these disorders. In the long term, this model could enable the development of more targeted therapeutic planning, focusing on the most influential factors of neuron development and treatment strategy. This approach will enhance our understanding of potential NDDs triggering mechanisms and broaden the potential reach for effective treatment strategies by incorporating the effect of parameters with biophysical processes in simulating realistic neurodevelopmental progress.

7.3 Validation with experiements

A crucial component of our NDDs model study involves comparing it with experimental observations. We apply the external-cue guided mechanism [32] to simulate the existence of external attractive cues in the extracellular matrix for several simulation cases and set the results against experimental neuron growth patterns (Fig. 8-10). By selectively placing external cues around the neuron, our model can effectively guide neurite outgrowth toward these external cues, allowing it to capture the complex and dynamic growth patterns observed in experimental cultures. These simulations explore how well our model reflects real-world biological behaviors, thereby validating its effectiveness for studying NDDs. This study compares simulation results with single neuron cultures to better understand individual neurite morphology. Although culturing denser neuron cultures is feasible, the densely overlap** neurite structures will significantly complicate neurite identifications and comparisons. Considering these limitations, efforts have been made to develop sparser neuron cultures to improve the clarity and reliability of neurite analysis.

Refer to caption
Figure 8: Comparison of simulated neuron growth patterns and experimental cultures of human iPSC-derived neurons. (A) Comparisons of single neuron cases up to day 5. (B) Comparisons of multiple-neuron cases involving two and three neurons. External cue placements are marked with cyan starts. These comparisons provide a visual validation of NDDs model to replicate key aspects of neurite morphology and developmental processes in real biological environments.

Comparison with healthy human iPSC-derived neurons. First, we validate the model by comparing single-neuron growth patterns from day 1 to day 5 (Fig. 8A). Experimental images of human iPSC-derived neurons cultured over five days are used as the reference. We incorporate the external cue-guided mechanism by placing external cues around the soma to direct neurite outgrowth. These results indicate that the NDDs model can capture the observed neurite growth patterns when external cues are placed according to experimental culture images. Next, we test the NDDs model with multiple neuron configurations (Fig. 8B). This setup involved comparing experimental images of multi-neuron growth with simulations where external cues were selectively placed to guide neurite outgrowth. The simulation outcomes closely aligned with the experimental images, demonstrating that the model can biomimetically capture complex growth behaviors in multi-neuron environments. Although measuring the exact biological conditions and translating them into parameters in the phase field model present inherent challenges, this comparison showcases the effectiveness and potential of our NDDs model in capturing essential aspects of the neurodevelopmental process, highlighting its potential as a powerful tool for studying NDDs. These results validate the biomimetic capabilities and applicability of the NDDs model in simulating realistic neurodevelopmental processes, providing valuable insights for targeted therapeutic strategies.

Comparison with rat hippocampus neurons undergoing excitotoxic apoptosis. Utilizing the same external cue placement strategy, we perform a detailed comparative analysis of two neuron growth cases with disorders by modifying Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT and compared with rat hippocampal neuron cultures undergoing excitotoxic apoptosis [32, 68] (Fig. 9). The NDDs model captures atrophy and retraction by adjusting Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, which controls the diffusion of neurotrophin concentration, and coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT, which sets the optimal neurotrophin concentration for neurodevelopmental processes. Based on experimental images of neuron growth undergoing excitotoxic apoptosis (Fig. 9A&E). Red outlines are empirically traced based on both experimental observations and accumulated expertise. External cues are placed to guide healthy neurite patterns (Fig. 9B&F). Subsequently, we decrease the magnitude of Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT from 6666 to 1111 to simulate reduced diffusion in neurotrophin concentration (Fig. 9C&G). This simulation reveals that neurites suffer significant atrophy and disconnections, showcasing a stark contrast to simulated healthy growth. In the other comparison, we lowered coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT magnitude from 1111 to 00 (Fig. 9D&H). The results show that neurites not only experience significant atrophy but also demonstrate retraction behaviors, which are influenced by the cascading energy balance at the interface. This study highlights the critical influence of Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT on pathological neuron growth affected by NDDs. These parameter adjustments show great potential to enable the NDDs model to simulate a variety of morphological abnormalities and provide insights into the dynamics of NDDs, aiding in the development of future targeted therapies.

Refer to caption
Figure 9: Comparison of experimental neuron cultures undergoing excitotoxic apoptosis and simulation results. (A&E) Experimental images of neurons exhibiting disorders are highlighted in red outlines. (B&F) Simulated healthy neurons using the NDDs model. (C&G) Simulated neuron morphology with disorders induced by reducing the diffusion rate Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT from 6 to 1. (D&H) Simulated neuron morphology with disorders induced by lowering the optimal concentration coptisubscript𝑐𝑜𝑝𝑡𝑖c_{opti}italic_c start_POSTSUBSCRIPT italic_o italic_p italic_t italic_i end_POSTSUBSCRIPT from 1 to 0.
Refer to caption
Figure 10: Additional comparison of experimental rat hippocampal neurons undergoing excitotoxic apoptosis and simulated growth patterns. (A&B) Showcases single neuron scenarios, illustrating the progression and diversity of growth patterns.

In addition to the above two in-depth comparison studies, we apply our NDDs model to simulate a range of cases following the same approach (Fig. 10). The first and third rows in Fig. 10 are the experimental images of neuron growth undergoing excitotoxic apoptosis (in red outlines). The second and fourth rows in Fig. 10 are the corresponding NDDs model simulation results of NDDs neurite growth patterns. Each simulation is able to simulate biomimetic growth behaviors under disorder conditions with atrophy and retraction, with the influence of external cues guided mechanism on neurites. These results not only demonstrate the capability of our NDDs model but also showcase its effectiveness in simulating realistic NDDs-associated neurite growth observed in experimental cultures.

8 Conclusion and future work

We present a novel computational model for NDDs, incorporating IGA, dynamic domain expansion and local refinement techniques into the phase field method. The NDDs model demonstrates an uplift in computational performance compared to existing models and provides insights into complex neurite behaviors. We conclude that:

  • 1.

    We have developed a PETSc-based NDDs model to utilize parallel processing with IGA and truncated T-splines. This approach allows for accurate simulations of neuron growth processes and disorders, significantly enhancing computational efficiency without sacrificing accuracy.

  • 2.

    We introduced optimal neurotrophin concentration and degradation of neurotrophin to the phase field model and simulated NDDs neurite morphological transformations including retraction and atrophy, capturing the inverse relationship between neurotrophin levels and neurite survival. It uncovers the significant impact of neurotrophin on NDDs.

  • 3.

    The NDDs model incorporates dynamic domain expansion and targeted local refinements at the phase field interface. This approach optimizes computational resources by expanding the domain dynamically based on the evolving neurite structures, ensuring high accuracy with reduced computational costs.

  • 4.

    We leveraged our computational NDDs model to investigate the intricate neurite morphological transformations associated with NDDs, including retraction, atrophy, branching, and thickness variations.

  • 5.

    The parameter study validated the NDDs model with an external-cue guided mechanism, by comparing it with observed experimental neuron growth patterns, demonstrating its potential to advance the NDDs research. Our study unveils the precise functional roles of related factors and parameters, offering critical insights into underlying mechanisms of NDDs.

The model accurately simulates neurite outgrowth morphologies through these novel techniques and expedites the study of NDDs. By capturing single- and multi-neuron dynamics, the model provides essential insights into the complex networks critical for understanding neurodevelopmental disorders. The parameter study reveals how the specific parameters influence neurite morphology and development. For example, how variations in parameters such as kp75subscript𝑘𝑝75k_{p75}italic_k start_POSTSUBSCRIPT italic_p 75 end_POSTSUBSCRIPT and Dcsubscript𝐷𝑐D_{c}italic_D start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT significantly alter neurite growth, affecting the thickness, branching, and survival of the neurites. These findings will aid in develo** more accurate target treatments and therapeutic planning. In addition, our NDDs model can simulate biomimetic neurite outgrowth patterns utilizing an external-cue-guided mechanism when compared with experimental observations. This places the model as a crucial and effective computational tool for studying NDDs.

Because the current implementation is limited to 2D simulations, our future endeavors include extending the NDDs model to 3D implementation utilizing truncated hierarchical B-spline [55, 102]. This improvement will facilitate a more precise representation of the complex 3D aspects of neuron morphology. We plan to conduct more experiments with our collaborators to validate the model and ensure its accuracy and reliability in capturing neurodevelopmental processes. To broaden our investigation, we intend to include unhealthy human iPSC-derived neurons currently being developed at the Mayo Clinic. Despite parallelization and optimizations, the model still requires substantial computational resources, rendering it less practical for rapid and precise predictions of neurite growth. Furthermore, we will explore the integration of advanced machine learning techniques, including convolutional recurrent neural networks [103], and transformers [104], to predict and analyze the time series evolution of neuron growth data. Incorporating machine learning models with physics-based simulations has proven successful for simple reaction-diffusion problems on 2D domain [105] and complex neurite tree structures [27]. In addition, physics-informed neural networks could significantly improve model performance [106] for complex problems, including neuron traffic jams [28]. These advancements will significantly improve the model and broaden its potential in NDDs study.

Code and data availability

The code and datasets generated and analyzed in this paper are accessible in the ”NNDs” GitHub repository. https://github.com/CMU-CBML/NNDs (10.5281/zenodo.12575160). Correspondence and requests for code and data should be addressed to K.Q. or Y.J.Z.

Declaration of competing interest

The authors declare no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgement

K. Qian and Y. J. Zhang were supported in part by the NSF Grant CMMI-1953323. K. Qian, V. A. Webster-Wood, and Y. J. Zhang were supported in part by the Pennsylvania Infrastructure Technology Alliance (PITA) and Pennsylvania Manufacturing Innovation Program (PAMIP) grants. A. S. Liao and V. A. Webster-Wood were supported in part by an NSF CAREER award ECCS-2044785. A. S. Liao was also supported by NSF Graduate Research Fellowship Grant DGE-1745016 and Carnegie Mellon University Jean-Francois and Catherine Heitz Scholarship. This work used RM-nodes on Bridges-2 Supercomputer at Pittsburgh Supercomputer Center [99, 100] through allocation ID eng170006p from the Advanced Cyberinfrastructure Coordination Ecosystem: Services & Support (ACCESS) program, which is supported by National Science Foundation grants #2138259, #2138286, #2138307, #2137603, and #2138296.

References

  • [1] A. Thapar, M. Cooper, M. Rutter, Neurodevelopmental disorders, The Lancet Psychiatry 4 (4) (2017) 339–346.
  • [2] H. Tager-Flusberg, Neurodevelopmental disorders, MIT press, 1999.
  • [3] M. Fujitani, Y. Otani, H. Miyajima, Pathophysiological roles of abnormal axon initial segments in neurodevelopmental disorders, Cells 10 (8) (2021) 2110.
  • [4] T. Yamamoto, Genomic aberrations associated with the pathophysiological mechanisms of neurodevelopmental disorders, Cells 10 (9) (2021) 2317.
  • [5] B. N. Dugger, D. W. Dickson, Pathology of neurodegenerative diseases, Cold Spring Harbor Perspectives in Biology 9 (2016) a028035.
  • [6] R. C. Brown, A. H. Lockwood, B. R. Sonawane, Neurodegenerative diseases: an overview of environmental risk factors, Environmental Health Perspectives 113 (9) (2005) 1250–1256.
  • [7] B. Connor, M. Dragunow, The role of neuronal growth factors in neurodegenerative disorders of the human brain, Brain Research Reviews 27 (1) (1998) 1–39.
  • [8] D. K. Berg, New neuronal growth factors, Annual Review of Neuroscience 7 (1984) 149–170.
  • [9] J. L. Elliott, W. D. Snider, Motor neuron growth factors, Neurology 47 (4 Suppl 2) (1996) 47S–53S.
  • [10] A. Datar, J. Ameeramja, A. Bhat, R. Srivastava, A. Mishra, R. Bernal, J. Prost, A. Callan-Jones, P. A. Pullarkat, The roles of microtubules and membrane tension in axonal beading, retraction, and atrophy, Biophysical Journal 117 (5) (2019) 880–891.
  • [11] S. Budday, P. Steinmann, E. Kuhl, Physical biology of human brain development, Frontiers in Cellular Neuroscience 9 (2015) 257.
  • [12] S. Wang, X. Wang, M. A. Holland, Multi-physics modeling and finite-element formulation of neuronal dendrite growth with electrical polarization, Brain Multiphysics 4 (2023) 100071.
  • [13] H. G. E. Hentschel, A. Fine, Instabilities in cellular dendritic morphogenesis, Physical Review Letters 73 (1994) 3592–3595.
  • [14] J. K. Krottje, A. Van Ooyen, A mathematical framework for modeling axon guidance, Bulletin of Mathematical Biology 69 (2007) 3–31.
  • [15] Y. E. Pearson, E. Castronovo, T. A. Lindsley, D. A. Drew, Mathematical modeling of axonal formation Part I: geometry, Bulletin of Mathematical Biology 73 (2011) 2837–2864.
  • [16] M. Aeschlimann, L. Tettoni, Biophysical model of axonal pathfinding, Neurocomputing 38-40 (2001) 87–92.
  • [17] G. J. Goodhill, M. Gu, J. S. Urbach, Predicting axonal response to molecular gradients with a computational model of filopodial dynamics, Neural Computation 16 (11) (2004) 2221–2243.
  • [18] S. M. Maskery, H. M. Buettner, T. Shinbrot, Growth cone pathfinding: a competition between deterministic and stochastic events, BMC Neuroscience 5 (22) (2004).
  • [19] R. A. Koene, B. Tijms, P. Van Hees, F. Postma, A. De Ridder, G. J. Ramakers, J. Van Pelt, A. Van Ooyen, NETMORPH: a framework for the stochastic generation of large scale neuronal networks with realistic neuron morphologies, Neuroinformatics 7 (2009) 195–210.
  • [20] H. Cuntz, F. Forstner, A. Borst, M. Häusser, One rule to grow them all: a general theory of neuronal branching and its practical application, PLoS Computational Biology 6 (8) (2010) e1000877.
  • [21] D. E. Donohue, G. A. Ascoli, A comparative computer simulation of dendritic morphology, PLoS Computational Biology 4 (6) (2008) e1000089.
  • [22] B. Torben-Nielsen, E. De Schutter, Context-aware modeling of neuronal morphologies, Frontiers in Neuroanatomy 8 (2014) 92.
  • [23] J. P. Eberhard, A. Wanner, G. Wittum, NeuGen: a tool for the generation of realistic morphology of cortical neurons and neural networks in 3D, Neurocomputing 70 (1-3) (2006) 327–342.
  • [24] A. van Ooyen, A. Carnell, S. de Ridder, B. Tarigan, H. D. Mansvelder, F. Bijma, M. de Gunst, J. van Pelt, Independently outgrowing neurons and geometry-based synapse formation produce networks with realistic synaptic connectivity, PLOS ONE 9 (1) (2014) e85858.
  • [25] M. O’Toole, P. Lamoureux, K. E. Miller, A physical model of axonal elongation: force, viscosity, and adhesions govern the mode of outgrowth, Biophysical Journal 94 (7) (2008) 2610–2620.
  • [26] B. P. Graham, A. Van Ooyen, Mathematical modelling and numerical simulation of the morphological development of neurons, BMC Neuroscience 7 (2006) S9.
  • [27] A. Li, A. Barati Farimani, Y. J. Zhang, Deep learning of material transport in complex neurite networks, Scientific Reports 11 (2021) 11280.
  • [28] A. Li, Y. J. Zhang, Isogeometric analysis-based physics-informed graph neural network for studying traffic jam in neurons, Computer Methods in Applied Mechanics and Engineering 403 (2023) 115757.
  • [29] T. Takaki, K. Nakagawa, Y. Morita, E. Nakamachi, Phase-field modeling for axonal extension of nerve cells, Mechanical Engineering Journal 2 (3) (2015) 15–00063.
  • [30] K. T. Nella, B. M. Norton, H.-T. Chang, R. A. Heuer, C. B. Roque, A. J. Matsuoka, Bridging the electrode–neuron gap: finite element modeling of in vitro neurotrophin gradients to optimize neuroelectronic interfaces in the inner ear, Acta Biomaterialia 151 (2022) 360–378.
  • [31] K. Qian, A. Pawar, A. Liao, C. Anitescu, V. Webster-Wood, A. W. Feinberg, T. Rabczuk, Y. J. Zhang, Modeling neuron growth using isogeometric collocation based phase field method, Scientific Reports 12 (2022) 8120.
  • [32] K. Qian, A. S. Liao, S. Gu, V. A. Webster-Wood, Y. J. Zhang, Biomimetic IGA neuron growth modeling with neurite morphometric features and CNN-based prediction, Computer Methods in Applied Mechanics and Engineering 417 (2023) 116213.
  • [33] A. Liao, W. Cui, Y. J. Zhang, V. Webster-Wood, Quantitative evaluation of neuron developmental morphology in vitro using the change-point test, Summer Biomechanics, Bioengineering and Biotransport Conference (2021).
  • [34] A. van Ooyen, Modeling Neural Development, MIT Press, 2003.
  • [35] Y. J. Zhang, Challenges and advances in image-based geometric modeling and mesh generation, in: Image-Based Geometric Modeling and Mesh Generation, Springer, 2013.
  • [36] Y. J. Zhang, Geometric Modeling and Mesh Generation from Scanned Images, Chapman and Hall/CRC, 2016.
  • [37] L. Piegl, W. Tiller, The NURBS Book, Springer Science & Business Media, 1996.
  • [38] W. J. Gordon, R. F. Riesenfeld, B-spline Curves and Surfaces (1974) 95–126.
  • [39] T. J. Hughes, J. A. Cottrell, Y. Bazilevs, Isogeometric analysis: CAD, finite elements, NURBS, exact geometry and mesh refinement, Computer Methods in Applied Mechanics and Engineering 194 (39-41) (2005) 4135–4195.
  • [40] J. A. Cottrell, T. J. Hughes, Y. Bazilevs, Isogeometric analysis: toward integration of CAD and FEA, John Wiley & Sons, 2009.
  • [41] H. Casquero, L. Liu, Y. J. Zhang, A. Reali, H. Gomez, Isogeometric collocation using analysis-suitable T-splines of arbitrary degree, Computer Methods in Applied Mechanics and Engineering 301 (2016) 164–186.
  • [42] T. W. Sederberg, J. Zheng, A. Bakenov, A. Nasri, T-splines and T-NURCCs, ACM Transactions on Graphics 22 (3) (2003) 477–484.
  • [43] T. W. Sederberg, D. L. Cardon, G. T. Finnigan, N. S. North, J. Zheng, T. Lyche, T-spline simplification and local refinement, ACM Transactions on Graphics 23 (3) (2004) 276–283.
  • [44] M. A. Scott, X. Li, T. W. Sederberg, T. J. Hughes, Local refinement of analysis-suitable T-splines, Computer Methods in Applied Mechanics and Engineering 213 (2012) 206–222.
  • [45] T. Dokken, T. Lyche, K. F. Pettersen, Polynomial splines over locally refined box-partitions, Computer Aided Geometric Design 30 (3) (2013) 331–356.
  • [46] K. A. Johannessen, T. Kvamsdal, T. Dokken, Isogeometric analysis using LR B-splines, Computer Methods in Applied Mechanics and Engineering 269 (2014) 471–514.
  • [47] H. Kang, F. Chen, J. Deng, Modified T-splines, Computer Aided Geometric Design 30 (9) (2013) 827–843.
  • [48] X. Wei, X. Li, K. Qian, T. J. Hughes, Y. J. Zhang, H. Casquero, Analysis-suitable unstructured T-splines: multiple extraordinary points per face, Computer Methods in Applied Mechanics and Engineering 391 (2022) 114494.
  • [49] L. Liu, Y. J. Zhang, X. Wei, Weighted T-splines with application in reparameterizing trimmed NURBS surfaces, Computer Methods in Applied Mechanics and Engineering 295 (2015) 108–126.
  • [50] L. Liu, Y. J. Zhang, X. Wei, Handling extraordinary nodes with weighted T-spline basis functions, Procedia Engineering 124 (2015) 161–173.
  • [51] J. Deng, F. Chen, X. Li, C. Hu, W. Tong, Z. Yang, Y. Feng, Polynomial splines over hierarchical T-meshes, Graphical Models 70 (4) (2008) 76–86.
  • [52] E. Evans, M. Scott, X. Li, D. Thomas, Hierarchical T-splines: analysis-suitability, Bézier extraction, and application as an adaptive basis for isogeometric analysis, Computer Methods in Applied Mechanics and Engineering 284 (2015) 1–20.
  • [53] X. Wei, Y. J. Zhang, T. J. Hughes, M. A. Scott, Truncated hierarchical Catmull–Clark subdivision with local refinement, Computer Methods in Applied Mechanics and Engineering 291 (2015) 1–20.
  • [54] X. Wei, Y. J. Zhang, T. J. Hughes, M. A. Scott, Extended truncated hierarchical Catmull–Clark subdivision, Computer Methods in Applied Mechanics and Engineering 299 (2016) 316–336.
  • [55] A. Pawar, Y. J. Zhang, Y. Jia, X. Wei, T. Rabczuk, C. L. Chan, C. Anitescu, Adaptive FEM-based nonrigid image registration using truncated hierarchical B-splines, Computers & Mathematics with Applications 72 (8) (2016) 2028–2040.
  • [56] B. Li, J. Fu, Y. J. Zhang, A. Pawar, A trivariate T-spline based framework for modeling heterogeneous solids, Computer Aided Geometric Design 81 (2020) 101882.
  • [57] B. Li, J. Fu, Y. J. Zhang, W. Lin, J. Feng, C. Shang, Slicing heterogeneous solid using octree-based subdivision and trivariate T-splines for additive manufacturing, Rapid Prototy** Journal 26 (1) (2019) 164–175.
  • [58] H. Casquero, L. Liu, Y. J. Zhang, A. Reali, J. Kiendl, H. Gomez, Arbitrary-degree T-splines for isogeometric analysis of fully nonlinear Kirchhoff–Love shells, Computer-Aided Design 82 (2017) 140–153.
  • [59] A. Li, Y. J. Zhang, Modeling intracellular transport and traffic jam in 3D neurons using PDE-constrained optimization, Journal of Mechanics 38 (2022) 44–59.
  • [60] A. Li, Y. J. Zhang, Modeling material transport regulation and traffic jam in neurons using PDE-constrained optimization, Scientific Reports 12 (2022) 3902.
  • [61] S. Balay, S. Abhyankar, M. F. Adams, S. Benson, J. Brown, P. Brune, K. Buschelman, E. Constantinescu, L. Dalcin, A. Dener, V. Eijkhout, J. Faibussowitsch, W. D. Gropp, V. Hapla, T. Isaac, P. Jolivet, D. Karpeev, D. Kaushik, M. G. Knepley, F. Kong, S. Kruger, D. A. May, L. C. McInnes, R. T. Mills, L. Mitchell, T. Munson, J. E. Roman, K. Rupp, P. Sanan, J. Sarich, B. F. Smith, S. Zampini, H. Zhang, H. Zhang, J. Zhang, PETSc/TAO users manual, Tech. Rep. ANL-21/39 - Revision 3.21, Argonne National Laboratory (2024).
  • [62] J. Zhang, J. Brown, S. Balay, J. Faibussowitsch, M. Knepley, O. Marin, R. T. Mills, T. Munson, B. F. Smith, S. Zampini, The PetscSF scalable communication layer, IEEE Transactions on Parallel and Distributed Systems 33 (4) (2022) 842–853.
  • [63] X. Wei, Y. J. Zhang, L. Liu, T. J. Hughes, Truncated T-splines: fundamentals and methods, Computer Methods in Applied Mechanics and Engineering 316 (2017) 349–372.
  • [64] G. Karypis, V. Kumar, A fast and high quality multilevel scheme for partitioning irregular graphs, SIAM Journal on Scientific Computing 20 (1) (1998) 359–392.
  • [65] H. Gomez, A. Reali, G. Sangalli, Accurate, efficient, and (iso) geometrically flexible collocation methods for phase-field models, Journal of Computational Physics 262 (2014) 153–171.
  • [66] D. Schillinger, M. J. Borden, H. K. Stolarski, Isogeometric collocation for phase-field fracture models, Computer Methods in Applied Mechanics and Engineering 284 (2015) 583–610.
  • [67] T. Takaki, Phase-field modeling and simulations of dendrite growth, ISIJ International 54 (2) (2014) 437–444.
  • [68] A. S. Liao, W. Cui, Y. J. Zhang, V. A. Webster-Wood, Semi-automated quantitative evaluation of neuron developmental morphology in vitro using the change-point test, Neuroinformatics 21 (1) (2023) 163–176.
  • [69] S. M. Allen, J. W. Cahn, A microscopic theory for antiphase boundary motion and its application to antiphase domain coarsening, Acta Metallurgica 27 (6) (1979) 1085–1095.
  • [70] T. Takaki, M. Asanishi, A. Yamanaka, Y. Tomita, Phase-field simulation during spherulite formation of polymer, Key Engineering Materials 345 (2007) 939–942.
  • [71] J. J. Eggleston, G. B. McFadden, P. W. Voorhees, A phase-field model for highly anisotropic interfacial energy, Physica D: Nonlinear Phenomena 150 (1-2) (2001) 91–103.
  • [72] T. Takaki, T. Hasebe, Y. Tomita, Two-dimensional phase-field simulation of self-assembled quantum dot formation, Journal of Crystal G rowth 287 (2) (2006) 495–499.
  • [73] B. Ren, J. Huang, M. C. Lin, S.-M. Hu, Controllable dendritic crystal simulation using orientation field, Computer Graphics Forum 37 (2) (2018) 485–495.
  • [74] D. R. McLean, A. van Ooyen, B. P. Graham, Continuum model for tubulin-driven neurite elongation, Neurocomputing 58 (2004) 511–516.
  • [75] D. R. McLean, B. P. Graham, Mathematical formulation and analysis of a continuum model for tubulin-driven neurite elongation, Proceedings of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences 460 (2048) (2004) 2437–2456.
  • [76] B. P. Graham, K. Lauchlan, D. R. Mclean, Dynamics of outgrowth in a continuum model of neurite elongation, Journal of Computational Neuroscience 20 (2006) 43–60.
  • [77] A. van Ooyen, B. P. Graham, G. J. A. Ramakers, Competition for tubulin between growing neurites during development, Neurocomputing 38-40 (2001) 73–78.
  • [78] R. Kobayashi, Modeling and numerical simulations of dendritic crystal growth, Physica D: Nonlinear Phenomena 63 (3-4) (1993) 410–423.
  • [79] H. Song, G. Ming, M. Poo, cAMP-induced switching in turning direction of nerve growth cones, Nature 388 (6639) (1997) 275–279.
  • [80] S. X. Bamji, M. Majdan, C. D. Pozniak, D. J. Belliveau, R. Aloyz, J. Kohn, C. G. Causing, F. D. Miller, The p75 neurotrophin receptor mediates neuronal apoptosis and is essential for naturally occurring sympathetic neuron death, The Journal of Cell Biology 140 (4) (1998) 911–923.
  • [81] G. L. Barrett, The p75 neurotrophin receptor and neuronal apoptosis, Progress in Neurobiology 61 (2) (2000) 205–229.
  • [82] R. Meeker, K. Williams, Dynamic nature of the p75 neurotrophin receptor in response to injury and disease, Journal of Neuroimmune Pharmacology 9 (2014) 615–628.
  • [83] R. B. Meeker, K. S. Williams, The p75 neurotrophin receptor: at the crossroad of neural repair and death, Neural Regeneration Research 10 (5) (2015) 721–725.
  • [84] L. Marchetti, F. Bonsignore, F. Gobbo, R. Amodeo, M. Calvello, A. Jacob, G. Signore, C. Schirripa Spagnolo, D. Porciani, M. Mainardi, et al., Fast-diffusing p75ntr monomers support apoptosis and growth cone collapse by neurotrophin ligands, Proceedings of the National Academy of Sciences 116 (43) (2019) 21563–21572.
  • [85] C. E. Krewson, W. M. Saltzman, Transport and elimination of recombinant human NGF during long-term delivery to the brain, Brain Research 727 (1-2) (1996) 169–181.
  • [86] B. Lu, P. T. Pang, N. H. Woo, The yin and yang of neurotrophin action, Nature Reviews Neuroscience 6 (8) (2005) 603–614.
  • [87] J. Piontek, C. C. Chen, M. Kempf, R. Brandt, Neurotrophins differentially regulate the survival and morphological complexity of human CNS model neurons, Journal of Neurochemistry 73 (1) (1999) 139–146.
  • [88] E. J. Huang, L. F. Reichardt, Neurotrophins: roles in neuronal development and function, Annual Review of Neuroscience 24 (1) (2001) 677–736.
  • [89] T. Lyche, E. Cohen, K. Mørken, Knot line refinement algorithms for tensor product B-spline surfaces, Computer Aided Geometric Design 2 (1-3) (1985) 133–139.
  • [90] A. Buffa, D. Cho, G. Sangalli, Linear independence of the T-spline blending functions associated with some particular T-meshes, Computer Methods in Applied Mechanics and Engineering 199 (23-24) (2010) 1437–1445.
  • [91] C. Giannelli, B. Jüttler, H. Speleers, Thb-splines: The truncated basis for hierarchical splines, Computer Aided Geometric Design 29 (7) (2012) 485–498.
  • [92] C. Giannelli, B. Jüttler, H. Speleers, Strongly stable bases for adaptively refined multilevel spline spaces, Advances in Computational Mathematics 40 (2014) 459–490.
  • [93] L. Liu, Y. J. Zhang, X. Wei, NURBS surface reparameterization using truncated T-splines, in: 23rd International Meshing Roundtable. London, UK. Oct. 12-15, 2014.
  • [94] J. L. Bentley, Multidimensional binary search trees used for associative searching, Communications of the ACM 18 (9) (1975) 509–517.
  • [95] J. Zhao, M. D. Davis, Y. A. Martens, M. Shinohara, N. R. Graff-Radford, S. G. Younkin, Z. K. Wszolek, T. Kanekiyo, G. Bu, APOE ε𝜀\varepsilonitalic_ε4/ε𝜀\varepsilonitalic_ε4 diminishes neurotrophic function of human iPSC-derived astrocytes, Human Molecular Genetics 26 (14) (2017) 2690–2700.
  • [96] K. Kawatani, M.-L. Holm, S. C. Starling, Y. A. Martens, J. Zhao, W. Lu, Y. Ren, Z. Li, P. Jiang, Y. Jiang, et al., Abca7 deficiency causes neuronal dysregulation by altering mitochondrial lipid metabolism, Molecular Psychiatry (2023) 1–11.
  • [97] Thermo Fisher Scientific, B-27 Plus Neuronal Culture System, https://assets.thermofisher.com/TFS-Assets/LSG/manuals/MAN0017319_B27_PlusNeuronalCultureSystem_UG.pdf (2018).
  • [98] W. Gropp, E. Lusk, N. Doss, A. Skjellum, A high-performance, portable implementation of the MPI message passing interface standard, Parallel Computing 22 (6) (1996) 789–828.
  • [99] N. Wilkins-Diehr, S. Sanielevici, J. Alameda, J. Cazes, L. Crosby, M. Pierce, R. Roskies, An overview of the XSEDE extended collaborative support program, in: High Performance Computer Applications - 6th International Conference, ISUM 2015, Vol. 595 of Communications in Computer and Information Science, 2016, pp. 3–13.
  • [100] J. Towns, T. Cockerill, M. Dahan, I. Foster, K. Gaither, A. Grimshaw, V. Hazlewood, S. Lathrop, D. Lifka, G. D. Peterson, R. Roskies, J. R. Scott, N. Wilkins-Diehr, XSEDE: accelerating scientific discovery, Computing in Science & Engineering 16 (5) (2014) 62–74.
  • [101] S. Diehl, E. Henningsson, A. Heyden, Efficient simulations of tubulin-driven axonal growth, Journal of Computational Neuroscience 41 (1) (2016) 45–63.
  • [102] A. Li, Y. J. Zhang, Intracellular material transport simulation in neurons using isogeometric analysis and deep learning, in: International Conference on Computational Science, Springer, 2023, pp. 486–493.
  • [103] X. Shi, Z. Chen, H. Wang, D.-Y. Yeung, W.-K. Wong, W.-C. Woo, Convolutional LSTM network: a machine learning approach for precipitation nowcasting, Advances in Neural Information Processing Systems 28 (2015).
  • [104] A. Vaswani, N. Shazeer, N. Parmar, J. Uszkoreit, L. Jones, A. N. Gomez, L. U. Kaiser, I. Polosukhin, Attention is all you need, Advances in Neural Information Processing Systems 30 (2017).
  • [105] A. Li, R. Chen, A. B. Farimani, Y. J. Zhang, Reaction diffusion system prediction based on convolutional neural network, Scientific Reports 10 (2020) 3894.
  • [106] M. Raissi, P. Perdikaris, G. E. Karniadakis, Physics-informed neural networks: a deep learning framework for solving forward and inverse problems involving nonlinear partial differential equations, Journal of Computational Physics 378 (2019) 686–707.