Skip to main content
Log in

Data-driven simulation of network-based tau spreading tailored to individual Alzheimer's patients

  • Original Article
  • Published:
Engineering with Computers Aims and scope Submit manuscript

Abstract

Tau tangles in the brain cortex spread along the brain network in distinct patterns among Alzheimer's patients. We aim to simulate their network-based spreading within the cortex, tailored to each individual along the Alzheimer's continuum, without assuming any assumptions about the network architecture. A group-level intrinsic spreading network was constructed to model the pathways for the proximal and distal spreading of tau tangles by optimizing the biophysical model based on a discovery dataset of longitudinal tau positron emission tomography images for 78 amyloid-positive individuals. Group-level spreading parameters were also obtained and subsequently adjusted to produce individuated tau trajectories. By simulating these individuated tau spreading models for every individual in the discovery dataset, we successfully captured proximal and distal tau spreading, allowing reliable inferences about the underlying mechanism of tau spreading. Simulating the models also allowed highly accurate prediction of future tau topography for both discovery and independent validation datasets.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

Data availability

The discovery dataset generated and/or analyzed during the current study is not publicly available due to confidentiality agreements but is available from the corresponding author upon reasonable request. The validation dataset is available in the Alzheimer’s Disease Neuroimaging Initiative (ADNI) repository (https://ida.loni.usc.edu/) and also available from the corresponding author upon reasonable request.

Code availability

All original code is available in this paper’s supplementary information.

References

  1. Brunello CA et al (2020) Mechanisms of secretion and spreading of pathological tau protein. Cell Mol Life Sci 77:1721–1744

    Article  Google Scholar 

  2. Spillantini MG, Goedert M (2013) Tau pathology and neurodegeneration. Lancet Neurol 12(6):609–622

    Article  Google Scholar 

  3. Bejanin A et al (2017) Tau pathology and neurodegeneration contribute to cognitive impairment in Alzheimer’s disease. Brain 140(12):3286–3300

    Article  Google Scholar 

  4. De Calignon A et al (2012) Propagation of tau pathology in a model of early Alzheimer’s disease. Neuron 73(4):685–697

    Article  Google Scholar 

  5. Hu W et al (2016) Hyperphosphorylation determines both the spread and the morphology of tau pathology. Alzheimers Dement 12(10):1066–1077

    Article  Google Scholar 

  6. Takeda S et al (2015) Neuronal uptake and propagation of a rare phosphorylated high-molecular-weight tau derived from Alzheimer’s disease brain. Nat Commun 6(1):8490

    Article  Google Scholar 

  7. DeVos SL et al (2018) Synaptic tau seeding precedes tau pathology in human Alzheimer’s disease brain. Front Neurosci 12:267

    Article  Google Scholar 

  8. Zanier ER et al (2018) Induction of a transmissible tau pathology by traumatic brain injury. Brain 141(9):2685–2699

    Google Scholar 

  9. Charil A et al (2019) Tau subtypes of Alzheimer’s disease determined in vivo using flortaucipir PET imaging. J Alzheimers Dis 71(3):1037–1048

    Article  Google Scholar 

  10. Vogel JW et al (2021) Four distinct trajectories of tau deposition identified in Alzheimer’s disease. Nat Med 27(5):871–881

    Article  Google Scholar 

  11. Braak H, Braak E (1991) Neuropathological stageing of Alzheimer-related changes. Acta Neuropathol 82(4):239–259

    Article  Google Scholar 

  12. Schäfer A, Mormino EC, Kuhl E (2020) Network diffusion modeling explains longitudinal tau pet data. Front Neurosci 14:566876

    Article  Google Scholar 

  13. Yang F et al (2021) Longitudinal predictive modeling of tau progression along the structural connectome. Neuroimage 237:118126

    Article  Google Scholar 

  14. Therriault J et al (2022) Intrinsic connectivity of the human brain provides scaffold for tau aggregation in clinical variants of Alzheimer’s disease. Sci Transl Med 14(659):eabc8693

    Article  Google Scholar 

  15. Cornblath EJ et al (2021) Computational modeling of tau pathology spread reveals patterns of regional vulnerability and the impact of a genetic risk factor. Sci Adv 7(24):eabg6677

    Article  Google Scholar 

  16. Fornari S et al (2019) Prion-like spreading of Alzheimer’s disease within the brain’s connectome. J R Soc Interface 16(159):20190356

    Article  Google Scholar 

  17. Putra P et al (2021) Braiding Braak and Braak: Staging patterns and model selection in network neurodegeneration. Netw Neurosci 5(4):929–956

    Article  Google Scholar 

  18. Raj A, Kuceyeski A, Weiner M (2012) A network diffusion model of disease progression in dementia. Neuron 73(6):1204–1215

    Article  Google Scholar 

  19. Wu JW et al (2013) Small misfolded Tau species are internalized via bulk endocytosis and anterogradely and retrogradely transported in neurons. J Biol Chem 288(3):1856–1870

    Article  Google Scholar 

  20. Franzmeier N et al (2020) Functional brain architecture is associated with the rate of tau accumulation in Alzheimer’s disease. Nat Commun 11(1):347

    Article  Google Scholar 

  21. Franzmeier N et al (2019) Functional connectivity associated with tau levels in ageing, Alzheimer’s, and small vessel disease. Brain 142(4):1093–1107

    Article  Google Scholar 

  22. Ossenkoppele R et al (2019) Tau covariance patterns in Alzheimer’s disease patients match intrinsic connectivity networks in the healthy brain. NeuroImage Clin 23:101848

    Article  Google Scholar 

  23. Brown JA et al (2019) Patient-tailored, connectivity-based forecasts of spreading brain atrophy. Neuron 104(5):856-868.e5

    Article  Google Scholar 

  24. Franzmeier N et al (2020) Patient-centered connectivity-based prediction of tau pathology spread in Alzheimer’s disease. Sci Adv 6(48):eabd1327

    Article  Google Scholar 

  25. Schäfer A, et al (2022) Correlating tau pathology to brain atrophy using a physics-based Bayesian model. Eng Comput, pp 1–11

  26. Schäfer A et al (2021) Bayesian physics-based modeling of tau propagation in Alzheimer’s disease. Front Physiol 12:702975

    Article  Google Scholar 

  27. Rao AV (2014) Trajectory optimization: a survey. In: Waschl H, Kolmanovsky I, Steinbuch M, del Re L. (eds) Optimization and Optimal Control in Automotive Systems. Lecture Notes in Control and Information Sciences. Springer, Cham. 455:3–21

  28. Balaji V, et al (2002) A graph neural network model for the prediction of longitudinal tau aggregation. J Nucl Med 63 (supplement 2) 2233

  29. Weickenmeier J et al (2019) A physics-based model explains the prion-like features of neurodegeneration in Alzheimer’s disease, Parkinson’s disease, and amyotrophic lateral sclerosis. J Mech Phys Solids 124:264–281

    Article  Google Scholar 

  30. Meisl G et al (2021) In vivo rate-determining steps of tau seed accumulation in Alzheimer’s disease. Sci Adv 7(44):eabh1448

    Article  Google Scholar 

  31. Lowe VJ et al (2018) Widespread brain tau and its association with ageing, Braak stage and Alzheimer’s dementia. Brain 141(1):271–287

    Article  Google Scholar 

  32. Lozes F, Elmoataz A, Lézoray O (2014) Partial difference operators on weighted graphs for image processing on surfaces and point clouds. IEEE Trans Image Process 23(9):3896–3909

    Article  MathSciNet  Google Scholar 

  33. Elmoataz A, Lezoray O, Bougleux S (2008) Nonlocal discrete regularization on weighted graphs: a framework for image and manifold processing. IEEE Trans Image Process 17(7):1047–1060

    Article  MathSciNet  Google Scholar 

  34. Grigor’yan A (2018) Introduction to analysis on graphs, vol 71. American Mathematical Soc

    Google Scholar 

  35. Kondor RI, Lafferty J (2002) Diffusion kernels on graphs and other discrete structures. In: Proceedings of the 19th International Conference on Machine Learning

  36. Zheng Y-Q et al (2019) Local vulnerability and global connectivity jointly shape neurodegenerative disease propagation. PLoS Biol 17(11):e3000495

    Article  Google Scholar 

  37. Vogel JW et al (2020) Spread of pathological tau proteins through communicating neurons in human Alzheimer’s disease. Nat Commun 11(1):2612

    Article  Google Scholar 

  38. Iturria-Medina Y et al (2014) Epidemic spreading model to characterize misfolded proteins propagation in aging and associated neurodegenerative disorders. PLoS Comput Biol 10(11):e1003956

    Article  Google Scholar 

  39. Desikan RS et al (2006) An automated labeling system for subdividing the human cerebral cortex on MRI scans into gyral based regions of interest. Neuroimage 31(3):968–980

    Article  Google Scholar 

  40. Cho H et al (2019) Progressive tau accumulation in Alzheimer disease: 2-year follow-up study. J Nucl Med 60(11):1611–1621

    Article  MathSciNet  Google Scholar 

  41. Kennan J (2006) A note on discrete approximations of continuous distributions. Univ. Wisconsin-Madison, Madison, WI, USA

  42. Krogh A, Hertz J (1991) A simple weight decay can improve generalization. Adv Neural Inform Process Syst 4: 950–957

  43. Kirkpatrick S, Gelatt CD Jr, Vecchi MP (1983) Optimization by simulated annealing. Science 220(4598):671–680

    Article  MathSciNet  Google Scholar 

  44. Tsallis C, Stariolo DA (1996) Generalized simulated annealing. Phys A 233(1–2):395–406

    Article  Google Scholar 

  45. Xiang Y, Gong X (2000) Efficiency of generalized simulated annealing. Phys Rev E 62(3):4473

    Article  Google Scholar 

  46. Nahar S, Sahni S, Shragowitz E (1986). Simulated annealing and combinatorial optimization. In: 23rd ACM/IEEE Design Automation Conference. 1986. IEEE

  47. Obuchi T, Kabashima Y (2016) Sparse approximation problem: how rapid simulated annealing succeeds and fails. J Phys Conf Ser 699 012017

  48. Stoer J, Bulirsch R (1991) Introduction to Numerical Analysis (Applied Mathematics vol 12). Springer, Berlin

    Google Scholar 

  49. Pan SJ, Yang Q (2010) A survey on transfer learning. IEEE Trans Knowl Data Eng 22(10):1345–1359

    Article  Google Scholar 

  50. Hutter F, Hoos HH, Stützle T (2007) Automatic algorithm configuration based on local search. In: Aaai. 2007

  51. Levenberg K (1944) A method for the solution of certain non-linear problems in least squares. Q Appl Math 2(2):164–168

    Article  MathSciNet  Google Scholar 

  52. Marquardt DW (1963) An algorithm for least-squares estimation of nonlinear parameters. J Soc Ind Appl Math 11(2):431–441

    Article  MathSciNet  Google Scholar 

  53. Lyness JN (1967) Numerical algorithms based on the theory of complex variable. In: Proceedings of the 1967 22nd National Conference. 1967

  54. Ralston A (1962) Runge-Kutta methods with minimum error bounds. Math Comput 16(80):431–437

    Article  MathSciNet  Google Scholar 

  55. Achard S, Bullmore E (2007) Efficiency and cost of economical brain functional networks. PLoS Comput Biol 3(2):e17

    Article  Google Scholar 

  56. He Y, Chen ZJ, Evans AC (2007) Small-world anatomical networks in the human brain revealed by cortical thickness from MRI. Cereb Cortex 17(10):2407–2419

    Article  Google Scholar 

  57. Brin S, Page L (1998) The anatomy of a large-scale hypertextual web search engine. Comput Netw ISDN Syst 30(1–7):107–117

    Article  Google Scholar 

  58. Rubinov M, Sporns O (2011) Weight-conserving characterization of complex functional brain networks. Neuroimage 56(4):2068–2079

    Article  Google Scholar 

  59. Hawrylycz MJ et al (2012) An anatomically comprehensive atlas of the adult human brain transcriptome. Nature 489(7416):391–399

    Article  Google Scholar 

  60. Cho H et al (2016) In vivo cortical spreading pattern of tau and amyloid in the Alzheimer disease spectrum. Ann Neurol 80(2):247–258

    Article  Google Scholar 

  61. Insel PS et al (2020) Neuroanatomical spread of amyloid β and tau in Alzheimer’s disease: implications for primary prevention. Brain Commun 2(1):fcaa007

    Article  Google Scholar 

  62. Thompson TB et al (2020) Protein-protein interactions in neurodegenerative diseases: a conspiracy theory. PLoS Comput Biol 16(10):e1008267

    Article  Google Scholar 

  63. Van der Velden BH et al (2022) Explainable artificial intelligence (XAI) in deep learning-based medical image analysis. Med Image Anal 79:102470

    Article  Google Scholar 

  64. Berron D et al (2020) Medial temporal lobe connectivity and its associations with cognition in early Alzheimer’s disease. Brain 143(4):1233–1248

    Article  Google Scholar 

  65. Harrison TM et al (2019) Longitudinal tau accumulation and atrophy in aging and alzheimer disease. Ann Neurol 85(2):229–240

    Article  Google Scholar 

  66. Leng F et al (2023) Neuroinflammation is independently associated with brain network dysfunction in Alzheimer’s disease. Mol Psychiatry 28(3):1303–1311

    Article  Google Scholar 

  67. Buckner RL, Andrews-Hanna JR, Schacter DL (2008) The brain’s default network: anatomy, function, and relevance to disease. Ann N Y Acad Sci 1124(1):1–38

    Article  Google Scholar 

  68. Garcés P et al (2014) The Default Mode Network is functionally and structurally disrupted in amnestic mild cognitive impairment—A bimodal MEG–DTI study. NeuroImage Clin 6:214–221

    Article  Google Scholar 

  69. Wu JW et al (2016) Neuronal activity enhances tau propagation and tau pathology in vivo. Nat Neurosci 19(8):1085–1092

    Article  MathSciNet  Google Scholar 

  70. Yamada K et al (2014) Neuronal activity regulates extracellular tau in vivo. J Exp Med 211(3):387–393

    Article  Google Scholar 

  71. Busche MA et al (2008) Clusters of hyperactive neurons near amyloid plaques in a mouse model of Alzheimer’s disease. Science 321(5896):1686–1689

    Article  Google Scholar 

  72. Busche MA et al (2012) Critical role of soluble amyloid-β for early hippocampal hyperactivity in a mouse model of Alzheimer’s disease. Proc Natl Acad Sci 109(22):8740–8745

    Article  Google Scholar 

  73. Castanho I et al (2020) Transcriptional signatures of tau and amyloid neuropathology. Cell Rep 30(6):2040-2054.e5

    Article  Google Scholar 

  74. Zhou J et al (2012) Predicting regional neurodegeneration from the healthy brain functional connectome. Neuron 73(6):1216–1227

    Article  Google Scholar 

  75. Montal V et al (2022) Network Tau spreading is vulnerable to the expression gradients of APOE and glutamatergic-related genes. Sci Transl Med 14(655):eabn7273

    Article  Google Scholar 

  76. Gong H et al (2013) Lipoprotein lipase (LPL) is associated with neurite pathology and its levels are markedly reduced in the dentate gyrus of Alzheimer’s disease brains. J Histochem Cytochem 61(12):857–868

    Article  Google Scholar 

  77. Villavicencio Tejo F, Quintanilla RA (2021) Contribution of the Nrf2 pathway on oxidative damage and mitochondrial failure in Parkinson and Alzheimer’s disease. Antioxidants 10(7):1069

    Article  Google Scholar 

  78. Caron F (2012) Bayesian nonparametric models for bipartite graphs. Adv Neural Inform Process Syst 25

  79. Ni Y et al (2018) Bayesian graphical models for computational network biology. BMC Bioinforma 19(3):59–69

    MathSciNet  Google Scholar 

  80. Matkowski J (2012) Mean-value theorem for vector-valued functions. Math Bohem 137(4):415–423

    Article  MathSciNet  Google Scholar 

  81. Jaynes ET (1957) Information theory and statistical mechanics. Phys Rev 106(4):620

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

This research was supported by the Basic Research Program through the National Research Foundation of Korea (NRF) grant funded by the Ministry of Science and ICT (MSIT) (No. 2022R1A4A1033856), by the NRF grant funded by the MSIT (No. 2023R1A2C2006201, Development of simulation-based Digital-Brain editing technology), and by the Korea Disease Control and Prevention Agency (grant No. 2023-ER1003-01), funded by the Ministry of Health & Welfare, Republic of Korea. This research was also supported by a grant from the Korea Health Technology R&D Project through the Korea Health Industry Development Institute (KHIDI), funded by the Ministry of Health & Welfare, Republic of Korea (grant No. HU20C0164), and by a faculty research grant from Yonsei University College of Medicine (6-2022-0073).

Author information

Authors and Affiliations

Authors

Consortia

Contributions

J. K. Seong and C. H. Lyoo had full access to all the data in the study and took responsibility for the data's integrity and the data analysis's accuracy. S. W. Kim, H. Cho, J. K. Seong, and C. H. Lyoo planned the study design and concept. H. Cho and C. H. Lyoo collected image data. S. W. Kim and C. H. Lyoo preprocessed image data. S. W. Kim and Y. Lee performed analyses. S. W. Kim, H. Cho, C. H. Lyoo, and J. K. Seong interpreted the results and wrote the original manuscript. All authors reviewed and edited the manuscript.

Corresponding author

Correspondence to Joon-Kyung Seong.

Ethics declarations

Conflict of interest

The authors declare no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary file1 (PDF 5368 KB)

Supplementary file2 (AVI 50689 KB)

Supplementary file3 (AVI 49800 KB)

Supplementary file4 (AVI 26204 KB)

Supplementary file5 (AVI 90866 KB)

Supplementary file6 (AVI 88643 KB)

Appendices

Appendix

Appendix A. Derivation of the objective function for group-level optimization

We first note that the loss function \(\ell \left( {{{\varvec{\uptheta}}};\left\{ {\left( {t_{i,j} ,{\tilde{\mathbf{x}}}_{i,j} } \right)} \right\}_{j = 0}^{{m_{i} - 1}} ,H_{1} } \right)\) in the group-level optimization objective (3) quantifies the discrepancy as the difference between the average rate of change in observed tau topographic maps over \(\left[ {t_{i,0} ,t_{{i,m_{i} - 1}} } \right)\) and the expected instantaneous rates of changes evaluated by the tau spreading model. The average rate of change, denoted by \({\tilde{\mathbf{y}}}_{i}\), can be represented as follows:

\({\tilde{\mathbf{y}}}_{i} = \frac{{{\tilde{\mathbf{x}}}_{{i,m_{i} - 1}} - {\tilde{\mathbf{x}}}_{i,0} }}{{t_{{i,m_{i} - 1}} - t_{i,0} }} = \sum\limits_{j = 0}^{{m_{i} - 2}} {w_{i,j} \frac{{{\tilde{\mathbf{x}}}_{i,j + 1} - {\tilde{\mathbf{x}}}_{i,j} }}{{t_{i,j + 1} - t_{i,j} }}}\),

where \(w_{i,j} = \frac{{t_{i,j + 1} - t_{i,j} }}{{t_{{i,m_{i} - 1}} - t_{i,0} }}\) denotes the weight assigned to each sub-interval \(\left[ {t_{i,j} ,t_{i,j + 1} } \right)\). Let \({\dot{\mathbf{x}}}\left( {t;{{\varvec{\uptheta}}}} \right)\) denote the rate of instantaneous rates of changes evaluated by the tau spreading model \({\mathbf{x}}\left( {t;{{\varvec{\uptheta}}}} \right)\). By the mean-value theorem for vector-valued function [80], for a function differentiable \({\mathbf{x}}\left( t \right) = \left[ {\begin{array}{*{20}c} {x_{1} \left( t \right)} & {x_{2} \left( t \right)} & \ldots & {x_{k} \left( t \right)} \\ \end{array} } \right]^{{\text{T}}}\) over a bounded open interval, there exist constants \(t_{1} ,t_{2} ,...,t_{k}\) in the interval such that

$$\frac{{{\mathbf{x}}\left( b \right) - {\mathbf{x}}\left( a \right)}}{b - a} = \left[ {\begin{array}{*{20}c} {\dot{x}_{1} \left( {t_{1} } \right)} & {\dot{x}_{2} \left( {t_{2} } \right)} & \cdots & {\dot{x}_{k} \left( {t_{k} } \right)} \\ \end{array} } \right]^{{\text{T}}} ,$$

with real numbers \(a\) and \(b\) in the interval. The exact values of the constants cannot be determined, however, so we tried to get the expectation of \({\dot{\mathbf{x}}}\left( {t;{{\varvec{\uptheta}}}} \right)\), rather than the raw value of \({\dot{\mathbf{x}}}\left( {t;{{\varvec{\uptheta}}}} \right)\), close to \({\tilde{\mathbf{y}}}_{i}\).

The expected value of \({\dot{\mathbf{x}}}\left( {t;{{\varvec{\uptheta}}}} \right)\) over the interval \(\left[ {t_{i,0} ,t_{{i,m_{i} - 1}} } \right)\), denoted by \({\hat{\mathbf{y}}}_{i}\), can be represented as follows:

$${\hat{\mathbf{y}}}_{i} \left( {{\varvec{\uptheta}}} \right) = {\mathbb{E}}_{{T_{i} \sim p\left( {T_{i} } \right)}} \left[ {{\dot{\mathbf{x}}}\left( {T_{i} ;{{\varvec{\uptheta}}}} \right)} \right],$$

where \(T_{i}\) denotes a random variable that can have a value between \(\left[ {t_{i,0} ,t_{{i,m_{i} - 1}} } \right)\) with the probability density function \(p\left( {T_{i} } \right)\). Since there is no information about the time constants except that they are in \(\left[ {t_{i,j} ,t_{i,j + 1} } \right)\), we approximate \(p\left( {T_{i} } \right)\) as a piecewise uniform distribution as follows:

$$p\left( {T_{i} } \right) = \left\{ \begin{gathered} w_{i,0} \cdot {\text{Uni}}\left( {t;t_{i,0} ,t_{i,1} } \right)\;\;\;\;\;{\text{if}}\;t \in \left( {t_{i,0} ,t_{i,1} } \right] \hfill \\ w_{i,1} \cdot {\text{Uni}}\left( {t;t_{i,0} ,t_{i,1} } \right)\;\;\;\;\;{\text{if}}\;t \in \left( {t_{i,1} ,t_{i,2} } \right] \hfill \\ \;\;\;\;\;\;\;\;\;\;\;\;\; \vdots \hfill \\ w_{{i,m_{i} - 2}} \cdot {\text{Uni}}\left( {t;t_{i,0} ,t_{i,1} } \right)\;\;{\text{if}}\;t \in \left( {t_{{i,m_{i} - 2}} ,t_{{i,m_{i} - 1}} } \right] \hfill \\ \end{gathered} \right.,$$

according to the principle of maximum entropy [81]. Herein, \({\text{Uni}}\left( {t;t_{i,j} ,t_{i,j + 1} } \right)\) denotes a continuous uniform distribution with the time interval. The expected value still cannot be evaluated due to intractable integration; we divided each interval \(\left( {t_{i,j} ,t_{i,j + 1} } \right)\) into \(H_{1}\) sub-intervals and approximated the continuous uniform distribution for each interval to a discrete uniform distribution \({\text{Uni}}\left\{ {t;t_{i,j} ,t_{i,j + 1} } \right\}\) whose support contains midpoints of the \(H_{1}\) sub-intervals, according to the best discrete approximation having equally weighted support points [41]. The expected value is then computed as follows:

$$\begin{gathered} {\mathbb{E}}_{{T_{i} \sim p\left( {T_{i} } \right)}} \left[ {{\dot{\mathbf{x}}}\left( {T_{i} ;{{\varvec{\uptheta}}}} \right)} \right] \approx \frac{1}{{H_{1} }}\sum\limits_{j = 0}^{{m_{i} - 2}} {\sum\limits_{k = 1}^{{H_{1} }} {w_{i,j} \cdot {\dot{\mathbf{x}}}\left( {\left\langle {\tau_{i,j}^{{\left( {2H_{1} } \right)}} } \right\rangle_{2k - 1} ;{{\varvec{\uptheta}}}} \right)} } \\ = \sum\limits_{j = 0}^{{m_{i} - 2}} {w_{i,j} \cdot {\text{Avg}}_{k} \left( {{\dot{\mathbf{x}}}\left( {\left\langle {\tau_{i,j}^{{\left( {2H_{1} } \right)}} } \right\rangle_{2k - 1} ;{{\varvec{\uptheta}}}} \right)} \right)} . \\ \end{gathered}$$

It is worth noting that for taking midpoints of the \(H_{1}\) sub-intervals, we divided each interval \(\left( {t_{i,j} ,t_{i,j + 1} } \right)\) into \(2H_{1}\) sub-intervals and took all odd-numbered points. Each value of \({\dot{\mathbf{x}}}\left( {\left\langle {\tau_{i,j}^{{\left( {2H_{1} } \right)}} } \right\rangle_{2k - 1} ;{{\varvec{\uptheta}}}} \right)\) was computed using the numerical simulation with the step size \(h_{1} = {{\left( {t_{i,j + 1} - t_{i,j} } \right)} \mathord{\left/ {\vphantom {{\left( {t_{i,j + 1} - t_{i,j} } \right)} {2H_{1} }}} \right. \kern-0pt} {2H_{1} }}\). Finally, the loss function for the group-level optimization problem was defined as the squared \(L_{2}\) norm of the average difference between \({\tilde{\mathbf{y}}}_{i}\) and \({\hat{\mathbf{y}}}_{i} \left( {{\varvec{\uptheta}}} \right)\), or equivalently formulated as follows:

$$\begin{gathered} \ell \left( {{{\varvec{\uptheta}}};\left\{ {\left( {t_{i,j} ,{\tilde{\mathbf{x}}}_{i,j} } \right)} \right\}_{j = 0}^{{m_{i} - 1}} ,H_{1} } \right) = \sum\limits_{j = 0}^{{m_{i} - 2}} {\left\| {{\tilde{\mathbf{y}}}_{i} - {\hat{\mathbf{y}}}_{i} \left( {{\varvec{\uptheta}}} \right)} \right\|_{2}^{2} } \\ \approx \left\| {{\text{Avg}}_{j} \left( {w_{i,j} \left\{ {\frac{{{\tilde{\mathbf{x}}}_{i,j + 1} - {\tilde{\mathbf{x}}}_{i,j} }}{{t_{i,j + 1} - t_{i,j} }} - {\text{Avg}}_{k} \left( {{\dot{\mathbf{x}}}\left( {\left\langle {\tau_{i,j}^{{\left( {2H_{1} } \right)}} } \right\rangle_{2k - 1} ;{{\varvec{\uptheta}}}} \right)} \right)} \right\}} \right)} \right\|_{2}^{2} . \\ \end{gathered}$$

Appendix B. Snapshot analysis

The snapshot analysis of tau tangle spreading encompassed two approaches: evaluating the relative importance of factors in the system and constructing vector fields to illustrate the proximal spreading patterns.

The relative influence of local reaction and diffusion is calculated for each node by dividing the absolute value of each term in Eq. (2) by their sum. Similarly, the relative influence of proximal and distal spreading on the diffusion of tau tangles was determined based on the decomposition \({\mathbf{L}} \cdot {\mathbf{x}}\left( t \right) = {\mathbf{L}}_{{{\text{prox}}}} \cdot {\mathbf{x}}\left( t \right) + {\mathbf{L}}_{{{\text{dist}}}} \cdot {\mathbf{x}}\left( t \right)\), where \({\mathbf{L}}_{{{\text{prox}}}}\) and \({\mathbf{L}}_{{{\text{dist}}}}\) represent weighted graph Laplacians of \({\mathbf{C}}_{{{\text{prox}}}}\) and \({\mathbf{C}}_{{{\text{dist}}}}\), respectively. The sign of the relative influence indicates whether tau SUVRs increase or decrease due to a specific term at that particular time.

The individualized vector field was constructed by assigning a gradient vector to each side of triangles within the pial surface mesh. For a triangle defined by vertices \(v_{1}\), \(v_{2}\), \(v_{3}\) and sides \(v_{1} v_{2}\), \(v_{2} v_{3}\), \(v_{3} v_{1}\), a gradient vector \(\kappa \cdot \sqrt {c_{{v_{1} v_{2} }} } \cdot \left( {\mathbf{x}_{{v_{1} }} \left( t \right) - \mathbf{x}_{{v_{2} }} \left( t \right)} \right) \cdot \left( {{\mathbf{v}}_{1} - {\mathbf{v}}_{2} } \right)\), where \({\mathbf{v}}_{1}\) and \({\mathbf{v}}_{2}\) represent the corresponding Cartesian coordinate points, was assigned to the side \(v_{1} v_{2}\). Similar gradient vectors were assigned to the sides \(v_{2} v_{3}\) and \(v_{3} v_{1}\). The representative vector of the triangle was determined as the Euclidean vector originating from its incenter and calculated as the average of the three gradient vectors of its sides.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kim, SW., Cho, H., Lee, Y. et al. Data-driven simulation of network-based tau spreading tailored to individual Alzheimer's patients. Engineering with Computers (2024). https://doi.org/10.1007/s00366-024-01988-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00366-024-01988-y

Keywords

Navigation