\(\def\loading{......LOADING......Please Wait} \def\RR{\bf R} \def\real{\mathbb{R}} \def\bold#1{\bf #1} \def\d{\mbox{Cord}} \def\hd{\widehat \mbox{Cord}} \DeclareMathOperator{\cov}{cov} \DeclareMathOperator{\var}{var} \DeclareMathOperator{\cor}{cor} \newcommand{\ac}[1]{\left\{#1\right\}} \DeclareMathOperator{\Ex}{\mathbb{E}} \DeclareMathOperator{\diag}{diag} \newcommand{\bm}[1]{\boldsymbol{#1}} \def\wait{......LOADING......Please Wait}\)

Some Machine Learning Methods for NeuroImaging Analysis


Xi (Rossi) LUO


The University of Texas
Health Science Center
School of Public Health
Dept of Biostatistics
and Data Science


Houston, Texas
September 17, 2019

Funding: NIH R01EB022911, P01AA019072, P20GM103645, P30AI042853; NSF/DMS (BD2K) 1557467

Slides viewable on web:


BigComplexData.com

Overall Theme

  • Machine learning: develop new models
  • Optimization: fast and large-scale computation
  • Theory: some theoretic properties
  • Neuroimaging: big and complex data

Granger Mediation Analysis

Co-Author

Yi Zhao

Yi Zhao
Tenure-track Assistant Prof at Indiana Univ

fMRI Experiments

  • Task fMRI: performs tasks under brain scanning
  • Randomized stop/go task:
    • press button if "go";
    • withhold pressing if "stop"
  • Not resting-state: "do nothing" during scanning


Goal: infer brain activation and connectivity

fMRI data: blood-oxygen-level dependent (BOLD) signals from each cube/voxel (~millimeters), $10^5$ ~ $10^6$ voxels in total.

Multilevel fMRI Studies

Sub 1, Sess 1

Time 1

2

~200


Sub i, Sess j

Sub ~100, Sess ~4



Large, multilevel (subject, session, voxel) data
e.g. $1000 \times 4 \times 300 \times 10^6 \approx 1 $ trillion data points

Raw Data: Motor Region


Time (seconds)

Black: fMRI BOLD activity
Blue: stop onset times; Red: go onset times

$Z_t$: Stimulus onsets convoluted with Canonical HRF

$M_t$, $R_t$: fMRI time series from two brain regions

Review: Granger Causality/VAR

  • Given two (or more) time series $x_t$ and $y_t$ $$\begin{align*} x_t &= \sum_{j=1}^p \psi_{1j} x_{t-j} + \sum_{j=1}^p \phi_{1j} y_{t-j} + \epsilon_{1t} \\ y_t &= \sum_{j=1}^p \psi_{2j} y_{t-j} + \sum_{j=1}^p \phi_{2j} x_{t-j} + \epsilon_{2t} \end{align*}$$
  • Also called vector autoregressive models
  • $y$ Granger causes $x$ if $\phi_{1j} \ne 0$ Granger, 69
  • Models pair-wise connections not pathways

Granger Causality/VAR

  • Granger Causality (VAR) popular for fMRI
    • Over 10,000 google scholar results on "granger causality neuroimaging", as of May 29, 2019
  • Models multiple stationary time series
  • AR($p$) (small $p$) fits fMRI well Lingdquist, 08
  • Not for non-stationary/task fMRI
  • Cannot model stimulus effects

Conceptual Brain Model with Stimulus

Goal: quantify effects stimulipreSMAPMC regions Duann, Ide, Luo, Li (2009). J of Neurosci

Model: Mediation Analysis and SEM

$$\begin{align*}M &= Z a + \overbrace{U + \epsilon_1}^{E_1}\\ R &= Z c + M b + \underbrace{U g + \epsilon_2}_{E_2}, \quad \epsilon_1 \bot \epsilon_2\end{align*}$$
  • Indirect effect: $a \times b$; Direct effect: $c$
  • Correlated errors: $\delta = \cor(E_1, E_2) \ne 0$ if $U\ne 0$

Mediation Analysis in fMRI

  • Mediation analysis (usually assuming $U=0$)Baron&anp;Kenny, 86; Sobel, 82; Holland 88; Preacher&Hayes 08; Imai et al, 10; VanderWeele, 15;...
  • Parametric Wager et al, 09 and functional Lindquist, 12 mediation, under (approx.) independent errors
    • Stimulus $\rightarrow$ brain $\rightarrow$ user reported ratings, one brain mediator
    • Assuming $U=0$ between ratings and brain
  • Multiple mediator and multiple pathways
    • Dimension reduction by arXiv1511.09354Chen, Crainiceanu, Ogburn, Caffo, Wager, Lindquist, 15
    • Pathway Lasso penalization Zhao, Luo, 16
  • This talk: integrating Granger causality and mediation analysis

Model & Method

Our Mediation Model

$$\begin{align*}M_{t} = Z_{t} a + E_{1t},\quad R_t = Z_t c + M_t b + E_{2t}\end{align*}$$
  • Temporal VAR errors $$\begin{align*} E_{1t}&=& \sum_{j=1}^{p}\left(\omega_{11_{j}}E_{1,t-j}+\omega_{21_{j}}E_{2,t-j}\right)+\epsilon_{1t} \\ E_{2t}&=&\sum_{j=1}^{p}\left(\omega_{12_{j}}E_{1,t-j}+\omega_{22_{j}}E_{2,t-j}\right)+\epsilon_{2t} \end{align*}$$
  • Spatial errors: $\epsilon_{1t}, \epsilon_{2t}$ $$ \begin{pmatrix} \epsilon_{1t} \\ \epsilon_{2t} \end{pmatrix}\sim\mathcal{N}\left(\boldsymbol{\mathrm{0}},\boldsymbol{\Sigma}\right), \quad \boldsymbol{\Sigma}=\begin{pmatrix} \sigma_{1}^{2} & \delta\sigma_{1}\sigma_{2} \\ \delta\sigma_{1}\sigma_{2} & \sigma_{2}^{2} \end{pmatrix} $$

Equivalent Form

$$\begin{align*} \scriptsize M_{t}& \scriptsize =Z_{t}A+\sum_{j=1}^{p}\left(\phi_{1j}Z_{t-j}+\psi_{11_{j}}M_{t-j}+\psi_{21_{j}}R_{t-j}\right)+\epsilon_{1t} \\ \scriptsize R_{t}& \scriptsize =Z_{t}C+M_{t}B+\sum_{j=1}^{p}\left(\phi_{2j}Z_{t-j}+\psi_{12_{j}}M_{t-j}+\psi_{22_{j}}R_{t-j}\right)+\epsilon_{2t} \end{align*} $$
  • Nonzero $\phi$'s and $\psi$'s denote the temporal influence from stimulus to mediator/outcome and etc
  • $A$, $B$, $C$ are causal following a similar proof in Sobel, Lindquist, 04

Causal Conditions

  • The treatment randomization regime is the same across time and participants.
  • Models are correctly specified, and no treatment-mediator interaction.
  • At each time point $t$, the observed outcome is one realization of the potential outcome with observed treatment assignment $\mathbf{\bar S}_{t}$, where $\mathbf{ \bar S}_{t}=( \mathbf{S}_{1},\dots,\mathbf{S}_{t})$.
  • The treatment assignment is random across time.
  • The causal effects are time-invariant.
  • The time-invariant covariance matrix is not affected by the treatment assignments.

Estimation: Conditional Likelihood

  • The full likelihood for our model is too complex
  • Given the initial $p$ time points, the conditional likelihood is $$ \begin{align*} & \tiny \ell\left(\boldsymbol{\Theta},\delta~|~\mathbf{Z},\mathcal{I}_{p}\right) = \sum_{t=p+1}^{T}\log f\left((M_{t},R_{t})~|~\mathbf{X}_{t}\right) \\ & \tiny = -\frac{T-p}{2}\log\sigma_{1}^{2}\sigma_{2}^{2}(1-\delta^{2})-\frac{1}{2\sigma_{1}^{2}}\|\mathbf{M}-\mathbf{X}\boldsymbol{\theta}_{1}\|_{2}^{2} \\ & \tiny -\frac{1}{2\sigma_{2}^{2}(1-\delta^{2})}\|(\mathbf{R}-\mathbf{M}B-\mathbf{X}\boldsymbol{\theta}_{2})-\kappa(\mathbf{M}-\mathbf{X}\boldsymbol{\theta}_{1})\|_{2}^{2} \end{align*} $$

Multilevel Data: Two-level Likelihood

  • Second level model, for each subject $i$ $$(A_i,B_i,C_i) = (A,B,C) + (\eta^A_i, \eta^B_i, \eta^C_i)$$ where errors $\eta$ are normally distributed
  • The two level likelihood is conditional convex
  • Two-stage fitting: plug-in estimates from the first level
  • Block coordinate fitting: jointly optimize first level likelihood + second level likelihood
Theorem: Assume assumptions (A1)-(A6) are satisfied. Assume $\mathbb{E}(Z_{i_{t}}^{2})=q\lt \infty$, for $i=1,\dots,N$. Let $T=\min_{i}T_{i}$.
1. If $\boldsymbol{\Lambda}$ is known, then the two-stage estimator $\hat{\delta}$ maximizes the profile likelihood of model asymptotically, and $\hat{\delta}$ is $\sqrt{NT}$-consistent.
2. If $\boldsymbol{\Lambda}$ is unknown, then the profile likelihood of model has a unique maximizer $\hat{\delta}$ asymptotically, and $\hat{\delta}$ is $\sqrt{NT}$-consistent, provided that $1/\varpi=\bar{\kappa}^{2}/\varrho^{2}=\mathcal{O}_{p}(1/\sqrt{NT})$, $\kappa_{i}=\sigma_{i_{2}}/\sigma_{i_{1}}$, $\bar{\kappa}=(1/N)\sum\kappa_{i}$, and $\varrho^{2}=(1/N)\sum(\kappa_{i}-\bar{\kappa})^{2}$.
Using the two-stage estimator $\hat{\delta}$, the CMLE of our model is consistent, as well as the estimator for $\mathbf{b}=(A,B,C)$.

Theory: Summary

  • Under regularity conditions, $N$ subs, $T$ time points
  • Our $\hat \delta$ is $\sqrt{NT}$-consistent
    • This relaxes the unmeasured confounding assumption in mediation analysis
  • Our $(\hat{A},\hat{B}, \hat{C})$ is also consistent

Simulations & Real Data

Comparison

  • Our methods: GMA-h and GMA-ts
  • Previous methods: BK Baron & Kenny, MACC Zhao and Luo, KKB Kenny et al
  • Other methods do not model the temporal correlations or time series like ours

Simulations

Low bias for $AB$

Low bias for temporal cor

Gray dash lines are the truth

GMA performs the best, and recovers the temporal correlations

Real Data Experiment

  • Public data: OpenFMRI ds30
  • Stop-go experiment: withhold (STOP) from pressing buttons
  • Expect "STOP" stimuli to deactivate brain region M1
  • Goal: quantify the role of region preSMA

Result

Result

  • STOP deactivates M1 directly ($C$) and indirectly ($AB$)
  • preSMA mediates a good portion of the total effect
    • Help resolve the debates among neuroscientists
  • Other methods under-estiamte the effects
  • Novel feedback findings: M1 → preSMA after lag 1 and 2 (not shown)

Discussion

  • Mediation analysis for multiple time series
  • Method: Granger causality + mediation
    • Optimizing complex likelihood
  • Theory: identifiability, consistency
  • Result: low bias and improved accuracy
  • Extension: functional mediation
  • Paper in Biometrics 2019
  • CRAN pkg: gma and references within

Covariate Assisted Principal regression

Co-Authors

Xuefei Cao

Yi Zhao
Indiana Univ Biostat

B Sandstede

Bingkai Wang
Johns Hopkins Biostat

B Sandstede

Stewart Mostofsky
Johns Hopkins Medicine

B Sandstede

Brian Caffo
Johns Hopkins Biostat

Statistics/Data Science Focuses

Motivating Example

Brain network connections vary by covariates (e.g. age/sex)


Goal: model how covariates change network connections

$$\textrm{function}(\textbf{graph}) = \textbf{age}\times \beta_1 + \textbf{sex}\times \beta_2 + \cdots $$

Resting-state fMRI Networks

  • fMRI measures brain activities over time
  • Resting-state: "do nothing" during scanning


  • Brain networks constructed using cov/cor matrices of time series

Mathematical Problem

  • Given $n$ (semi-)positive matrix outcomes, $\Sigma_i\in \real^{p\times p}$
  • Given $n$ corresponding vector covariates, $x_i \in \real^{q}$
  • Find function $g(\Sigma_i) = x_i \beta$, $i=1,\dotsc, n$
  • In essense, regress positive matrices on vectors

Some Related Problems

  • Heterogeneous regression or weighted LS:
    • Usually for scalar variance $\sigma_i$, find $g(\sigma_i) = f(x_i)$
    • Goal: to improve efficiency, not to interpret $x_i \beta$
  • Covariance models Anderson, 73; Pourahmadi, 99; Hoff, Niu, 12; Fox, Dunson, 15; Zou, 17
    • Model $\Sigma_i = g(x_i)$, sometimes $n=i=1$
    • Goal: better models for $\Sigma_i$
  • Multi-group PCA Flury, 84, 88; Boik 02; Hoff 09; Franks, Hoff, 16
    • No regression model, cannot handle vector $x_i$
    • Goal: find common/uncommon parts of multiple $\Sigma_i$
  • Tensor-on-scalar regression Li, Zhang, 17; Sun, Li, 17
    • No guarantees for positive matrix outcomes

Massive Edgewise Regressions

  • Intuitive method by mostly neuroscientists
  • Try $g_{j,k}(\Sigma_i) = \Sigma_{i}[j,k] = x_i \beta$
  • Repeat for all $(j,k) \in \{1,\dotsc, p\}^2$ pairs
  • Essentially $O(p^2)$ regressions for each connection
  • Limitations: multiple testing $O(p^2)$, failure to accout for dependencies between regressions

Our CAP in a Nutshell


$\mbox{PCA}(\Sigma_i) = x_i \beta$

  • Essentially, we aim to turn unsupervised PCA to a supervised PCA
  • Ours differs from existing PCA methods:
    • Supervised PCA Bair et al, 06 models scalar-on-vector

Model and Method

Model

  • Find principal direction (PD) $\gamma \in \real^p$, such that: $$ \log({\gamma}^\top\Sigma_{i}{\gamma})=\beta_{0}+x_{i}^\top{\beta}_{1}, \quad i =1,\dotsc, n$$

Example (p=2): PD1 largest variation but not related to $x$

PCA selects PD1, Ours selects PD2

Advantages

  • Scalability: potentially for $p \sim 10^6$ or larger
  • Interpretation: covariate assisted PCA
    • Turn unsupervised PCA into supervised
  • Sensitivity: target those covariate-related variations
    • Covariate assisted SVD?
  • Applicability: other big data problems besides fMRI

Method

  • MLE with constraints: $$\scriptsize \begin{eqnarray}\label{eq:obj_func} \underset{\boldsymbol{\beta},\boldsymbol{\gamma}}{\text{minimize}} && \ell(\boldsymbol{\beta},\boldsymbol{\gamma}) := \frac{1}{2}\sum_{i=1}^{n}(x_{i}^\top\boldsymbol{\beta}) \cdot T_{i} +\frac{1}{2}\sum_{i=1}^{n}\boldsymbol{\gamma}^\top \Sigma_{i}\boldsymbol{\gamma} \cdot \exp(-x_{i}^\top\boldsymbol{\beta}) , \nonumber \\ \text{such that} && \boldsymbol{\gamma}^\top H \boldsymbol{\gamma}=1 \end{eqnarray}$$
  • Two obvious constriants:
    • C1: $H = I$
    • C2: $H = n^{-1} (\Sigma_1 + \cdots + \Sigma_n) $

Choice of $H$

Proposition: When (C1) $H=\boldsymbol{\mathrm{I}}$ in the optimization problem, for any fixed $\boldsymbol{\beta}$, the solution of $\boldsymbol{\gamma}$ is the eigenvector corresponding to the minimum eigenvalue of matrix $$ \sum_{i=1}^{n}\frac{\Sigma_{i}}{\exp(x_{i}^\top\boldsymbol{\beta})} $$

Will focus on the constraint (C2)

Algoirthm

  • Iteratively update $\beta$ and then $\gamma$
  • Prove explicit updates
  • Extension to multiple $\gamma$:
    • After finding $\gamma^{(1)}$, we will update $\Sigma_i$ by removing its effect
    • Search for the next PD $\gamma^{(k)}$, $k=2, \dotsc$
    • Impose the orthogonal constraints such that $\gamma^{k}$ is orthogonal to all $\gamma^{(t)}$ for $t\lt k$

Theory for $\beta$

Theorem: Assume $\sum_{i=1}^{n}x_{i}x_{i}^\top/n\rightarrow Q$ as $n\rightarrow\infty$. Let $T=\min_{i}T_{i}$, $M_{n}=\sum_{i=1}^{n}T_{i}$, under the true $\boldsymbol{\gamma}$, we have \begin{equation} \sqrt{M_{n}}\left(\hat{\boldsymbol{\beta}}-\boldsymbol{\beta}\right)\overset{\mathcal{D}}{\longrightarrow}\mathcal{N}\left(\boldsymbol{\mathrm{0}},2 Q^{-1}\right),\quad \text{as } n,T\rightarrow\infty, \end{equation} where $\hat{\boldsymbol{\beta}}$ is the maximum likelihood estimator when the true $\boldsymbol{\gamma}$ is known.

Theory for $\gamma$

Theorem: Assume $\Sigma_{i}=\Gamma\Lambda_{i}\Gamma^\top$, where $\Gamma=(\boldsymbol{\gamma}_{1},\dots,\boldsymbol{\gamma}_{p})$ is an orthogonal matrix and $\Lambda_{i}=\mathrm{diag}\{\lambda_{i1},\dots,\lambda_{ip}\}$ with $\lambda_{ik}\neq\lambda_{il}$ ($k\neq l$), for at least one $i\in\{1,\dots,n\}$. There exists $k\in\{1,\dots,p\}$ such that for $\forall~i\in\{1,\dots,n\}$, $\boldsymbol{\gamma}_{k}^\top\Sigma_{i}\boldsymbol{\gamma}_{k}=\exp(x_{i}^\top\boldsymbol{\beta})$. Let $\hat{\boldsymbol{\gamma}}$ be the maximum likelihood estimator of $\boldsymbol{\gamma}_{k}$ in Flury, 84. Then assuming that the assumptions are satisfied, $\hat{ \boldsymbol{\beta}}$ from our algorithm is $\sqrt{M_{n}}$-consistent estimator of $\boldsymbol{\beta}$.

Simulations


PCA and common PCA do not find the first principal direction, because they don't model covariates

Resting-state fMRI

Regression Coefficients


Age

Sex

Age*Sex



No statistical significant changes were found by massive edgewise regression

Statistical Power

Brain Map of $\gamma$

Discussion

  • Regress matrices on vectors
  • Method to identify covariate-related directions
  • Theorectical justification
  • Manuscript: DOI: 10.1101/425033 (Revision for Biostatistics)
  • R pkg: cap

Thank you!


Comments? Questions?




BigComplexData.com

or BrainDataScience.com