Skip to main content
Log in

Tucker Tensor Regression and Neuroimaging Analysis

  • Published:
Statistics in Biosciences Aims and scope Submit manuscript

Abstract

Neuroimaging data often take the form of high-dimensional arrays, also known as tensors. Addressing scientific questions arising from such data demands new regression models that take multidimensional arrays as covariates. Simply turning an image array into a vector would both cause extremely high dimensionality and destroy the inherent spatial structure of the array. In a recent work, Zhou et al. (J Am Stat Assoc, 108(502):540–552, 2013) proposed a family of generalized linear tensor regression models based upon the CP (CANDECOMP/PARAFAC) decomposition of regression coefficient array. Low-rank approximation brings the ultrahigh dimensionality to a manageable level and leads to efficient estimation. In this article, we propose a tensor regression model based on the more flexible Tucker decomposition. Compared to the CP model, Tucker regression model allows different number of factors along each mode. Such flexibility leads to several advantages that are particularly suited to neuroimaging analysis, including further reduction of the number of free parameters, accommodation of images with skewed dimensions, explicit modeling of interactions, and a principled way of image downsizing. We also compare the Tucker model with CP numerically on both simulated data and real magnetic resonance imaging data, and demonstrate its effectiveness in finite sample performance.

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

Similar content being viewed by others

References

  1. ADHD (2017) The ADHD-200 sample. http://fcon$XXSlahUndXX$1000.projects.nitrc.org/indi/adhd200/. Accessed Mar 2017

  2. ADNI (2017) Alzheimer’s disease neuroimaging initiative. http://adni.loni.ucla.edu. Accessed Mar 2017

  3. Caffo B, Crainiceanu C, Verduzco G, Joel S, Mostofsky SH, Bassett S, Pekar J (2010) Two-stage decompositions for the analysis of functional connectivity for fMRI with application to Alzheimer’s disease risk. Neuroimage 51(3):1140–1149

    Article  Google Scholar 

  4. Chen SS, Donoho DL, Saunders MA (2001) Atomic decomposition by basis pursuit. SIAM Rev. 43(1):129–159

    Article  MathSciNet  Google Scholar 

  5. de Leeuw J (1994) Block-relaxation algorithms in statistics. In: Information systems and data analysis. Springer, Berlin, pp 308–325

  6. Fan J, Li R (2001) Variable selection via nonconcave penalized likelihood and its oracle properties. J Am Stat Assoc 96(456):1348–1360

    Article  MathSciNet  Google Scholar 

  7. Frank IE, Friedman JH (1993) A statistical view of some chemometrics regression tools. Technometrics 35(2):109–135

    Article  Google Scholar 

  8. Friston K, Ashburner J, Kiebel S, Nichols T, Penny W (eds) (2007) Statistical parametric mapping: the analysis of functional brain images. Academic Press, London

    Google Scholar 

  9. Goldsmith J, Huang L, Crainiceanu C (2014) Smooth scalar-on-image regression via spatial bayesian variable selection. J Comput Graph Stat 23:46–64

    Article  MathSciNet  Google Scholar 

  10. Kolda TG, Bader BW (2009) Tensor decompositions and applications. SIAM Rev 51(3):455–500

    Article  MathSciNet  Google Scholar 

  11. Lange K (2004) Optimization. Springer texts in statistics. Springer, New York

    Google Scholar 

  12. Lange K (2010) Numerical analysis for statisticians. Statistics and computing, second edn. Springer, New York

    Book  Google Scholar 

  13. Lehmann EL, Romano JP (2005) Testing statistical hypotheses. Springer texts in statistics, third edn. Springer, New York

    Google Scholar 

  14. Li Y, Zhu H, Shen D, Lin W, Gilmore JH, Ibrahim JG (2011) Multiscale adaptive regression models for neuroimaging data. J R Stat Soc 73:559–578

    Article  MathSciNet  Google Scholar 

  15. Li F, Zhang T, Wang Q, Gonzalez M, Maresh E, Coan J (2015) Spatial Bayesian variable selection and grouping in high-dimensional scalar-on-image regressions. Ann Appl Stat (in press)

  16. McCullagh P, Nelder JA (1983) Generalized linear models. Monographs on statistics and applied probability. Chapman & Hall, London

    Book  Google Scholar 

  17. Reiss P, Ogden R (2010) Functional generalized linear models with images as predictors. Biometrics 66:61–69

    Article  MathSciNet  Google Scholar 

  18. Rothenberg TJ (1971) Identification in parametric models. Econometrica 39(3):577–91

    Article  MathSciNet  Google Scholar 

  19. Tibshirani R (1996) Regression shrinkage and selection via the lasso. J R Stat Soc 58(1):267–288

    MathSciNet  MATH  Google Scholar 

  20. Tucker LR (1966) Some mathematical notes on three-mode factor analysis. Psychometrika 31:279–311

    Article  MathSciNet  Google Scholar 

  21. van der Vaart AW (1998) Asymptotic statistics, volume 3 of Cambridge series in statistical and probabilistic mathematics. Cambridge University Press, Cambridge

    Google Scholar 

  22. Wang X, Nan B, Zhu J, Koeppe R (2014) Regularized 3D functional regression for brain image data via haar wavelets. Ann Appl Stat 8:1045–1064

    Article  MathSciNet  Google Scholar 

  23. Yue Y, Loh JM, Lindquist MA (2010) Adaptive spatial smoothing of fMRI images. Stat Interface 3:3–14

    Article  MathSciNet  Google Scholar 

  24. Zhou H, Li L (2014) Regularized matrix regression. J R Stat Soc 76:463–483

    Article  MathSciNet  Google Scholar 

  25. Zhou H, Li L, Zhu H (2013) Tensor regression with applications in neuroimaging data analysis. J Am Stat Assoc 108(502):540–552

    Article  MathSciNet  Google Scholar 

  26. Zou H, Hastie T (2005) Regularization and variable selection via the elastic net. J R Stat Soc 67(2):301–320

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lexin Li.

Appendix

Appendix

1.1 Proof of Lemma 1

We rewrite the array inner product

$$\begin{aligned} \langle {\varvec{B}},{\varvec{X}}\rangle= & {} \langle {\varvec{B}}_{(d)},{\varvec{X}}_{(d)} \rangle = \langle {\varvec{B}}_d {\varvec{G}}_{(d)} ({\varvec{B}}_D \otimes \cdots \otimes {\varvec{B}}_{d+1} \otimes {\varvec{B}}_{d-1} \otimes \cdots \otimes {\varvec{B}}_1) ^{\tiny {\text{ T }}}, {\varvec{X}}_{(d)} \rangle \\= & {} \langle {\varvec{G}}_{(d)}, {\varvec{B}}_d ^{\tiny {\text{ T }}}{\varvec{X}}_{(d)}({\varvec{B}}_D \otimes \cdots \otimes {\varvec{B}}_{d+1} \otimes {\varvec{B}}_{d-1} \otimes \cdots \otimes {\varvec{B}}_1) \rangle \\= & {} \langle {\varvec{G}}_{(d)}, \tilde{\varvec{X}}_{(d)} \rangle = \langle {\varvec{G}}, \tilde{\varvec{X}}\rangle , \end{aligned}$$

where the second and fourth equalities follow from (4) and the third follows from the invariance of trace function under cyclic permutation.

1.2 Proof of Proposition 1

It is easy to see that the block relaxation algorithm monotonically increases the objective values, i.e., \(\ell (\varvec{\theta }^{(t+1)}) \ge \ell (\varvec{\theta }^{(t)})\) for all \(t \ge 0\). Therefore its global convergence property follows from the standard theory for monotone algorithms [5, 11, 12]. Specifically global convergence is guaranteed under the following conditions: (i) \(\ell \) is coercive, (ii) the stationary points of \(\ell \) are isolated, (iii) the algorithmic mapping is continuous, (iv) \(\varvec{\theta }\) is a fixed point of the algorithm if and only if it is a stationarity point of \(\ell \), and (v) \(\ell (\varvec{\theta }^{(t+1)}) \ge \ell (\varvec{\theta }^{(t)})\) with equality if and only if \(\theta ^{(t)}\) is a fixed point of the algorithm. Condition (i) is guaranteed by the compactness of the set \(\{\varvec{\theta }: \ell (\varvec{\theta }) \ge \ell (\varvec{\theta }^{(0)})\). Condition (ii) is assumed. Condition (iii) follows from the strict concavity assumption and implicit function theorem. By Fermat’s principle, \(\varvec{\theta }= ({\varvec{G}}, {\varvec{B}}_1, \ldots , {\varvec{B}}_D)\) is a fixed point of the block relaxation algorithm if \(D\ell ({\varvec{G}}) = \mathbf{0}\) and \(D\ell ({\varvec{B}}_d)=\mathbf{0}\) for all d. Thus \(\varvec{\theta }\) is a fixed point if and only if it is a stationarity point of \(\ell \), i.e., condition (iv) is satisfied. Condition (v) follows from the monotonicity of the block relaxation algorithm. Local convergence follows from the classical Ostrowski theorem, which states that the algorithmic sequence \(\varvec{\theta }^{(t)}\) is local attracted to strictly local minimum \(\varvec{\theta }^{(\infty )}\) if the spectral radius of the differential of the algorithmic map \(\rho [dM(\varvec{\theta }^{(\infty )})]\) is strictly less than one. This follows from the strict concavity assumption of the block updates. See Zhou et al. [25] for more details.

1.3 Proof of Lemma 2

Assume \({\varvec{B}}\) admits the Tucker decomposition (3). By (4),

$$\begin{aligned} {\varvec{B}}_{(d)} = {\varvec{B}}_d {\varvec{G}}_{(d)} ({\varvec{B}}_D \otimes \cdots \otimes {\varvec{B}}_{d+1} \otimes {\varvec{B}}_{d-1} \otimes \cdots \otimes {\varvec{B}}_1) ^{\tiny {\text{ T }}}. \end{aligned}$$

Using the well-known fact that \(\mathrm {vec}({\varvec{X}}{\varvec{Y}}{\varvec{Z}}) = ({\varvec{Z}}^{\tiny {\text{ T }}}\otimes {\varvec{X}}) \mathrm {vec}({\varvec{Y}})\),

$$\begin{aligned} \mathrm {vec}{\varvec{B}}_{(d)} = [({\varvec{B}}_D \otimes \cdots \otimes {\varvec{B}}_{d+1} \otimes {\varvec{B}}_{d-1} \otimes \cdots \otimes {\varvec{B}}_1) {\varvec{G}}_{(d)} ^{\tiny {\text{ T }}}\otimes {\varvec{I}}_{p_d}] \mathrm {vec}({\varvec{B}}_d). \end{aligned}$$

Thus by the chain rule we have

$$\begin{aligned} {\varvec{J}}_d= & {} D{\varvec{B}}({\varvec{B}}_d) = D{\varvec{B}}({\varvec{B}}_{(d)}) \cdot D{\varvec{B}}_{(d)}({\varvec{B}}_d) = \varvec{\Pi }_d \frac{\partial \mathrm {vec}{\varvec{B}}_{(d)}}{\partial (\mathrm {vec}{\varvec{B}}_d) ^{\tiny {\text{ T }}}} \\= & {} \varvec{\Pi }_d \{[{\varvec{B}}_D \otimes \cdots \otimes {\varvec{B}}_{d+1} \otimes {\varvec{B}}_{d-1} \otimes \cdots \otimes {\varvec{B}}_1) {\varvec{G}}_{(d)} ^{\tiny {\text{ T }}}] \otimes {\varvec{I}}_{p_d} \}. \end{aligned}$$

Again by the chain rule, \(D\eta ({\varvec{B}}_d) = D\eta ({\varvec{B}}) \cdot D{\varvec{B}}({\varvec{B}}_d) = (\mathrm {vec}{\varvec{X}}) ^{\tiny {\text{ T }}}{\varvec{J}}_d\). For the derivative in \({\varvec{G}}\), the duality Lemma 1 implies \(\langle {\varvec{B}}, {\varvec{X}}\rangle = \langle {\varvec{G}}, \tilde{\varvec{X}}\rangle \) for \(\tilde{\varvec{X}}= \llbracket {\varvec{X}}; {\varvec{B}}_1 ^{\tiny {\text{ T }}}, \ldots , {\varvec{B}}_D ^{\tiny {\text{ T }}}\rrbracket \). Then, by (4), we have

$$\begin{aligned} D\eta ({\varvec{G}}) = (\mathrm {vec}\tilde{\varvec{X}}) ^{\tiny {\text{ T }}}= (\mathrm {vec}{\varvec{X}}) ^{\tiny {\text{ T }}}({\varvec{B}}_D \otimes \cdots \otimes {\varvec{B}}_1). \end{aligned}$$

Combining these results gives the gradient displayed in Lemma 2.

Next we consider the Hessian \(d^2 \eta \). Because \({\varvec{B}}\) is linear in \({\varvec{G}}\), the block \({\varvec{H}}_{{\varvec{G}},{\varvec{G}}}\) vanishes. For the block \({\varvec{H}}_{{\varvec{B}},{\varvec{B}}}\), the \((i_d,r_d,i_{d'},r_{d'})\)-entry is

$$\begin{aligned} h_{(i_d,r_d),(i_{d'},r_{d'})}&= \sum _{j_1,\ldots ,j_D} x_{j_1,\ldots ,j_D} \frac{\partial ^2 b_{j_1,\ldots ,j_D}}{\partial \beta _{i_d}^{(r)} \partial \beta _{i_{d'}}^{(r')}} \\&= \sum _{j_1,\ldots ,j_D} x_{j_1,\ldots ,j_D} \sum _{s_1,\ldots ,s_D} g_{s_1,\ldots ,s_D} \frac{\partial ^2 \beta _{j_1}^{(s_1)} \cdots \beta _{j_D}^{(s_D)}}{\partial \beta _{i_d}^{(r)} \partial \beta _{i_{d'}}^{(r')}}. \end{aligned}$$

The second derivative in the summand is nonzero only if \(j_d=i_d\), \(j_{d'}=i_{d'}\), \(s_d=r_d\), \(s_{d'}=r_{d'}\), and \(d \ne d'\). Therefore

$$\begin{aligned} h_{(i_d,r_d),(i_{d'},r_{d'})}&= 1_{\{d\ne d'\}} \sum _{j_d=i_d,j_{d'}=i_{d'}} x_{j_1,\ldots ,j_D} \sum _{s_d=r_d,s_{d'}=r_{d'}} g_{s_1,\ldots ,s_D} \prod _{d''\ne d, d'} \beta _{j_{d''}}^{(s_{d''})}. \end{aligned}$$

The first sum is over \(\prod _{d''\ne d,d'}p_{d''}\) terms and the second term is over \(\prod _{d''\ne d,d'} R_{d''}\) terms. A careful inspection reveals that the sub-block \({\varvec{H}}_{dd'}\) shares the same entries as the matrix

$$\begin{aligned} {\varvec{X}}_{(dd')} ({\varvec{B}}_D \otimes \cdots \otimes {\varvec{B}}_{d+1} \otimes {\varvec{B}}_{d-1} \otimes \cdots \otimes {\varvec{B}}_{d'+1} \otimes {\varvec{B}}_{d'-1} \otimes \cdots \otimes {\varvec{B}}_1) {\varvec{G}}_{(dd')} ^{\tiny {\text{ T }}}. \end{aligned}$$

Finally, for the \({\varvec{H}}_{{\varvec{G}},{\varvec{B}}}\) block, the \(\{(r_1,\ldots ,r_D),(i_d,r_d)\}\)-entry is

$$\begin{aligned} h_{(r_1,\ldots ,r_D),(i_{d},s_{d})}= & {} \sum _{j_1,\ldots ,j_D} x_{j_1,\ldots ,j_D} \frac{\partial ^2 b_{j_1,\ldots ,j_D}}{\partial g_{r_1,\ldots ,r_D} \partial \beta _{i_d}^{(s_d)}} \\= & {} \sum _{j_1,\ldots ,j_D} x_{j_1,\ldots ,j_D} \sum _{t_1,\ldots ,t_D} \frac{\partial ^2 g_{t_1,\ldots ,t_D} \beta _{j_1}^{(t_1)} \cdots \beta _{j_D}^{(t_D)}}{\partial g_{r_1,\ldots ,r_D} \partial \beta _{i_{d}}^{(s_d)}} \\= & {} \sum _{j_1,\ldots ,j_D} x_{j_1,\ldots ,j_D} \frac{\partial \beta _{j_1}^{(r_1)} \cdots \beta _{j_D}^{(r_D)}}{\partial \beta _{i_{d}}^{(s_d)}} \\= & {} 1_{\{r_d=s_d\}} \sum _{j_d=i_d} x_{j_1,\ldots ,j_D} \prod _{d'\ne d} \beta _{j_{d'}}^{(r_{d'})}, \end{aligned}$$

where the sum is over \(\prod _{d' \ne d} p_{d'}\) terms. The sub-block \({\varvec{H}}_d \in \mathrm {I \! R} ^{\prod _d R_d \times p_dR_d} \) has at most \(p_d \prod _d R_d\) nonzero entries. A close inspection suggests that the nonzero entries coincide with those in the matrix

$$\begin{aligned} {\varvec{X}}_{(d)} ({\varvec{B}}_D \otimes \cdots \otimes {\varvec{B}}_{d+1} \otimes {\varvec{B}}_{d-1} \otimes \cdots \otimes {\varvec{B}}_1). \end{aligned}$$

1.4 Proof of Proposition 2

Since \(\mu = b'(\theta )\), \(d\mu /d\theta = b''(\theta ) = \sigma ^2 /a(\phi )\) and

$$\begin{aligned} \nabla \ell ({\varvec{G}}, {\varvec{B}}_1,\ldots ,{\varvec{B}}_D)&= \frac{y - b'(\theta )}{a(\phi )} \frac{d\theta }{d\mu } \frac{d\mu }{d\eta } \nabla \eta ({\varvec{G}}, {\varvec{B}}_1,\ldots ,{\varvec{B}}_D) \\&= \frac{(y - \mu ) \mu '(\eta )}{\sigma ^2} [{\varvec{B}}_D \otimes \cdots \otimes {\varvec{B}}_1 \, {\varvec{J}}_1 \ldots {\varvec{J}}_D] ^{\tiny {\text{ T }}}(\mathrm {vec}{\varvec{X}}) \end{aligned}$$

by Lemma 2. Further differentiating shows

$$\begin{aligned}&d^2 \ell ({\varvec{G}}, {\varvec{B}}_1,\ldots ,{\varvec{B}}_D) \\&\quad = - \frac{1}{\sigma ^2} \nabla \mu ({\varvec{G}}, {\varvec{B}}_1,\ldots ,{\varvec{B}}_D) d\mu ({\varvec{G}}, {\varvec{B}}_1,\ldots ,{\varvec{B}}_D) + \frac{y - \mu }{\sigma ^2} d^2 \mu ({\varvec{G}}, {\varvec{B}}_1,\ldots ,{\varvec{B}}_D) \\&\quad = - \frac{[\mu '(\eta )]^2}{\sigma ^2} ([{\varvec{B}}_D \otimes \cdots \otimes {\varvec{B}}_1 \, {\varvec{J}}_1 \ldots {\varvec{J}}_D] ^{\tiny {\text{ T }}}\mathrm {vec}{\varvec{X}}) ([{\varvec{B}}_D\\&\qquad \otimes \cdots \otimes {\varvec{B}}_1 \, {\varvec{J}}_1 \ldots {\varvec{J}}_D] ^{\tiny {\text{ T }}}\mathrm {vec}{\varvec{X}}) ^{\tiny {\text{ T }}}\\&\qquad + \frac{(y-\mu )\theta ''(\eta )}{\sigma ^2} ([{\varvec{B}}_D \otimes \cdots \otimes {\varvec{B}}_1 \, {\varvec{J}}_1 \ldots {\varvec{J}}_D] ^{\tiny {\text{ T }}}\mathrm {vec}{\varvec{X}}) ([{\varvec{B}}_D\\&\qquad \otimes \cdots \otimes {\varvec{B}}_1 \, {\varvec{J}}_1 \ldots {\varvec{J}}_D] ^{\tiny {\text{ T }}}\mathrm {vec}{\varvec{X}}) ^{\tiny {\text{ T }}}\\&\qquad + \frac{(y-\mu )\theta '(\eta )}{\sigma ^2} d^2 \eta ({\varvec{B}}). \end{aligned}$$

It is easy to see that \(\mathbf {E}[\nabla \ell ({\varvec{G}}, {\varvec{B}}_1,\ldots ,{\varvec{B}}_D)] = \mathbf{0}\). Moreover, \(\mathbf {E}[-d^2\ell ({\varvec{G}}, {\varvec{B}}_1,\ldots ,{\varvec{B}}_D)] = {\varvec{I}}({\varvec{G}}, {\varvec{B}}_1,\ldots ,{\varvec{B}}_D)\). Then (8) follows.

1.5 Proof of Proposition 3

The proof follows from a classical result [18] that states that, if \(\theta _0\) be a regular point of the information matrix \(I(\theta )\), then \(\theta _0\) is locally identifiable if and only if \(I(\theta _0)\) is nonsingular. The regularity assumptions are satisfied by Tucker regression model: (1) the parameter space \(\mathcal {{\varvec{B}}}\) is open, (2) the density \(p(y,{\varvec{x}}|{\varvec{B}})\) is proper for all \({\varvec{B}}\in \mathcal {{\varvec{B}}}\), (3) the support of the density \(p(y,{\varvec{x}}|{\varvec{B}})\) is same for all \({\varvec{B}}\in \mathcal {{\varvec{B}}}\), (4) the log-density \(\ell ({\varvec{B}}|y,{\varvec{x}}) = \ln p(y,{\varvec{x}}|{\varvec{B}})\) is continuously differentiable, and (5) the information matrix

$$\begin{aligned} {\varvec{I}}({\varvec{B}})&= [{\varvec{B}}_D \otimes \cdots \otimes {\varvec{B}}_1 \, {\varvec{J}}_1 \ldots {\varvec{J}}_D] ^{\tiny {\text{ T }}}\left[ \sum _{i=1}^n \frac{\mu '(\eta _i)^2}{\sigma _i^2} (\mathrm {vec}\, {\varvec{x}}_i) (\mathrm {vec}\, {\varvec{x}}_i) ^{\tiny {\text{ T }}}\right] [{\varvec{B}}_D \\&\quad \otimes \cdots \otimes {\varvec{B}}_1 \, {\varvec{J}}_1 \ldots {\varvec{J}}_D] \end{aligned}$$

is continuous in \({\varvec{B}}\) by Proposition 2. Therefore \({\varvec{B}}\in \mathcal {{\varvec{B}}}\) is locally identifiable if and only if \({\varvec{I}}({\varvec{B}})\) is nonsingular.

1.6 Proof of Theorem 1

The asymptotics for tensor regression follow from the standard theory of M-estimation. The key observation is that the nonlinear part of tensor model (4) is a degree-\((D+1)\) polynomial of parameters \({\varvec{G}}\) and \({\varvec{B}}_d\) and the collection of polynomials \(\{\langle {\varvec{B}}, {\varvec{X}}\rangle , {\varvec{B}}\in \mathcal {{\varvec{B}}}\}\) form a Vapnik–C̆ervonenkis (VC) class. Then the classical uniform convergence theory applies [21]. The arguments in [25] extends the classical argument for GLM [21, Example 5.40] to the CP tensor regression model. The same proof also applies to the Tucker model with little changes and thus is omitted here. For the asymptotic normality, we need to establish that the log-likelihood function of Tucker regression model is quadratic mean differentiable (q.m.d.) [13]. By a well-known result [13, Theorem 12.2.2] or [21, Lemma 7.6], it suffices to verify that the density is continuously differentiable in parameter for \(\mu \)-almost all x and that the Fisher information matrix exists and is continuous. The derivative of density is

$$\begin{aligned} \nabla p({\varvec{B}}_1,\ldots ,{\varvec{B}}_D) = \nabla e^{\ell ({\varvec{B}}_1,\ldots ,{\varvec{B}}_D)} = p({\varvec{B}}_1,\ldots ,{\varvec{B}}_D) \nabla \ell ({\varvec{B}}_1,\ldots ,{\varvec{B}}_D), \end{aligned}$$

which is well defined and continuous by Proposition 2. The same proposition shows that the information matrix exists and is continuous. Therefore the Tucker regression model is q.m.d. and the asymptotic normality follows from the classical result for q.m.d. families [21, Theorem 5.39].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, X., Xu, D., Zhou, H. et al. Tucker Tensor Regression and Neuroimaging Analysis. Stat Biosci 10, 520–545 (2018). https://doi.org/10.1007/s12561-018-9215-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12561-018-9215-6

Keywords

Navigation