• Keine Ergebnisse gefunden

Sparse principal component regression via singular value decomposition approach

N/A
N/A
Protected

Academic year: 2022

Aktie "Sparse principal component regression via singular value decomposition approach"

Copied!
29
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

https://doi.org/10.1007/s11634-020-00435-2 REGULAR ARTICLE

Sparse principal component regression via singular value decomposition approach

Shuichi Kawano1

Received: 17 March 2020 / Revised: 30 November 2020 / Accepted: 11 December 2020 / Published online: 8 February 2021

© The Author(s) 2021

Abstract

Principal component regression (PCR) is a two-stage procedure: the first stage per- forms principal component analysis (PCA) and the second stage builds a regression model whose explanatory variables are the principal components obtained in the first stage. Since PCA is performed using only explanatory variables, the principal com- ponents have no information about the response variable. To address this problem, we present a one-stage procedure for PCR based on a singular value decomposition approach. Our approach is based upon two loss functions, which are a regression loss and a PCA loss from the singular value decomposition, with sparse regularization.

The proposed method enables us to obtain principal component loadings that include information about both explanatory variables and a response variable. An estimation algorithm is developed by using the alternating direction method of multipliers. We conduct numerical studies to show the effectiveness of the proposed method.

Keywords ADMM·Lasso·One-stage procedure·Singular value decomposition· Principal component analysis

Mathematics Subject Classification 62H25·62J07·62J05 1 Introduction

Principal component regression (PCR), invented by Jolliffe (1982) and Massy (1965), is widely used in various fields of research, including chemometrics, bioinformatics, and psychology, and has been extensively studied (Chang and Yang 2012; Dicker

Supplementary Information The online version contains supplementary material available athttps://doi.

org/10.1007/s11634-020-00435-2.

B

Shuichi Kawano skawano@ai.lab.uec.ac.jp

1 Department of Computer and Network Engineering, Graduate School of Informatics and Engineering, The University of Electro-Communications, 1-5-1 Chofugaoka, Chofu, Tokyo 182-8585, Japan

(2)

et al. 2017; Febrero-Bande et al.2017; Frank and Friedman 1993; Hartnett et al.

1998; Reiss and Ogden2007; Rosipal et al.2001; Wang and Abbott2008). PCR is a two-stage procedure: one first performs principal component analysis (PCA) (Jolliffe 2002; Pearson1901), and then performs regression in which the explanatory variables are the selected principal components. However, the principal components have no information on the response variable. Because of this, the prediction accuracy of the PCR could be low, if the response variable is related to principal components having small eigenvalues.

To address this problem, a one-stage procedure for PCR was proposed in Kawano et al. (2015). This one-stage procedure was developed by combining a regression squared loss function with the sparse PCA (SPCA) loss function in Zou et al. (2006).

The estimate of the regression parameter and loading matrix in the PCA is obtained as the minimizer of the combination of two loss functions with sparse regularization.

By virtue of sparse regularization, sparse estimates of the parameters can be obtained.

Kawano et al. (2015) referred to the one-stage procedure as sparse principal component regression (SPCR). Kawano et al. (2018) also extended SPCR within the framework of generalized linear models. However, it is unclear whether the PCA loss function in Zou et al. (2006) is the best choice for building SPCR, as there exist several formulae for PCA.

This paper proposes a novel formulation for SPCR. As a PCA loss for SPCR, we adopt a loss function based on a singular value decomposition approach (Shen and Huang2008). Using the basic loss function, a combination of the PCA loss and the regression squared loss, with sparse regularization, we derive an alternative formula- tion for SPCR. We call the proposed method as sparse principal component regression based on a singular value decomposition approach (SPCRsvd). An estimation algo- rithm of SPCRsvd is developed using an alternating direction method of multipliers (Boyd et al.2011) and a linearized alternating direction method of multipliers (Li et al.

2014; Wang and Yuan2012). We show the effectiveness of SPCRsvd through numeri- cal studies. Specifically, the performance of SPCRsvd is shown to be competitive with or better than that of SPCR.

As an alternative approach, partial least squares (PLS) (Frank and Friedman1993;

Wold1975) is a widely used statistical method that regresses a response variable on composite variables built by combining a response variable and explanatory variables.

In Chun and Kele¸s (2010), sparse partial least squares (SPLS) was proposed, which enables the removal of irrelevant explanatory variables when constructing the compos- ite variables. PLS and SPLS are similar to SPCR and SPCRsvd in terms of using new explanatory variables with information relating the response variable to the original explanatory variables. Herein, these methods are compared using simulated data and real data.

The remainder of the paper is organized as follows. In Sect.2, we review SPCA in Zou et al. (2006) and Shen and Huang (2008), and SPCR in Kawano et al. (2015).

We present SPCRsvd in Sect.3. Section4derives two computational algorithms for SPCRsvd and discusses the selection of tuning parameters. Monte Carlo simulations and real data analyses are presented in Sect.5. Conclusions are given in Sect.6.

(3)

2 Preliminaries

2.1 Sparse principal component analysis

PCA finds a loading matrix that induces a low-dimensional structure in the data. As an easy way to interpret the principal component loading matrix, SPCA has been proposed. To date, several formulae for SPCA have been proposed (Bresler et al.

2018; Chen et al. 2020; d’Aspremont et al. 2007; Erichson et al.2020; Shen and Huang2008; Vu et al.2013; Witten et al.2009; Zou et al.2006). For an overview of SPCA, we refer the reader to Zou and Xue (2018) and the references therein. In this subsection, we review the two formulae for SPCA in Zou et al. (2006) and Shen and Huang (2008).

LetX =(x1, . . . ,xn)denote ann×pdata matrix, wherenandpare the number of observations and the number of variables, respectively. Without loss of generality, we assume that the columns of the matrixXare centered. In Zou et al. (2006), SPCA was proposed as

minA,B

⎧⎨

n

i=1

xiA Bxi22+λ k

j=1

βj22+ k

j=1

λ1,jβj1

⎫⎬

subject to AA=Ik, (1)

where A = 1, . . . ,αk)and B = 1, . . . ,βk) are p ×k principal component (PC) loading matrices,kdenotes the number of principal components,Ik is thek× k identity matrix,λ, λ1,1, . . . , λ1,k are non-negative regularization parameters, and · q is theLq norm for an arbitrary finite vectors. This SPCA formulation can be regarded as a least squares approach. The first term represents performing PCA by least squares. The second and third terms represent sparse regularization similar to elastic net regularization (Zou and Hastie2005). These terms enable us to set some of the estimates of B to zero. Ifλ =0, then the regularization terms reduce to the adaptive lasso (Zou2006).

A simple calculation gives

minA,B

k j=1

jXβj22+λβj22+λ1,jβj1

subject to AA=Ik. (2)

Optimizing the parametersAandBfor this minimization problem is straightforward.

Given a fixed A, the SPCA problem (2) turns out to be a simple elastic net problem.

Thus, the estimate ofBcan be obtained by the least angle regression algorithm (Efron et al.2004) or the coordinate descent algorithm (Friedman et al.2007; Wu and Lange 2008). Given a fixed B, an estimate of Acan be obtained by solving the reduced rank Procrustes rotation problem (Zou et al.2006). By alternating procedures, we can obtain the final estimatesAˆandBˆ ofAandB, respectively. Note that onlyBˆ is used as the principal component loading matrix.

(4)

Alternately, Shen and Huang (2008) proposed another formulation of SPCA, which can be regarded as a singular value decomposition (SVD) approach. Consider a low- rank approximation of the data matrixXobtained by SVD in the form

U DV= r k=1

dkukvk, (3)

whereU = (u1, . . . ,ur)is ann×r matrix withUU = Ir,V =(v1, . . . ,vr)is anr×rorthogonal matrix,D=diag(d1, . . . ,dr), andr <min(n,p). The singular values are assumed to be ordered such thatdr ≥ · · · ≥dp≥0. Using the connection between PCA and SVD, Shen and Huang (2008) obtained the sparse PC loading by estimatingV with sparse regularization.

To achieve sparseness ofV, Shen and Huang (2008) adopted the rank-one approx- imation procedure. First, the first PC loading vector v˜1 is obtained by solving the minimization problem

min˜ u1v1

X− ˜u1v˜12F+λP(˜v1)

subject to ˜u12=1. (4)

Here u˜1,v˜1 are defined as rescaled vectors such that u˜1v˜1 = d1u1v1, P(·) is a penalty function that induces the sparsity of v˜1, and · F is the Frobenius norm defined byAF =

tr(AA)for an arbitrary matrix A. As the penalty function, Shen and Huang (2008) used the lasso penalty (Tibshirani1996), the hard-thresholding penalty (Donoho and Johnstone1994), or the smoothly clipped absolute deviation (SCAD) penalty (Fan and Li2001). The rank-one approximation problem is easy to solve (4); see Algorithm 1 in Shen and Huang (2008). The remaining PC loading vectors are obtained by performing rank-one approximations of the corresponding residual matrices. For example, to derive the second PC loading vectorv˜2, we solve the minimization problem

˜min

u2v2

X− ˜u2v˜22F+λP(˜v2)

subject to ˜u22=1,

whereX=X− ˜u1v˜1. The regularization parameterλis selected by cross-validation.

2.2 Sparse principal component regression

For a one-dimensional continuous response variableY and ap-dimensional explana- tory variable x, suppose we have obtained a dataset{(yi,xi);i = 1, . . . ,n}. We assume that the response variable is explained by variables composed by PCA of X =(x1, . . . ,xn). Traditional PCR uses a regression model with a few PC scores corresponding to large eigenvalues. Note that these PC scores are derived by PCA prior to the regression. This two-stage procedure might then fail to predict the response if the response variable is related to PCs corresponding to small eigenvalues.

(5)

To attain a one-stage procedure for PCR, the SPCR proposed in Kawano et al.

(2015) was formulated as the following minimization problem:

A,minB0

n i=1

yiγ0γBxi

2

+w n i=1

xiA Bxi22 +λβξ

k j=1

βj22+λβ(1−ξ) k

j=1

βj1+λγγ1

(5) subject to AA=Ik,

whereγ0is an intercept,γ =1, . . . , γk)comprises coefficients for regression,λβ andλγ are non-negative regularization parameters,wis a positive tuning parameter, andξ in[0,1]is a tuning parameter. The first term in Formula (5) is the regression squared loss function including the PCs Bx as explanatory variables, while the second term is the PCA loss function used in SPCA in Zou et al. (2006). Sparse regularization in SPCR has two roles: sparseness and identifiability of parameters. For the identifiability by sparse regularization, we refer the reader to Choi et al. (2010), Jennrich (2006), Kawano et al. (2015). Kawano et al. (2018) also extended SPCR from the viewpoint of generalized linear models, which can deal with binary, count, and multi-categorical data for the response variable.

3 SVD-based sparse principal component regression

SPCR uses two basic loss functions: the regression squared loss function and the PCA loss function in Zou et al. (2006). However, it is unclear whether the PCA loss is the best choice for building SPCR. To investigate this issue, we propose another formulation for SPCR using the SVD approach in Shen and Huang (2008).

We consider the following minimization problem:

β0,β,minZ,V

1

nyβ01nX Vβ22+w

nX−Z V2F+λVV1+λββ1

subject to VV =Ik, (6)

where β0 is an intercept,kis the number of PCs, β is ak-dimensional coefficient vector, Z is ann×kmatrix of PCs,V is a p×kPC loading matrix, and1nis an n-dimensional vector of ones. In addition,wis a positive tuning parameter andλV, λβ are non-negative regularization parameters.

The first term is the regression squared loss function relating the response and the PCs X V. The second term is the PCA loss function in the SVD approach in Shen and Huang (2008). Although the formula is seemingly different from the first term in Formula (4), they are essentially equivalent: we estimate thekPCs simultaneously, while Shen and Huang (2008) estimates them sequentially. The third and fourth terms constitute the lasso penalty that induces zero estimates of the parametersV andβ,

(6)

respectively. The tuning parameterwcontrols the degree of the second term. A smaller value for w is used when our aim is to obtain better prediction accuracies, while a larger value for w is used when we want to obtain exact expressions of the PC loadings. The minimization problem (6) allows us to perform regression analysis and PCA simultaneously. We call this method SPCRsvd. In Sect.5, we will observe that SPCRsvd is competitive with or better than SPCR through numerical studies.

We remark on two points here. First, it is possible to use Z in the first term of (6) instead ofX V, sinceZ is also the PCs. However, the formulation withZ instead of X V did not perform well in numerical studies, so we adopt the formulation with X V here. Second, SPCR imposes a ridge penalty for the PC loading but SPCRsvd does not. The ridge penalty basically comes from SPCA in Zou et al. (2006). Because SPCRsvd is not based on SPCA in Zou et al. (2006), a ridge penalty does not appear in Formula (6). It is possible to add a ridge penalty and replace the lasso penalty with other penalties that induce sparsity, e.g., the adaptive lasso penalty, the SCAD penalty, or minimax concave penalty (Zhang2010), but the our aim of this paper is to establish the basic procedure of Formula (6).

4 Implementation 4.1 Computational algorithm

To obtain the estimates of the parameters β,Z,V in Formula (6), we employ the alternating direction method of multipliers (ADMM) and the linearized alternating direction method of multipliers (LADMM). ADMM and LADMM have recently been used in various models with sparse regularization; see, for example, Boyd et al. (2011);

Danaher et al. (2014); Li et al. (2014); Ma and Huang (2017); Price et al. (2019); Tan et al. (2014); Wang et al. (2018); Yan and Bien (2020) and Ye and Xie (2011).

To solve the minimization problem (6) by using ADMM, we rewrite the problem as

β0,β,β0min,Z,V,V0,V1

1

nyβ01nX V1β22+w nX

Z V2F+λVV01+λββ01

subject to VV =Ik, V =V0=V1, β=β0. (7) The scaled augmented Lagrangian for the problem (7) is then given by

1

nyβ01nX V1β22+w

nX−Z V2F+λVV01+λββ01

+ρ1

2 V−V0+12F+ρ2

2 V1V0+22F+ρ3

2 β−β0+λ322

subject to VV =Ik,

(7)

where1, 2,λ3are dual variables andρ1, ρ2, ρ3(>0)are penalty parameters. This gives rise to the following ADMM algorithm:

Step 1 Set the values of the tuning parameterw, the regularization parametersλV, λβ, and the penalty parametersρ1, ρ2, ρ3.

Step 2 Initialize all the parameters asβ0(0),β(0),β(00),Z(0),V(0),V0(0),V1(0), (10), (20),λ(30).

Step 3 Form=0,1,2, . . ., repeat from Steps 4 to 11 until convergence.

Step 4 UpdateV1as follows:

vec(V1(m+1))= 1

nβ(m)β(m)XX+ρ2

2 IkIp

1

vec 1

nX(yβ0(m)1n(m) +ρ2

2 (V0(m)(2m))

, where⊗represents the Kronecker product.

Step 5 UpdateV as follows:

V(m+1)=P Q, where PandQare the matrices given by the SVD

w

n XZ(m)+ρ1

2

V0(m)(1m)

=PΩQ.

Step 6 UpdateV0as follows:

v0i j(m+1)=S

ρ1(vi j(m+1)+λ(1i jm))+ρ2(vi j(m+1)+λ(2i jm))

ρ1+ρ2 , λV

ρ1+ρ2

,

i=1, . . . ,p, j=1, . . . ,k,

where v(0i jm) = (V0(m))i j, vi j(m) = (V(m))i j, λi j ( = 1,2)is the (i,j)-th element of the matrix ( =1,2), andS(·,·)is the soft-thresholding operator defined byS(x, λ)=sign(x)(|x| −λ)+.

Step 7 UpdateZ byZ(m+1)=X V(m+1). Step 8 Updateβas follows:

β(m+1)= 1

nV1(m+1)XX V1(m+1)+ρ3

2 Ik

1 1

nV1(m+1)TX(yβ0(m)1n) +ρ3

2 (0m)λ(3m))

.

(8)

Step 9 Updateβ0as follows:

β0(mj+1)=S

β(jm+1)+λ(3mj)β

ρ3

, j=1, . . . ,k,

whereλ(3mj)andβ(jm)are thej-th elements of the vectorsλ(3m)andβ(m), respec- tively.

Step 10 Updateβ0as follows:

β0(m+1)= 1

n1n(yX V1(m+1)β(m+1)).

Step 11 Update1, 2,λ3as follows:

(1m+1) =(1m)+V(m+1)V0(m+1), (2m+1) =(2m)+V1(m+1)V0(m+1), λ(3m+1) =λ(3m)+β(m+1)β(0m+1). The derivations of the updates are given in “Appendix A”.

To apply LADMM to the minimization problem (6), we consider the following problem:

β0,β,βmin0,Z,V,V0

1

nyβ01nX V0β22+w

nX−Z V2F+λVV01+λββ01

subject to VV =Ik, V =V0, β=β0. (8)

The augmented Lagrangian for this problem is given by 1

nyβ01nX V0β22+w

nXZ V2F+λVV01+λββ01

+ρ1

2 V0V +2F+ρ2

2 β−β0+λ22

subject to VV =Ik,

where,λare dual variables andρ1, ρ2(>0)are penalty parameters.

The updates of the LADMM algorithm are almost the same as those of the ADMM algorithm. We summarize the updates and the derivations in “Appendix B”.

Here we remark on the main differences between ADMM and LADMM. LADMM has two penalty parameters (ρ1, ρ2), while ADMM has three penalty parameters (ρ1, ρ2, ρ3). This means that the total number of tuning parameters in LADMM is only one less than that in ADMM. This is an advantage of LADMM regardless of whether the user tunes the penalty parameters subjectively or objectively. On the other hand, approximation by Taylor expansion is used in LADMM. If this approximation is inappropriate, LADMM may fail to estimate parameters. In terms of running times,

(9)

ADMM seems to be faster than LADMM, based on several numerical studies. These results will be presented in Sect.6when discussing the limitations of the current study.

4.2 Determination of tuning parameters

We have the six tuning parameters: w, λV, λβ, ρ1, ρ2, ρ3. The penalty parameters ρ1, ρ2, ρ3are fixed asρ1=ρ2=ρ3=1 in accordance with Boyd et al. (2011). The tuning parameterwis set according to the purpose of the analysis. A small value is allocated towwhen the user considers the regression loss to be more important than the PCA loss. This idea follows Kawano et al. (2015,2018).

The two regularization parametersλV, λβare objectively selected byK-fold cross- validation. For the original dataset divided into theKdatasets(y(1),X(1)), . . . , (y(K), X(K)), the criterion for theK-fold cross-validation in ADMM is given by

CV= 1 K

K k=1

1 n

y(k)− ˆβ0(−k)1(k)X(k)Vˆ1(−k)βˆ(−k)2

2, (9)

whereβˆ0(−k),Vˆ1(−k),βˆ(−k)are the estimates ofβ0,V1,β, respectively, computed with the data excluding thek-th dataset. We omit the CV criterion for LADMM, since we only replaceVˆ1(−k)in (9) withVˆ0(−k).

We choose the values of the regularization parametersλV, λβfrom the minimizers of CV in (9).

5 Numerical study 5.1 Monte Carlo simulations

We conducted Monte Carlo simulations to investigate the effectiveness of SPCRsvd.

The simulations had six cases, which were the same as those in Kawano et al. (2015) except for Case 6. These six cases are given as follows.

Case 1 The 10-dimensional covariate vectorx=(x1, . . . ,x10)follows a multivariate normal distribution having a zero mean vector and variance-covariance matrix Σ. The response was obtained by

yi =ζ1e1xi+ζ2e2xi+εi, i=1, . . . ,n, wheree1=(1,0, . . . , 0

9

),e2=(0,1,0, . . . , 0

8

), andεi are independently distributed as a normal distribution with mean zero and varianceσ2. We used ζ1=2, ζ2=1, Σ=I10. Then we note thate1ande2are eigenvectors ofΣ.

Case 2 This case is the same as Case 1 except with ζ1 = 8, ζ2 = 1, Σ = diag(1,32,1 , . . . ,1

8

). Then e2 becomes the first eigenvector. In addition,

(10)

Cov(y,x1) = 8 and Cov(y,x2) = 9. For more details of this setting, we refer to p. 196 in Kawano et al. (2015).

Case 3 The 20-dimensional covariate vectorx =(x1, . . . ,x20)has multivariate nor- mal distributionN20(0, Σ). The response was obtained as

yi =4ζxi+εi, i =1, . . . ,n,

where εi are independently distributed as N(0, σ2). We used ζ = (ν,0 , . . . ,0

11

)andΣ=block diag1,I11), whereν=(−1,0,1,1,0,−1,−1,0,1) and1)i j =0.9|ij|(i,j,=1, . . . ,9). Note thatνis a sparse approximation of the fourth eigenvector ofΣ1. This case deals with the situation where the response is associated with the fourth principal component.

Case 4 The 30-dimensional covariate vectorx =(x1, . . . ,x30)has multivariate nor- mal distributionN30(0, Σ). The response was obtained as

yi =4ζ1xi+4ζ2xi+εi, i =1, . . . ,n,

where εi are independently distributed as N(0, σ2). We used ζ1 = 1,0 , . . . ,0

21

),ζ2=(0 , . . . ,0

9

,ν2,0 , . . . ,0

15

), Σ =block diag1, Σ2,I15). Here ν1 = (−1,0,1,1,0,−1,−1,0,1),ν2 = (1, . . . , 1

6

), and 2)i j = 0.9|ij| (i,j,= 1, . . . ,6). Note that ν1 is a sparse approximation of the third eigenvector ofΣ1andν2is the first eigenvector ofΣ2. This case deals with the situation where the response is associated with the third principal component fromΣ1and the first principal component fromΣ2.

Case 5 This case is the same as Case 4 except withν2=(1,0,−1,−1,0,1). Note that ν2is a sparse approximation of the third eigenvector ofΣ2. This case deals with the situation where the response is associated with the third principal components fromΣ1andΣ2.

Case 6 This case is the same as Case 2 except with x =(x1, . . . ,x100). This is a high-dimensional case of Case 2.

The sample size was set ton=50,200. The standard deviation was set toσ =1,2.

We considered the two algorithms given in Sect.4.1: ADMM for SPCRsvd (SPCRsvd- ADMM) and LADMM for SPCRsvd (SPCRsvd-LADMM). SPCRsvd was fitted to the simulated data with one or five components(k=1,5)except for Case 6 and one or two components(k=1,2)for Case 6. We set the value of the tuning parameterwto 0.1 and employed five-fold cross-validation for selecting the regularization parameters λV,λβ. We used a two-dimensional grid and evaluated the CV in (9) on the grid, as illustrated in Fig.1. The cross-validation surface was obtained by SPCRsvd-ADMM withk=1 and was estimated by data generated from Case 1 withn=50,σ =1. The minimum is achieved for the combination of the first candidate ofλV and the seventh candidate ofλβ.

(11)

candida te n

umbers of lambda_V 2

4 6

8 10

candidate n

umbers of lambda_beta 2 4 6 8 10

CV

0.095 0.100 0.105 0.110 0.115 0.120

Fig. 1 Cross-validation surface in SPCRsvd-ADMM estimated by data generated from Case 1

SPCRsvd was compared with SPCR, PCR, SPLS, and PLS. SPCR was computed by the packagespcr, SPLS by spls, and PLS and PCR bypls. These packages are included in the software R (R Core Team2020). We used the default settings of the packages when determining the values of tuning parameters in SPCR, PCR, SPLS, and PLS. The values of the tuning parameterswandξ in SPCR were set to 0.1 and 0.01, respectively, and then the regularization parameters were selected by five-fold cross-validation. The value of the regularization parameter in SPLS was selected by 10-fold cross-validation. The number of components in SPLS, PLS, and PCR was also selected by 10-fold cross-validation from ranges from one to five when SPCRsvd- ADMM, SPCRsvd-LADMM, and SPCR employk =5 and from one to two when k=2. The performance was evaluated in terms of MSE=E[(y− ˆy)2]. The simulation was conducted 100 times. MSE was estimated from 1,000 random samples.

We summarize the means and standard deviations of MSEs in Tables1,2,3,4,5and 6. The results forσ =1,2 had similar tendencies. PCR and PLS were worst in almost all cases, so we will focus on comparing the other methods. SPCRsvd-LADMM and SPCRsvd-ADMM were competitive with SPCR. In particular, SPCRsvd-LADMM and SPCRsvd-ADMM provided smaller MSEs than SPCR in almost all cases when k=1. Compared to SPLS, SPCRsvd-LADMM and SPCRsvd-ADMM were slightly inferior in many cases whenk=5. However, SPLS produced so large values of MSEs in many cases whenk=1.

The true positive rate (TPR), the true negative rate (TNR), and the Matthews correla- tion coefficient (MCC) (Matthews1975) were also computed for SPCRsvd-LADMM, SPCRsvd-ADMM, SPCR, and SPLS. TPR and TNR are respectively defined by

TPR= TP

j:ζj =0 = 1 100

100 k=1

j : ˆζ(jk)=0∧ζj =0 j :ζj =0 ,

(12)

Table1Mean(standarddeviation)ofMSEforCase1 σnkSPCRsvd-LADMMSPCRsvd-ADMMSPCRSPLSPLSPCR 15011.5841.1622.1301.4551.9995.663 (1.322)(0.119)(1.932)(0.472)(0.433)(0.646) 51.3241.1741.2911.1241.2843.791 (0.878)(0.114)(0.712)(0.134)(0.139)(1.083) 20011.6981.0333.8631.0301.2565.598 (1.715)(0.061)(2.474)(0.052)(0.122)(0.559)) 51.2401.0391.7611.0211.0543.568 (1.001)(0.059)(1.736)(0.050)(0.052)(0.978) 25015.2884.7365.4695.0045.4268.765 (1.394)(0.448)(1.470)(0.647)(0.613)(0.746) 54.7954.7334.9364.6925.1297.091 (0.848)(0.427)(0.661)(0.556)(0.550)(1.118) 20014.8564.1277.2164.1074.3868.606 (1.782)(0.221)(2.347)(0.238)(0.245)(0.633) 54.3764.1394.5304.0864.2176.631 (0.995)(0.214)(1.285)(0.200)(0.208)(0.995) TheboldvaluescorrespondtothesmallestmeansamongSPCRsvd-LADMM,SPCRsvd-ADMM,andSPCR

(13)

Table2Mean(standarddeviation)ofMSEforCase2 σnkSPCRsvd-LADMMSPCRsvd-ADMMSPCRSPLSPLSPCR 15011.9571.2231.27841.43647.11867.487 (7.338)(0.161)(0.141)(18.707)(11.370)(4.646) 52.0461.2301.8871.1091.30639.232 (7.332)(0.145)(7.363)(0.120)(0.151)(13.636) 20011.0371.0411.05440.50246.29565.608 (0.063)(0.061)(0.051)(15.652)(5.246)(3.078) 51.0551.0391.0301.0231.05435.782 (0.098)(0.055)(0.052)(0.049)(0.052)(13.074) 25016.4145.6645.09843.83050.34670.566 (10.200)(7.298)(0.546)(19.342)(11.393)(4.921) 55.8485.0185.3804.4245.14842.498 (7.273)(0.582)(7.302)(0.470)(0.553)(13.706) 20014.1784.1894.21742.87649.25968.664 (0.2383)(0.233)(0.207)(16.265)(5.550)(3.317) 54.2284.1894.1144.0914.21638.880 (0.241)(0.227)(0.205)(0.202)(0.208)(13.069) TheboldvaluescorrespondtothesmallestmeansamongSPCRsvd-LADMM,SPCRsvd-ADMM,andSPCR

(14)

Table3Mean(standarddeviation)ofMSEforCase3 σnkSPCRsvd-LADMMSPCRsvd-ADMMSPCRSPLSPLSPCR 15011.5641.5811.79320.62520.84721.404 (0.314)(0.331)(2.160)(1.924)(2.012)(1.295) 51.6631.9331.5631.9983.39822.244 (0.437)(0.602)(0.316)(1.192)(1.442)(1.475) 20011.0851.0981.09615.25916.81720.642 (0.068)(0.072)(0.069)(4.717)(2.886)(0.863) 51.1141.1441.0961.0891.15820.759 (0.083)(0.105)(0.070)(0.240)(0.080)(0.917) 25016.4126.4086.56224.35324.42324.520 (1.279)(1.247)(2.057)(2.389)(2.342)(1.441) 56.6156.8296.3496.5258.00025.519 (1.591)(1.832)(1.258)(2.178)(2.183)(1.730) 20014.5794.6104.76619.07820.22023.627 (1.963)(1.961)(2.632)(4.390)(2.733)(1.002) 54.6544.4514.7634.2724.43023.776 (1.963)(0.300)(2.632)(0.361)(0.272)(1.063) TheboldvaluescorrespondtothesmallestmeansamongSPCRsvd-LADMM,SPCRsvd-ADMM,andSPCR

(15)

Table4Mean(standarddeviation)ofMSEforCase4 σnkSPCRsvd-LADMMSPCRsvd-ADMMSPCRSPLSPLSPCR 15012.5952.3022.30721.54047.460433.826 (1.542)(0.593)(0.619)(1.389)(23.355)(114.041) 52.7202.6462.2496.015711.93933.604 (1.903)(0.819)(0.558)(5.308)(3.919)(7.875) 20011.1601.1761.15821.01824.899477.828 (0.075)(0.077)(0.076)(0.991)(5.165)(37.972) 51.1651.2011.1581.1831.70123.414 (0.079)(0.103)(0.077)(0.106)(0.261)(1.582) 25019.6959.5119.66724.98350.747437.040 (3.176)(2.290)(2.413)(1.946)(23.481)(114.199) 510.7349.5529.51112.71217.23736.904 (4.010)(2.480)(2.320)(6.581)(4.258)(7.903) 20014.7054.6954.66224.10327.978480.882 (0.304)(0.303)(0.310)(1.213)(5.196)(37.853) 54.7644.7444.6604.4585.21926.522 (0.390)(0.319)(0.312)(0.305)(0.462)(1.730) TheboldvaluescorrespondtothesmallestmeansamongSPCRsvd-LADMM,SPCRsvd-ADMM,andSPCR

(16)

Table5Mean(standarddeviation)ofMSEforCase5 σnkSPCRsvd-LADMMSPCRsvd-ADMMSPCRSPLSPLSPCR 15012.1552.2072.14435.28335.09434.654 (0.501)(0.577)(0.524)(3.264)(2.726)(1.806) 52.5743.1712.11310.19016.03335.537 (1.142)(1.654)(0.512)(6.852)(5.439)(2.125) 20011.1511.4931.50630.62930.87634.297 (0.076)(3.318)(3.491)(2.614)(2.497)(1.602) 51.2081.2201.1561.2361.81434.208 (0.112)(0.105)(0.076)(0.167)(0.304)(1.580) 25018.9498.9859.23638.65938.61237.800 (2.151)(2.133)(3.420)(3.369)(3.104)(2.004) 59.6719.4058.84817.13321.49538.780 (2.993)(2.447)(2.170)(8.875)(6.056)(2.355) 20014.6544.6754.99934.09334.30137.374 (0.300)(0.306)(3.575)(2.778)(2.681)(1.812) 54.8054.7194.6354.5555.30637.343 (0.376)(0.322)(0.307)(0.458)(0.471)(1.790) TheboldvaluescorrespondtothesmallestmeansamongSPCRsvd-LADMM,SPCRsvd-ADMM,andSPCR

(17)

Table6Mean(standarddeviation)ofMSEforCase6 σnkSPCRsvd-LADMMSPCRsvd-ADMMSPCRSPLSPLSPCR 15014.2311.7962.38841.87153.04767.579 (8.020)(0.336)(0.445)(18.629)(4.839)(4.285) 23.9853.7141.4551.06244.02266.887 (7.576)(5.460)(0.451)(0.070)(4.491)(4.119) 20011.1321.1611.94940.59747.64165.833 (0.087)(0.103)(0.191)(15.581)(4.672)(2.973) 21.1381.1621.0581.01821.01465.233 (0.121)(0.121)(0.053)(0.045)(2.891)(3.132) 250112.0098.3599.31345.06256.41070.652 (8.929)(1.560)(1.723)(18.692)(4.828)(4.361) 212.18312.9676.3914.25047.68169.978 (8.387)(6.057)(1.666)(0.280)(4.499)(4.233) 20015.1885.2417.78442.87850.61968.876 (0.369)(0.393)(0.783)(16.130)(4.790)(3.053) 25.1935.2614.2714.07324.59168.280 (0.400)(0.355)(0.216)(0.182)(2.938)(3.193) TheboldvaluescorrespondtothesmallestmeansamongSPCRsvd-LADMM,SPCRsvd-ADMM,andSPCR

(18)

Table 7 Mean (standard deviation) of TPR, TNR, and MCC for Case 1

σ n k SPCRsvd-LADMM SPCRsvd-ADMM SPCR SPLS

1 50 1 TPR 0.980 1 0.810 0.930

(0.272) (0) (0.394) (0.174)

TNR 0.461 0.553 0.387 0.951

(0.339) (0.271) (0.325) (0.130)

MCC 0.327 0.470 0.188 0.881

(0.295) (0.232) (0.144) (0.184)

5 TPR 0.970 1 0.980 1

(0.171) (0) (0.140) (0)

TNR 0.512 0.532 0.273 0.905

(0.288) (0.249) (0.201) (0.220)

MCC 0.412 0.460 0.238 0.879

(0.253) (0.222) (0.137) (0.235)

200 1 TPR 0.870 1 0.430 1

(0.337) (0) (0.497) (0)

TNR 0.480 0.700 0.711 1

(0.381) (0.311) (0.357) (0)

MCC 0.306 0.637 0.123 1

(0.347) (0.305) (0.183) (0)

5 TPR 0.960 1 0.850 1

(0.196) (0) (0.358) (0)

TNR 0.577 0.630 0.441 0.916

(0.333) (0.298) (0.300) (0.152)

MCC 0.474 0.557 0.255 0.880

(0.323) (0.292) (0.176) (0.196)

2 50 1 TPR 0.890 1 0.870 0.795

(0.314) (0) (0.337) (0.247)

TNR 0.356 0.347 0.238 0.942

(0.288) (0.219) (0.313) (0.126)

MCC 0.214 0.301 0.122 0.775

(0.203) (0.169) (0.109) (0.212)

5 TPR 0.970 1 0.990 0.940

(0.171) (0) (0.100) (0.163)

TNR 0.387 0.412 0.142 0.878

(0.249) (0.224) (0.151) (0.215)

MCC 0.309 0.354 0.146 0.795

(0.189) (0.172) (0.112) (0.239)

200 1 TPR 0.860 1 0.370 1

(0.348) (0) (0.485) (0)

Referenzen

ÄHNLICHE DOKUMENTE

Decomposition of the European GDP based on Singular Spectrum Analysis.

CASE Network Studies &amp; Analyses No.457 – Post-crisis Lesson for EMU Governance from the .... They can be subject to further

Die zweite Orthonormalmatrix ist die Transponierte der Matrix, die die orthonormalisierten Eigenvektoren von A T · A, nach Eigenwertgr ¨osse geordnet, als Spalten enth

 Unsupervised Learning: Finde “interessante” Strukturen in Daten (z.B. Gruppen); es gibt keine Zielgrösse?. Ergebnis kann nicht validiert

• PC 1: Straight line with smallest orthogonal distance to all points. • PC 1 &amp; PC 2: Plane with smallest orthogonal distance to

• PC 1: Straight line with smallest orthogonal distance to all points. • PC 1 &amp; PC 2: Plane with with smallest orthogonal distance to

A decomposition method for large-scale convex optimization problems with block- angular structure and many linking constraints is analysed.. The method is based on a

M. Theoretical foundations of the potential function method in pattern recognition learning. A training algorithm for optimal margin classiers. In Fifth Annual Workshop on COLT