1 Introduction

Functional data that both predictor and response are random functions are often encountered in meteorology, medicine, biology, economy and finance (Ramsay and Silverman 2005). Due to its flexibility and interpretability, functional regression analysis has received a lot of attention in past years. For example, see Cardot et al. (1999, 2003), Chiou et al. (2003), Ramsay and Silverman (2005), Yao et al. (2005), Cai and Hall (2006), Hall and Horowitz (2007), Ferraty and Vieu (2006) and Baíllo and Grané (2009).

To improve power of prediction and interpretation of functional regression models, some additional real-valued predictors were introduced into functional regression models, which leaded to some new functional linear regression models. For example, Aneiros-Pérez and Vieu (2006) proposed a semi-functional partial linear regression model by combining the feature of a linear model together with the methodology for nonparametric treatment of functional data; Aneiros-Pérez and Vieu (2008) presented an extended semi-functional partial linear regression model for dependent data; Aneiros-Pérez and Vieu (2011) further proposed a fully automatic estimation procedure in a partial linear model with functional data; Zhang et al. (2007) developed a partial functional linear model by incorporating a parametric linear regression into functional linear regression models; Wong et al. (2008) proposed a functional-coefficient partially linear regression model by combining nonparametric and functional-coefficient regression model; Dabo-Niang and Guillas (2010) introduced a functional semiparametric model in which a real-valued random variable was explained by the sum of an unknown linear combination of the components of a multivariate random variable and an unknown transformation of a functional random variable and the random error was autocorrelated; Lian (2011) considered a functional partial linear model by taking advantage of both parametric and nonparametric functional models; Lian (2012) proposed an empirical likelihood approach to nonparametric functional regression and semi-functional partially linear model; Zhou and Chen (2012) introduced a semi-functional linear model by combining the feature of a functional linear regression model and a nonparametric regression model.

To broaden the applicability of functional linear regression models, a varying-coefficient functional linear regression model (VCFLRM) by allowing the slope function to depend on some additional scalar covariates was also proposed by Cardot and Sarda (2008) and has only received a little attention in recent years. For example, Wu et al. (2010) discussed estimation of the slope function in a VCFLRM based on functional principal components for sparse and irregular data, and investigated the asymptotic properties of the proposed estimators; Müller and Sentürk (2011) presented a review of statistical inference on VCFLRM. Inspired by the work of Cardot and Sarda (2008) and Wu et al. (2010), we consider a varying-coefficient partially functional linear regression model (VCPFLRM) by relaxing the linearity assumption in Zhang et al. (2007), which is an extension of partial functional linear regression models and varying-coefficient functional linear regression models.

Polynomial spline is a very popular smoothing technique in a nonparametric regression, and it enables us to extend the standard methods for parametric models to nonparametric settings and is easy to implement in applications, hence it is employed to approximate functional coefficients in our considered VCPFLRM. Based on the polynomial spline approximations to functional coefficients, we first use the least squares approach to estimate parameters in polynomial spline approximations and then obtain estimations of functional coefficients. Under some regular conditions, we discuss the global and uniform convergence rates of the proposed estimators.

The rest of this paper is organized as follows. Section 2 describes varying coefficient partially functional linear regression models and presents the polynomial spline estimators of functional coefficients. In Sect. 3, we study asymptotic properties of the proposed estimators. Simulation studies are conducted to investigate the performance of the proposed methods in Sect. 4. Technique details are given in the Appendix.

2 Model and estimation

Let \(Y\) be a real-valued response variable defined on a probability space \((\Omega ,\mathfrak {B},\mathcal {P})\), let \(U\) and \(Z=(Z_1,\ldots ,Z_p)^{T}\) be one-dimensional and \(p\)-dimensional vectors of explanatory variables defined on the same probability space, respectively. Also, let \(\{X(t){:}\, t\in \mathcal {T}\}\) be a zero mean, second-order (i.e., \(E|X(t)|^2 <\infty \) for all \(t\in \mathcal {T}\)) stochastic process defined on the probability space \((\Omega ,\mathfrak {B},\mathcal {P)}\) with sample paths in \(L_2(\mathcal {T})\), which represents the Hilbert space containing square integrable functions defined on \( \mathcal {T}\) with inner product \(\langle x, y\rangle =\int _ \mathcal {T} x(t)y(t)dt\) for \(\forall x, y \in L_2(\mathcal {T})\) and norm \(\Vert x\Vert _2= \langle x, x\rangle ^{1/2}\). We assume that the relationship between \(Y\) and \((X,U,Z)\) is given by

$$\begin{aligned} Y=\int _{\mathcal {T}} X(t)a_0(t)dt+a_1(U)Z_1+\cdots +a_p(U)Z_p+\varepsilon , \end{aligned}$$
(1)

where \(a_0(t)\) and \(a_j(U)\)’s are unknown smooth functions for \(j=1,\ldots ,p\), and \(\varepsilon \) is a random error with mean zero and finite variance \(\sigma ^2\) and is independent of \((X,U,Z)\). Without loss of generality, we assume \(\mathcal {T}=[0,1]\). Clearly, the above defined model includes varying-coefficient models and functional linear regression models, which correspond to the cases that \(a_0(t)=0\) and \(a_j(U)=0\) for \(j=1,\ldots ,p\), respectively. Also, it includes partial functional linear regression models (Zhang et al. 2007) when \(a_j(U)\equiv \beta _j\) for \(j=1,\ldots ,p\). Hence, the above defined model (1) is an extension of partial functional linear regression models and varying-coefficient models, and is referred to as a varying-coefficient partially functional linear regression model.

Let the data set \(\{X_i,U_i,Z_i,Y_i\}_{i=1}^n\) be \(n\) independent realizations of \(\{X,U,Z,Y\}\) generated from model (1), i.e.,

$$\begin{aligned} Y_i=\int _0^1 X_i(t)a_0(t)dt+a_1(U_i)Z_{i1}+\cdots +a_p(U_i)Z_{ip}+\varepsilon _i \quad \mathrm{for} \quad i=1,\ldots ,n,\nonumber \\ \end{aligned}$$
(2)

where the random errors \(\varepsilon _i\)’s are independent and identically distributed with \(E\varepsilon _i=0\) and \(E \varepsilon _i^2=\sigma ^2\), and are independent of \((X_i,U_i,Z_i)\).

In what follows, we consider estimations of functional parameters \(a_j\)’s for \(j=0,1,\ldots ,p\). Let \(S_{k_0,N_{0n}}\) be the space of polynomial splines on \([0,1]\) with degree \(k_0\) and \(N_{0n}\) knots \(u_{0,1},\ldots ,u_{0,N_{0n}}\) satisfying

$$\begin{aligned} 0=u_{0,0}<u_{0,1}<\cdots <u_{0,N_{0n}}<u_{0,N_{0n}+1}=1, \end{aligned}$$
(3)

and \(S_{k_j,N_{jn}}\) \((j=1,\ldots ,p)\) be the space of polynomial splines on \([a, b]\), which is the compact support of the density of \(U\), with degree \(k_j\) and \(N_{jn}\) knots \(u_{j,1},\ldots ,u_{j,N_{jn}}\) satisfying

$$\begin{aligned} a=u_{j,0}<u_{j,1}<\cdots <u_{j,N_{jn}}<u_{j,N_{jn}+1}=b, \end{aligned}$$
(4)

where the numbers of knots \(N_{jn}~ (j=0,1,\ldots ,p)\) increase when sample size \(n\) increases. The space of polynomial splines \(S_{k_j,N_{jn}}\) is a linear space of \(K_{jn}\)-dimension, where \(K_{jn}\equiv N_{jn}+k_j+1\) for \(j=0,1,\ldots ,p\). Generally, we can choose the truncated power basis and B-spline function as a basis of the above defined linear space. Due to some good numerical properties of B-spline function, we use the B-spline basis to approximate functional parameters \(a_j\) as follows. More details for spline function and spline space can refer to de Boor (2001) and Schumaker (1981).

Following the arguments of de Boor (2001), if the unknown function \(a_j\) \((j=0,1,\ldots ,p)\) is sufficiently smooth, there exists a spline function \(\bar{a}_j\) in the linear space \(S_{k_j,N_{jn}}\) such that

$$\begin{aligned} a_j\approx \bar{a}_j=\sum _{s=1}^{K_{jn}} b_{js}B_{js}, \end{aligned}$$
(5)

where \(B_{js}\) denotes the B-spline function in the linear space \(S_{k_j,N_{jn}}\) for \(j=0,1,\ldots ,p\). Thus, the model (2) can be approximated by

$$\begin{aligned} Y_i\approx \int _0^1 X_i(t)\bar{a}_0(t)dt+\bar{a}_1(U_i)Z_{i1}+\cdots +\bar{a}_p(U_i)Z_{ip}+\varepsilon _i. \end{aligned}$$
(6)

Denote

$$\begin{aligned} l(b)=\sum _{i=1}^n \left\{ Y_i-\sum _{s=1}^{K_{0n}} b_{0s}<X_i,B_{0s}>-\sum _{j=1}^p\sum _{s'=1}^{K_{jn}}b_{js'}B_{js'}(U_i)Z_{ij}\right\} ^2, \end{aligned}$$
(7)

where \(b=(b_0^T, b_1^T,\ldots ,b_p^T)^T\) with \(b_j=(b_{j1},\ldots ,b_{jK_{jn}})^T\) for \(j=0,1,\ldots ,p\). By minimizing \(l(b)\) given in Equation (7), we can obtain the least squares estimator \(\widehat{b}=(\widehat{b}_0^T,\widehat{b}_1^T,\ldots ,\widehat{b}_p^T)^T\) of \(b\), where \(\widehat{b}_j=(\widehat{b}_{j1},\ldots ,\widehat{b}_{jK_{jn}})^T\) for \(j=0,1,\ldots ,p\). Thus, the polynomial spline estimators of functional parameters \(a_j\)’s are given by \(\widehat{a}_j=\sum _{s=1}^{K_{jn}} \widehat{b}_{js}B_{js}\). In this case, we can also define an estimator of variance \(\sigma ^2\) as \(\widehat{\sigma }^2_n=n^{-1}\sum _{i=1}^n\{Y_i-<X_i,\widehat{a}_0>-\sum _{j=1}^p \widehat{a}_j(U_i)Z_{ij}\}^2\).

3 Asymptotic properties

In this section, we investigate asymptotic properties of the above proposed estimators. For simplicity, we first introduce the following notation. For two sequences of positive numbers \(c_n\) and \(d_n\), \(c_n\lesssim d_n\) represents that \(c_n/d_n\) is uniformly bounded, and \(c_n\asymp d_n\) if and only if \(c_n\lesssim d_n\) and \(d_n\lesssim c_n\). The covariance operator \(\Gamma \) of a random function \(X\) is defined as \(\Gamma x(t)=\int _0^1 EX(t)X(s)x(s)ds\) for \(x\in L_2(\mathcal {T})\). \(\Vert \cdot \Vert _{\infty }\) denotes the super-norm of function on some region \(D\), that is, \(\Vert r\Vert _{\infty }=\sup _{x\in D} |r(x)|\). Denote \(\mathcal {K}_n=\max _{j\in \{0,1,\cdots ,p\}} K_{jn}\), \(h_j=\max _{l=0,\cdots ,N_{jn}}(u_{j,l+1}-u_{j,l})\), and let \(C^{\,q}([a,b])\) be the collection of all functions that are \(q\) times continuously differentiable on \([a,b]\).

To study asymptotic properties of the above proposed estimators, we here assume that the degrees \(k_j\)’s are fixed and the numbers of knots \(N_{jn}\)’s depend on sample size \(n\). In addition, we also require the following assumptions.

  1. (A1)

    For the knot sequences given in Eqs. (3) and (4), there exists some positive constant \(C_1\) such that

    $$\begin{aligned} \max _{j=0,1,\cdots ,p}\frac{\max _{l=0,\cdots ,N_{jn}}(u_{j,l+1}-u_{j,l})}{\min _{l=0,\cdots ,N_{jn}}(u_{j,l+1}-u_{j,l})} \le C_1. \end{aligned}$$

    Also, \(\mathcal {K}_n\asymp n^r\) for \(0<r<1/3\), and \(h_j\asymp \mathcal {K}_n^{-1}\) for \(j=0,1,\ldots ,p\).

  2. (A2)

    The density function \(f_U(u)\) of random variable \(U\) has a compact support \(D_u=[a,b]\) and \(f_U(u)\) is bounded away from zero and infinity on \(D_u\).

  3. (A3)

    \(a_0(t)\in C^q([0,1])\), \(a_j(u)\in C^q([a,b])\) for \(j=1,\ldots ,p\) and \(1<q\le k\), where \(k=\min _{j=0,1,\ldots ,p} k_j\).

  4. (A4)

    \(\Vert X\Vert _2\le C_2<\infty \) a.s., and there is a positive constant \(C_3\) such that \(\langle \Gamma a_0^*, a_0^*\rangle \ge C_3 \Vert a_0^*\Vert ^2\) for any \(a_0^* \in S_{k_0,N_{0n}}\), where \(C_2\) is a positive constant and \(C_3\) does not depend on \(n\).

  5. (A5)

    The eigenvalues of \(E(Z_i^* {Z_i^*}^T|X_i=x,U_i=u)\) are uniformly bounded away from zero and infinity for all \((x,u)\in L_2(\mathcal {T})\times D_u\), where \(Z_i^*=(1,Z_{i1},\ldots ,Z_{ip})^T\).

  6. (A6)

    For some \(m_0 >2\), \(E|Z_{1j}|^{m_0}<\infty \) for \(j=1,\ldots ,p\).

Remark 1

Assumption (A1) is similar to Eq. (3) of Zhou et al. (1998) and Assumption (C3) of Xue and Yang (2006). Also, \(\mathcal {K}_n\) stands for the growth rate of the dimension of the spline spaces relative to sample size. Assumption (A2) is very common in nonparametric regression, for example, see Condition 1 of Stone (1985) and Condition 2 of Chen (1991). Assumption (A3) ensures that \(a_0(t)\) and \(a_j(U)\) for \(j=1,\ldots ,p\) are sufficiently smooth so that they can be approximated by spline functions. Assumption (A4) is a stronger condition than that given in functional linear regression models. Assumption (A5) is a generalization of Condition (ii) of Huang and Shen (2004) and Assumption (C2) of Xue and Yang (2006). Assumption (A6) is similar to Condition (v) of Huang and Shen (2004).

Under Assumptions (A1)–(A6), we obtain the following global and uniform convergence rates of the polynomial spline estimators.

Theorem 1

Suppose that Assumptions (A1)–(A6) hold. Then, we have

$$\begin{aligned} \Vert \widehat{a}_j-a_j\Vert _2^2=O_p\left( \frac{\mathcal {K}_n}{n}\right) +O_p\left( \mathcal {K}_n^{-2q}\right) \quad \mathrm{for} \quad j=0,1,\ldots ,p. \end{aligned}$$

Theorem 2

Under Assumptions (A1)–(A6), we have

$$\begin{aligned} ||\widehat{a}_j-a_j||_{\infty }=O_p\left( \mathcal {K}_n n^{-1/2}\right) +O_p(\mathcal {K}_n^{1/2-q}) \quad \mathrm{for} \quad j=0,1,\ldots ,p. \end{aligned}$$

Remark 2

Theorem 1 shows the global convergence rates of the polynomial spline estimators of \(a_0(t)\) and \(a_j(U)\) for \(j=1,\ldots ,p\), which are similar to Theorem 1 of Huang and Shen (2004) and Newey (1997) and Theorem 3.2 of Huang et al. (2004). Particularly, when \(\mathcal {K}_n\asymp n^{1/(1+2q)}\), we have \(\Vert \widehat{a}_j-a_j\Vert _2^2=O_p(\mathcal {K}_n^{-2q/(1+2q)})\), which is the optimal global convergence rate given in Stone (1982). The uniform convergence rate in Theorem 2 is the same as Theorem 7 of Newey (1997). The above results indicate that the existence of a random function does not affect the convergence rates of the polynomial spline estimators of functional coefficients.

4 Simulation study

Experiment 1

To investigate the finite sample performance of our proposed methodologies, we conducted the first simulation study. In this simulation study, we generated data \(\{X_i,U_i,Z_i,Y_i\}_{i=1}^n\) from the following model

$$\begin{aligned} Y_i=\int _0^1 X_i(t)a_0(t)dt+a_1(U_i)Z_{i1}+a_2(U_i)Z_{i2}+\varepsilon _i \quad \mathrm{for} \quad i=1,\ldots ,n. \end{aligned}$$

For functional linear components, similar to Lian (2011), we took \(a_0(t)=\sum _{j=1}^{50} \kappa _j\phi _j(t)\) and \(X_i(t)=\sum _{j=1}^{50} \xi _{ij} \iota _j\phi _j(t)\) with \(\kappa _1=0.5\) and \(\kappa _j=4/j^2\) for \(j=2,\ldots ,50\), \(\phi _1(t)=1\) and \(\phi _j(t)=\sqrt{2}\cos ((j-1)\pi t)\) for \(j=2,\ldots ,50\), and \(\iota _j=1/j\) and \(\xi _{ij}\) was independently and uniformly distributed on the interval \([-\sqrt{3},\sqrt{3}]\) for \(j=1,\ldots ,50\). For varying coefficient components, we set \(a_1(U)=0.138+(0.316+0.982U)\exp (-3.89U^2)\), \(a_2(U)=-0.437-(0.659+1.260U)\exp (-3.89U^2)\), and \(U_i\), \(Z_{i1}\) and \(Z_{i2}\) were simulated from the uniform distribution on the interval \([-0.5,0.5]\) (Wong et al. 2008). Random errors \(\varepsilon _i\)’s were independently generated from the normal distribution with mean zero and variance \(0.2^2\), i.e., \(\varepsilon _i\sim N(0,0.2^2)\).

To implement our proposed methods, we took \(k_0=2\), \(k_1=3\) and \(k_2=3\) in using B-spline functions to approximate functions \(a_0(t)\), \(a_1(U)\) and \(a_2(U)\), respectively. For the knot positions, we can uniformly take knots on the considered interval of random variable \(t\) (or \(U\)) or take the sample quantiles of random variable \(t\) (or \(U\)) to be their corresponding knots. For simplicity, we uniformly took knots \(u_{0,h}\) on the interval \([0,1]\) (i.e., \(u_{0,h}=h/(N_{0n}+1)\)) for \(h=0,1,\ldots ,N_{0n}+1\), and knots \(u_{j,h}\) on the interval \([-0.5,0.5]\) (i.e., \(u_{j,h}=-0.5+h/(N_{jn}+1)\)) for \(j=1,2\) and \(h=0,1,\ldots ,N_{jn}+1\). Thus, selecting the numbers of knots is equivalent to choosing the numbers of B-spline functions \(K_{0n},K_{1n}\) and \(K_{2n}\) when \(k_0\), \(k_1\) and \(k_2\) are fixed. Generally, AIC, BIC, the “leave-one-out” cross-validation (Rice and Silverman 1991) and the modified multi-fold cross-validation (Cai et al. 2000) can be used to select the required numbers of B-spline functions. Here, we used the “leave-one-out” cross-validation technique to choose the numbers of B-spline functions. Following Rice and Silverman (1991), \(K_{0n}, K_{1n}\) and \(K_{2n}\) can be selected by minimizing the following cross-validation score:

$$\begin{aligned} \displaystyle CV(K_{0n},K_{1n},K_{2n})=\frac{1}{n}\sum \limits _{i=1}^n \left\{ Y_i-<X_i, \widehat{a}_0^{-i}>-\widehat{a}_1^{-i}(U_i)Z_{i1}-\widehat{a}_2^{-i}(U_i)Z_{i2}\right\} ^2, \end{aligned}$$

where \(\widehat{a}_0^{-i}(t)\), \(\widehat{a}_1^{-i}(U)\) and \(\widehat{a}_2^{-i}(U)\) are respectively estimators of \({a}_0(t)\), \({a}_1(U)\) and \({a}_2(U)\), and are evaluated by deleting the \(i\)th observation \(\{X_i, U_i, Z_i, Y_i\}\) from the full data set \(\{(X_j,U_j,Z_j,Y_j)\!: j=1,\ldots ,n\}\).

To assess the performance of our proposed estimators, we computed the mean square error of prediction (MSEP) of response variable \(Y\) (Cardot et al. 2003), which is defined by

$$\begin{aligned} \text {MSEP}=\frac{1}{n}\sum \limits _{i=1}^n \left\{ \widehat{Y}_i-<X_i, a_0>- a_1(U_i)Z_{i1}- a_2(U_i)Z_{i2}\right\} ^2, \end{aligned}$$

and the square-root of average squared error (RASE) of functional parameters \(a_0(\cdot )\), \(a_1(\cdot )\) and \(a_2(\cdot )\) (Huang and Shen 2004), which is defined by

$$\begin{aligned} \text {RASE}=\sum _{j=0}^2 \text {RASE}_j \quad \mathrm{with}\quad \text {RASE}_j=\left\{ \frac{1}{n_j}\sum _{h=1}^{n_j}(\widehat{a}_j(t_h)-a_j(t_h))^2\right\} ^{1/2} \end{aligned}$$

where \(t_h\)’s are the regular grid points for \(h=1,\ldots , n_j\).

In this simulation study, we considered two different sample sizes: \(n=200\) and \(500\). For each sample size, results were obtained via 1000 replications. Table 1 presented sample means, medians and variances of RASE, \(\text {RASE}_j\) and MSEP. Figures 1, 2, and 3 displayed the polynomial spline estimates of \(a_0(t)\), \(a_1(U)\) and \(a_2(U)\) for a special replication corresponding to minimum of \(\text {RASE}\). From Table 1, we observed that the sample mean, median and variance of \(\text {RASE}\), \(\text {RASE}_j\) and \(\text {MSEP}\) decrease as sample size increases. Also, from Figures 1, 2, and 3, we observed that estimates of \(a_0(t)\), \(a_1(U)\) and \(a_2(U)\) become more and more accurate as sample size increases. All these findings showed that our proposed estimation procedure performs well under our considered settings.

Fig. 1
figure 1

The true \(a_0(t)\) (solid curve) and its polynomial spline estimation \(\widehat{a}_0(t)\) (dash curve) under sample sizes: a \(n=200\) and b \(n=500\).

Fig. 2
figure 2

The true \(a_1(U)\) (solid curve) and its polynomial spline estimation \(\widehat{a}_1(U)\) (dash curve) under sample sizes: a \(n=200\) and b \(n=500\).

Fig. 3
figure 3

The true \(a_2(U)\) (solid curve) and its polynomial spline estimation \(\widehat{a}_2(U)\) (dash curve) under sample sizes: a \(n=200\) and b \(n=500\).

Table 1 Sample means, medians and variances of RASE and MSEP

Experiment 2

To compare the performance of our proposed estimators with those obtained from a partial functional linear model, we conducted the following two simulation studies in this experiment.

In the second simulation study, the observed data \(\{X_i,U_i,Z_i,Y_i\}_{i=1}^n\) were generated from the following partial functional linear regression model (PFLRM):

$$\begin{aligned}Y_i=\int _0^1 X_i(t)a_0(t)dt+a_1(U)Z_{i1}+a_2(U)Z_{i2}+\varepsilon _i\end{aligned}$$

with the same settings as given in the first simulation study except that \(a_1(U)\) and \(a_2(U)\) were taken to be \(a_1(U)\equiv 1.5\) and \(a_2(U)\equiv -1\) and \(\varepsilon _i\mathop {\sim }\limits ^\mathrm{i.i.d.} N(0,0.5^2)\) for \(i=1,\ldots ,n\).

In the third simulation study, the observed data \(\{X_i,U_i,Z_i,Y_i\}_{i=1}^n\) were generated from the following VCPFLRM:

$$\begin{aligned} Y_i=\int _0^1 X_i(t)a_0(t)dt+a_1(U_i)Z_{i1}+a_2(U_i)Z_{i2}+\varepsilon _i \end{aligned}$$

with the same settings as given in the first simulation study except that \(a_1(U)\) and \(a_2(U)\) were taken to be \(a_1(U)=\sin (2 \pi U)\) and \(a_2(U)=4U(1-U)\), \(U_i\mathop {\sim }\limits ^\mathrm{i.i.d.} \mathrm{Uniform}(0,1)\), \(Z_{ij}\mathop {\sim }\limits ^\mathrm{i.i.d.} \mathrm{Uniform}(-1,1)\) for \(j=1\) and \(2\), and \(\varepsilon _i\mathop {\sim }\limits ^\mathrm{i.i.d.} N(0,0.5^2)\) for \(i=1,\ldots ,n\).

For each of the above two simulation studies, \(500\) data sets were generated and were fitted to PFLRM and our proposed VCPFLRM with the same selections of \(k_0\), \(k_1\), \(k_2\) and knots as given in the first simulation study, respectively. Results corresponding to MSEP for \(n=200\) and \(500\) were presented in Table 2. From Table 2, we observed that (i) using our proposed VCPLRM to fit PFLRM data has the same performance as using PFLRM to fit PFLRM data in terms of their means and standard deviations of MSEP; (ii) using PFLRM to fit VCPLRM data may yield a relatively large mean value of MSEP, which indicated that our proposed VCPFLRM behaves better than PFLRM under misspecified functional linear models; (iii) increasing sample size can reduce standard deviation of MSEP.

Table 2 Mean and standard deviation (SD) of MSEP in experiment 2