1 Introduction

With the development of technology of computation and measurement, scientists usually confront the data providing information about curves, surfaces or anything else varying with continuous variables. Such type of data structure, called functional data, attracted great interests in various fields. For example, in chemometrics the spectrometric data consists of hundreds of different wavelength spectra, fMRI data can recover the contours of invisible human organs and spatial data is used to study the topological, geometric, or geographic properties of entities. Due to the infinite dimensionality and the strong mutual relations of predictors, the traditional multivariate statistical methods fail to analyze functional data. To overcome these problems, Ramsay and Dalzell (1991), Ramsay and Silverman (1997, 2005) introduced some fundamental models and tools for functional data analysis.

Regression analysis is very popular in statistical analysis. As an extension of ordinary linear models, Ramsay and Silverman (1997, 2005) introduced the functional linear model to model the relationship between a scalar response and a functional predictor. Further, Cardot et al. (1999), Cai and Hall (2006), Hall and Horowitz (2007) and Li and Hsing (2007) proposed estimation methods based on functional principal component analysis and investigated the asymptotic properties of the estimators. On the other hand, Cardot et al. (2003) and Crambes et al. (2009) employed penalized B-spline and smoothing spline to estimate the functional slope parameter. As an extension of nonparametric model, functional nonparametric regression was also studied in literatures. Kernel regression (Ferraty and Vieu 2006), local linear regression (Baíllo and Grané 2009) and K-nearest neighbours method (Burba et al. 2009) are used to deal with the functional nonparametric models.

In order to improve the power of prediction and interpretation of the functional regression model, some additional real-valued predictors could be introduced. There is some recent literature focusing on this situation. For example, Aneiros-Pérez and Vieu (2006) introduced a semi-functional partial linear regression model to predict the fat content of the chopped pure meat. Further, Aneiros-Pérez and Vieu (2008) extended this model to dependent data. Zhang et al. (2007) introduced the partial functional linear model to assess the effect of women’s hormone on the total hip bone mineral density and Shin (2009) proposed a new estimation method based on functional principal component analysis. Cardot and Sarda (2008) generalized the functional linear model to a varying coefficient functional linear model in which an additional random variable influenced smoothly the functional coefficient. Zhou and Chen (2012) introduced a semi-functional linear model which combined the functional linear regression model and the nonparametric regression model.

In functional linear regression, because of its orthogonality, interpretability and best representation, functional principal component analysis approach has been extensively used to estimate the slope function (see Cardot et al. 1999; Cai and Hall 2006; Hall and Horowitz 2007; Li and Hsing 2007; Shin 2009). As a very popular smooth technique, polynomial spline or regression spline method can produce a smooth function estimate and can be operated easily, so it has received considerable attention in nonparametric/semiparametric regression (see Chen 1991; Stone 1994; Stone et al. 1997; Zhou et al. 1998; Huang 2003a, b; Huang et al. 2004a; Huang and Shen 2004b and so on). However, there is limited literature discussed the polynomial spline method in the functional data case. We only noted that Ramsay and Silverman (1997, 2005) applied polynomial spline to estimate the functional linear model, but they didn’t investigate the asymptotic behaviors of the estimator.

In this paper, we focus on the polynomial spline estimators for the partial functional linear models. We employ the polynomial spline basis to approximate the functional coefficients. Using profile least squares technique, we obtain the optimal convergence rate and asymptotic normality for estimators of parameters. Based on these estimators, we also have the limiting distribution of Wald test statistic for linear hypothesis of parameters. The numerical studies indicate our proposed procedure can enjoy more smoothness for functional coefficients in finite samples.

The rest of the paper is organized as follows. In Sect. 2, we introduce the polynomial spline estimate for partial functional linear models. Section 3 investigates the asymptotic properties of estimators and discusses statical inference problem. Simulation studies are presented in Sect. 4. Conclusion and further research are given in Sect. 5. All technical details and proofs are given in the “Appendix”.

2 Polynomial spline estimation

Let the observed data \((X_i,\mathbf{Z}_i,Y_i)\), \(i=1,\ldots , n\), which are independent and identically distributed (i.i.d.), be generated from the following partial functional linear model

$$\begin{aligned} Y_i=\int \limits _0^1 X_i(t)\alpha (t)dt+ \mathbf{Z}_i^{T}\beta +\varepsilon _i, \quad i=1,\ldots ,n, \end{aligned}$$
(1)

where \(Y_i\) and \(\mathbf{Z}_i=(Z_{i1},\ldots ,Z_{ip})^{T}\) are the scalar response variable and the p-dimensional predictor vector, respectively. The predictor variable \(X_i\) is a random function valued in \(H=L^2([0,1])\), the Hilbert space containing square integrable functions defined on the unit interval. Let \(\langle \phi ,\varphi \rangle =\int ^1_0 \phi (t)\varphi (t)dt\) denote the usual inner product of function \(\phi \) and \(\varphi \) and let \(\Vert \phi \Vert =\langle \phi ,\phi \rangle ^{1/2}\) denote the norm of H. The random errors \(\varepsilon _i\) are independent and identically distributed with mean 0 and finite variance \(\sigma ^2\) and are independent of \((X_i,\mathbf{Z}_i)\). Let \(\beta \) be an unknown p-dimensional parameter vector and \(\alpha (t)\) be an unknown smoothing function belonging to H.

Before introducing the polynomial spline estimation, we recall simply the polynomial spline function. Let \(k\ge 0\). The sequence \(0=t_0<t_1<\cdots <t_{N_n}<t_{N_{n}+1}=1\) is a partition of interval [0, 1], which is called knot sequence. Suppose a function is a polynomial of degree k on each of the intervals \([t_i,t_{i+1}](i=0,1,\ldots ,N_{n})\), and it has \(k-1\) continuous derivatives for \(k\ge 1\) on the interval [0, 1], then it is called spline function of degree k.

We next consider the polynomial spline estimate \(\widehat{\alpha }\) of \(\alpha \). Let \(S_{k,N_n}\) be the space of polynomial splines defined on interval [0, 1] with degree k and \(N_n\) interior knots. The space \(S_{k,N_n}\) is a \(K_n\)-dimensional linear space, \(K_n=N_n+k+1\). From Theorem XII.1 of de Boor (2001), we can conclude that, if the slope function \(\alpha (t)\) is sufficiently smooth, there is a spline function \(a(t)\in S_{k,N_n} \) such that

$$\begin{aligned} \alpha (t)\approx a(t)=\sum _{s=1}^{K_n}b_sB_s(t), \end{aligned}$$
(2)

where \(B_j, j=1,\ldots , K_n\) are the B-spline basis functions. Plugging the approximation (2) into model (1), we have

$$\begin{aligned} Y_i \approx \sum _{s=1}^{K_n}b_s \langle X_i, B_s \rangle +\mathbf{Z}_i^{\,T}\beta +\varepsilon _i, \quad i=1, \ldots , n, \end{aligned}$$
(3)

where the spline coefficient vector \(b=(b_1,\ldots , b_{K_n})^T\) and parameter vector \(\beta \) are to be estimated. Then, the semiparametric estimation problem in model (1) turns into the ordinary parametric estimation problem.

Let the square loss function

$$\begin{aligned} l(\beta ,b)=\sum _{i=1}^n\left( Y_i-\mathbf{Z}_i^{T}\beta -\sum _{s=1}^{K_n}b_s \langle X_i,B_s \rangle \right) ^2. \end{aligned}$$
(4)

The estimators of b and \(\beta \) can be obtained by minimizing (4). For ease to discuss the asymptotic properties, we apply the profile least squares procedure to estimate unknown spline coefficients and parameters. The estimators for \(\beta \) and b are given by

$$\begin{aligned} \widehat{\beta }=(\mathbf{Z}^T (I-A) \mathbf{Z})^{-1} \mathbf{Z}^T(I-A)Y, \quad \widehat{b}=(B^T B)^{-1}B^T( Y- \mathbf{Z} \widehat{\beta }), \end{aligned}$$
(5)

where \(Y = (Y_1, \ldots , Y_n)^T\), \(\mathbf{Z}=(\mathbf{Z}_1, \ldots , \mathbf{Z}_n)^T\), \(B=\big \{\langle X_i,B_{j}\rangle \big \}_{\mathop {i=1,\dots ,n}\limits _{j=1,\dots ,K_n}}\) and \(A = B(B^T B)^{-1} B^T\). Then, the polynomial spline estimator of \(\alpha (t)\) and the estimator of \(\sigma ^2\) can be respectively defined by

$$\begin{aligned} \widehat{\alpha }(t)=\sum _{s=1}^{K_n} \widehat{b}_s B_s(t), \quad ~~~\widehat{\sigma }^{\,2}_n = \frac{1}{n}\sum \limits _{i=1}^n \Big (Y_i- \langle X_i,\widehat{\alpha }\rangle -\mathbf{Z}_i^{\,T}\widehat{\beta }\Big )^2. \end{aligned}$$
(6)

3 Asymptotic properties

In this section, we investigate the asymptotic properties of the polynomial spline estimators. For ease to discuss the asymptotic behaviors of our proposed estimators, the following notation is needed. For two sequences of positive numbers \(a_n\) and \(b_n\), \(a_n\lesssim b_n\) signifies \(a_n/b_n\) is uniformly bounded and \(a_n\asymp b_n\) if \(a_n\lesssim b_n\) and \(b_n\lesssim a_n\). The covariance operator \(\Gamma \) of the random function X is defined as \(\Gamma x(t)=\int _0^1 EX(t)X(s)x(s)ds, x\in H\). The norm \(\Vert \cdot \Vert \) of a function \(f\in C^{k+1}([0,1])\) is defined as \(\Vert f\Vert =\big (\int _0^1 f(t)^2dt\big )^{1/2}\).

In order to establish the theoretical properties of polynomial spline estimation, the following assumptions are required:

  1. (C1)

    There are some positive constants M and \(\frac{1}{4(k+1)}<r<\frac{1}{2}\) such that

    $$\begin{aligned} h=\max _{j=0,\ldots ,N_n}(t_{j+1}-t_j)\asymp n^{-r}, \quad K_n\asymp n^r,\quad h/\min _{j=0,\ldots ,N_n}(t_{j+1}-t_j)\le M. \end{aligned}$$
  2. (C2)

    \(E||X||^4<\infty \) and the eigenvalues of the covariance operator \(\Gamma \) of X are strictly positive.

  3. (C3)

    \(E|Z_{11}|^4+\cdots +E|Z_{1p}|^4+E|\varepsilon _1|^4<\infty .\)

  4. (C4)

    For \(j=1,\ldots ,p\), \(E(Z_{1j}|X_1)\) is a continuous linear functional, that is, there exists a function \(g_j\in H\) such that \(E(Z_{1j}|X_1)=\langle X_1,g_j \rangle \). Further, we assume \(g_j, j=1,\ldots ,p\) and slope function \(\alpha \) are smooth enough, that is, \(g_j\in C^{k+1}([0,1])\), \(\alpha \in C^{k+1}([0,1])\).

  5. (C5)

    Let \(\eta _{1j}=Z_{1j}-E(Z_{1j}|X_1)=Z_{1j}-\langle X_1,g_j \rangle , j=1,\ldots ,p\), \(\eta _1=(\eta _{11},\ldots ,\eta _{1p})^T\). Furthermore, we assume that \(\Sigma =E\eta _1\eta _1^T\) is a positive definite matrix.

Remark 1

Conditions (C1)–(C5) are very general in polynomial spline estimation and functional linear model. In fact, condition (C1) is similar to (3) in Zhou et al. (1998). For the number of spline basis \(K_n\), the requirement is similar to (16) in Shin (2009). Condition (C2) is very common in functional linear model (see H1 and H2 in Cardot et al. 1999 and (12) in Shin 2009). However, we don’t need additional assumption on eigenvalues of the covariance operator \(\Gamma \) like (14) in Shin (2009). Condition (C3) is similar to (11) in Aneiros-Pérez and Vieu (2006) and (17) in Shin (2009). Condition (C4) requires the dependence between the covariate \(Z_{1j}, (j=1,\ldots ,p)\) and the random function \(X_1\) is a continuous linear functional, which is a special case of conditional expectation operators \(E(X_{ij}|T_i=t)\) in Aneiros-Pérez and Vieu (2006). Furthermore, to assure the validity of the polynomial spline estimation, we need a restricted smooth condition on each functional coefficient \(g_j\) and \(\alpha \). Condition (C5) is similar to (12) in Aneiros-Pérez and Vieu (2006) and (20) in Shin (2009).

Under the above assumption conditions, we have the following results.

Theorem 1

If conditions (C1)–(C5) hold, as \(n\rightarrow \infty \), we have

$$\begin{aligned} \sqrt{n}(\widehat{\beta }-\beta )\mathop {\longrightarrow }\limits ^{D} N(0,\sigma ^2\Sigma ^{-1}). \end{aligned}$$

Theorem 2

Suppose that conditions (C1)–(C5) are satisfied, then

$$\begin{aligned} \Vert \widehat{\alpha }-\alpha \Vert ^2=O_p\Big (\frac{K_n}{n}+K_n^{-2(k+1)}\Big ). \end{aligned}$$

Remark 2

For the estimation of the parameter vector, Theorem 1 shows that the asymptotic result is similar to Theorem 1(i) in Aneiros-Pérez and Vieu (2006) and Theorem 3.1 in Shin (2009). For the estimation of functional coefficient, Theorem 2 indicates that, under smoother conditions (\(C^{k+1}\) in particular), the global convergence rate is similar to those given in Newey (1997) and Huang and Shen (2004b) in nonparametric regression setting, which shows that the existence of a random vector as a predictor does not change the rate of convergence of the estimated functional coefficient. Moreover, if we take \( r=(a+1)/(a+2b)\) and \(k=(2b-1)/2(a+1)-1\), then we can obtain the same rate of convergence of the estimated functional coefficient as Shin (2009).

For the estimator of variance \(\sigma ^2\), we have the following theorem

Theorem 3

If conditions (C1)–(C5) hold, then we have

$$\begin{aligned} \sqrt{n}\left( \widehat{\sigma }_n^2-\sigma ^2\right) \mathop {\longrightarrow }\limits ^{D} N(0,\Lambda ^2), \end{aligned}$$

where \(\Lambda ^2=E(\varepsilon _1^2-\sigma ^2)^2\).

Further, let \(\widehat{\Sigma }_n=n^{-1}{\mathbf{Z}^T(I-A)\mathbf{Z}}\). In the light of the above theorems, we can obtain the following corollary.

Corollary 1

Under conditions (C1)–(C5), as \(n\rightarrow \infty \), we have

$$\begin{aligned} \widehat{\chi }^2_{n,p}=\frac{n}{\widehat{\sigma }^2_n}(\widehat{\beta }-\beta )^T\widehat{\Sigma }_n(\widehat{\beta }-\beta )\mathop {\longrightarrow }\limits ^{D} \chi ^2_p. \end{aligned}$$

Remark 3

According to the corollary 1, we can obtain an approximate \((1-\gamma )\) asymptotic confidence region for parameter vector \(\beta \), that is,

$$\begin{aligned} \Big \{a\in R^p: \frac{n}{\widehat{\sigma }^2_n}(\widehat{\beta }-a)^T\widehat{\Sigma }_n(\widehat{\beta }-a)\le \chi ^2_{p,1-\gamma }\Big \}. \end{aligned}$$

Also, we can get an approximate \((1-\gamma )\) asymptotic confidence interval for every parameter \(\beta _j, j=1,\ldots ,p\), that is,

$$\begin{aligned} \left[ \widehat{\beta }_j+z_{\gamma /2}\frac{\widehat{\sigma }_n(\widehat{\Sigma }_n)^{-1}_{jj}}{\sqrt{n}}, \widehat{\beta }_j-z_{\gamma /2}\frac{\widehat{\sigma }_n(\widehat{\Sigma }_n)^{-1}_{jj}}{\sqrt{n}}\right] , \end{aligned}$$

where \(\widehat{\sigma }_n(\widehat{\Sigma }_n)^{-1}_{jj}\) is the jth diagonal element of \(\widehat{\sigma }_n(\widehat{\Sigma }_n)^{-1}\).

4 Simulation studies

In this section, we present some simulation results to illustrate the finite sample behaviors of the polynomial spline estimation and compare our method with the Shin (2009)’s.

4.1 Models for generating simulation data

In this subsection we specify four models to generate simulation data \(\big \{(X_i,\mathbf{Z}_i,Y_i)\big \}_{i=1}^n\). In the first three models we take the same form as Lian (2011) to generate \(X_i\), that is,

$$\begin{aligned} X_i=\sum _{j=1}^{50} \xi _{ij} j^{-1}\phi _j(t), \end{aligned}$$

where \(\phi _1(t)=1, \phi _j(t)=\sqrt{2}\cos ((j-1)\pi t)\) for \( j\ge 2\) and \(\xi _{ij}\) is independent and identical distribution with \(U[-\sqrt{3},\sqrt{3}]\).

Model 1: \(Y_i=1.5Z_{i1}-Z_{i2}+2Z_{i3}+\int _0^1 X_i(t)\alpha (t)dt+\varepsilon _i\), where \(\mathbf{Z}_i=(Z_{i1}, Z_{i2}, Z_{i3})^T\) is from a multivariate normal distribution \(N(0,\Phi )\) with covariance matrix \(\Phi =[0.9, 0.2,0.3;0.2,0.5,0.1;0.3,0.1,1]\). The functional coefficient \(\alpha (t)=\sum _{j=1}^{50} b_j\phi _j(t)\), where \(b_1=0.5, b_j=4j^{-2}\), for \( j\ge 2\) and the error variable \(\varepsilon _i \) is N(0, 1).

Model 2: \(Y_i=2Z_{i1}-Z_{i2}+\int _0^1 X_i(t)\alpha (t)dt+\varepsilon _i\), \( Z_{i1}=\int _0^1 X_i(t)\alpha _1(t)dt+\varepsilon _{i1}\), \( Z_{i2}=\int _0^1 X_i(t)\alpha _2(t)dt+\varepsilon _{i2} \), where the functional coefficient \(\alpha (t)\) is similarly defined in Model 1. In addition, functional coefficients \(\alpha _1(t)=\sum _{j=1}^{50} b_{1j}\phi _j(t)\) and \(\alpha _2(t)=\sum _{j=1}^{50} b_{2j}\phi _j(t)\), where \(b_{11}=1, b_{21}=-0.5, b_{1j}=2j^{-2}, b_{2j}=3j^{-2}\) for \(j\ge 2\). Random error variables \(\varepsilon _i\) and \(\varepsilon _{i1}\) are N(0, 0.25) and \(\varepsilon _{i2}\) is N(0, 0.64).

Model 3: \(Y_i=1.5Z_{i1}+5Z_{i2}-1.7Z_{i3}+\int _0^1 X_i(t)\alpha (t)dt+\varepsilon _i\), where \(\mathbf{Z}_i=(Z_{i1}, Z_{i2}, Z_{i3})^T\) is from a multivariate normal distribution \(N(0,\mathbf{I}_3)\). The functional coefficient is given by

$$\begin{aligned} \alpha (t)=2\sin (0.5\pi t)+4\sin (1.5\pi t)+5\sin (2.5\pi t), \end{aligned}$$

which is similar to example (a) in Cardot et al. (2003). The error variable \(\varepsilon _i\) is N(0, 0.36).

Model 4: We take the same example in Shin (2009), that is,

$$\begin{aligned} Y_i=2Z_{i1}-Z_{i2}+1.5Z_{i3}+5Z_{i4}-1.7Z_{i5}+\int \limits _0^1 X_i(t)\alpha (t)dt+\varepsilon _i, \end{aligned}$$

where \(X_i(t)\) is a standard Brownian motion and \(\alpha (t)=\sqrt{2}\sin (\pi t/2)+3\sqrt{2}\sin (3\pi t/2)\). Random vector \(\mathbf{Z}_i=(Z_{i1}, Z_{i2}, Z_{i3}, Z_{i4}, Z_{i5})^T\) is from a multivariate normal distribution \(N(0,\mathbf{I}_5)\), and error variable \(\varepsilon _i\) is N(0, 1).

For practicality, the random functions \(X_i(t)\) in Models 1–4 are all only observed at 100 equally spaced points on [0, 1].

4.2 Implementation

In this subsection, we specifically illustrate implementation of our method and Shin (2009)’s method. To implement Shin (2009)’s method, we need to turn discrete observation data of \(X_i(t)\) into functional data objects. In this paper, we utilize the method mentioned in Chapter 4 of Ramsay et al. (2009) and choose 25 B-spline functions to build functional data. We also use pca.fd function mentioned in Chapter 7 of Ramsay et al. (2009) to carry out a functional principal components analysis. For our procedure, we have to choose the degrees of spline functions, the positions and the number of knots. Similarly to Huang and Shen (2004b), we choose B-spline basis with equally spaced knots and the fixed degree 2 in this paper. Then, we only need to select the number of the B-spline basis and eigenfunctions \(K_n\). Many methods can be used to select \(K_n\), for example, AIC (Akaike 1974), BIC (Schwarz 1978), “leave-one-subject-out” cross-validation (Rice and Silverman 1991) and modified multi-fold cross-validation (Cai et al. 2000). In this paper, we use “leave-one-subject-out” cross-validation technique to choose the number of B-spline basis and eigenfunctions. Specifically, we select \(K_n\) by minimizing the following cross-validation score:

$$\begin{aligned} CV(K_n)=\sum _{i=1}^n \Big (Y_i- \langle X_i, \widehat{\alpha }^{-i} \rangle -\mathbf{Z}_i^T\widehat{\beta }^{-i}\Big )^2, \end{aligned}$$

where \(\widehat{\alpha }^{-i}\) and \(\widehat{\beta }^{-i}\) are estimators computed by deleting the ith observation \((X_i, \mathbf{Z}_i, Y_i)\). In our procedure, the number of B-spline basis ranges from 3 to 12 and the number of eigenfunctions ranges from 1 to 10. For the integrals involved in matrix B, we approximate them by the trapezoidal rule.

Two risk functions are used to assess the performances of our estimators and the Shin (2009)’s: the mean square prediction error of the response variable Y, which is similar to (26) in Cardot et al. (2003),

$$\begin{aligned} \text {MSPE}=n^{-1}\sum _{i=1}^n \left( \widehat{Y}_i-\mathbf{Z}_i^T\beta -\int _0^1 X_i(t)\alpha (t)dt\right) ^2, \end{aligned}$$

and the square-root of average squared error (RASE) of functional coefficient \(\alpha (t)\), which is similar to (6) in Huang and Shen (2004b),

$$\begin{aligned} \text {RASE}=\left[ n^{-1}_{grid}\sum _{k=1}^{n_{grid}}(\widehat{\alpha }(t_k)-\alpha (t_k))^2\right] ^{1/2}, \end{aligned}$$

where \(\{t_k, k=1,\ldots , n_{grid}\}\) are grid points chosen to be equally spaced on the interval [0, 1]. In this paper, the number of grid points \(n_{grid}=101\).

We use Matlab to implement our procedure. For each simulation model above-mentioned, we consider two different sample sizes: \(n=100\) and \(n=500\), and each simulation experiment has been repeated 500 times.

Fig. 1
figure 1

The empirical distribution function (real line) of \(\widehat{\chi }_{n,3}^2\) from 500 simulated samples under Model 1 with different sample sizes: (left) n = 100; (right) n = 500

4.3 Simulation results

In this subsection, we present some simulation results of 4 simulation models mentioned in (4.1). Notations \(_s\) and \(_p\) denote our estimation method and the Shin (2009)’s, respectively.

Figure 1 displays the empirical distribution function of \(\widehat{\chi }_{n,3}^2\) from 500 simulated samples under Model 1. For Models 2–4, the empirical distribution functions of \(\widehat{\chi }_{n,p}^2\) have same performance, so we omit them to save space. We can see from this figure that as the sample size n increases, the empirical distribution more and more approach the theoretical distribution, which also reveals the validity of asymptotic normality in Sect. 3.

Table 1 The MSE of estimators \(\widehat{\beta }\) and \(\widehat{\sigma }_n^2\) under different models and sample sizes
Table 2 The mean(sd) of RASE and MSPE for Models 1–4
Fig. 2
figure 2

The true \(\alpha (t)\) (solid curve), the polynomial spline estimator (dashed curve) and the Shen (2009)’s (dotted curve) under different models and sample sizes: (left) n = 100; (right) n = 500

Table 1 summarizes mean squared errors (MSE) of estimators \(\widehat{\beta }\) and \(\widehat{\sigma }_n^2\) under Models 1–4. Table 2 presents the mean and standard deviation of RASE and MSPE to evaluate the performance of our estimation procedure. Figure 2 shows our estimate (dashed curve) and the Shin (2009)’s (dotted curve) from the typical samples which correspond to the minimum of \(\text {RASE}_s\) and \(\text {RASE}_p\) under Models 1–4, respectively. From these results in our simulation examples, we can know that the two estimation methods are very close for the parameter component \(\beta \) and \(\sigma \). However, from the prediction perspective and the estimation effect of function coefficient \(\alpha (t)\), if the functional coefficient \(\alpha (t)\) can be expressed linear combination of eigenfunctions of covariance operator \(\Gamma \), the Shin (2009)’s method is superior to ours, if not, our method seems to perform better than Shin (2009)’ method. At the same time, the differences between the two estimation methods become smaller and smaller as the sample size n increases.

Table 3 displays the mean and standard deviation of running cpu time in a Dell personal computer with Inter(R) Core(TM)2 Duo CPU. We seem to infer from Table 3 that our method is more computationally expedient at least in the examples studied when the sample size is small, while the Shin (2009)’s is more computationally expedient if the sample size is large.

Table 3 The mean(sd) of running cpu time for Models 3–4

5 Conclusion and further research

In this paper, we propose the polynomial spline estimation for the partial functional linear model. Some asymptotic results are established, including asymptotic normality for the parameter vector and the global rate of convergence for the functional coefficient. By simulation studies, we verify the validity of theoretical results. On the one hand, from the prediction perspective and the estimation effect of function coefficient \(\alpha (t)\), we detect if the functional coefficient \(\alpha (t)\) can be expressed linear combination of eigenfunctions of covariance operator \(\Gamma \), the Shin (2009)’s method is superior to ours, if not, our method seems to perform better than Shin (2009)’ method. While the differences between the two estimation methods become smaller and smaller as the sample size n increases. On the other hand, from computational time, we can draw a conclusion that our method is more computationally expedient at least in the examples studied when the sample size is small, while the Shin (2009)’s is more computationally expedient if the sample size is large. From our limited study, we only consider the functional predictor would be observed fully. However, we can usually obtain some sparsely discrete observations for each functional observation in practice. For this case, we can use smooth techniques to approximate the functional observations. And then, the polynomial spline method can also be used to estimate the partial functional linear model.