24
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 Submitted to Biometrika (2009), xx, x, pp. 1–21 C 2009 Biometrika Trust Printed in Great Britain Functional Quadratic Regression BY FANG YAO Department of Statistics, University of Toronto, 100 Saint George Street, Toronto, Ontario M5S 3G3, Canada [email protected] HANS-GEORG ULLER Department of Statistics, University of California, Davis, One Shields Avenue, Davis, California, 95616, U.S.A. [email protected] SUMMARY We extend the common linear functional regression model to the case where the dependency of a scalar response on a functional predictor is of polynomial rather than linear nature. Focusing on the quadratic case, we demonstrate the usefulness of the polynomial functional regression model which encompasses linear functional regression as a special case. Our approach works under mild conditions for the case of densely spaced observations and also can be extended to the important practical situation where the functional predictors are derived from sparse and irregular measurements, as is the case in many longitudinal studies. A key observation is the equivalence of the functional polynomial model with a regression model that is a polynomial of the same order in the functional principal component scores of the predictor processes. Theo- retical analysis as well as practical implementations are then based on this equivalence and on basis representations of predictor processes. We also obtain an explicit representation of the re- Corresponding author

Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

Submitted to Biometrika (2009), xx, x, pp. 1–21

C© 2009 Biometrika Trust

Printed in Great Britain

Functional Quadratic Regression

BY FANG YAO

Department of Statistics, University of Toronto, 100 Saint George Street, Toronto,

Ontario M5S 3G3, Canada

[email protected]

HANS-GEORG MULLER ∗

Department of Statistics, University of California, Davis, One Shields Avenue, Davis,

California, 95616, U.S.A.

[email protected]

SUMMARY

We extend the common linear functional regression model to the case where the dependency

of a scalar response on a functional predictor is of polynomial rather than linear nature. Focusing

on the quadratic case, we demonstrate the usefulness of the polynomial functional regression

model which encompasses linear functional regression as a special case. Our approach works

under mild conditions for the case of densely spaced observations and also can be extended to

the important practical situation where the functional predictors are derived from sparse and

irregular measurements, as is the case in many longitudinal studies. A key observation is the

equivalence of the functional polynomial model with a regression model that is a polynomial of

the same order in the functional principal component scores of the predictor processes. Theo-

retical analysis as well as practical implementations are then based on this equivalence and on

basis representations of predictor processes. We also obtain an explicit representation of the re-

∗ Corresponding author

Page 2: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

2 F. YAO AND H. G. MULLER

gression surface that defines quadratic functional regression and provide functional asymptotic

results for an increasing number of model components as sample size (number of subjects in the

study) increases. The improvements that can be gained by adopting quadratic as compared to

linear functional regression are illustrated with a case study which includes absorption spectra as

functional predictors.

Some key words: Absorption Spectra; Asymptotics; Functional Data Analysis; Polynomial Regression; Prediction;

Principal Component.

1. INTRODUCTION

Data which include a functional predictor in the form of a smooth random trajectory are increas-

ingly common. Typical scenarios in which such data arise include frequently monitored trajecto-

ries such as in movement tracking (Faraway, 1997) and longitudinal studies where the trajectory

is probed through noisy and often sparse and irregularly spaced measurements. Regression mod-

els that can handle functional predictors are therefore needed for a variety of settings and ap-

plications, and many aspects of these functional regression relations remain open problems. We

consider here the case where a functional predictor is paired with a scalar response. Examples

of such situations include various biological trajectories with regular observations as predictors

(Kirkpatrick & Heckman, 1989) and are also commonly encountered in biodemographic appli-

cations (Muller & Zhang, 2005). In many longitudinal studies, measurements of longitudinal

trajectories are recorded only intermittently and often at time points that do not conform with a

regular grid (Muller, 2005).

As an example of a typical functional regression problem, consider the sample of trajectories

in Figure 1. Displayed are 50 randomly selected 100-channel absorption spectra which are used

to predict the composition of food samples. The spectra shown are from meat specimen and the

goal is to predict fat contents. These spectra have been densely sampled at 100 support points

and are seen to be quite smooth, with only small measurement errors. Taking advantage of the

smoothness of these trajectories is key to the efficient modeling of regression relationships that

Page 3: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

97

98

99

100

101

102

103

104

105

106

107

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125

126

127

128

Biometrika-08-035.R2 3

include functional predictors. The functional nature of the predictors and the measurements need

to be adequately reflected in the statistical modeling of such data (Rice, 2004; Zhao et al., 2004).

In this paper, we consider both densely and sparsely sampled longitudinal predictor data. In

many longitudinal studies for which one wishes to apply functional regression, the predictor pro-

cess must be inferred from noisy and sparse measurements (James et al., 2000; Yao et al., 2005c).

For the most appropriate analysis of a given set of data with functional predictors, one desires a

variety of readily available models, from which the data analyst can choose the most appropriate

approach. The situation is analogous to the case of ordinary regression models involving vector

data, where quadratic models are commonplace and occupy a well-established place in many

applications such as response-surface analysis (see e.g. Melas et al., 2003).

Within the field of functional data analysis, with excellent overviews provided in Ramsay

& Silverman (2002, 2005), an emphasis has been the functional linear model. Various aspects

of this model have been studied, including implementations, computing and asymptotic theory

(Cardot et al., 2003; Shen & Faraway, 2004; Cardot & Sarda, 2005; Cai & Hall, 2006; Hall &

Horowitz, 2007). The functional linear model has been brought to the fore in Ramsay & Dalzell

(1991). It imposes a structural linear constraint on the regression relationship which may or may

not be satisfied. The linear constraint is helpful for addressing the problem that the elements of a

finite sample of random functions are very far away from each other in the infinite-dimensional

function space, but it may be too restrictive for some applications. It is therefore of interest to

consider extensions of the functional linear model that convey a reasonable structural constraint,

but at the same time enhance its flexibility. We propose here the extension of the linear model to

the case of a polynomial functional relationship, analogous to the extension of linear regression

to polynomial regression in traditional regression settings and highlight the important special

case of a quadratic regression.

To achieve the regularization that is necessary for any functional regression model, we project

predictor processes on a suitable basis of the underlying function space, which is then truncated

at a reasonable number of included components. We implement this regularization through the

Page 4: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

170

171

172

173

174

175

176

4 F. YAO AND H. G. MULLER

eigenbasis of the predictor processes which leads to parsimonious representations. A key finding

is that the functional polynomial regression model can be represented as a polynomial regression

model in the functional principal component scores of the predictor process. Accordingly, it is

convenient to implement the model as polynomial regression in the principal components of

predictor processes. The representation in terms of functional principal components makes it

possible to include both densely and sparsely observed predictor trajectories, allows for simple

numerical implementation, and enables us to obtain asymptotic consistency results within the

framework of a general measurement model.

2. FUNCTIONAL LINEAR AND POLYNOMIAL REGRESSION

2·1. From functional linear to quadratic regression

The functional regression models we consider include a functional predictor paired with a scalar

response. The predictor process is assumed to be square integrable and is defined on a finite

domain T , with mean function E{X(t)} = μX(t) and covariance function cov{X(s),X(t)} =

G(s, t) for s, t ∈ T . The covariance function G can be decomposed by means of the eigenval-

ues and eigenfunctions of the autocovariance operator of X. Denoting eigenvalue/eigenfunction

pairs by {(λ1, φ1), (λ2, φ2), . . .}, ordered according to λ1 ≥ λ2 ≥ . . ., one obtains G(s, t) =∑k λkφk(s)φk(t). The well-established linear functional regression model with scalar response

(Ramsay & Dalzell, 1991) is given by

E(Y |X) = μY +∫T

β(s)Xc(s) ds, (1)

where Xc(t) = X(t) − μX(t) denotes the centered predictor process. The regression parameter

function β is assumed to be smooth and square integrable.

To estimate the function β, some form of regularization is needed, for which we employ trun-

cated basis representations of predictor processes X. We choose the (orthonormal) eigenfunc-

tions of predictor processes X as basis; alternative choices such as the wavelet basis may prove

convenient in some applications (Morris & Carroll, 2006). When selecting the eigenbasis, one

Page 5: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

214

215

216

217

218

219

220

221

222

223

224

Biometrika-08-035.R2 5

takes advantage of the equivalence between the predictor process and the countable sequence of

uncorrelated functional principal components. These scores are the random coefficients ξk in the

Karhunen-Loeve representation

X(t) = μX(t) +∑

k

ξkφk(t), t ∈ T , (2)

with ξk =∫

Xc(t)φk(t)dt. They are uncorrelated with zero mean and var(ξk) = λk.

While the functional linear model in (1) has been well investigated and has proven useful

in many applications, it is desirable to develop a class of more general “parametric” functional

regression models for situations where the functional linear model is inadequate. If a functional

linear model does not provide an appropriate fit, a natural alternative is to move from a linear

to a quadratic functional regression model, similarly to the situation in ordinary regression. This

approach follows the classical strategy to embed an ill-fitting model into a larger class of models.

It is thus natural to consider a quadratic regression relationship when moving one step beyond

the functional linear model, or on occasion a functional relation that involves a polynomial of

order higher than 2. The functional linear model is always included as a special case.

The quadratic functional regression relationship involves a square integrable univariate linear

parameter function β(t) and a square integrable bivariate quadratic parameter function γ(s, t)

and is given by

E(Y |X) = α +∫T

β(t)Xc(t)dt +∫T

∫T

γ(s, t)Xc(s)Xc(t)dsdt, (3)

where α is an intercept. The linear part is seen to be the same as in model (1), while a quadratic

term has been added. This term reflects that beyond the effect that the ensemble of the values of

Xc(t), t ∈ T , has on the response, the products {Xc(s)Xc(t)}, s, t ∈ T , and in particular the

square terms {Xc(t)}2, t ∈ T , are included as additional predictors.

Since the eigenfunctions {φk}k=1,2,... of the process X form a complete basis, the regression

parameter functions in (3) can be represented in this basis,

β(t) =∞∑

k=1

βkφk(t), γ(s, t) =∞∑

k,�=1

γk�φk(s)φ�(t), (4)

Page 6: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

241

242

243

244

245

246

247

248

249

250

251

252

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

268

269

270

271

272

6 F. YAO AND H. G. MULLER

for suitable sequences (βk)k=1,2,... and (γk�)k,�=1,2,... with∑

k β2k < ∞ and

∑k,� γ2

k� < ∞. Sub-

stituting representations (2), (4) for the components in the quadratic model (3) and applying the

orthonormality property of the eigenfunctions, one finds that the functional quadratic model in

(3) can be alternatively expressed as a function of the scores ξk of predictor processes X,

E(Y |X) = α +∞∑

k=1

βkξk +∞∑

k=1

k∑�=1

γk�ξkξ�, (5)

where γk� = 2γk� for k �= � and γk� = γk� for k = �. We also note that model (3) implies the

constraint μY = E(Y ) = α +∑

k γkkλk, i.e., the intercept α has the representation α = μY −∑k γkkλk.

2·2. Functional polynomial regression

Considering the more general case of a polynomial regression, we define the p-th order (p ≥ 3)

functional polynomial model in analogy to (3) as follows,

E(Y |X) = α +∫T

β(t)Xc(t)dt +∫T 2

γ(s, t)Xc(s)Xc(t)dsdt (6)

+∫T 3

γ3(t1, t2, t3)Xc(t1)Xc(t2)Xc(t3)dt1dt2dt3 + . . .

+∫T p

γp(t1, . . . , tp)Xc(t1) . . . Xc(tp)dt1 . . . dtp,

where again α is the intercept, and β, γ, γj , 3 ≤ j ≤ p, are the linear, quadratic, and jth order

regression parameter functions, defining the effects of the corresponding interactions. Using the

same arguments as those leading to (5), this model can also be written in terms of the predictor

functional principal components,

E(Y |X) = α +∑j1≥1

βj1ξj1 +∑

j1≤j2

γj1j2ξj1ξj2 +∑

j1≤j2≤j3

γj1j2j3ξj1ξj2ξj3 (7)

+ . . . +∑

j1≤...≤jp

γj1...jpξj1 . . . ξjp,

where the terms in this representation are self-explanatory.

The interpretation of these polynomial models is complex. The presence of a jth order inter-

action terms means that the joint values of the predictor process at j time points have an effect on

the outcome, in addition to the joint effects of the process values at � time points for all � < j. For

Page 7: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

289

290

291

292

293

294

295

296

297

298

299

300

301

302

303

304

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

Biometrika-08-035.R2 7

the quadratic model, the interaction effects at two time points are added to the effects at a single

time point. The interaction effects are perhaps easier to understand in terms of the functional

principal components as in version (7), where the interpretation is the same as for the conven-

tional polynomial regression model which includes all possible interaction terms, The functional

principal components themselves are projections of the predictor process in the directions deter-

mined by the eigenfunctions and accordingly are interpreted in terms of the shape of their corre-

sponding eigenfunctions, often as contrasts between positively and negatively weighted parts of

the predictor process (Castro et al., 1986; Jones & Rice, 1992; Izem & Kingsolver, 2005).

For the models that are expressed in terms of the functional principal components of the forms

(5), (7), one can easily introduce variations by omitting some of the interaction terms. For exam-

ple, a noteworthy variation of the functional quadratic model is

E(Y |X) = α +∑

k

βkξk +∑

k

γkkξ2k. (8)

If expressed in the form of (3), model (8) imposes a restriction on the quadratic parameter func-

tion γ(s, t), which in this case will be of “diagonal” form γ(s, t) =∑

k γkkφk(s)φk(t). This

version of the functional quadratic regression model does not include interaction terms.

2·3. Explicit representations

The functional population normal equations provide solutions for functional regression mod-

els under certain regularity conditions (He et al., 2000). The functional least squares deviation,

expressed in terms of the parameters βk and γk� in representation (4) is given by

Q {(βk), (γk�), k, � = 1, 2, . . . , k ≤ �} = E

⎛⎝Y − α −∑

k

βkξk −∑k≤�

γklξkξ�

⎞⎠2

. (9)

One then obtains the normal equations by differentiating Q with respect to the sets of parameters.

Using condition (A1) in the Appendix, the functional normal equations lead to the solutions

α = μY −∑

k

γkkλk, βk = ηk/λk, for ηk = E(ξkY ),

γk� = ρk�/(λkλ�), for k < �, ρk� = E(ξkξ�Y ), (10)

γkk = (ρkk − μY λk)/(τk − λ2k), for τk = E(ξ4

k).

Page 8: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

337

338

339

340

341

342

343

344

345

346

347

348

349

350

351

352

353

354

355

356

357

358

359

360

361

362

363

364

365

366

367

368

8 F. YAO AND H. G. MULLER

We note that the representation of the linear parameter function β(t) that results from (10) is

identical to that in the functional linear model, which is β(t) =∑∞

k=1 ηkλ−1k φk(t).

3. INFERENCE FOR FUNCTIONAL QUADRATIC REGRESSION

3·1. Estimation

Our starting point for modeling are the actual observations, which consist either of densely

spaced and non-random (dense design) or alternatively of sparse and randomly (irregularly)

spaced repeated measurements (sparse design) of the predictor trajectories Xi, contaminated

with additional measurement errors. The dimension of the vector of noisy recordings U i =

(Ui1, . . . , UiNi)T of the random trajectory Xi made at (fixed or random time) points T i =

(Ti1, . . . , TiNi)T is subject-specific but non-random for dense designs, and is a random variable

for sparse designs, assumed to be independently and identically distributed as N and independent

of all other random variables. The measurement errors εij are assumed to be independent and

identically distributed with Eεij = 0, E(ε2ij) = σ2, and independent of the functional principal

components ξik in (2), leading to the representation

Uij = Xi(Tij) + εij = μX(Tij) +∞∑

k=1

ξikφk(Tij) + εij , Tij ∈ T . (11)

Estimates μX , G, λk, φk and σ2 of the underlying population mean function μX , covariance

function G, eigenvalues λk, eigenfunctions φk and error variance σ2 are easily obtained by apply-

ing a nonparametric functional approach (Yao et al., 2005a), implemented in the PACE package,

available at http://www.stat.ucdavis.edu/∼mueller/; compare also Rice & Silverman (1991).

A key step is estimation of the regression parameter functions β and γ, based on representa-

tions (4) and (10). The cross-covariance surfaces

C1(t) = cov{X(t), Y } =∞∑

k=1

ηkφk(t), t ∈ T ,

C2(s, t) = E{X(s)X(t)Y } =∞∑

k,�=1

ρk�φk(s)φk(t), s, t ∈ T , (12)

Page 9: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

385

386

387

388

389

390

391

392

393

394

395

396

397

398

399

400

401

402

403

404

405

406

407

408

409

410

411

412

413

414

415

416

Biometrika-08-035.R2 9

can be estimated by using “raw” covariances C(1)i (Tij) = {Uij − μX(Tij)}Yi, 1 ≤ j ≤ Ni,

and C(2)i (Tij , Til) = {Uij − μX(Tij)}{Uil − μX(Til)}Yi, 1 ≤ j �= l ≤ Ni, as input for one- and

two-dimensional smoothing steps; see (29) in the Appendix for further details. Note that one

needs to remove the diagonal elements C(2)i (Tij , Tij) prior to this smoothing step in order not

to contaminate the estimates with the measurement error in Uij , in analogy to the situation for

auto-covariance surface estimation (Yao et al., 2005a). The resulting estimates are denoted by

C1 and C2.The bandwidths for the one- and two-dimensional smoothing steps needed to obtain

C1 and C2 are chosen by generalized cross-validation, similarly to the choices implemented in

PACE; in related studies, the resulting estimation errors were found to be not overly sensitive to

these choices, see e.g. Liu & Muller (2009).

From (12) one then obtains estimates of the quantities ηk and ρk� in (10), k, � = 1, . . . ,K,

where K is the number of eigenfunctions included for approximating the predictor process X,

ηk =∫T

C1(t)φk(t)dt, ρk� =∫T

∫T

C2(s, t)φk(s)φ�(t)dsdt, (13)

by observing that the relations in (13) hold for the corresponding population quantities. Using

Y = n−1∑n

i=1 Yi and plugging in the estimates (13), one then obtains estimates of the regres-

sion coefficients in (5),

α = Y −K∑

k=1

γkkλk, βk = λ−1k ηk, γk� = (λkλ�)−1ρk�, 1 ≤ � < k ≤ K. (14)

Regarding estimation of γkk, for dense designs, we use the moment estimates τDk =

n−1∑n

i=1 ξI 4ik for fourth moments τk, where ξI

ik is based on the integral method (17).

Neither the proposed estimation schemes nor the consistency results require Gaussianity for

the dense design case. The situation is different for the sparse case, where the integration-based

estimates ξIik, used for the estimation of γkk, are not consistent, due to the sparseness of the

design. This difficulty can be overcome by making the assumption that τk = 3λ2k, k = 1, 2, . . .,

which then makes it possible to extend the above estimation scheme to the sparse case, yielding

γDkk = (ρkk − Y λk)/(τk − λ2

k), γSkk = λ−2

k (ρkk − Y λk)/2, (15)

Page 10: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

433

434

435

436

437

438

439

440

441

442

443

444

445

446

447

448

449

450

451

452

453

454

455

456

457

458

459

460

461

462

463

464

10 F. YAO AND H. G. MULLER

where superscripts D and S denote dense and sparse designs. The resulting estimates for the

regression functions are

β(t) =K∑

k=1

βkφk(t) and γ(s, t) =K∑

k=1

k∑�=1

γk�φk(s)φ�(t), s, t ∈ T , (16)

where βk and γk� are as in (14), and γkk as in (15).

In the sparse case, for the simpler functional linear model (e.g. Yao et al., 2005b), Gaussianity

or other restrictive assumptions are not needed for consistent estimation of the regression param-

eter function (corresponding to β with the component γ absent). However, in the quadratic case,

sparse designs require the additional assumption τk = 3λ2k, k = 1, 2, . . . for consistent estima-

tion of the parameter τk; this assumption is satisfied under Gaussianity (which is not required at

this stage). For the choice of the number of included components K one may use cross-validation

or model selection criteria such as pseudo-BIC (Bayesian information criterion); we adopt the

latter, see (30) in the appendix.

3·2. Prediction

We next aim for the prediction of an unknown response Y ∗, based on noisy observations U∗ =

(U∗1 , . . . , U∗

N∗)T of a new predictor trajectory X∗(·), taken at T ∗ = (T ∗1 , . . . , T ∗

N∗)T . For the

dense design case, the traditional integral estimates of functional principal components ξ∗k, based

on the definition ξ∗k =∫{X∗(t) − μX(t)}φk(t)dt, are

ξ∗Ik =N∗∑j=2

{U∗j − μX(T ∗

j )}φk(T ∗j )(T ∗

j − T ∗j−1), k = 1, . . . ,K. (17)

The interaction and quadratic terms, as needed for the functional quadratic model, are obtained

directly by ξ∗Ik ξ∗I� , k, � = 1, . . . ,K.

Considering the sparse design case, define ξ∗K = (ξ∗1 , . . . , ξ∗K)T , φ∗k =

{φk(T ∗1 ), . . . , φk(T ∗

N∗)}T , the K × N∗ matrix H = (λ1φ∗1, . . . , λKφ∗

K)T and Λ =

diag{λ1, . . . , λK}. The best prediction of the elements of the matrix ξ∗Kξ∗TK , given U∗, is

the conditional expectation E(ξ∗Kξ∗TK |U∗), which under Gaussian assumptions is found to be(

ξ∗kξ∗�

P )1≤k,�≤K

= E(ξ∗Kξ∗TK |U∗) = ξ∗K ξ

∗TK + Λ − HΣ−1

U∗HT . (18)

Page 11: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

481

482

483

484

485

486

487

488

489

490

491

492

493

494

495

496

497

498

499

500

501

502

503

504

505

506

507

508

509

510

511

512

Biometrika-08-035.R2 11

Here ξ∗K = (ξ∗P1 , . . . , ξ∗PK )T is a vector of estimates ξ∗Pk as in (28), obtained in a conditioning

step. The estimate ΣU∗ of ΣU∗ is obtained by substituting estimates G and σ2 for G and σ2.

For both dense and sparse designs, the functional quadratic prediction of the response Y ∗ from

the measurements U∗ is then given by

Y ∗ = α +K∑

k=1

βk ξ∗k +

K∑k=1

k∑�=1

γk�ξ∗kξ

∗� , (19)

where ξ∗k, ξ∗kξ∗� refer to ξ∗Ik , ξ∗Ik ξ∗I� as in (17) for dense designs and to ξ∗Pk (28), ξ∗kξ

∗�

P(18) for

sparse designs and α, βk and γk� are obtained as in (14) and (15).

In many applications a simple empirical measure to gauge the strength of the regression re-

lation is useful. One such measure that coincides with the usual coefficient of determination in

a simple linear regression, and in general provides a comparison of the prediction error when

using a simple sample mean of the responses for prediction with that using a proposed predictor

is the following “quasi”-R2

R2Q = 1 −

∑ni=1(Yi − Yi)2∑ni=1(Yi − Y )2

, (20)

where the predicted response Yi for the ith subject is as in (19). This quasi-R2 does not automat-

ically increase when predictors are added to a model and permits straightforward interpretation

and model comparison. We note that the estimation scheme we have outlined for functional

quadratic regression can be analogously extended to the case of functional polynomial mod-

els. Such an extension is particularly straightforward for polynomial models that do not include

interaction terms such as the variant of the quadratic model in (8).

4. THEORETICAL PROPERTIES

We consider the asymptotic consistency of the estimated regression functions in a functional set-

ting where the number of functional principal components depends on the sample size n, i.e.,

K = K(n), and tends to infinity as the sample size increases, n → ∞. This allows for increas-

ingly better approximations of the underlying functions for larger samples, adequately reflecting

the nonparametric nature of the problem. In practice the choice of K will depend on the intrinsic

Page 12: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

529

530

531

532

533

534

535

536

537

538

539

540

541

542

543

544

545

546

547

548

549

550

551

552

553

554

555

556

557

558

559

560

12 F. YAO AND H. G. MULLER

structural complexity and also estimation accuracy of the covariance structure. Writing ‖β‖ =

{∫T β2(t)dt}1/2, ‖γ‖T 2 = {

∫T

∫T γ2(s, t)dsdt}1/2, βK(t) =

∑Kk=1 βkφk(t) and γK(s, t) =∑K

k=1

∑k�=1 γk�φk(s)φ�(t), we note that the square integrability properties imply, as n → ∞,

Rβ,K ≡ ‖β − βK‖ = (∞∑

k=K+1

β2k)1/2 → 0,

Rγ,K ≡ ‖γ − γK‖T 2 = (∞∑

k=K+1

k∑�=1

γ2k�)

1/2 → 0. (21)

In the following, h denotes the bandwidth for estimating the surface C2 in (29). The technical

assumptions are listed in Appendix A2, where also the sequence K = K(n) and the constants

πk, which play a role in the following results, are defined in equation (31). The increasing se-

quence K(n) places an upper bound on the number K of predictor components that are included

in the functional quadratic regression. The following result provides consistency of the estimated

model components defined at (14). For conciseness, let ID denote an indicator for terms that are

to be included for dense designs, i.e. ID = 1 for dense and = 0 for sparse designs.

THEOREM 1. Under (A1), (A2), (A3.1)-(A3.3) (see Appendix for all assumptions) for dense

designs or under (A1), (A2), (A4.1) and (A4.2) for sparse designs, as n → ∞, for any sequence

K = K(n) ≤ K(n), with πk = 1/λk + 1/min1≤j≤k(λj − λj+1),

α − α = Op

( 1n1/2h2

K∑k=1

πk

λ2k

+IDK

N1/2

)+ op

{( ∞∑k=K+1

γ2kk

)1/2}

(22)

‖β − β‖ = Op

( 1n1/2h2

K∑k=1

πk

λk+ Rβ,K

), (23)

‖γ − γ‖T 2 = Op

( 1n1/2h2

∑1≤�≤k≤K

πk + π�

λkλ�+

IDK2

N1/2+ Rγ,K

). (24)

We next consider the consistency of the prediction of Y ∗ for a new subject or sampling unit.

For dense designs, the prediction given the data (U∗1 , . . . , U∗

N∗) targets E(Y ∗|X∗) as in (5). For

sparse designs, due to the sparsity of the available measurements U∗ for the predictor trajectory

X∗, the target of the prediction is conditional on these measurements and thus becomes

Y ∗ = E{E(Y ∗|X∗)|U ∗} = α +∞∑

k=1

βkE(ξ∗k|U ∗) +∑

1≤�≤k≤∞γk�E(ξ∗kξ∗� |U∗). (25)

Page 13: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

577

578

579

580

581

582

583

584

585

586

587

588

589

590

591

592

593

594

595

596

597

598

599

600

601

602

603

604

605

606

607

608

Biometrika-08-035.R2 13

THEOREM 2. Let Y ∗ be the prediction (19) for both dense and sparse designs, E(Y ∗|X∗) as

in (5), and Y ∗ as in (25).

(i) Under (A1)-(A3) for dense designs, as n → ∞,

Y ∗ − E(Y ∗|X∗) = Op

( 1n1/2h2

∑1≤�≤k≤K

πk + π�

λkλ�+

K2

N∗1/2

)+ op(Rβ,K + Rγ,K). (26)

(ii) Under (A2) and (A4) for sparse designs, as n → ∞,

Y ∗ − Y ∗ = Op

( 1n1/2h2

∑1≤�≤k≤K

πk + π�

λkλ�

)+ op(Rβ,K + Rγ,K). (27)

This result establishes the consistency of the prediction of the response, given the data for a new

subject. Extensions to more general polynomial models are analogous.

5. SIMULATION STUDIES

We studied the Monte Carlo performance of the functional quadratic model (3) in compar-

isons with the functional linear model (FLM) (1) for both dense and sparse designs. Each of

400 simulation runs consisted of a sample of n = 100 predictor trajectories Xi, with mean

function μX(s) = s + sin (s), 0 ≤ s ≤ 10, and a covariance function derived from two eigen-

functions, φ1(s) = −5−1/2 cos (πs/10) and φ2(s) = 5−1/2 sin (πs/10), 0 ≤ s ≤ 10. The cor-

responding eigenvalues were chosen as λ1 = 4, λ2 = 1 and λk = 0, k ≥ 3, the measurement

errors in (11) as εij∼N(0, 0.52). To study the effect of Gaussianity, which is of interest espe-

cially for the sparse design case, we considered two settings: (i) ξik ∼ N (0, λk) (Gaussian); (ii)

ξik are generated from the mixture of two normals, N{(λk/2)1/2, λk/2) with probability 1/2

and N{−(λk/2)1/2, λk/2} with probability 1/2 (mixture distribution). The number of measure-

ments where each trajectory was sampled was selected from {30, . . . , 34}, respectively from

{5, . . . , 9} with equal probability for dense respectively sparse cases.

The response variables were generated from the regression model Yi = m(ξi1, ξi2, . . .) + εi,

where m(·) characterizes functional relationship of the predictor trajectories Xi with the re-

sponses Yi, with errors εi ∼ N(0, 0.1), so that μY = 0 and σ2Y = 0.1. We compared the per-

formance of functional quadratic and linear models under two situations: (a) quadratic with

Page 14: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

625

626

627

628

629

630

631

632

633

634

635

636

637

638

639

640

641

642

643

644

645

646

647

648

649

650

651

652

653

654

655

656

14 F. YAO AND H. G. MULLER

m(ξ1, ξ2, . . .) = ξ1 + ξ2 + ξ21 + ξ2

2 + ξ1ξ2; (b) linear with m(ξ1, ξ2, . . .) = 2ξ1 + ξ2/2. The

functional quadratic model was implemented as described in Section 3, with number of model

components for the predictor processes automatically selected as the smallest number explaining

at least 90% of total variation. To evaluate the prediction of new responses for future subjects,

we generated 100 new predictor trajectories and responses X∗i and Y ∗

i , respectively, with mea-

surements U∗ij taken at the same time points as Uij . The performance measure tabulated in Ta-

ble 1 is relative prediction error RPE =∑

i(Y∗i − Y ∗

i )2/Y ∗2i . The results suggest that functional

quadratic regression leads to similar prediction errors as functional linear regression when the

underlying regression function m(·) is linear, while functional quadratic regression is associated

with dramatically improved errors across all scenarios when the underlying function is quadratic.

6. APPLICATION TO SPECTRAL DATA

To demonstrate the usefulness of the proposed methods, we explore the functional relation-

ship between absorbance trajectories (longitudinally measured functional predictor) and the fat

content (response) of meat. The data are recorded on a Tecator Infratec Food and Feed Analyzer,

a spectrometer that works in the wavelength range 850 - 1050 nm. Each of the n = 199 included

sampling units consists of the set of measurements made on a specific finely chopped pure meat

specimen (Borggaard & Thodberg, 1992). These measurements include the fat content, obtained

analytically, and a 100 channel spectrum of absorbances, obtained as log transform of the trans-

mittance measured by the spectrometer, yielding 100 equidistant points at which the spectrum

is recorded, giving rise to a dense design. The task is to predict fat contents from the spectrum,

setting the stage for a functional regression analysis.

A subsample of 50 randomly selected spectra (predictor trajectories) displayed in Figure 1

indicates that the predictor trajectories are smooth. The estimated mean function is in the left

panel of Figure 2. Four eigenfunctions are selected for modeling which explain more than 99.8%

of the total variation (right panel of Figure 2). The estimated univariate linear function β(t) and

the bivariate quadratic surface γ(s, t) in Figure 3 each exhibit several broad peaks and valleys;

Page 15: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

673

674

675

676

677

678

679

680

681

682

683

684

685

686

687

688

689

690

691

692

693

694

695

696

697

698

699

700

701

702

703

704

Biometrika-08-035.R2 15

especially spectral values near 40 units are strongly weighted. The quadratic response surface

obtained from the fitted quadratic regression model is presented in Figure 4.

For prediction, one typically will choose additional components, guided by one-leave-out pre-

diction errors. We compare the prediction performance of the proposed functional quadratic

model with functional linear regression and also with partial least squares, a popular approach

in chemometrics; we refer to Xu et al. (2007) and the references therein. The results for predic-

tion errors and quasi-R2 (20) in dependence on the number of included components in Table 2

demonstrate that for more than three components (as required for reasonably good prediction)

the error of functional quadratic model is consistently smaller than that for partial least squares

which in turn is smaller than that of functional linear regression.

ACKNOWLEDGEMENTS

We wish to thank two referees for constructive comments. This research was supported by the

National Science Foundation and a NSERC Discovery grant.

APPENDIX

A1. Estimation Procedures

Using the notation introduced in Section 3, the PACE estimates (best linear estimates) (Yao et al.,

2005a) of ξ∗k, conditional on the observations, are given by

ξ∗Pk = λkφ∗Tk Σ−1

U∗(U ∗ − μ∗X), k = 1, . . . ,K, (28)

where μ∗X = (μX(T ∗

1 ), . . . , μX(T ∗m))T . It follows from results in Muller (2005) that, as designs

become dense, ξ∗Ik (17) and ξ∗Pk (28) can be considered asymptotically equivalent. Smoothing

kernels κ1, κ2 are compactly supported smooth densities with zero means and finite variances

and are implemented with suitable bandwidth sequences b and h. With f{θ, (s, t), (Tij , Til)} =

θ0 + θ11(s − Tij) + θ12(t − Til), the one- and two-dimensional smoothers to estimate C1(t) and

C2(s, t) in (12) are obtained by minimizing

Page 16: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

721

722

723

724

725

726

727

728

729

730

731

732

733

734

735

736

737

738

739

740

741

742

743

744

745

746

747

748

749

750

751

752

16 F. YAO AND H. G. MULLER

n∑i=1

Ni∑j=1

κ1(Tij − t

b){C(1)

i (Tij) − α0 − α1(t − Tij)}2,

n∑i=1

∑1≤j �=l≤Ni

κ2(Tij − s

h,Til − t

h)[C

(2)i (Tij , Til) − f{θ, (s, t), (Tij , Til)}

]2, (29)

with respect to α = (α0, α1)T and θ = (θ0, θ11, θ12)T , which yields C1(t) = α0(t) and

C2(s, t) = θ0(s, t). The number of included components is chosen by minimizing

BIC(K) ∝n∑

i=1

Ni∑j=1

[− 1

2σ2

{Uij − μ(Tij) −

K∑k=1

ξikφk(Tij)}2

]+ K log(

n∑i=1

Ni). (30)

A2. Technical Assumptions and Proofs

For model (3) or (5) to be well defined in the least squares sense, we require the following

moment conditions for predictor processes. Let ν1 and ν2 be positive integers.

(A1)∑∞

k=1 Eξ4k < ∞, E(ξν1

k ξν2� ) = Eξν1

k Eξν2� for ν1 + ν2 = 3 and 1 ≤ k, � < ∞; E(ξν1

k ξν2� ) =

Eξν1k Eξν2

� for ν1 + ν2 = 4 and 1 ≤ k �= � < ∞.

Let b∗ = b∗(n), h∗ = h∗(n), h = h(n) denote the bandwidths for estimating μX (26), G (27)

and σ (2) in Yao et al. (2005a). The Fourier transforms of κ1 and κ2 are given by κF1 (u) =∫

exp(−iut)κ1(t)dt and κF2 (u, v) =

∫exp(−(iut + ivs)κ2(s, t)ds dt respectively. The follow-

ing assumptions are needed for both fixed and random designs,

(A2.1) max(b∗, h∗, h, b, h) → 0, min(nb∗4, nh4, nb4) → ∞, max(nb∗6, nh6, nb6) < ∞,

min(nh∗6, nh6) → ∞, max(nh∗8, nh8) < ∞, as n → ∞. Furthermore, assume that

h = O{min(b1/2, b1/21 , h1/2, h∗1/2)} for simplicity.

(A2.2) κF1 and κF

2 are absolutely integrable,∫|κF

1 (u)|du < ∞,∫ ∫

|κF2 (u, v)|dudv < ∞.

(A2.3) n−1/2h−2∑K

k=1

∑k�=1(πk + π�)(λkλ�)−1 → 0, as n → ∞, where

DX =∫T 2{G(s, t) − G(s, t)}2dsdt, δk = min1≤j≤k(λj − λj+1),

K = inf{j ≥ 1 : λj − λj+1 ≤ 2DX} − 1, πk = 1/λk + 1/δk.(31)

To obtain consistent functional principal component estimates for dense designs, we re-

quire both the pooled data across all subjects and the data from each subject to be dense

in T . Denote the sorted time points across all subjects by a0 ≤ T(1) ≤ . . . ≤ T(N)

≤ b0, and

let Δ = max{T(m) − T(m−1) : m = 1, . . . , N + 1}, where N =∑n

i=1 Ni, T = [a0, b0], t(0) =

Page 17: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

769

770

771

772

773

774

775

776

777

778

779

780

781

782

783

784

785

786

787

788

789

790

791

792

793

794

795

796

797

798

799

800

Biometrika-08-035.R2 17

a0, and t(N+1) = b0. For the ith subject, suppose that the time points Tij have been or-

dered non-decreasingly. Let Δi = max{Tij − Ti,j−1 : j = 1, . . . , Ni + 1} and Δ∗ = max{Δi :

i = 1, . . . , n}, where ti0 = a0 and ti,ni+1 = b0, and N = N/n. Put Nmax = max{Ni : i =

1, . . . , n} and Nmin = min{Ni : i = 1, . . . , n}. Denote the distribution that generates Uij for

the ith subject at Tij by Ui(t)∼U(t) with density gU (u; t). Let g∗U (u1, u2; t1, t2) be the density

of (U(s1), U(t2)) and ‖f‖∞ = supt∈T |f(t)| for any function with support T . The following

assumptions are for the case of dense designs, where (A3.3) is needed for consistent estimation

of τ = E(ξ4k) and (A3.4) for consistency of the prediction.

(A3.1) Δ = O(min{n−1/2b∗−1, n−1/2h−1, n−1/2b−1, n−1/4h∗−1, n−1/4h−1}), Δ∗ = O(1/N ),

and C1N ≤ Nmin ≤ Nmax ≤ C2N for some C1, C2 > 0, supt∈T E{U4(t)} < ∞.

(A3.2) (d2/dt2)gU (u; t) is uniformly continuous on �× T ; {d2/(dt�11 dt�22 )}g∗U (u1, u2; t1, t2) is uni-

formly continuous on �2 × T 2, for �1 + �2 = 2, 0 ≤ �1, �2 ≤ 2.

(A3.3) K2 = o(N1/2), maxk≤K

‖φ′k‖∞ = O(N1/2); E(‖X ′‖∞) < ∞, E(‖X ′2‖2

∞) = o(N).

(A3.4) K2 = o(N∗1/2), maxk≤K

‖φ′k‖∞ = O(N∗1/2), where K is as in (31).

For sparse designs, denote the marginal and joint densities of T , (T,U) and (T1, T2, U1, U2)

by gT (t), gU (t, u), g∗U (t1, t2, u1, u2) g(t, u) and g2(t1, t2, u1, u2). The following assumptions

are only needed for sparse designs; (A4.1) and (A4.2) guarantee basic regularity and smoothness

requirements, while the Gaussian assumption (A4.3) is needed for consistency of predictions.

(A4.1) (Tij , Uij) and (T ∗j , U∗

j ) are distributed as (T,U), Tij , T∗j are distributed as T (i = 1, . . . , n,

j = 1, . . . , Ni). The random variables N1, . . . , Nn, are distributed as N∗∼N and are inde-

pendent of all other random variables. The centered fourth moment is finite, i.e., E[{U −

μX(T )}4] < ∞.

(A4.2) gT (t) > 0 and (d�/dt�)gT (t) is continuous on T ; (d2/dt2)gU (t;u) is uniformly continuous

on T × �; {d2/(dt�11 dt�22 )}g∗U (t1, t2, u1, u2) is uniformly continuous on T 2 ×�2, for �1 +

�2 = 2, 0 ≤ �1, �2 ≤ 2.

(A4.3) The predictor process X ∈ L2(T ) is a Gaussian process.

Weak uniform convergence rates for estimated mean μX and covariance functions G for sparse

designs were derived in Yao et al. (2005a) and analogous arguments apply to C1 and C2 in

Page 18: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

817

818

819

820

821

822

823

824

825

826

827

828

829

830

831

832

833

834

835

836

837

838

839

840

841

842

843

844

845

846

847

848

18 F. YAO AND H. G. MULLER

(29). Minor modifications of these results yield analogous results for fixed designs. Combining

Theorem 1 of Hall & Hosseini-Nasab (2006) and arguments used in the proof of Theorem 2 of

Yao et al. (2005a) sets the stage for the proofs of the main results.

LEMMA 1. Under (A2), (A3.1) and (A3.2) for dense designs or under (A2), (A4.1) and (A4.2)

for sparse designs, it holds that

supt∈T

|μX(t) − μX(t)| = Op(1

n1/2b∗), sup

s,t∈T|G(s, t) − G(s, t)| = Op(

1n1/2h∗2 ),

supt∈T

|C1(t) − C1(t)| = Op(1

n1/2b), sup

s,t∈T|C2(s, t) − C2(s, t)| = Op(

1n1/2h2

), (32)

and as a consequence, σ2 − σ2 = Op(n−1/2h∗−2 + n−1/2h−1). Considering eigenvalues λk of

multiplicity one, φk can be chosen such that

pr( sup1≤k≤K

|λk − λk| ≤ DX) = 1, supt∈T

|φk(t) − φk(t)| = Op(πk

n1/2h∗2 ), k = 1, . . . , K.(33)

LEMMA 2. Under (A1), (A2), (A3.1)-(A3.3) for dense designs or under (A1), (A2), (A4.1) and

(A4.2) for sparse designs, it holds for any K ≤ K that

βk − βk = Op(πk

n1/2h2λk), γk� − γk� = Op(

πk + π�

n1/2h2λkλ�), 1 ≤ � < k ≤ K, (34)

γDkk − γkk = Op(

πk

n1/2h2λ2k

+1

N1/2), γS

kk − γkk = Op(πk

n1/2h2λ2k

) 1 ≤ k ≤ K, (35)

where βk, γk� are as in (14) for � < k, and γDkk, γS

kk are as in (15).

Proof of Lemma 2. It is easy to show the rate for βk, by observing that λ−1k < πk from (31)

and ηk − ηk = Op(πkn−1/2h−2), λ−1

k − λ−1k = Op(n−1/2h∗−2λ−2

k ) from Lemma 1. Regard-

ing γk� for k > �, note that ρk� − ρk� = Op{(φk + π�)n−1/2h−2} and (λkλ�)−1 − (λkλ�) =

Op{n−1/2h∗−2λ−1k λ−1

� (λk + λ�)−1} from Lemma 1 and again λ−1k < πk, leading to (34). For

(35), the rate of γSkk can be obtained by the same argument used for γk� with � < k. To char-

acterize the rate of γDkk, it is necessary to quantify (ξI

ik − ξik), where ξIik is as in (17) for the

ith subject. Using similar arguments as in the proof of Theorem 1 in Yao & Lee (2006), one

finds, with Lemma 1 and (A3.3), that n−1(∑n

i=1 ξI4ik ) − E(ξ4

k) = Op{πk/(n1/2h∗2) + N−1/2}.

Analogous arguments for γk�, � < k, complete the proof. �

Page 19: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

865

866

867

868

869

870

871

872

873

874

875

876

877

878

879

880

881

882

883

884

885

886

887

888

889

890

891

892

893

894

895

896

Biometrika-08-035.R2 19

Proof of Theorem 1. Results (23) and (24) follow immediately from the representations in

(10), (14) and Lemma 2. One can easily show (22) by applying the Cauchy-Schwarz inequality

to∑∞

k=K+1 γkkλk and noting that, as n → ∞,∑∞

k=K+1 λ2k converges even faster to 0 than∑∞

k=K+1 λk. �

Proof of Theorem 2. We first show (26) concerning dense designs. To quantify the prediction

error of (ξ∗Ik − ξ∗k), where ξ∗Ik is as in (17), similar arguments as in the proof of Theorem 1 in Yao

& Lee (2006) imply that ξ∗Ik − ξ∗k = Op{πk/(n1/2h∗2) + N∗−1/2} and analogously ξ∗Ik ξ∗I� −

ξ∗kξ∗� = Op{(πk + π�)/(n1/2h∗2) + N∗−1/2}. Applying Lemma 2, the discrepancy between Y ∗

and Y ∗K =

∑Kk=1 βkξ

∗k +

∑1≤�≤k≤K γk�ξ

∗kξ∗� is seen to be Op

{(n1/2h2)−1

∑1≤�≤k≤K(πk +

π�)(λk + λ�)−1 + K2N∗−1/2}. To find the bound for the remainder term QK = E(Y ∗|X∗) −

Y ∗K =

∑∞k=K+1 βkξ

∗k +

∑∞k=K+1

∑k�=1 γk�ξ

∗kξ

∗� , it is sufficient to show E(Q2

K) = o(R2β,K +

R2γ,K), where Rβ,K , Rγ,K are as in (21). Since E(QK) =

∑∞k=K+1 γkkλk, with the Cauchy-

Schwarz inequality and (A1)

E(Q2K) = var

( ∞∑k=K+1

βkξ∗k +

∞∑k=K+1

k∑�=1

γk�ξ∗kξ∗�

)+

( ∞∑k=K+1

γkkλk

)2

≤∞∑

k=K+1

β2kλk +

∞∑k=K+1

k−1∑�=1

γ2k�λkλ� +

∞∑k=K+1

γ2kkτk +

( ∞∑k=K+1

γ2kk

)( ∞∑k=K+1

λ2k

), (36)

which implies (26). For the case of sparse designs, supt∈T |λkφk(t) − λkφk(t)| =

Op{πk/(n1/2h∗2)} according to the proof of Theorem 2 in Yao et al. (2005a). For ξ∗k =

E(ξ∗k|U ∗) and ξ∗kξ∗� = E(ξ∗kξ∗� |U∗), it is easy to verify that ξ∗Pk − ξ∗k = Op{πk/(n1/2h∗2)}

and ξ∗kξ∗�P− ξ∗kξ

∗� = Op{(πk + π�)/(n1/2h∗2)}. For the truncated quantity Y ∗

K =∑K

k=1 βk ξ∗k +∑

1≤�≤k≤K γk�ξ∗kξ

∗� , we find Y ∗ − Y ∗

K = Op

{(n1/2h2)−1

∑1≤�≤k≤K(πk + π�)(λk + λ�)−1

}.

Considering Y ∗ − Y ∗K = E

( ∑∞k=K+1 βkξ

∗k +

∑∞k=K+1

∑k�=1 γk�ξ

∗kξ∗� |U∗), it suffices to show

that E{(Y ∗ − Y ∗K)2} = o(R2

β,K + R2γ,K), where the expectation is unconditional (with re-

spect to both U∗ and X∗). Noting that E{(Y ∗ − YK)2} ≤ var(Y ∗ − Y ∗K) + (

∑∞k=K+1 γkkλk)2,

E[var

{E(Y ∗ − Y ∗

K |X∗)|U ∗}]≥ 0 and (A3.3), one obtains var(Y ∗ − Y ∗

K) ≤ var{E(Y ∗ −

YK |X∗)}. An upper bound for the right hand side has been obtained in (36), as (A4.3) implies

(A1). This completes the proof of (27). �

Page 20: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

913

914

915

916

917

918

919

920

921

922

923

924

925

926

927

928

929

930

931

932

933

934

935

936

937

938

939

940

941

942

943

944

20 F. YAO AND H. G. MULLER

REFERENCES

BORGGAARD, C. & THODBERG, H. (1992). Optimal minimal neural interpretation of spectra. Analytical Chemistry

64, 545–551.

CAI, T. & HALL, P. (2006). Prediction in functional linear regression. The Annals of Statistics 34, 2159–2179.

CARDOT, H., FERRATY, F., MAS, A. & SARDA, P. (2003). Testing hypotheses in the functional linear model.

Scandinavian Journal of Statistics. Theory and Applications 30, 241–255.

CARDOT, H. & SARDA, P. (2005). Estimation in generalized linear models for functional data via penalized likeli-

hood. Journal of Multivariate Analysis 92, 24–41.

CASTRO, P. E., LAWTON, W. H. & SYLVESTRE, E. A. (1986). Principal modes of variation for processes with

continuous sample curves. Technometrics 28, 329–337.

FARAWAY, J. J. (1997). Regression analysis for a functional response. Technometrics 39, 254–261.

HALL, P. & HOROWITZ, J. L. (2007). Methodology and convergence rates for functional linear regression. The

Annals of Statistics 35, 70–91.

HALL, P. & HOSSEINI-NASAB, M. (2006). On properties of functional principal components analysis. Journal of

the Royal Statistical Society: Series B (Statistical Methodology) 68, 109–126.

HE, G., MULLER, H. G. & WANG, J. L. (2000). Extending correlation and regression from multivariate to functional

data. In Asymptotics in Statistics and Probability, M. L. Puri, ed. VSP International Science Publishers, pp. 301–

315.

IZEM, R. & KINGSOLVER, J. (2005). Variation in continuous reaction norms: Quantifying directions of biological

interest. American Naturalist 166, 277–289.

JAMES, G. M., HASTIE, T. J. & SUGAR, C. A. (2000). Principal component models for sparse functional data.

Biometrika 87, 587–602.

JONES, M. C. & RICE, J. A. (1992). Displaying the important features of large collections of similar curves. The

American Statistician 46, 140–145.

KIRKPATRICK, M. & HECKMAN, N. (1989). A quantitative genetic model for growth, shape, reaction norms, and

other infinite-dimensional characters. Journal of Mathematical Biology 27, 429–450.

LIU, B. & MULLER, H. (2009). Estimating derivatives for samples of sparsely observed functions, with application

to on-line auction dynamics. Journal of the American Statistical Association .

MELAS, V., PEPELYSHEV, A. & CHENG, R. (2003). Designs for estimating an extremal point of quadratic regression

models in a hyperball. Metrika 58, 193–208.

MORRIS, J. S. & CARROLL, R. J. (2006). Wavelet-based functional mixed models. Journal of the Royal Statistical

Society: Series B (Statistical Methodology) 68, 179–199.

MULLER, H.-G. (2005). Functional modelling and classification of longitudinal data. Scandinavian Journal of

Statistics. Theory and Applications 32, 223–240.

Page 21: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

961

962

963

964

965

966

967

968

969

970

971

972

973

974

975

976

977

978

979

980

981

982

983

984

985

986

987

988

989

990

991

992

993

994

Biometrika-08-035.R2 21

MULLER, H.-G. & ZHANG, Y. (2005). Time-varying functional regression for predicting remaining lifetime distri-

butions from longitudinal trajectories. Biometrics 61, 1064–1075.

RAMSAY, J. O. & DALZELL, C. J. (1991). Some tools for functional data analysis. Journal of the Royal Statistical

Society: Series B (Statistical Methodology) 53, 539–572.

RAMSAY, J. O. & SILVERMAN, B. W. (2002). Applied functional data analysis. Springer Series in Statistics. New

York: Springer-Verlag.

RAMSAY, J. O. & SILVERMAN, B. W. (2005). Functional data analysis. Springer Series in Statistics. New York:

Springer, 2nd ed.

RICE, J. A. (2004). Functional and longitudinal data analysis: Perspectives on smoothing. Statistica Sinica , 631–647.

RICE, J. A. & SILVERMAN, B. W. (1991). Estimating the mean and covariance structure nonparametrically when

the data are curves. Journal of the Royal Statistical Society: Series B (Statistical Methodology) 53, 233–243.

SHEN, Q. & FARAWAY, J. (2004). An F test for linear models with functional responses. Statistica Sinica 14,

1239–1257.

XU, L., JIANG, J.-H., WU, H.-L. & YU, G.-L. S. R.-Q. (2007). Variable-weighted pls. Chemometrics and

Intelligent Laboratory Systems 85, 140–143.

YAO, F. & LEE, T. C. M. (2006). Penalized spline models for functional principal component analysis. Journal of

the Royal Statistical Society, Series B 68, 3–25.

YAO, F., MULLER, H.-G. & WANG, J.-L. (2005a). Functional data analysis for sparse longitudinal data. Journal of

the American Statistical Association 100, 577–590.

YAO, F., MULLER, H. G. & WANG, J. L. (2005b). Functional data analysis for sparse longitudinal data. Journal of

the American Statistical Association 100, 577–590.

YAO, F., MULLER, H.-G. & WANG, J.-L. (2005c). Functional linear regression analysis for longitudinal data. The

Annals of Statistics 33, 2873–2903.

ZHAO, X., MARRON, J. S. & WELLS, M. T. (2004). The functional data analysis view of longitudinal data. Statistica

Sinica 14, 789–808.

[Received June 2008. Revised April 2009]

Page 22: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

1009

1010

1011

1012

1013

1014

1015

1016

1017

1018

1019

1020

1021

1022

1023

1024

1025

1026

1027

1028

1029

1030

1031

1032

1033

1034

1035

1036

1037

1038

1039

1040

1041

1042

22 F. YAO AND H. G. MULLER

Table 1. Monte Carlo estimates of the 25th, 50th and 75th percentiles of relative prediction error,

comparing predictions obtained by the Functional Quadratic Model (FQM) and Functional

Linear Model (FLM) for both dense and sparse designs, based on 400 Monte Carlo runs with

sample size n = 100. The underlying regression function is quadratic or linear, and the principal

components of the predictor process are generated from Gaussian or mixture distributions.

Gaussian MixtureDesign True Fitted

25th 50th 75th 25th 50th 75th

FQM 0.037 0.201 1.403 0.033 0.166 0.950Quadratic

FLM 0.204 1.207 30.02 0.162 0.860 17.39Dense

FQM 0.047 0.228 1.168 0.031 0.151 0.808Linear

FLM 0.049 0.269 1.573 0.029 0.162 0.946

FQM 0.067 0.358 3.950 0.040 0.207 1.341Quadratic

FLM 0.233 1.340 29.16 0.157 0.845 16.01Sparse

FQM 0.032 0.167 0.934 0.032 0.159 0.840Linear

FLM 0.038 0.193 1.063 0.037 0.178 0.936

Table 2. Medians of cross-validated relative prediction errors (PE×104) and of quasi-R2 (20)

(R2Q × 100), for varying numbers of components, comparing functional quadratic model (FQM),

functional linear model (FLM) and Partial Least Squares (PLS) for the spectral data.

Components 1 2 3 4 5 6 7 8 9 10

PE 101 84.4 16.9 6.58 2.30 1.79 1.08 1.18 0.60 0.50FQM

R2Q 24.6 32.5 77.8 94.1 98.5 97.8 99.0 99.3 99.6 99.8

PE 81.3 64.7 35.0 11.9 5.31 5.72 5.05 5.57 4.75 5.58FLM

R2Q 22.3 27.9 70.2 90.1 93.8 94.1 94.4 94.4 94.5 94.5

PE 80.1 26.0 12.5 10.1 6.40 5.87 5.68 5.05 4.10 4.04PLS

R2Q 23.1 76.5 86.9 90.8 93.8 94.2 94.6 94.8 96.4 96.5

Page 23: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

1057

1058

1059

1060

1061

1062

1063

1064

1065

1066

1067

1068

1069

1070

1071

1072

1073

1074

1075

1076

1077

1078

1079

1080

1081

1082

1083

1084

1085

1086

1087

1088

1089

1090

Biometrika-08-035.R2 23

0 10 20 30 40 50 60 70 80 90 1002

2.5

3

3.5

4

4.5

Spectrum Channel

Abso

rbanc

e

Fig. 1. The functional predictor trajectories, consisting of

100 channel spectra of log-transformed absorbances, for

50 randomly selected meat specimen.

0 20 40 60 80 1002.7

2.8

2.9

3

3.1

3.2

3.3

3.4

3.5

3.6

Spectrum Channel

Abso

rban

ce

0 20 40 60 80 100−0.25

−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

0.2

Spectrum Channel

Eige

nfun

ctio

ns

2 4 6 8 10

0

1

2

3

4

5

6

7

8x 10

5

Number of FPCs

BIC

Fig. 2. Smooth estimates of the predictor mean func-

tion (left panel) and first (solid), second (dashed), third

(dashed-dot) and fourth (dotted) eigenfunctions (right

panel), as well as the values of the Bayesian Information

Criterion (30), plotted against number of included princi-

pal components (right panel).

Page 24: Functional Quadratic Regressionanson.ucdavis.edu/~mueller/quadreg-r2.pdf · From functional linear to quadratic regression The functional regression models we consider include a functional

1105

1106

1107

1108

1109

1110

1111

1112

1113

1114

1115

1116

1117

1118

1119

1120

1121

1122

1123

1124

1125

1126

1127

1128

1129

1130

1131

1132

1133

1134

1135

1136

1137

1138

24 F. YAO AND H. G. MULLER

0 10 20 30 40 50 60 70 80 90 100−12

−10

−8

−6

−4

−2

0

2

4

6

8

Spectrum Channel

Line

ar R

eg. T

erm

0

10

20

30

40

50

60

70

80

90

100

0

10

20

30

40

50

60

70

80

90

100

−2

−1

0

1

2

Spectrum ChannelSpectrum Channel

Qua

drat

ic Re

g. T

erm

Fig. 3. Estimated parameter functions of functional

quadratic regression, with the linear part β(t) (left panel) ,

and the quadratic part γ(s, t) (right panel) for spectral data.

−100

10

−1−0.500.5150

55

60

65

70

75

80

85

1st FPC2nd FPC

Fitte

d su

rface

−10

0

10

−1−0.5

00.5

1

30

40

50

60

70

80

90

100

3rd FPC1st FPC

Fitte

d su

rface

−10−5

05

10

−0.6

−0.4

−0.2

0

0.2

0.4

40

60

80

100

4rd FPC1st FPC

Fitte

d su

rface

−1−0.5

00.5

1

−1

−0.5

0

0.5

140

50

60

70

80

90

2nd FPC3rd FPC

Fitte

d su

rface

−1

0

1

−0.6−0.4−0.200.20.450

55

60

65

70

75

80

85

90

2nd FPC4th FPC

Fitte

d su

rface

−1

0

1

−0.50

0.5

30

40

50

60

70

80

90

100

4th FPC3rd FPC

Fitte

d su

rface

Fig. 4. Fitted quadratic response surface for bY (fat con-

tent), obtained by varying two predictor principal compo-

nents at a time and fixing the other two at 0. Arranged from

top left to right and then bottom left to right: bY versus

(ξ1, ξ2, 0, 0), (ξ1, 0, ξ3, 0), (ξ1, 0, 0, ξ4), (0, ξ2, ξ3, 0),

(0, ξ2, 0, ξ4) and (0, 0, ξ3, ξ4), respectively.