arXiv:1609.06421v4 [math.ST] 26 Sep 2019 iiat tB,Idaa I,TxsAM B,Vnebl n part and Vanderbilt UBC, comments. A&M, useful Texas for Econometrics MIT, in Indiana, Identification Pedr BC, Porter, at Jack Newey, ticipants M¨uller, referen Whitney Ulrich Conocimiento, Generaci´on Jansson, de de Michael Programa Spanish the by ∗ † eateto cnmc,UiesddCro I eMdi,eal j email: Madrid, de III Carlos Universidad Economics, of 2016. Department 20th, September version: First eiaaercIetfiainadFse Information Fisher and Identification Semiparametric Keywords: no consumption. a in in error aversion measurement risk nonparametric of Finall with measures average studies. and valuation factor contingent discount irregular in show pay I to unemployme willingness Second, for heterogeneity. model nonparametric necessar structural and find a I in First, identification results. regular estimati general of the rates of on applicability results identificatio impossibility irregular) implies “genera (possibly it novel zero, mo implies a in it introduce cases then positive, these I If characterize parameter. first nonparametric f I a information Fisher respectively. the zero, whether or on “qualit depending its of irregular, measure or a as information statistical on based E classification: JEL hspprpoie ytmtcapoc osemiparametri to approach systematic a provides paper This dnicto;Smprmti oes ihrInformatio Fisher Models; Semiparametric Identification; 1;C1 3;C35 C33; C31; C14; nvria alsIId Madrid de III Carlos Universidad unCro Escanciano Carlos Juan etme 5h 2019 25th, September Abstract 1 hnohrcniin od If hold. conditions other when n enme G21-972BI0 thank I PGC2018-096732-B-I00. number ce n he xmlsilsrt the illustrate examples Three on. atAn,Rl io n eia par- seminar and Xiao, Ruli Sant’Anna, o odtosfrsemiparametric for conditions y ” dnicto a eregular be can Identification y”. ,Isuyietfiaino the of identification study I y, † [email protected] funded Research [email protected]. eswt este ierin linear densities with dels rteprmtri positive is parameter the or dnicto ftemedian the of identification prmti ue Equation Euler nparametric spells two with duration nt cpnso h 08Cneec on Conference 2018 the of icipants ie ihrinformation”. Fisher lized dnicto htis that identification c n. ∗ 1 Introduction

Nonparametric identification is considered the benchmark for reliable empirical analysis in eco- nomics. Unfortunately, many nonparametric economic models of interest are unidentified with weak assumptions; see, for example, discrete choice models with nonparametric unobserved het- erogeneity. Yet, certain interesting aspects of such models might be point-identified by the same weak assumptions, a situation henceforth referred to as semiparametric identification. Although this observation has long been recognized in economics (see the early discussion in Hurwicz 1950), no systematic method is currently available for assessing which aspects (i.e. functionals) of a nonparametric structural parameter are identified and which are not. Furthermore, even when nonparametric point-identification holds, there could be many parameters that are only “irregularly identified” (in the sense of being identified, but having a zero Fisher information; see e.g. Chamberlain 1986, Heckman 1990, and Khan and Tamer 2010). This paper aims to establish general conditions for regular and irregular semiparametric identification (or lack thereof) and to relate these conditions to the concept of statistical information or the generaliza- tions proposed herein. The results obtained for irregular identification have important practical implications, as any inferences on such parameters are expected to be unstable in empirical analysis. In particular, if a parameter is irregularly identified, then no regular with a parametric rate of convergence exists (see Chamberlain 1986).1 An important observation for relating identification and information is that identification depends on both linear and nonlinear effects—see Sargan (1983) and Chen, Chernozhukov, Lee and Newey (2014)—while statistical information pertains only to linear effects. To establish a useful link between the two some structure is thus necessary. Moreover, I show that it is hard to give sufficient conditions for identification in nonlinear models allowing for plausibly high levels of irregularity (see Section 4.1). These arguments motivate an initial focus on linear models, i.e. models with densities that are linear in a nonparametric parameter. In these models a more complete and transparent analysis of semiparametric identification is possible, permitting both nonparametric unidentification and high degrees of irregularity. Specifically, I establish necessary and sufficient conditions for regular and irregular semiparametric identification. Because many important economic models are not linear but can be written as linear after reparametrization, or by fixing some parameters, the results obtained for linear models are widely applicable. The analysis of linear models already makes explicit that the separation of irregular identi-

1Notable examples of irregularly identified parameters in econometrics include densities, regression functions, and their derivatives evaluated at fixed points of continuous variables; regression discontinuity parameters, see Cattaneo and Escanciano (2017); binary choice coefficients under Manski’s (1975) conditions, see Chamberlain (1986, 2010); sample selection models, see Heckman (1990), Andrews and Schafgans (1998) and Goh (2017); mixed proportional models, see Hahn (1994) and Ridder and Woutersen (2003); average treatment effects, see Khan and Tamer (2010); or interaction parameters in triangular systems, see Khan and Nekipelov (2018).

2 fication from no identification is a rather delicate issue. The classical Fisher information is not useful when identification is not regular, because it cannot distinguish between irregular identi- fication and no identification (it is zero in both cases). This paper introduces a new “generalized Fisher information” that seems well-suited for irregular cases. If positive, it implies semipara- metric irregular identification when the classical Fisher information is zero and other conditions hold. If zero, it implies impossibility results on rates of convergence for , extending Chamberlain’s (1986) impossibility result to slower rates of convergence than parametric. The identification results are then extended to semiparametric models that are nonlinear in the parameter of interest but linear in nuisance parameters. Examples include commonly used linear and nonlinear panel data models and structural models of unemployment duration (see e.g. Heckman and Singer 1984a, 1984b), among many others. In this setting, it is possible to present simple sufficient conditions for identification of the main parameter based on the gener- alized Fisher information, allowing for moderate irregularity of the main parameter, arbitrary irregularity for nuisance parameters, and nonparametric unidentification. As a general rule, the impossibility results derived in this paper on regular identification and rates of convergence hold for general linear and nonlinear models. The sufficient conditions for identification require more structure, though, because nonlinearities can overwhelm linear effects (cf. Chen et al. 2014). Section 9.3 in the Supplemental Appendix presents sufficient con- ditions for semiparametric identification in nonlinear models, noting that the problem becomes particularly challenging when the model is nonparametrically unidentified. I illustrate the usefulness of the theory by deriving new identification results in the structural model of unemployment duration with two spells and nonparametric heterogeneity recently pro- posed by Alvarez, Borovickov´aand Shimer (2016). These authors first establish that their model is unidentified, and then discuss a prior sign restriction on the parameters that leads to non- parametric identification. Complementing their results, I characterize the identified set without prior sign restrictions and obtain necessary conditions for regular semiparametric identification. As an implication of my results, I show that the proportion of individuals at risk of severe long term unemployment is irregularly identified. A second example shows irregular identification of the median Willingness to Pay (WTP) in contingent valuation studies. The median WTP is identified under weak support conditions, as shown below, and it is an important parameter in this literature. See Carson and Hanemann (2005) for a detailed survey and Lewbel, McFadden and Linton (2011) and references therein for semiparametric estimation of moments of WTP. Using the semiparametric identification strategy of Lewbel (1997), Khan and Tamer (2010) have provided sufficient conditions for the mean WTP to be irregularly identified. I revisit the identification of moments of WTP, and then establish irregular identification of the median WTP, which is a new result. The operator-based approach followed in the present paper is very different from that taken in Khan and Tamer

3 (2010), see Section 6.2 for further discussion. A third example is a nonlinear model for a consumption-based asset pricing Euler equation with a nonparametric measurement error in consumption. This example demonstrates how the results of this paper can be applied to conditional moment models. Nonparametric and semi- parametric treatments of consumption-based asset pricing models, including Newey and Powell (1988), Chen and Ludvigson (2009), Escanciano and Hoderlein (2010), Lewbel, Linton and Srisuma (2011), Chen et al. (2014) and Escanciano et al. (2015), do not deal with measurement error in consumption. Yet, accounting for measurement error is vital for empirical studies that use household-level data, as shown in, e.g., Shapiro (1984), Altonji and Siow (1987), Runkle (1991), and Alan, Attanasio and Browning (2009). I obtain new primitive conditions for regular identification of the discount factor and measures of risk aversion under more general specifica- tions of the marginal utility and the measurement error mechanism than previously considered. In particular, I show that identification of the discount factor is more robust to assumptions about the measurement error than is identification of risk aversion measures. In summary, this paper provides general semiparametric identification and impossibility results for regular identification and irregular rates of estimation, and it shows their utility in some example applications. Inference on irregular parameters is challenging. In particular, establishing rates of convergence can be a cumbersome task, but fortunately, the literature has proposed rate-adaptive estimation and inference methods; see, e.g., Andrews and Shafgans (1998), Khan and Tamer (2010), Chen and Liao (2014), and Chen and Pouzo (2015). Rate- adaptive methods are recommended for inference about irregularly identified parameters. The question of whether irregular identification holds or not, and if so, to what degree, is still of first-order importance, because with irregular identification all estimation methods, including rate-adaptive methods, are expected to be sensitive to the (unknown) data generating process. This paper demonstrates that the strength of this sensitivity, and thus the quality of identification, can be measured by the Fisher information or the generalizations proposed herein. The rest of the paper is organized as follows. After a literature review, Section 3 sets the statistical framework and introduces three examples that will be used throughout the paper. Section 4 characterizes identification in linear models and defines the generalized Fisher infor- mation. Section 5 analyzes semiparametric models. Section 6 studies three examples in detail: the nonparametric unemployment duration model of Alvarez et al. (2016), the median WTP, and the consumption-based Euler equation with measurement error. Section 7 concludes. An Appendix contains proofs of the main results, and a Supplemental Appendix includes further discussion on identification conditions for linear and nonlinear models.

4 2 Literature Review

The identification problem has a long history in economics; see the seminal studies by Koopmans (1949), Hurwicz (1950), Koopmans and Reirsol (1950), Fisher (1966) and Rothenberg (1971). Bekker and Wansbeek (2001) and Dufour and Liang (2014) provide more recent contributions as well as a survey of existing results in parametric settings. Chamberlain (1986) shows that a positive semiparametric Fisher information is necessary for regular estimation in semiparametric models. Under the explicit assumption of nonparametric identification, Van der Vaart (1991) shows the equivalence between a positive semiparametric information and a differentiability condition that is necessary for regular estimation. He briefly discusses an “intuitive” local identification condition, but does not recognize that this condition may be neither sufficient for identification, as shown in Chen et al. (2014), nor necessary, as shown in Sargan (1983). Bickel et al. (1998, Chapter 6) and Ishwaran (1999) present impossibility results on regularity in some exponential and uniform mixture models. Newey (1990) provide further impossibility theorems. There are, of course, many papers reporting sufficient conditions for identification in spe- cific nonparametric models; see the comprehensive reviews in Matzkin (2007, 2013) and Lewbel (2016). The closest to my paper is Chen et al. (2014), who provide sufficient conditions for nonparametric local identification and for regular semiparametric identification for conditional moment models. These authors recognize the difficulty of studying semiparametric irregular identification (see Chen et al. 2014, pg. 796), and do not analyze that case, which is the focus of this paper. The results of the present study also help in interpreting their nonparametric identi- fication conditions in terms of statistical information (see Section 4.1). Chen and Santos (2018) investigate local nonparametric regular overidentification. Khan and Tamer (2010) show irregu- lar identification in two important examples and investigate rate-adaptive inference. Chen and Liao (2014) and Chen and Pouzo (2015) provide general inference results for irregular function- als. Independently of this paper, Bonhomme (2011) studies regular and irregular identification and estimation of average marginal effects in nonlinear panel data models with fixed effects. Also related is the identification analysis of Severini and Tripathi (2006, 2012) in nonparametric IV. These authors characterize the set of linear continuous functionals that are regularly and irregularly identified when the nonparametric structural regression is not necessarily identified. Santos (2011) investigates semiparametric regular estimation in the IV setting. The present paper deals broadly with semiparametric identification in likelihood and condi- tional moment models. Furthermore, it follows the tradition of the seminal work by Rothenberg (1971) in linking the identification problem to the concept of statistical information (and gen- eralizations proposed herein), albeit in a nonparametric setting.

5 3 Setting and Examples

The data is an independent and identically distributed (iid) sample Z1, ..., Zn from a distribution P that belongs to a class of probability measures = P : λ Λ , where Λ is a subset of a P { λ ∈ } Hilbert space (H, , H), with inner product , H and norm H . For example, in parametric m h· ·i h· ·i k·k models Λ R and H = is the Euclidean norm. This paper focusses on nonparametric ⊂ k·k |·| models where Λ is infinite-dimensional, e.g. a subset of a space of probability densities. The nonparametric parameter that generates the data is denoted by λ Λ, i.e. P = P . The goal 0 ∈ λ0 is to find sufficient and necessary conditions for identification of φ(λ) at φ(λ0), for a functional φ(λ):Λ Rp, allowing for the full model to be unidentified at λ . That is, the equation P = 7→ P 0 λ P may have more than one solution in Λ. This setting includes as a special case semiparametric models where λ = (θ, η) Λ=Θ H, Θ Rp and H is a subset of another Hilbert Space ∈ × ⊂ ( , , ). A leading example of functional is the finite-dimensional parameter, i.e. φ(θ, η)= θ. H h· ·iH However, the setting also includes functionals of the nuisance parameter φ(λ) = χ(η), where χ : H Rp, which, despite the name, may be of interest. For example, η can measure 7→ unobserved heterogeneity, and one might be interested in average marginal effects or policy counterfactuals that involve averaging across a heterogeneous population.

To introduce the definition of identification, let fλ be the density of Pλ with respect to (wrt) a σ-finite measure µ. Denote by (λ ) = λ Λ: λ λ H < δ a ball of radius δ around Bδ 0 { ∈ k − 0k } λ0.

Definition (Semiparametric Identification): φ(λ) is locally identified in at φ(λ ) if there P 0 exists δ > 0 such that for all λ (λ ), f = f µ-almost surely (µ-a.s.) implies φ(λ)= φ(λ ). ∈ Bδ 0 λ λ0 0 If this implication holds for all λ Λ, then φ(λ) is (globally) identified at φ(λ ). ∈ 0

To simplify the exposition, I simply write “φ(λ0) is locally identified” rather than “φ(λ) is locally identified in at φ(λ )”, and if “locally” is dropped then identification is meant to P 0 be global. For parametric models, i.e. Λ Rm, Fisher (1966) and Rothenberg (1971) show ⊂ that, with sufficient smoothness of the model, non-singularity of the Fisher information matrix is necessary and sufficient for local identification of λ0. In nonparametric models a positive information is not necessary for identification anymore, and this leads to the classification of identification in regular and irregular (cf. Khan and Tamer 2010).

Definition (Regular and Irregular Semiparametric Identification): φ(λ0) is (locally) regularly (respectively, irregularly) identified if it is (locally) identified and its Fisher Information is positive (respectively, zero).

Identification and regularity/irregularity are separate concepts. So, for example, the negation of regular identification, which is used extensively throughout the paper, entails two possibilities: irregular identification or no identification at all.

6 The usefulness of the results will be illustrated with several examples. In all these applica- tions, what distinguishes this paper from others in the literature is the focus on semiparametric identification and its degree (regular or irregular), rather than on nonparametric identification and whether it holds or not.

Example 1: Unemployment Duration with Heterogeneity. Alvarez, Borovickov´aand Shimer (2016) propose a structural model for transitions in and out of employment that implies a du- ration of unemployment given by the first passage time of a Brownian motion with drift, a random variable with an inverse Gaussian distribution. The parameters of the inverse Gaussian distribution are allowed to vary in arbitrary ways to account for unobserved heterogeneity in workers. These authors investigate nonparametric identification of the distribution of unob- served heterogeneity, which has a density λ0 wrt a σ-finite measure π, when two unemployment spells Z = (t , t ) are observed on the set 2, [0, ). The reduced form parameters i i1 i2 T T ⊆ ∞ (α, β) R [0, ) are functions of structural parameters. The distribution of Z is absolutely ∈ × ∞ i continuous with Lebesgue density fλ0 (t1, t2) given, up to a normalizing constant, by

2 αt −β 2 αt −β 2 β − ( 1 ) − ( 2 ) 2t1 2t2 fλ0 (t1, t2)= 3/2 3/2 e λ0(α, β)dπ(α, β). (1) ZR×[0,∞) t1 t2

Alvarez, Borovickov´aand Shimer (2016) show that λ0 is nonparametrically identified up to the sign of α, but do not investigate semiparametric (regular or irregular) identification, which

is the focus of study here. Specifically, I show that the cdf of λ0 at a point will be irregularly

identified when λ0 is identified. This functional is an important parameter. For example, φ(λ0)= E [1 (α α )1(β β )] , for a fixed α < 0 < β , quantifies the proportion of individuals at risk ≤ 0 ≤ 0 0 0 of severe long term unemployment (an individual with parameters α and β, α α and β β , ≤ 0 ≤ 0 has a probability larger or equal than 1 exp(2α β ) of remaining unemployed forever). This − 0 0 example is studied in detail in Section 6.1. N

Example 2: Willingness to Pay. In contingent valuation studies one observes Zi =(Yi,Vi,Xi), where Y = 1(W V ) , i.e. Y = 1 if W V , and zero otherwise, V is a continuous random i i ≤ i i i ≤ i i variable chosen by the researcher, with known distribution F , and X a d dimensional vector V i − of covariates. Here, Wi is willingness-to-pay of individual i for a new product or resource, which

is an unobserved continuous non-negative random variable. It is assumed that Wi and Vi are

conditional independent given Xi. The density of Zi, with respect to a suitable measure µ, is

f (y,v,x) = [G (v, x)]y [1 G (v, x)]1−y , λ0 0 − 0 where G (v, x) = P [W v X = x] and λ (v, x) = ∂G (v, x) /∂v. Here, one parameter of 0 i ≤ | i 0 0 interest is the median of the distribution of W,

φ(λ0)= Median(W ).

7 Lewbel (1997) and Lewbel, McFadden and Linton (2011) investigate nonparametric and semi-

parametric estimation of moments φ(λ0) = E [r(Wi)] and φ(λ0) = E [r(Wi,X)] , respectively.

Khan and Tamer (2010) show that E [Wi] is irregularly identified when the support of W is unbounded, and discuss rates of convergence for this functional. There is also an extensive ′ literature for the related binary choice model when Wi has the representation Wi = θ0X + εi. In Section 6.2 of this paper I study identification of the median WTP. This example is also useful to illustrate the systematic aspect of the proposed method, i.e. a single approach can be used for different functionals. N

In the following example, I obtain new regular identification results by an application of the characterization of regular semiparametric identification to a conditional moment model.

Example 3: Consumption-based Asset Pricing Models with Measurement Error. Consumption- based asset pricing Euler Equations are important models in economics. When applied to microeconomic data it is vital to account for measurement error in consumption, as in

E θ u˙ (C∗ )R u˙ (C∗) =0, 0 0 t+1 t+1 − 0 t Ft   ∗ where θ0 is the discount factor,u ˙ 0 is the marginal utility of consumption Ct , Rt+1 is the gross return of an asset and t denotes the σ-field generated by the agent’s information set at time t. F ∗ The econometrician observes Ct, which is a noisy measure of Ct , following the specification

∗ Ct = m(Ct ,εt), where m is unknown and εt is the measurement error. The primitives of the model, λ0, are θ , u˙ , m and the distribution of (C∗ , R ,ε ,C∗,ε ) given . Of particular interest are 0 0 t+1 t+1 t+1 t t Ft θ0 andu ˙ 0. The observed data is Zi = (Ct+1,i,Ct,i, Rt+1,i,Xt,i), for a sample of households, and where X is a vector of household characteristics (e.g. family size) in . The results of this t,i Ft,i paper are applied to this example to obtain sufficient conditions for identification of the discount

factor, φ(λ0)= θ0, and the Average Arrow-Pratt coefficient of Absolute Risk Aversion (AARA), φ(λ )= E [( ∂u˙ (C∗)/∂C∗) /u˙ (C∗)]. This example is studied in detail in Section 6.3. N 0 − 0 t t 0 t

4 Linear Nonparametric Models

This section first introduces some notation that will be used throughout the paper. For a

generic measure ν, let Lq(ν), q 1, denote the Banach space of (equivalence classes of) real- ≥ 1/q valued measurable functions h such that h := h q dν < (henceforth I drop the k kq,ν | | ∞ sets of integration in integrals and the qualification ν almost surely for simplicity of notation). R−  So, for example, a function in Lq(ν) is discontinuous when there is no continuous function in

8 its equivalence class. Define the Hilbert space L of P square integrable measurable functions 2 − with inner product h, f = hfdP and norm h 2 = h, h (I drop the dependence on q = 2 h i 0 0 k k h i and ν = P in this case). TheR set Lq(ν) (L2) is the subspace of zero mean functions in Lq(ν) (respectively, L ). Henceforth, for a generic linear operator K : , (K) := f : 2 G1→ G2 N { ∈ G1 Kf =0 denotes its kernel. Define B = b H : λ + b Λ and let T (λ ) denote the linear } 0 { ∈ 0 ∈ } 0 span of elements in B0. The identification results given below will be relative to the tangent set

T (λ0) (i.e, relative to Λ). The goal is to relate identification with the concept of statistical information in a nonpara- metric setting. To that end, let us consider the score operator, see e.g. Begun, Hall, Huang and Wellner (1983), which, for linear models defined as in Assumption 1 below, is the operator S : T (λ ) L given by 0 7→ 2 fλ0+b fλ0 Sb Sλ0 b := − 1(fλ0 > 0). (2) ≡ fλ0

Note that under linearity of S, this operator has a unique extension from B0 to T (λ0), see Deb- nath and Mikusinski (2005, pg. 26), and hence Sb is well-defined when b T (λ ) B . More ∈ 0 \ 0 generally, existence of the score operator is necessary for the classical mean square differentia- bility assumption, which means that for every path λ Λ with t−1(λ λ ) b T (λ ) H t ∈ t − 0 → ∈ 0 ⊂ the following holds, 1/2 1/2 fλt fλ0 1 1/2 − Sbfλ 0 as t 0. (3) t − 2 0 → ↓ 2,µ

The definition of Sb in (3) is the most commonly used in the literature and applies equally

to linear and nonlinear models. I will use this latter definition for nonlinear models, but keep the more natural definition in (2) for linear models. Often the path can be taken of the form

λt = λ0 + tb and Sb = ∂ log fλ0+tb/∂t is simply the score associated to the parametric submodel

fλ0+tb at the “truth” t = 0, where, henceforth, derivatives wrt to t are one-sided and evaluated at zero. 2 This section investigates identification when the density fλ and the functional φ are linear. To simplify the exposition, it is assumed that the functional is a scalar, with the understanding that all the results below have straightforward extensions to multivariate functionals.

Assumption 1: (i) The map φ˙ : T (λ ) H R defined by φ˙(b) = φ(λ + b) φ(λ ) is 0 ⊆ 7→ 0 − 0 linear; (ii) the score operator S : T (λ ) H L in (2) is well-defined and linear; (iii) for each 0 ⊆ 7→ 2 b (S), there exists c c(b) R, c =0, such that λ + cb Λ; (iv) φ˙ and S are continuous. ∈N ≡ ∈ 6 0 ∈ Assumption 1(i) holds for the leading example of the finite-dimensional parameter in a semi- , i.e. φ(θ, η) = θ. Section 9.3 in the Supplemental Appendix relaxes 1(i). Assumption 1(ii) holds in Examples 1 and 2. In other models the linearity assumption holds

2For the sake of exposition, I refer to these as linear, although a more mathematically precise name is affine.

9 after a suitable reparametrization. For example, consider a simple model of a binary outcome Y ∗ that is only observed when D =1. That is, the available data is Z =(Y,D,X), where Y = Y ∗D and Y ∗ is independent of D given X. Define q(x)= E [Y ∗ X] and p(x)= E [D X] . The density | | of (Y,D,X) is a nonlinear function of p and q, since for example P (Y =1,D =1 X = x) = | q(x)p(x), but if we reparametrize the density in terms of λ0 =(λ01,λ02) with λ01(x)= q(x)p(x) and λ02(x) = p(x), then the density becomes linear in λ0. Section 5 below and Section 9.3 in the Supplemental Appendix relax Assumption 1(ii). Assumption 1(iii) is only used to prove the necessity of the main identification condition below, and it can be dropped altogether by restricting attention only to b B , see Remark 2 below, although it facilitates exposition. ∈ 0 Overall, Assumption 1 is convenient because, with this assumption, identification can be fully characterized. Thus, the identification results under Assumption 1 provide a benchmark for what can be achieved in more complicated situations. ∗ Assumption 1(iv) guarantees the existence of the adjoint operator S : L2 T (λ0) of S, ∗ 7→ satisfying g,Sb = S g, b H for all g L and b T (λ ). The following definition extends the h i h i ∈ 2 ∈ 0 Fisher information matrix to a nonparametric context (cf. Koˇsevnik and Levit 1976).

∗ Definition (Fisher Information): The information operator is defined as Iλ0 := S S.

Roughly, I b measures the Fisher information of λ in the direction b T (λ ), i.e. the λ0 0 ∈ 0 classical Fisher information corresponding to fλ0+tb at t =0. To establish a link between Iλ0 and

identification, note that, under Assumption 1, semiparametric identification of φ(λ0) will hold if, for all b T (λ ), ∈ 0 f f f 1(f > 0)Sb =0= φ(λ + b) φ(λ ) φ˙(b)=0, λ0+b − λ0 ≡ λ0 λ0 ⇒ 0 − 0 ≡ or, since f > 0 P a.s. and (S)= (I ), λ0 − N N λ0 (I ) (φ˙). (4) N λ0 ⊂N The following proposition proves that (4), which involves the nonparametric Fisher information,

is necessary and sufficient for semiparametric identification of φ(λ0) under Assumption 1.

Proposition 4.1 Under Assumption 1, identification of φ(λ0) holds iff (4) holds.

Without Assumption 1(ii) both implications of Proposition 4.1 fail, which motivates the initial focus on linear models. That (4) is not sufficient for identification follows from a counterexample given in Chen et al. (2014), while that it is not necessary by Sargan (1983). Some structure is thus needed for the intuitive identifiability condition (4) to be useful for identification. Assump- tion 1 is a natural starting point, because with this assumption identification is characterized.

10 Remark 1: (i) When Λ is a subset of densities wrt a σ-finite measure π, it is natural to define the score operator as an operator from L (π) to L (P) since λ L (π) and Sb L (P) without 1 1 ∈ 1 ∈ 1 additional assumptions. Proposition 4.1 holds with these extended definitions as well. (ii) If a Hilbert space approach is preferred, it is convenient to introduce directions b L0(G ), where ∈ 2 0 G0 is the measure associated to λ0, and re-define the score operator in (2) and the functional φ˙ with b replaced by λ b (and T (λ ) as the linear span of B = b L0(G ): λ + λ b Λ ). The 0 0 0 { ∈ 2 0 0 0 ∈ } specific definition used for the operator S and functional φ˙ will be clear from the context.

Remark 2: Assumption 1(iii) can be dropped if the identification condition (4) is replaced by

(S) B (φ˙) B . (5) N ∩ 0 ⊂N ∩ 0

The identification condition (4) is based on the nonparametric Fisher information and, as such, it is not useful if the goal is to disentangle regular and irregular semiparametric identification. To introduce a more useful characterization, I first define the semiparametric Fisher information for

φ. The information for estimating the parameter ψ(t)= φ(λt),λt := λ0+tb, under the density fλt at t = 0 is, by the delta method, equal to [∂ψ(t)/∂t]−1 Sb 2 [∂ψ(t)/∂t]−1 = Sb 2/[φ˙(b)]2. The || || || || semiparametric Fisher information is the infimum of the informations over all such parametric submodels (cf. Stein 1956) and is given by Sb 2 Iφ = inf || || 2 , (6) b∈Bφ φ˙(b) h i where := b T (λ ): φ˙(b) =0, φ˙(b) 1 . Bφ ∈ 0 6 ≤ n o ˙ By the continuity in Assumption 1(iv), S and φ are uniquely extended to T (λ0), where henceforth, for a subspace V, V denotes the closure of V in the norm topology. Moreover, there exists an r T (λ ), called the Riesz’s representer of φ,˙ such that for all b T (λ ), φ ∈ 0 ∈ 0

φ˙(b)= b, r H. h φi

I can then identify φ˙ with rφ, and provide identification results in terms of rφ using duality. I will

make clear in the text when S is viewed as acting on T (λ0) or on its extension T (λ0). Henceforth, whenever identification is discussed in terms of rφ, like in the next theorem, I mean S and φ˙ to be defined on T (λ0) (so, for example, rφ is guaranteed to exist by the Riesz representation theorem). Let (S∗) := f T (λ ): g L ,S∗g = f . Then, the following result provides a R { ∈ 0 ∃ ∈ 2 } full characterization of semiparametric regular and irregular identification in linear models.

Theorem 4.1 Under Assumption 1: (i) φ(λ ) is regularly identified iff r (S∗); (ii) φ(λ ) is 0 φ ∈ R 0 irregularly identified iff r (S∗) (S∗); and (iii) φ(λ ) is unidentified iff r / (S∗). φ ∈ R \ R 0 φ ∈ R 11 The results that seem to be novel here are the “identification” parts and the separation of irregular identification from no identification. The fact that r (S∗) is equivalent to I > 0 φ ∈ R φ is due to Van der Vaart (1991, Theorem 4.1). As discussed earlier, the characterizations of Theorem 4.1 do not necessarily hold without Assumption 1, and some care must be exercised to extrapolate these results to nonlinear models. Section 9.3.1 in the Supplemental Appendix presents a nonlinear counterexample, building on that given in Chen et al. (2014), that shows

that Iφ > 0 may hold while φ(λ0) is not identified.

Remark 3: Assumption 1(iii) can be dropped for part (i) of Theorem 4.1 because a positive Fisher information implies r (S∗). φ ∈ R

4.1 Irregular Identification and Generalized Information

When the Fisher information Iφ is zero, it does not provide information on identification (it can- not distinguish between irregular identification and no identification). This section introduces a “generalized Fisher information” that extends, in a sense described later, the classical Fisher information to irregular cases, and which is given by

Sb 2 Iφ,ρ = inf || ||2ρ , (7) b∈Bφ φ˙(b)

where 1 ρ < . The classical Fisher information corresponds to ρ = 1, i.e. I I . ≤ ∞ φ ≡ φ,1 Furthermore, it can be shown that I I for 1 <ρ< . φ,1 ≤ φ,ρ ∞ The sense in which the generalized Fisher information provides a generalization of the clas- sical Fisher information is shown in the next two results. The first result extends the sufficient condition for identification in Theorem 4.1(i) to the irregular case. Under Assumption 1, Iφ,1 =0 and I > 0 for some ρ> 1 corresponds to irregular identification. Because φ˙(b) I−2ρ Sb 1/ρ, φ,ρ ≤ φ,ρ || || for b , an interpretation of a positive generalized information is that φ˙(b) is continuous in ∈ Bφ wrt the Fisher semi-norm Sb , with a modulus of continuity quantified by 1/ρ (smaller Bφ || || ρ corresponding to more regularity). The inequality above directly gives identification on a restricted set, and the next result proves that this set can be extended to the whole parameter space Λ under Assumption 1.

Theorem 4.2 Let Assumption 1(i-ii) hold. If I > 0, 1 ρ< , then φ(λ ) is identified. φ,ρ ≤ ∞ 0 As discussed at the end of this section, there is an extensive literature on nonparametric identi-

fication that requires, implicitly or explicitly, the condition Iφ,ρ > 0 to establish rates of estima-

tion. Theorem 4.2 shows Iφ,ρ > 0 is sufficient for identification, while the next result shows it is necessary for achieving certain rates of convergence. Let denote a class of sequences in Λ, and A

12 let Pn denote the n fold probability of P . I provide a formal definition of rate of convergence λ − λ (see e.g. Ishwaran 1996, Definition 7).

Definition (Rate of Convergence): The estimator T has a rate of convergence r on for n n A estimating φ(λ ) if for each ε> 0 there exists K(ε) > 0 such that for each λ , 0 { n}∈A n lim sup Pλn ( Tn φ(λn) >K(ε)rn) < ε. n→∞ | − | Theorem 4.3 Suppose that for each ε > 0 there exists a path λ Λ passing through λ such t ∈ 0 that for all t sufficiently small the following holds: (i) Ct φ(λt) φ(λ0) 1, for some C > 0, 2 2ρ ≤| − | ≤ and (ii) (fλt fλ0 ) /fλ0 < εt , for 1 ρ < . Suppose contains all sequences λn for k − k −1/2ρ ≤ ∞ A { } which φ(λn) = φ(λ0)+ O(n ). Then, the rate of convergence for any estimator of φ(λ0) on must be slower than O(n−1/2ρ). A

Conditions (i) and (ii) in Theorem 4.3 correspond to Iφ,ρ = 0, since under these conditions

bt (λt λ0) φ and ≡ − ∈ B Sb 2 ε Iφ,ρ = inf || || 2ρ , b∈Bφ [φ˙(b)]2ρ ≤ C and because ε > 0 is arbitrary, it must hold that Iφ,ρ = 0. I stress that Assumption 1 is not required here, so Theorem 4.3 holds for linear and nonlinear models/parameters. Theorem 4.3 extends the impossibility result of Chamberlain (1986) to the irregular case ρ > 1. Its proof relies on techniques, originally due to Lecam (1973), that bound the total variation distance between the distribution under fλn and that under fλ0 , for a suitable sequence λn. Previously, Donoho and Liu (1987) have used similar techniques based on Hellinger distance to provide lower bounds for convergence of estimators. The modest contribution of Theorem 4.3 is to connect the zero generalized Fisher information introduced in this paper with these lower bounds, extending important work by Chamberlain (1986) to the irregular case.

It is important to stress that there is nothing that prevents the possibility that Iφ,ρ = 0 for all ρ 1, which, under Theorem 4.3, implies impossibility of polynomial rates. Indeed, in several ≥ important models logarithmic rates are common, see, e.g., Fan (1991) for classical measurement error problems. This possibility suggests that the definition of the generalized information and the conditions of Theorem 4.3 should be modified to accommodate severely irregular cases. For a function ψ that is increasing, non-negative, right continuous at 0 and with ψ(0) 0, one can ↓ define the generalized Fisher information Sb 2 Iφ,ψ = inf || || . b∈Bφ ψ [φ˙(b)]2   With this modification different degrees of irregularity, including severe irregularity, are allowed. The case ψ (ǫ) = ǫρ, 1 < ρ 2, corresponds to mild or moderate irregularity, while ψ (ǫ) = ≤ 13 exp(ǫ) 1 or ψ (ǫ) = exp( 1/(ǫa)), with a> 0, is suitable for severe irregularity with possibility − − of logarithmic rates. A version of Theorem 4.3 that allows for severe irregularity follows mutatis mutandis, simply replacing t2ρ by ψ (t2) . At this point, it is useful to compare the results of this paper with the general nonpara- metric local identification results in Chen et al. (2014) for conditional moment restrictions models. These authors obtain sufficient conditions for nonparametric identification of linear and nonlinear conditional moments by suitably restricting the parameter space. When conditional moments are only Frechet differentiable, they consider the parameter space to have tangents in 2 2 b : mb˙ > C b H , for the derivativem ˙ of a conditional mean operator m and a positive { || || k k } constant C. In our setting, an analog that allows comparison with statistical information would 3 be m(λ)=(fλ fλ0 ) /fλ0 , with derivativem ˙ (λ) = S(λ λ0). On the parameter space with − 2 2 − tangents b : Sb >C b H the nonparametric information is positive (i.e. regular nonpara- { || || k k } metric identification), which implies a positive semiparametric Fisher information for all linear continuous functionals. Chen et al. (2014) also consider conditions corresponding to higher order differentiability and these conditions do allow for irregular semiparametric identification. In their general case, they restrict tangents to the set b : Sb 2 > C b 2ρ , for ρ > 1, which { || || k kH } implies a positive generalized Fisher information Iφ,ρ for all continuous linear functionals φ. To 2 2ρ prove such result, use φ˙(b) r b H and their assumption b : Sb >C b H to ≤k φkH k k Bφ ⊂{ || || k k } bound 2 Sb C Iφ,ρ = inf || || 2ρ > 0. b∈Bφ [φ˙(b)]2ρ ≥ r k φkH This shows that the restrictions on neighborhoods in Chen et al. (2014) have a statistical interpretation in terms of the generalized Fisher information introduced here (for all continuous linear functionals). The parameter ρ is also linked to the nonlinearity permitted in the model (see Assumption 2 in Chen et al. 2014 or Assumption 2 below), which typically restricts its values to 1 ρ 2. The case ρ > 2 only holds when the second derivative of m(λ) is zero. ≤ ≤ From this discussion, it seems that it may be hard to accomodate nonlinear cases for severely ill-posed problems where I = 0 for all ρ 1 and some functional φ. Examples of such severe φ,ρ ≥ irregularity for certain functionals φ include nonparametric IV models or measurement error models with Gaussian errors. To give sufficient conditions for irregular identification that allow for a variety of degrees of irregularity for different functionals consider adding the following condition to Assumption 1:

Assumption 1: (v) the score operator S is compact.

3 More precisely, the effective score operator in their conditional moment restriction model is S(λ λ0) = −1 − Σ m˙ (λ λ0) for a conditional variance Σ. Under the standard assumption that the eigenvalues of Σ are bounded − away from zero, there is a one to one relation between identification in conditional moment models as in Chen et al. (2014) and identification in terms of statistical information as considered here; see e.g. Chamberlain (1992).

14 Assumption 1(v) guarantees the existence of a sequence λ ,ϕ , ψ ∞ such that (cf. Kress, { j j j}j=1 1999, Theorem 15.16) ∗ Sϕj = λjψj and S ψj = λjϕj. (8) This is the so called Singular Value Decomposition (SVD) of S. The elements ϕ ∞ and { j}j=1 ψ ∞ are complete orthonormal bases for (S∗) and (S), respectively, and the singular { j}j=1 R R values λ are the squared-root eigenvalues of the information operator I = S∗S : T (λ ) j λ0 0 7→ T (λ ). Furthermore, defining for β R, 0 ∈ ∞ := b T (λ ) such that b 2 := λ−2β b, ϕ 2 < , Mβ ∈ 0 k kβ j h jiH ∞ ( j=1 ) X it is well known (see e.g. Carrasco, Florens and Renault 2007) that

∞ (S∗) = b T (λ ) such that b, ϕ 2 < , R ≡M0 ∈ 0 h jiH ∞ ( j=1 ) X whereas ∞ (S∗) = b T (λ ) such that λ−2 b, ϕ 2 < . R ≡M1 ∈ 0 j h jiH ∞ ( j=1 ) X The following result gives sufficient, and in some cases necessary, conditions for regular and irregular identification in terms of r . k φkβ Theorem 4.4 Let Assumption 1 hold. Then (i) φ(λ ) is regularly identified iff r < , and 0 k φk1 ∞ in that case I = r −2; (ii) If r = but r < for some 0 <β< 1, then φ,1 k φk1 k φk1 ∞ k φkβ ∞ Sb 2 1 inf || || > 0, 2ρ 2ρ b∈Bφ,||b||H≤1 [φ˙(b)] ≥ r k φkβ

for ρ =1/β, so φ(λ0) is locally irregularly identified.

In part (i) it is possible to give another expression for the efficiency bound as a variance. The condition r < implies there exists i L such that S∗i = r and I−1 = E[i2 (Z)]. The k φk1 ∞ φ ∈ 2 φ φ φ,1 φ efficiency bound is then E[i2 (Z)]. It is known that in many cases bounds on r correspond to φ k φkβ imposing smoothing conditions on rφ (see Kress 1999, Chapter 8). Hence, in these cases one can

index the level of irregularity by the level of smoothness of the influence function rφ. Related conditions have been extensively used in the literature of ill-posed inverse problems in and econometrics as “source conditions”, see e.g. Carrasco, Florens and Renault (2007) and Chen and Reiss (2011). The novelty of Theorem 4.4 is in the relation between “smoothness” of

rφ and identification and information.

15 5 Semiparametric Models

This section studies the important class of semiparametric models, where = P : θ Θ, η P { θ,η ∈ ∈ H . The parameter space Λ = (θ, η): θ Θ, η H is a subset of a Hilbert space H = Rp . } {′ ∈ ∈ } ×H Define (θ , η ), (θ , η ) H := θ θ + η , η . For semiparametric models the score operator, h 1 1 2 2 i 1 2 h 1 2iH defined as in (3), has the representation (by the chain rule)

S (b , b )= l˙′ b + l˙ b , b =(b , b ) T (λ ) H, (9) θ η θ θ η η θ η ∈ 0 ⊆ ˙ p ˙ where lθ L2 is the ordinary score function of θ and lη is a continuous linear operator from ∈ ˜ ˙ ˙ T (η0) to L2. Let lθ := lθ Π ˙ lθ be the so-called efficient score function for θ, where ΠV ⊂ H − R(lη ) denotes the orthogonal projection operator onto V. The efficient Fisher information matrix for ˜ ˜ ˜′ θ is Iθ := E lθlθ . The following result provides a characterization of the main identification condition forh thei finite-dimensional parameter of a semiparametric model. For simplicity, I consider the case p =1, the extension to p> 1 follows from applying the result to the functionals φ(λ)= α′θ for α Rp. Section 9.2 in the Supplemental Appendix provides a parallel result for ∈ linear continuous functionals of the nuisance parameter, allowing for θ to be infinite-dimensional and possibly unidentified. Next proposition appears to be a new characterization of the main condition for local identification in semiparametric models.

Proposition 5.1 For the functional φ(λ) = θ R: (I ) (φ˙) holds iff (i) l˙ / (l˙ ) ∈ N λ0 ⊂ N θ ∈ R η (positive information I˜ > 0) or (ii) l˙ (l˙ ) (l˙ ) (zero information I˜ =0). θ θ ∈ R η R η θ In the remainder of this section I extend some of the previous results to a class of models that are nonlinear in the parameter of interest but linear in nuisance parameters.

Assumption 2: For some ρ 1 and for all ε > 0, there exists δ > 0 and a continuous linear ≥ operator S such that, (f f ) /f S(λ λ ) <ε θ θ ρ , for all λ =(θ, η) (λ ). k λ − λ0 λ0 − − 0 k | − 0| ∈ Bδ 0 Assumption 2 is a mean-square differentiability condition with a Lipschitz property on the derivative. It generally holds for models that are nonlinear and smooth in the parameter of interest θ, but linear in the nuisance parameters. Examples of models satisfying Assumption 2 include, among others, structural models of unemployment duration in Heckman and Singer (1984a, 1984b); linear and nonlinear panel data models with fixed effects (see e.g. Bonhomme 2012); incomplete and complete games with multiple equilibria (see e.g. Bajari, Hahn, Hong and Ridder 2011); semiparametric measurement error models (see e.g. Hu and Schennach 2008); dynamic discrete choice models (see e.g. Hu and Shum 2012); and binary discrete choice models with single and multiple agents (see e.g. Chamberlain 1986, and more recently, Khan and

Nekipelov 2018). Importantly, Assumption 2 allows the nonparametric parameter λ0 to be

16 unidentified and the parameter θ0 to be locally irregularly identified, as it occurs in many of the aforementioned applications. The latter feature differentiates our analysis from Chen et al.’s (2014) setting. In most cases 1 ρ 2 in Assumption 2, which will limit the degree of ≤ ≤ irregularity permitted in identifying θ0, but functionals of the nuisance parameter are allowed to have arbitrary degrees of irregularity, which can be important to accommodate many economic applications with smooth densities (e.g. Heckman and Singer 1984a, 1984b). The generalized Fisher information for θ is

Sb 2 Iθ,ρ = inf || || 2ρ , b∈Bθ θ θ | − 0| where := b T (λ ): b = (θ θ , b ), θ = θ , θ θ 1 . It is straightforward to show Bθ { ∈ 0 − 0 η 6 0 | − 0| ≤ } that I˜θ = Iθ,1. Next theorem extends Theorem 4.2 to the nonlinear setting of Assumption 2.

Theorem 5.1 Let Assumption 2 hold. If I > 0 for some ρ, 1 ρ < , then θ is locally φ,ρ ≤ ∞ 0 identified: regularly if ρ =1 and irregularly if ρ> 1 and I˜θ =0.

Theorem 5.1 extends Theorem 7 in Chen et al. (2014) to the semiparametric irregular case ρ> 1.

If the generalized information Iθ,ρ is zero, Theorem 4.3 implies impossibility results on rates of convergence. Assumption 2 facilitates the verification of the conditions for Theorem 4.3 to hold. To see this, consider a path λ Λ passing through λ such that S (λ λ ) = o(tρ) and t ∈ 0 k t − 0 k θ θ = Ct. For such a path, Assumption 2 yields the conditions of Theorem 4.3. Nevertheless, | t − 0| Assumption 2 is not necessary for Theorem 4.3 to hold.

6 Examples

6.1 Unemployment Duration with Nonparametric Heterogeneity

Nonparametric heterogeneity has played a critical role in rationalizing unemployment duration ever since the seminal contributions by Heckman and Singer (1984a, 1984b). Recent work by Alvarez et al. (2016) is motivated from this perspective. These authors have shown nonpara-

metric identification of the distribution of unobserved heterogeneity, denoted by G0, in their nonparametric structural model for unemployment with two-spells under a sign restriction on the parameter α, either α 0 or α 0. As discussed by these authors, assuming either case ≥ ≤ imposes unattractive restrictions on the economic model. For example, α 0 implies that all ≥ workers return to work eventually. With this background in mind, I characterize the nonpara- metric identified set without prior restrictions. The characterization of the identified set can be used to investigate alternative, more attractive, conditions for nonparametric identification or to engage in sensitivity identification analysis or partial identification bounds.

17 Let Λ denote the parameter space for λ, which is a subset of densities wrt π. By Remark 1(i), it is natural to consider the score operator defined on T (λ ) L (π), where T (λ ) is the 0 ⊂ 1 0 linear span of elements in B = b L0(π): λ + b Λ . The score operator is, up to an 0 { ∈ 1 0 ∈ } irrelevant constant, given by

2 αt −β 2 αt −β 2 1 β − ( 1 ) − ( 2 ) Sb = e 2t1 2t2 b(α, β)dπ(α, β). f (t , t ) 3/2 3/2 λ0 1 2 Z t1 t2 The following proposition, which builds on the nonparametric identification results with sign restrictions in Alvarez et al. (2016), characterizes (S) under the following mild assumption. N Assumption 3: (i) The set [0, ) be a convex set with a non-empty interior; (ii) the T ⊆ ∞ measure π is such that dπ( α, β)= dπ(α, β) and has no atom at α = 0. − −

Proposition 6.1 Under Assumption 3, (S)= b T (λ ) L0(π): b(α, β)= e−4αβb( α, β) . N ∈ 0 ⊂ 1 −  A corollary of this proposition is that the identified set for λ is the set λ + b, where b 0 { 0 ∈ (S) B . An equivalent characterization of (S) is N ∩ 0} N C(α, β) (S)= b T (λ ): b(α, β)= , where C(α, β) is an odd function of α . N ∈ 0 1 e4αβ  −  Proposition 4.1 can then be used to check if a given linear functional is identified or not. The characterization of (S) also can be used to find new point-identification results. The following N corollary, which follows directly from Proposition 6.1 illustrates this point.

Corollary 6.1 Under Assumption 3, if Λ is a set of symmetric densities in α, i.e. for all λ Λ, ∈ λ(α, β)= λ( α, β) π a.s, then λ is nonparametrically identified. − − 0 The analysis of (S) does not reveal, however, the degree of identification for a given N functional, whether regular or irregular. To understand this, one must analyze the adjoint score operator. For this, it is convenient to consider the approach of Remark 1(ii), which, as shown below, leads to the score adjoint operator

S∗g = E [g(Z) α, β] , g L0. (10) | ∈ 2 0 Recall that under Remark 1(ii) T (λ0) is the linear span of B0 = b L2(G0): λ0 + λ0b Λ , {0 ∈ 0 ∈ } and one say heterogeneity is nonparametric if T (λ0) is dense in L2(G0) (i.e. T (λ0) = L2(G0)). The following result provides a necessary condition for regular identification of continuous linear functionals in this example.

18 Proposition 6.2 Let Assumption 3(i) hold. If heterogeneity is nonparametric, then S∗ has the representation in (10) and

(S∗) b(α, β) L0(G ): b(α, β)= C + C β2e2αβh(α2, β2) , R ⊂ ∈ 2 0 1 2  for constants C and C and a continuous function h(u, v) defined on (0, )2 that, if is 1 2 ∞ T bounded, is an infinite number of times differentiable at u (0, ), for all v (0, ). ∈ ∞ ∈ ∞

This result has important implications on which functionals are regularly identified in this model. Take, for example, a moment of unobserved heterogeneity

φ(λ0)= E [r(α, β)] ,

where r L (G ). From the reparametrization of Remark 1(ii), with H = L (G ), ∈ 2 0 2 0

φ˙(b)= r(α, β)b(α, β)λ (α, β)dπ(α, β)= r, b H, 0 h i Z and hence r = r φ(λ ) L0(G ). Thus, only smooth moment functions r satisfying the φ − 0 ∈ 2 0 conclusions of Proposition 4.1 will be regularly identified. Alvarez et al.’s (2016) nonparametric identification result must imply (S)= 0 . In turn, this result and Theorem 3 in Luenberger N { } (1997, p.157) yield that (S∗)= L0(G ). Thus, the set (S∗) (S∗)= L0(G ) (S∗) is rather R 2 0 R \R 2 0 \R large by Proposition 6.2, which shows that the class of irregularly identified functionals is large

in this model. Intuitively, this follows because the density fz/α,β(t1, t2) is very smooth in the parameters (α, β), so that (S∗) only contains very smooth functions. A direct implication of R Proposition 6.2 is that the cumulative distribution function (cdf) of unobserved heterogeneity at the fixed point (α , β ), i.e. φ(λ )= E [1(α α )1(β β )] , is not regularly identified because 0 0 0 ≤ 0 ≤ 0 1(α α )1(β β ) is not continuous when (α , β ) is in the interior of the support of λ . ≤ 0 ≤ 0 0 0 0 Recall, φ(λ )= E [1 (α α )1(β β )] for a fixed α < 0 < β is the proportion of individuals 0 ≤ 0 ≤ 0 0 0 at risk of severe long term unemployment, and this proportion is thus non-regularly identified, as an implication of Proposition 6.2. We formalize the discussion in the following corollary.

Corollary 6.2 Let Assumption 3(i) hold. If heterogeneity is nonparametric and (α0, β0) is in the interior of the support of λ , then φ(λ )= E [1 (α α )1(β β )] is not regularly identified. 0 0 ≤ 0 ≤ 0

6.2 Willingness-to-Pay

The observed data is Z =(Y ,V ,X ), where Y =1(W V ) , W is the unobserved willingness- i i i i i i ≤ i i to-pay of individual i, Vi is a continuous observed random variable with known and absolutely continuous conditional distribution F , and X a d dimensional vector of covariates. The V/X i −

19 support of W is := [0,w ] and that of V is := [0, v ], with 0 < w , v SW max SV max max max ≤ . It is assumed that W and V are conditional independent given X . Assume G (v, x) = ∞ i i i 0 P [W v X = x] is differentiable at v for each x, and define λ (v, x) = ∂G (v, x) /∂v. The i ≤ | i 0 0 conditional density λ0 belongs to the parameter space Λ, a subset of Lebesgue densities on

W X . Define the measure µ ( 0 B)= µ ( 1 B)= µVX (B), where B is a Borel set of S d+1× S { } × { } × R and µVX (B) is the probability measure for (V,X). The density of Zi wrt µ is

f (y,v,x) = [G (v, x)]y [1 G (v, x)]1−y . (11) λ0 0 − 0

Suppose one is interested in identification of the moment functional φ(λ0)= E [r(Wi,X)] , as in Lewbel, McFadden and Linton (2011). I consider first this case, and then analyze identification of the median WTP,

φ(λ )= Median(W ) = inf w : G (w) 1/2 , 0 { ∈ SW 0 ≥ }

where G0 (w)= P [Wi w] is the cdf of Wi. ≤ 0 Following Remark 1(ii), the score operator is defined on L2(π), where π denotes the proba-

bility measure for (W, X) with conditional cdf G0 (v, x) . From (11) the score operator is given for y 0, 1 , v and x by ∈{ } ∈ SV ∈ SX 1 v S(b)= 1(f > 0)[2y 1] b(w, x)λ (w, x) dw. (12) f (z) λ0 − 0 λ0 Z0 By the Fundamental Theorem of Calculus (S)= b T (λ ) L0(π): b(w, x)=0 π a.s. on N { ∈ 0 ⊂ 2 − 0 w v . If v w then (S) = 0 and all linear functionals of λ are identified ≤ ≤ max} max ≥ max N { } 0 by Proposition 4.1.

If vmax < wmax identification is more complicated and depends on the functional. For the

functional φ(λ0)= E [r(Wi,X)] ,

φ˙(b) := φ(λ + bλ ) φ(λ ) 0 0 − 0

= r(w, x)b(w, x)λ0 (w, x) dµX dw ZSW ×SX = r, b H, h i

where H = L2(π) and µX is the measure of X. The unique Riesz’s representer of φ˙ in T (λ0) is given by r = Π r. φ T (λ0) A sufficient condition for identification (cf. Proposition 4.1) is

wmax rφ(w, x)b(w, x)λ0 (w, x) dµX dw =0 Zvmax ZSX for all b T (λ ) L0(π). Taking b(w, x)= r (w, x) leads to the following result: ∈ 0 ⊂ 2 φ

20 Proposition 6.3 If vmax

0 Consider henceforth the general case of a nonparametric model for WTP, so T (λ0)= L2(π). In this case r = r φ(λ ). φ − 0 To investigate the degree of identification, whether irregular or regular, let us compute the adjoint score operator. First, note by conditional independence 1 S(b)= 1(fλ0 > 0)[2y 1] E [1(W V )b(W, X) V = v,X = x] . fλ0 (z) − ≤ | Considering separately the case y = 1 and y =0, we see that

S(b)= E [b(W, X) Y = y,V = v,X = x]1(f > 0). | λ0 It is then well-known that the adjoint operator of a conditional mean operator is

S∗g(w, x)= E [g(Y,V,X) W = w,X = x] , | which after some simple algebra gives

vmax S∗g(w, x)= 1(w v)g(1,v,x)f (v)dv + 1(w > v)g(0,v,x)f (v)dv, ≤ V/X=x V/X=x Z0 Z where fV/X=x(v) is the known conditional Lebesgue density of V given X = x. A direct conse- quence of this representation of S∗ and the Fundamental Theorem of Calculus is that S∗g(w, x) is absolutely continuous in w, for each x. This result can be used to characterize when mo- ment functionals such as φ(λ0)= E [r(Wi,X)] are regularly identified. The regularity condition r = S∗g, for some g L , implies by the Fundamental Theorem of Calculus ∈ 2 ∂r(w, x) =(g(0,w,x) g(1,w,x)) f (w), ∂w − V/X=x which has a solution ∂r(v, x) 1 g(y,v,x) = [1 2y] 1(fV/X=x(v) > 0), (13) − ∂v fV/X=x(v) in L2, provided vmax ∂r(v, x) 2 1 dµ (x)dv < . (14) ∂v f (v) X ∞ Z0 Z   V/X=x Next result summarizes the findings on moments.

21 Proposition 6.4 Assume T (λ )= L0(π) and v w . Let r L (π) be a moment function 0 2 max ≥ max ∈ 2 such that r(w, x) is differentiable at w, for each x. Then, (14) is necessary and sufficient for

regular identification of φ(λ0)= E [r(Wi,X)] . This result has interesting implications for known results in the literature. First, the regular estimator that results from the moment representation based on the solution (13)

φ(λ0)= E [g(Y,V,X)] ∂r(V,X) 1 = E [1 2Y ] − ∂v f (V )  V/X  is related to (but different from) the estimator proposed in Lewbel (1997) without covariates, which is given by ∂r(V ) 1 E [1(V 0) Y ] . ≥ − ∂v f (V )  V  The arguments above then show that the sufficient finite variance condition derived in Lewbel (1997) for asymptotic normality of his estimators turns out to be also necessary. When applied to the mean of W, i.e. r(w, x)= w, which is a leading example in Khan and Tamer (2003), the necessary and sufficient condition for regular identification becomes simply vmax 1 dµ (x)dv < . (15) f (v) X ∞ Z0 Z V/X=x If supports of W and V are unbounded, i.e. w = v = , so f (u) vanishes in max max ∞ V/X=x the tails, the last condition does not hold, which gives Khan and Tamer’s (2010) irregularity result using a different method of proof (they compute least favorable distributions and Fisher information). Note that irregularity can also happen with bounded supports, depending on the density fV/X=x(v), and that the condition is necessary (results which were not considered in Khan and Tamer (2003) for the mean r(w, x)= w). I now turn into the more difficult problem of semiparametric identification of the median

WTP, φ(λ0)= Median(W ). This functional is nonlinear, and hence the results of Theorem 4.1 are not readily applicable. For local identification, we require the following condition:

Assumption 4: (i) v >φ(λ ) (ii) G (v) is continuous on and differentiable in a neigh- max 0 0 SV borhood of v = φ(λ0) with a positive derivative E [λ0 (v,Xi)] .

Assumption 4(i) is a support condition that is weaker than that for identification of moments of WTP. Assumption 4(ii) is typical in quantile identification. Under Assumption 4, the median WTP has an influence function

1(w<φ(λ0)) 0.5 rφ(w, x)= −{ − }. E [λ0 (φ(λ0),Xi)] As the result above suggests, the discontinuity of the influence function implies irregular iden- tification. Next result, formalizes this finding.

22 0 Proposition 6.5 Let Assumption 4 hold, and assume T (λ0)= L2(π). Then, φ(λ0)= Median(W ) is irregularly identified.

6.3 Asset Pricing Euler Equation with Measurement Error

The goal of this example is to provide primitive conditions based on the results of this paper for identification of the discount factor θ0 and measures of risk aversion. These are important parameters in these models. For example, discount factors are a key determinant of individual’s intertemporal decisions such as asset accumulation (Venti and Wise, 1998, Samwick, 2006), labor supply decisions (MaCurdy, 1981) and job search (Dellavigna and Paserman, 2005).

6.3.1 Identification of the Discount Factor

The Euler equation with measurement error is a nonlinear conditional moment restriction model. The first step in our analysis is to parametrize the model in a way that makes it amenable to the ∗ results of this paper. To that end, I consider the following assumption. Recall Ct = m(Ct ,εt).

∗ Assumption 5: (i) Ct+1 is independent of Ct, conditional on Ct+1, and measurable wrt

σ(Ct+1,εt+1); (ii) (εt+1,εt) is independent of Rt+1, given (Ct+1,Ct); (iii) the distribution of ∗ Ct conditional on Ct does not depend on t.

∗ Assumption 5(i) can be relaxed to: Ct+1 is independent of Ct−1, conditional on (Ct+1,Ct), at the cost of increasing the dimension of the arguments in the nonparametric component given ∗ below. If m is monotone in Ct , then 5(i) can be written in terms of εt+1. Assumption 5(i-iii) is less restrictive than typical assumptions considered in the literature, which assume, in addition

to functional form assumptions on m, that εt is independent of “everything”; see, for example, Altonji and Siow (1987), Runkle (1991), Dynam (2000), and Alan et al. (2009).

For the sake of exposition, I consider the case without household’s characteristics Xt,i. The presence of X in adds additional moment restrictions, so it is simpler for identification. t,i Ft,i All the arguments below can be easily adapted to the presence of Xt,i. It is also straightforward to extend the identification results to models with more than one asset, habit formation or other observable variables in the marginal utility.

Assumption 5 ensures the following parametrization in terms of observables (Ct+1, Rt+1,Ct),

E θ u˙(C∗ )R u˙(C∗) C = E θ u˙(C∗ )E [R C ,C ,ε ,ε ] u˙(C∗) C 0 t+1 t+1 − t t 0 t+1 t+1| t+1 t t+1 t − t t ∗ ∗   = E θ0u˙(Ct+1)E [Rt+1 Ct+1,Ct] u˙(Ct ) Ct ,  | − ∗ ∗ = E θ0E u˙(Ct+1) Ct+1 E [Rt+1 Ct+1,C t] u˙(Ct ) Ct | − = E [θ0η0(Ct+1)Rt+1 η0(Ct) Ct] ,  − |

23 ∗ where η0(Ct+1) = E u˙(Ct+1) Ct+1 and the first equality uses 4(i), the second 4(ii), the third 4(i) and the fourth 4(iii). This new parametrization is a nonlinear conditional moment indexed by the discount factor θ0 and the projected marginal utility η0. Let µ be the probability measure of C . Then, I denote λ =(θ , η ) Λ=Θ H, Θ (0, 1) and H L (µ). t 0 0 0 ∈ × ⊂ ⊂ 2 The following condition guarantees that the conditional mean operator

Aη(c)= E [η(C )R C = c] , (16) t+1 t+1| t is well-defined and compact when viewed as A : L (µ) L (µ). This is a standard assumption 2 → 2 in the literature, see e.g. Carrasco et al. (2007). Let g(Ct+1,Ct) be the joint Lebesgue density

of (Ct+1,Ct), and let ft+1(Ct+1) and ft(Ct) denote its marginals, respectively.

Assumption 5: (iv) 0 < E R2 g(C ,C )/f (C )f (C ) < . t+1 { t+1 t t t+1 t t } ∞   Define, for λ =(θ, η) Λ=Θ H, ∈ × M = θAη(c) η(c). λ − The identification results of Section 5 can be adapted to this moment model replacing (f f ) /f λ − λ0 λ0 by M M and the norm by . I then proceed to verify Assumption 2 in this ex- λ0+b − λ0 k·k k·k2,µ ample. It is straightforward to show that for all b =(b , b ) R L (µ), with b = θ θ θ η ∈ × 2 θ − 0 M M S(b) θ Ab (c) b (c) b , (17) k λ0+b − λ0 − k2,µ ≤k 0 η − η k2,µ | θ| where

S (bθ, bη)= l˙θbθ + l˙ηbη, l˙ b = b Aη and l˙ b = θ Ab (c) b (c). θ θ θ 0 η η 0 η − η Since A is bounded, for each ε> 0, one can make θ Ab (c) b (c) <ε by choosing δ small k 0 η − η k2,µ enough and b < δ. Thus, Assumption 2 holds with ρ =1. k ηk2,µ It follows from the previous parametrization and Assumption 5 that local identification of the discount factor is regular.4 Formally, Theorem 3.2 in Kress (1999, p. 29) implies that the range of l˙ , (l˙ ), is closed. It follows from Rudin (1973, 4.14) that (l˙∗) is also closed, and η R η R η by the expression above, (S∗) is also closed. Then, by the results of this paper, all locally R identified linear continuous functionals of λ0 =(θ0, η0) are regularly locally identified. Note that

this does not apply to functionals of the marginal utilityu ˙ 0.

4The actual effective score operator of the model is proportional to S, with a proportionality “constant” 2 given by the inverse of the conditional variance E Vt Ct = c , where Vt+1 = θ0η0(Ct+1)Rt+1 η0(Ct). This +1 − conditional variance is assumed to be bounded and bounded  away from zero. Thus, statements related to regularity can be given in terms of S.

24 Positive information for the discount factor θ holds iff (1, 0) (S∗), which means there 0 ∈ R exists g L (µ) such that ∈ 2 Aη ,g =1, l˙∗g =0. (18) h 0 i η ˙∗ ∗ ˙∗ ∗ Since lηg = θ0A g(c) g(c), the equation lηg = 0 means that g is an eigenfunction of A with −1 − ∗ eigenvalue θ0 . Such eigenfunction exists because eigenvalues of A are complex conjugates of those of A and θ0 is real-valued. Then, a sufficient condition for local identification of the discount factor is that for such eigenfunction, say g : η ,g = 0. Note that η ,g = 0 and 0 h 0 0i 6 h 0 0i 6 θ > 0 guarantee (18) by choosing g = cg with c =1/ Aη ,g (since multiples of eigenfunctions 0 0 h 0 0i are eigenfunctions). The discussion is summarized in the following result.

Proposition 6.6 Let Assumption 5 hold and assume η0,g0 = 0 for g0 one of the eigenfunc- ∗ −1 h i 6 tions of A corresponding to the eigenvalue θ0 . Then, θ0 is locally regularly identified. The condition η ,g = 0 is mild, and holds, for example, when η and g are positive. h 0 0i 6 0 0 Escanciano and Hoderlein (2010) present primitive conditions for nonparametric identification

of positive η0 and g0 based on Perron-Frobenius theory. Chen et al (2014) and Escanciano et al. (2015) also use Perron-Frobenius to obtain identification of related but different Euler equation models. See also Hansen and Scheinkman (2009) and Christensen (2017) for other applications of Perron-Frobenius theory. These identification results are nonparametric and for models without measurement error. In contrast, a simple semiparametric identification condition for the discount factor is presented here, in a model with measurement error, namely η ,g =0. h 0 0i 6 Proposition 6.6 thus shows that regular local identification of the discount factor holds under

rather general conditions on the measurement error mechanism (m and the distribution of εt are nonparametric and unidentified under our conditions). An important empirical literature has provided estimation and inference results on Euler equations accounting for measurement error. Papers within this literature use functional form assumptions for utilities and for the measurement error mechanism. The identification result of Proposition 6.6 opens the door for more robust empirical strategies for inference on the discount factor in microeconomic applications based on the Euler Equation. For example, Altonji and Siow (1987), Runkle (1991), Dynam (2000), and Alan, Attanasio and Browning (2009) assume ∗ ∗ parametric marginal utilities and m(Ct ,εt)= Ct εt, i.e. ∗ Ct = Ct εt, (19)

with εt independent of everything and, in some cases, assumed to be log normally distributed. The identification result above shows that regular local identification of the discount factor follows under more general assumptions than previously recognized, including situations where the marginal utility, the measurement equation and other nonparametric parameters are not identified. This point illustrates the concept of semiparametric identification emphasized in this paper.

25 6.3.2 Identification of Average Risk Aversion

The Average Arrow-Pratt coefficient of Absolute Risk Aversion (AARA) parameter is given by ∂u˙ (C∗)/∂C∗ χ(u ˙ )= E − 0 t t . 0 u˙ (C∗)  0 t  The following conditions guarantee that this parameter is well-defined, and satisfies some prop- erties given below. Let µ∗ denote the probability measure of C∗, with density f ∗( ), and assume t t · the parameter space for marginal utilities U˙ satisfies, for a small positive number ǫ,

U˙ u˙ L (µ∗) :u ˙(c∗) ǫ> 0 and u˙ 2dµ∗ =1 . ⊂ ∈ 2 ≥  Z  The scale normalization in U˙ is necessary to achieve identification in nonparametric Euler equa- tions.

∗ Assumption 5: (v) The functions log (u ˙ 0( )) and log(ft ( )) are continuously differentiable on ∗ · · the convex support of Ct (possibly unbounded) and these functions and their derivatives are in L (µ∗). The true marginal utility satisfiesu ˙ U.˙ 2 0 ∈ By Assumption 5(v) and integration by parts,

∗ ∗ χ(u ˙ 0)= E [log (u ˙ 0(Ct )) d(Ct )] , where ∗ ∗ ∗ ∂ft (c ) 1 d(c ) ∗ ∗ ∗ . (20) ≡ ∂c ft (c )

The functional χ(u ˙ 0), although nonlinear, is concave and differentiable, with pathwise derivative d(C∗) χ˙(b)= E b(C∗) t . t u˙ (C∗)  0 t  ∗ Thus, the AARA parameter as a functional on L2(µ ) has the Riesz’s representer

∗ ∗ d(Ct ) rχ(Ct )= ∗ . (21) u˙ 0(Ct ) To link the marginal utility with the projected marginal utility we need to define the conditional mean operator L : L (µ∗) L (µ), 2 → 2 Lu˙(C )= E [u ˙(C∗) C ] , t t | t which has an adjoint operator L∗ : L (µ) L (µ∗) given by 2 → 2 L∗w(C )= E [w(C ) C∗] . t t | t Applying the results of this paper, one obtains the following identification result for the AARA.

26 Proposition 6.7 Let Assumption 5 hold and assume θ0 is locally identified. Then, the following condition is sufficient for regular local identification of the AARA: there exists g L (µ) such ∈ 2 that, with r given in (21) and l˙∗g = θ A∗g(c) g(c), χ η 0 − ∗ ˙∗ ∗ rχ(Ct )= E lηg(Ct) Ct . (22) h i It is convenient to decompose (22) into two parts: (i) existence of w L (µ) such that ∈ 2 r (C∗)= E [w(C ) C∗]; (23) χ t t | t and (ii) conditions that guarantee that such w belongs to (l˙∗). I provide simple primitive R η conditions for (23) to hold in the multiplicative measurement error model (19), when the error

density is known and given by fε, e.g. log-normal as in e.g. Alan, Attanasio and Browning (2009). In this model, the regularity condition (23) is,

∗ ∗ rχ(c )= fε(c/c )w(c)dc. Z The following Lemma provides a sufficient condition for existence of w L (µ) satisfying this ∈ 2 equation. Let L1(R) and L2(R) denote the set of integrable and squared integrable functions, respectively. For f L (R), define the Fourier transform fˆ = (2π)−1/2 e−itxf(t)dt, where ∈ 1 i = √ 1. Define K(u) = exp(u)f (exp(u)) and x(τ) = exp( τ)r (exp(τ)). − ε − χ R Lemma 6.1 If K(u) is symmetric in u, x L (R) and x/ˆ Kˆ L (R), then, there exists a ∈ 2 ∈ 2 solution w L (µ) of (23). Moreover, a solution is given by ∈ 2 1 xˆ(t) w(c)= Re eit log(c) dt, 2π Kˆ (t) Z where Re denotes the real part.

The symmetry condition on K is satisfied by the log normal distribution used in the empirical literature. I now provide primitive conditions for w (l˙∗). Note that if w / (l˙∗) the AARA ∈ R η ∈ R η is not identified. By duality, w (l˙∗) = (l˙ )⊥ has a simple interpretation: w is orthogonal ∈ R η N η to all projected marginal utilities solving the Euler equation, i.e.

E [w(Ct)η(Ct)] = 0 for all η such that θ0Aη = η. (24)

By compactness of A, the space of such η′s is finite-dimensional (see Kress 1999), which means

that (24) can be tested. Importantly, (24) holds for η0 under (23), since by iterated expectations

∗ E [w(Ct)η0(Ct)] = E [rχ(Ct )η0(Ct)] ∗ ∗ = E [rχ(Ct )u ˙ 0(Ct )] ∗ = E [d(Ct )] =0.

27 A primitive condition for identification of η0 is r(Ct+1,Ct) > 0 and g(Ct+1,Ct) > 0, where r(C ,C ) = E [R C ,C ] . Thus, these primitive conditions and the mild integrability t+1 t t+1| t+1 t conditions of Lemma 6.1 imply regular identification of the ARRA by virtue of Proposition 6.7.

7 Conclusions

This paper provides tools for investigating semiparametric identification, with a particular em- phasis on irregular identification. First, it considers semiparametric identification for linear models and obtains necessary and sufficient conditions for regular and irregular identification. I then show that semiparametric irregular identification is a common feature of many economic models of practical interest. The application to the structural model for unemployment in Al- varez, Borovickov´aand Shimer (2016) illustrates this point. The Euler Equation application illustrates the usefulness of the characterization of regular identification and its applicability to the important class of conditional moment models. Regular identification of the discount factor and measures of risk aversion can be obtained under relatively simple conditions, despite the nonlinearity of both the model and the AARA functional. The question of whether zero information corresponds to a lack of identification is a rather delicate question, as was first pointed out in Chamberlain (1986). Indeed, I show here that irregularity corresponds mathematically to a boundary case in an infinite-dimensional space. Regular identification is, however, easier to characterize and, under mild smoothness conditions, a positive semiparametric Fisher information for the parameter implies its local identification. When the Fisher information is zero, positivity of a new generalized Fisher information intro- duced in this paper implies irregular identification. When the generalized Fisher information is zero, I obtain impossibility results on rates of convergence. The impossibility results on regular identification and rates apply to both linear and nonlinear models and parameters. A number of issues deserve further study. For example, it will be useful to investigate primitive conditions for positive or zero generalized Fisher information in specific economic applications and, using the tools provided here, to see how these conditions translate into specific rates of convergence for estimators. Likewise, Section 9.3 in the Supplemental Appendix provides sufficient conditions for regular and irregular semiparametric identification in nonlinear models and for nonlinear functionals. Applying these results to specific examples, and establishing connections with attainability of rates of convergence for semiparametric estimators, remain topics for future research.

28 8 Appendix: Proofs of Main Results

Proof of Proposition 4.1: I first show that (I ) = (S). From the definition of I , the N λ0 N λ0 implication (S) (I ) trivially holds. The other implication follows from I b, b H = N ⊂ N λ0 h λ0 i Sb 2 . Having (I ) = (S), that (4) is a sufficient condition for identification of φ(λ ) k k N λ0 N 0 follows from the definition of identification. If f = f , then b = λ λ (S) (φ˙), and λ λ0 − 0 ∈N ⊂N hence φ(λ0 + b)= φ(λ0). To prove the necessity, suppose that (4) does not hold, i.e.

(S) (φ˙), N N then there exists b T (λ ) such that b (S) but b / (φ˙). This means by linearity that ∈ 0 ∈ N ∈ N for all c R, c = 0, cb (S) but cb / (φ˙). By Assumption 1(iii) there exists c such that ∈ 6 ∈ N ∈ N λ + cb Λ, cb (S) and cb / (φ˙). This implies that f = f but φ(λ + cb) = φ(λ ). 0 ∈ ∈N ∈N λ0+cb λ0 0 6 0 That is, φ(λ0) is not identified. 

Proof of Theorem 4.1: By Proposition 1 in Luenberger (1997, p.52)

(S) (φ˙) N ⊂N is equivalent to (φ˙)⊥ (S)⊥, N ⊂N since both (S) and (φ˙) are closed linear subspaces, and where henceforth V ⊥ denotes the N N orthocomplement of the subspace V . However, since

φ˙(b)= b, r H, h φi for all b T (λ ), it follows that (φ˙)⊥ = span r . On the other hand, by Theorem 3 in ∈ 0 N { φ} Luenberger (1997, p.157) (S)⊥ = (S∗). N R The identification part follows from Proposition 4.1. The qualification of regular or irregular follows from Theorem 4.1 in van der Vaart (1991), which shows that r (S∗) I > 0.  φ ∈ R ⇐⇒ φ Proof of Theorem 4.2: Suppose φ(λ ) is not identified. That means we can find λ Λ such 0 ∈ that φ(λ) = φ(λ ) and f = f . Then, b = λ λ (S) but φ˙(b) = 0. Choose c sufficiently 6 0 λ λ0 − 0 ∈ N 6 small so that φ˙(cb) 1, and hence cb and ≤ ∈ Bφ

2 Sb Iφ,ρ = inf || || =0, b∈Bφ [φ˙(b)]2ρ

contradicting the positivity of Iφ,ρ. 

29 −1/(2ρ) Proof of Theorem 4.3: Fix K > 0 and take t tn = (2K/C)n and λn = λtn , so −1/(2ρ) ≡ 2 2ρ −1 that φ(λn) φ(λ0) 2Kn and (fλn fλ0 ) /fλ0 < ε(2K/C) n . Let Pn denote the | − | ≥ k − kn n probability measure of fλn , and P0 that of fλ0 , and Pn and P0 their corresponding n fold 2 − product. Lemma 1 in Lecam (1973) and the basic inequality √a √b (a b)2/b for − ≤ − a,b > 0, imply that for each δ > 0 we can find γ > 0 sufficiently small such that

v(Pn, Pn) < δ if (f f ) /f 2 < γ/n, n 0 k λn − λ0 λ0 k where v(P, Q) denotes the total variation distance between P and Q. The details are given as follows. Lemma 1 in Lecam (1973) gives the inequality

v(Pn, Pn) y(2 y2)1/2 n 0 ≤ − 2 if H(P , P ) y/√n and y 1. From the basic inequality √a √b (a b)2/b for a,b > 0, n 0 ≤ ≤ − ≤ − we have   1 2 H2(P , P )= f f dµ n 0 2 λn − λ0 Z 1 p p  (f f )2 /f dµ ≤ 2 λn − λ0 λ0 Z 1   = (f f ) /f 2 2 k λn − λ0 λ0 k

with y2 = ε(2K/C)2ρ/2 (where the last inequality follows from the assumptions of the theorem). Since we can always choose ε> 0 small enough so that y 1, Lemma 1 in Lecam (1973) can be n n ≤ applied to achieve v(Pn, P0 ) < δ by making y small enough. Next, Lemma 8 in Ishwaran (1996) −1/(2ρ) implies that the rate of any estimator cannot be better than OP (2Kn ). Since this holds for each fixed K > 0, it follows that the rate of any estimator must be slower than n−1/(2ρ). 

Proof of Theorem 4.4: The first part follows from Theorem 4.1(i) and the characterization of (S∗) as those elements b T (λ ) with b < . Any element b T (λ ) has the singular R ∈ 0 k k1 ∞ ∈ 0 value expansion (cf. Kress, 1999, Theorem 15.16)

∞ b = b, ϕ ϕ + Π b, h jiH j N (S) j=1 X which implies under semiparametric identification of the functional

∞ φ˙(b)= b, ϕ r ,ϕ h jiH h φ jiH j=1 X

30 and ∞ Sb = λ b, ϕ ψ . j h jiH j j=1 X By Cauchy-Schwarz, for b T (λ ), ∈ 0 ∞ 1/2 ∞ 1/2 ˙ −2 2 2 2 φ(b) λj rφ,ϕj H λj b, ϕj H ≤ h i ! h i ! j=1 j=1 X X = r Sb . k φk1 || || This yields I r −2 . φ,1 ≥k φk1 ∗ By regular identification there exists iφ with finite variance such that S iφ = rφ, and hence another version of Cauchy-Schwarz gives

φ˙(b) = Sb, i h φiH

iφ Sb . ≤k k|| || Indeed, it can be shown that i = r and equality in Cauchy-Schwarz holds when Sb and k φk k φk1 iφ are proportional. The formal argument requires constructing

k −1 j=1 λj iφ, ψj ϕj bk = h i . k i , ψ 2 P j=1 h φ ji Note φ˙(ϕ )= r ,ϕ = λ i , ψ (since SϕP = λ ψ ) and hence j h φ jiH j h φ ji j j j

φ˙(bk) =1,

and b is in the set = b T (λ ): φ˙(b) =0, φ˙(b) 1 . Moreover, k Bφ ∈ 0 6 ≤ n o

2 1 Sbk = || || k i , ψ 2 j=1 h φ ji which can get arbitrarily close to the bound Pi −2 = r −2 by choosing a large k. Thus, we k φk k φk1 conclude I = r −2 = i −2 . φ,1 k φk1 k φk For the second part, note that r < implies r (S∗), and hence identification. The k φkβ ∞ φ ∈ R condition r = then implies zero Fisher information by the first part of the Theorem. On k φk1 ∞ the other hand, by Holder inequality with p = 1/β and q = 1/(1 β), for any 0 <β< 1, and −

31 for all b with b H 1, || || ≤ ∞ 1/2 ∞ 1/2 φ˙(b) λ−2β r ,ϕ 2 λ2β b, ϕ 2 ≤ j h φ jiH j h jiH j=1 ! j=1 ! X X 1/2 ∞ r λ2β b, ϕ 2β b, ϕ 2−2β ≤k φkβ j h jiH h jiH j=1 ! X ∞ β/2 ∞ (1−β)/2 r λ2 b, ϕ 2 b, ϕ 2 ≤k φkβ j h jiH h jiH j=1 ! j=1 ! X X r Sb β. ≤k φkβ || || Thus, for ρ =1/β, Sb 2 1 inf || || > 0, 2ρ 2ρ b∈Bφ,||b||H≤1 [φ˙(b)] ≥ r k φkβ and local identification follows. 

Proof of Proposition 5.1: For the functional φ(λ) = θ it holds (φ˙) = (b , b ): b = 0 . N { θ η θ } Then, by orthogonality

(S)= (b , b ): (l˙′ b + l˙ b )2dP =0 N θ η θ θ η η θ0,η0  Z  ˜′ ˙′ ˙ 2 = (bθ, bη): (l bθ + Π l bθ + lηbη) dPθ0,η0 =0 θ R(l˙η) θ  Z  2 ˜′ ˙′ ˙ 2 = (bθ, bη): l bθ dPθ0,η0 =0, (Π l bθ + lηbη) dPθ0,η0 =0 θ R(l˙η ) θ  Z   Z  ′ ˜ ˙′ ˙ = (bθ, bη): bθIθbθ =0, Π ˙ lθbθ = lηbη . R(lη ) − n o Then, if I˜ > 0, then we have (S) (φ˙). If I˜ =0, then there are two cases: (1) l˙ (l˙ ) θ N ⊂N θ θ ∈ R η and (2) l˙ (l˙ ) (l˙ ). In case (1) the identification condition (S) (φ˙) does not hold, θ ∈ R η R η N ⊂N as we can find b = 0 such that l˙ b = l˙ b (so (b , b ) (S) but (b , b ) / (φ˙)). In case θ 6 θ θ − η η θ η ∈ N θ η ∈ N (2) (S) (φ˙) holds even though there is zero information for the parameter. Thus, if N ⊂ N l˙ (l˙ ) (l˙ ) we have “irregular identification”.  θ ∈ R η R η Proof of Theorem 5.1: Choose 0 <ε 0 such that for all λ (λ ) with θ,ρ ∈ Bδ 0 θ = θ 6 0 (f f ) /f S(λ λ ) (f f ) /f S(λ λ ) θ θ ρ k λ − λ0 λ0 − − 0 k = k λ − λ0 λ0 − − 0 k | − 0| S(λ λ ) θ θ ρ S(λ λ ) k − 0 k | − 0| k − 0 k ε I−1/2 ≤ × θ,ρ < 1, (25)

32 where have used Assumption 2 and the definition of the generalized Fisher information. The inequality (25) implies that (f f ) /f = 0, or equivalently f = f . That is, local k λ − λ0 λ0 k 6 λ 6 λ0 identification holds. 

Proof of Proposition 6.1: Define L : L (π) L (P) as 1 7→ 1

Lb = fz/α,β(t1, t2)b(α, β)dπ(α, β), Z where the conditional density of Z given (α, β) is

f (t , t ) f(t ; α, β)f(t ; α, β), z/α,β 1 2 ∝ 1 2 ( denotes equality up to multiplication by a normalizing constant) and f(t; α, β) denotes the ∝ inverse Gaussian density − 2 β − (αt β) f(t; α, β) e 2t . ∝ t3/2 Decompose L as

Lb = fz/α,β(t1, t2)b(α, β)dπ(α, β)+ fz/α,β(t1, t2)b(α, β)dπ(α, β) Zα≥0 Zα≤0 L b + L b. ≡ + −

Theorem 1 in Alvarez et al. (2016) and Proposition 4.1 here imply that both L+ and L− are −1 −1 linear injective operators, and therefore have inverses, L+ and L− , respectively. Define the normalizing positive constant

CL = f(t1; α, β)f(t2; α, β)λ0(α, β)dπ(α, β)dt1dt2. Z Then, using that the inverse Gaussian satisfies

f(t; α, β)= e2αβf(t; α, β), − it can be shown that

L b = C−1 e4αβf(t ; α, β)f(t ; α, β)b(α, β)dπ(α, β) − L 1 − 2 − Zα≤0 = C−1 e−4αβf(t ; α, β)f(t ; α, β)b( α, β)dπ(α, β) − L 1 2 − Zα≥0 = L (e−4αβb( α, β)). − + − Then, using these results, b (L), i.e. L b + L b = 0, is equivalent to ∈N + − b(α, β)= L−1 (L b) − + − = L−1L (e−4αβb( α, β)) + + − = e−4αβb( α, β). −

33 This concludes the proof after noticing that (S)= (L).  N N

Proof of Proposition 6.2: Let fz,α,β(t1, t2,α,β) denote the joint density of (Z,α,β) wrt the

product of the Lebesgue measure and π, and fα,β/z(t1, t2) the corresponding conditional density of (α, β) given Z. By the definition of S with the reparametrization of Remark 1(ii) 1 Sb = f (t , t )b(α, β)λ (α, β)dπ(α, β) f (t , t ) z/α,β 1 2 0 λ0 1 2 Z 1 = f (t , t )b(α, β)dπ(α, β) f (t , t ) z,α,β 1 2 λ0 1 2 Z

= fα,β/z(t1, t2)b(α, β)dπ(α, β) Z = E [b(α, β) Z] . |

By well-known properties of adjoint operators and substitution of fz/α,β(t1, t2), we obtain

S∗g = E [g(Z) α, β] . |

= g(t1, t2)fz/α,β(t1, t2)dt1dt2 2 ZT = Cβ2e2αβh(α2, β2),

where 1 h(u, v)= g(t1, t2) 3/2 3/2 s(u, v; t1)s(u, v; t2)dt1dt2 2 ZT t1 t2 and ut v s(u, v; t) = exp , t , (u, v) (0, ). − 2 − 2t ∈ T ∈ ∞   We check that the conditions for an application of the Leibniz’s rule hold. These conditions are

m m 1. The partial derivative ∂ s(u, v; t1)s(u, v; t2)/∂ u exists and is a continuous function on an open neighborhood B of (u, v), for a.s. (t , t ) 2. 1 2 ∈ T

2. There is a positive function hm(t1, t2) such that ∂ms(u, v; t )s(u, v; t ) sup 1 2 h (t , t ) (26) ∂mu ≤ m 1 2 (u,v)∈B

and 1 g(t1, t2) 3/2 3/2 hm(t1, t2)dt1dt2 < . (27) 2 ∞ ZT t1 t2 Simple differentiation and induction show that for any integer m 0 ≥ ∂ms(u, v; t )s(u, v; t ) 1 2 =2−m( 1)m(t + t )ms(u, v; t )s(u, v; t ). ∂mu − 1 2 1 2 34 Therefore, by monotonicity we can find u∗ and v∗ such that (26) holds with

−m m ∗ ∗ ∗ ∗ hm(t1, t2)=2 (t1 + t2) s(u , v ; t1)s(u , v ; t2).

Furthermore, by E [g(Z) α, β] < for all α and β, and the boundedness of , condition (27) | ∞ T holds. The continuity of h(u, v) is proved using similar arguments (i.e. for m = 0). 

Proof of Corollary 6.2: By Theorem 4.1 part (i) it is sufficient to check that r / (S∗), φ ∈ R provided Assumption 1(i,ii,iv) holds. By Proposition 6.2 the score operator is well-defined linear and continuous. Furthermore, because heterogeneity is nonparametric and the discussion before the Corollary, the corresponding functional φ˙ for φ(λ )= E [1(α α )1(β β )] is 0 ≤ 0 ≤ 0

φ˙(b)= r , b H, h φ i where H = L0(G ) and the Riesz representer r H is 2 0 φ ∈ r (α, β)=1(α α )1(β β ) G(α , β ). φ ≤ 0 ≤ 0 − 0 0 Clearly, this function r / (S∗) and the result follows.  φ ∈ R Proof of Proposition 6.3: The score operator and the functional satisfy Assumptions 1(i,ii,iv), since r L (π). The condition ∈ 2 r (w, x)=0 π a.s on v

Proof of Proposition 6.4: To show that (14) is sufficient for regular identification, note that (13) is a solution to r = S∗g with g L . The see that g L , note ∈ 2 ∈ 2 E g2(Z) = E g2(1,V,X)G (V,X)+ g2(0,V,X)(1 G (V,X)) 0 − 0 2 2   E g (1,V,X)+ g (0,V,X)  ≤ vmax ∂r(v, x) 2 1 2   dµ (x)dv < . ≤ ∂v f (v) X ∞ Z0 Z   V/X=x The necessity follows from the arguments prior to the statement of the Proposition, i.e. from

∂r(w, x) =(g(0,w,x) g(1,w,x)) f (w), ∂w − V/X=x and the fact that g L . Conclude by an application of Theorem 4.1(i).  ∈ 2

35 Proof of Proposition 6.5: I first verify that median WTP is identified. This is implied by continuity G (v) and the support condition v φ(λ ), since by the conditional independence 0 max ≥ 0 G (w)= E[E [Y X,V = w]] 0 | is identified on and G (v ) 1/2 by monotonicity of cdfs. Standard arguments on SV 0 max ≥ quantiles, see Corollary 21.5 in van der Vaart (1998), yield that the median WTP has an influence function 1(w<φ(λ0)) 0.5 rφ(w)= −{ − }. E [λ0 (φ(λ0),Xi)] The discontinuity of the influence function then implies that identification must be irregular, since S∗g(w, x) is absolutely continuous in w. 

Proof of Proposition 6.6: By(17) Assumption 2 holds with ρ = 1. By Theorem 5.1 it remains to check I > 0. This is, however, equivalent to (1, 0) (S∗), or existence of g L (µ) such θ,1 ∈ R ∈ 2 that Aη ,g =1, l˙∗g =0. h 0 i η If η ,g = 0 for g one of the eigenfunctions of A∗, we define g = cg with c =(θ η ,g )−1 . h 0 0i 6 0 0 0h 0 0i Note that, l˙∗g = cl˙∗g = 0 and Aη ,g = cθ η ,g =1.  η η 0 h 0 i 0h 0 0i

Proof of Proposition 6.7: The functional χ(u ˙ 0), although nonlinear, is concave and differen- tiable, with pathwise derivative d(C∗) χ˙(b)= E b(C∗) t , t u˙ (C∗)  0 t  which implies, with l˙ η = θ Aη(c) η(c), η 0 −

̟(ǫ) = sup χ(u ˙ 1) χ(u ˙ 0) ∗ ˙ | − | ku˙ 1−u˙ 0k2,µ ≤δ,||lηL(˙u1−u˙ 0)||2,µ≤ǫ

δ sup χ˙(u ˙ 1 u˙ 0) ≤ ∗ ˙ | − | ku˙ 1−u˙ 0k2,µ ≤δ,||lηL(˙u1−u˙ 0)||2,µ≤ǫ

= δ sup E l˙ηL(u ˙ 1(Ct) u˙ 0(Ct))g(Ct) ku˙ −u˙ k ∗ ≤δ,||l˙ L(˙u −u˙ )|| ≤ǫ − 1 0 2,µ η 1 0 2,µ h i

δ g ǫ, ≤ || ||2,µ where the first inequality uses concavity, the last equality uses (22), so

χ˙(u ˙ u˙ )= E [(u ˙ u˙ )(C∗)r (C∗)] 1 − 0 1 − 0 t χ t = E (u ˙ u˙ )(C∗)l˙∗g(C ) 1 − 0 t η t h i = E l˙ L(u ˙ (C ) u˙ (C ))g(C ) , η 1 t − 0 t t h i 36 and the last inequality follows by Cauchy-Schwarz. Lemma 9.3 in the Supplemental Appendix

then implies local identification of χ(u ˙ 0). It remains to show the regularity, but this follows from Cauchy-Schwarz, since

2 ˙ l˙ Lu˙ 2 lηLu˙ 1 inf || η || = > 0. 2 2 2 u˙:˙χ(˙u)6=0 χ˙(u ˙) ≥ 2 ˙ g 2,µ g 2,µ lη Lu˙ || || | | || ||



Proof of Lemma 6.1: By the change of variables with c = exp(z) and c∗ = exp(τ), and multiplying both sides by exp( τ) the integral equation −

∗ ∗ rχ(c )= fε(c/c )w(c)dc Z is transformed into a convolution-type problem

x(τ)= K(z τ)y(z)dz, (28) − Z where x(τ) = exp( τ)r (exp(τ)), K(u) = exp(u)f (exp(u)) and y(z)= w(exp(z)). By Polyanin − χ ε and Manzhirov (2008, p. 285) if x L (R), a necessary and sufficient condition for existence of ∈ 2 y L (R) satisfying (28) isx/ ˆ Kˆ L (R). The solution is given by ∈ 2 ∈ 2 1 xˆ(t) y(z)= Re eitz dt, 2π Kˆ (t) Z and in terms of w, 1 xˆ(t) w(c)= Re eit log(c) dt. 2π Kˆ (t) Z Note this solution is also in L (µ) ifx/ ˆ Kˆ L (R), since, by a change of variables and Fubini, 2 ∈ 2

2 1 xˆ(t) xˆ(s) i(t−s) log Ct E w(Ct) = 2 E[e ]dtds | | (2π) Kˆ (t) Kˆ (s) Z Z   2 1 xˆ(t)

2 dt < . ≤ (2π) Kˆ (t) ∞ Z

37 9 Supplemental Appendix

I will extensively use basic results from operator theory and Hilbert spaces in this Supplemental Material. See Carrasco, Florens and Renault (2007) for an excellent review of these results. This Appendix is organized as follows. Section 9.1 establishes sufficient conditions for local irregular identification in models linear in nuisance parameters. Section 9.2 characterizes identification of linear continuous functionals of nuisance parameters in semiparametric models. Section 9.3 establishes sufficient conditions for identification in general nonlinear models.

9.1 Models Linear in Nuisance Parameters

Define the nuisance score operator

fθ,η0+bη fθ,η0 l˙η(θ)bη = − , (29) fθ0,η0

and the (negative) approximated score for θ as

fθ0,η0 fθ,η0 sθ = − . fθ0,η0

I drop the dependence on θ and denote l˙ l˙ . Define the (negative) approximated efficient 0 η ≡ η(θ0) scores ˜θ := sθ Π ˙ sθ, and the approximated Fisher Information − R(lη(θ)) G(θ)= s˜ 2. || θ|| Let Ψ be the class of measurable functions ψ : [0, ) [0, ) that are increasing, right ∞ −→ ∞ continuous at 0 and with ψ(0) = 0. Then, consider the following assumption.

Assumption D: (i) The map l˙ : T (η ) L is linear for each θ in a neighborhood of η(θ) 0 ⊆ H 7→ 2 θ (ii) there exists a positive constant C such that G(θ) > Cψ( θ θ 2) in a neighborhood of 0 | − 0| θ , where ψ Ψ. 0 ∈ Assumption D(i) holds for many models of interest. Assumption D(ii) follows from conditions on

the derivative of G(θ) at θ0. For example, if G(θ) is differentiable at θ0 with full rank derivative

at θ0, then Assumption D(ii) holds with ψ(ǫ)= ǫ. This corresponds to the case of regular local identification. A necessary condition for Assumption D(ii) is that (l˙∗ ) =0, since otherwise N η(θ) 6 G(θ) = 0.

Theorem 9.1 Let Assumption D hold. Then, θ is locally identified at θ0.

38 Proof of Theorem 9.1: Write f f f f f f θ,η − θ0,η0 = θ,η − θ,η0 θ0,η0 − θ,η0 fθ0,η0 fθ0,η0 − fθ0,η0 = l˙ b s . η(θ) η − θ Note that by standard least squares theory for all b T (η ), and all θ in a neighborhood of θ , η ∈ 0 0 ˙ 2 2 lη(θ)bη sθ Π ˙ sθ sθ || − || ≥ || R(lη(θ)) − || > Cψ( θ θ 2). | − 0| This inequality implies local identification. 

9.2 Functionals of Nuisance Parameters in Semiparametric Models

Let χ : R be a linear continuous functional, and let r T (η ) be such that for all H 7→ χ ∈ 0 ⊂ H b T (η ), η ∈ 0 χ(b )= b ,r . η h η χiH To give a general result, I allow for θ to be infinite-dimensional, and ask the question: When lack of identification of one parameter, here θ, does not have an effect, at least locally, on identification on another parameter χ(η)? A similar characterization to that of Proposition 5.1 is obtained for φ(λ) = χ(η), allowing for singular information for both θ and the functional φ(λ)= χ(η). Define the operator

− ˙∗ ˙ ˙∗ ˙ Aηθ = lηlη lηlθ,   where B− denotes the generalized Moore-Penrose inverse of B.

Proposition 9.1 For the functional φ(λ)= χ(η) R: (i) if (l˙ ) (l˙ )= 0 , then (S) ∈ R θ ∩ R η { } N ⊂ (φ˙) holds iff r (l˙∗); (ii) if (l˙ ) (l˙ ) = 0 , then (S) (φ˙) holds if r N χ ∈ R η R θ ∩ R η 6 { } N ⊂ N χ ∈ (l˙∗) (A∗ ). R η ∩N ηθ Proof of Proposition 9.1: Note that for the functional φ(λ) = χ(η), where χ : H R is a 7→ linear continuous functional with χ(b )= b ,r , η h η χiH it holds that (φ˙)= (b , b ): b ,r =0 . Therefore, by the proof of Proposition 5.1 (which N { θ η h η χiH } is also valid for infinite-dimensional θ, with I˜ interpreted as an operator), (S) (φ˙) θ N ⊂ N ′ ˜ ˙′ ˙ ˜ iff bθIθbθ = 0 and Π ˙ lθbθ = lηbη implies bη,rχ H = 0. If Iθ is positive definite, then R(lη ) − h i 39 (b , b ) (S) iff b =0and 0= l˙ b . Therefore, (b , b ) (φ˙) iff (l˙ ) (χ), which is θ η ∈ N θ η η θ η ∈ N N η ⊂ N equivalent to r (l˙∗). If I˜ is semi-positive definite, there are two cases (i) (l˙ ) (l˙ ) = 0 χ ∈ R η θ R θ ∩R η 6 { } and (ii) (l˙ ) (l˙ ) (l˙ ). In case (i), l˙ b = l˙ b , and for all such b it must hold that R θ ⊂ R η R η θ θ − η η η b ,r = 0. All the solutions of l˙ b = l˙ b can be written as b = (l˙ ) A b . Thus, h η χiH θ θ − η η η N η − ηθ θ the orthogonality b ,r = 0 holds if r (l˙∗) (A∗ ). In case (ii) 0 = l˙ b must imply h η χiH χ ∈ R η ∩N ηθ η η that (b , b ) (φ˙), which holds if (l˙ ) (χ) or equivalently r (l˙∗). Therefore, θ η ∈ N N η ⊂ N χ ∈ R η if (l˙ ) (l˙ ) = 0 (I˜ is positive definite or case (ii) above) then (S) (φ˙) holds iff R θ ∩ R η { } θ N ⊂ N r (l˙∗); (ii) if (l˙ ) (l˙ ) = 0 (case (i) above) then (S) (φ˙) holds if (l˙∗) (A∗ ). χ ∈ R η R θ ∩R η 6 { } N ⊂N R η ∩N ηθ 

Remark 9.1 The conditions for local identification of χ(η0) depend on whether θ0 is locally

identified or not. The case (ii) corresponds to the situation of local unidentification of θ0, and it

is shown that despite this lack of local identification of θ0, χ(η0) might still be locally identified. To interpret the result, one can think of r (l˙∗) as the identification condition for χ(η ) that χ ∈ R η 0 would be needed if θ0 was known. If θ0 is not known, but is identified, one can treat it as known for the purpose of identifying χ(η0). However, if θ0 is not identified, an additional condition must be met to avoid the lack of identification of θ0 to spread out to χ(η0). Technically, this condition is that for all b = (b , b ) such that l˙ b = l˙ b (these b′s are directions that lead to θ η θ θ − η η zero nonparametric information), it must hold that b ,r = 0. Under r (l˙∗), a simple h η χiH χ ∈ R η condition for this orthogonality is r (A∗ ). χ ∈N ηθ

˙∗ ˙∗ Remark 9.2 In both cases rχ (lη) (lη) corresponds to the case of zero information for ∈ R R ∗ φ(λ)= χ(η) at φ(λ0)= χ(η0). Regular identification of χ(η) in case (ii) requires that for all rχ that solve r = l˙∗r∗ it holds that r∗ (l˙∗). Under this condition, lack of identification of θ χ η χ χ ∈ N θ 0 does not affect regular identification of χ(η0).

Van der Vaart (1991) has shown that a positive information of χ(η ) is equivalent to r (l˙∗) 0 χ ∈ R η when θ0 is locally regularly identified and η0 is identified. Proposition 9.1 characterizes local

regular and irregular identification of χ(η0), allowing for θ0 to be locally regular or irregularly identified, or even unidentified. The results of Proposition 9.1 are applied to measures of risk aversion in Example 3 on the Euler Equation.

9.3 General Nonlinear Models

The following modulus of continuity is shown to be useful for the study of identification

̟(ǫ) = sup φ(λ) φ(λ0) . (30) −1 | − | λ∈Bδ (λ0):|| fλ−fλ f ||≤ǫ ( 0 ) λ0

40 I drop the dependence of ̟(ǫ) on δ for simplicity of notation. Lemma 9.3 below shows that ̟(ǫ) 0 as ǫ 0 is sufficient for local identification of φ(λ ). A related modulus of continuity was ↓ ↓ 0 introduced in Donoho and Liu (1987) for the purpose of obtaining bounds on the optimal rate of convergence for functionals of a density (they assume identification and use the Hellinger metric). −1 Using (fλ fλ ) f is convenient because we can exploit simultaneously the linearity of || − 0 λ0 || certain models and the Hilbert space structure.

Lemma If there exists δ > 0 such that ̟(ǫ) 0 as ǫ 0, then φ(λ ) is locally identified. → → 0 Proof of Lemma 9.3: Suppose that φ(λ0) is not locally identified. Then, for all δ > 0, we can find a λ∗ Λ (λ ) such that (f ∗ f ) /f = 0 and φ(λ∗) = φ(λ ), and therefore, for ∈ δ 0 k λ − λ0 λ0 k 6 0 all ǫ> 0, ̟(ǫ) φ(λ∗) φ(λ ) > 0, ≥| − 0 | showing that ̟(ǫ) does not converge to zero as ǫ 0.  → The following result provides a general local identification result. Recall Ψ is the class of measurable functions ψ : [0, ) [0, ) that are increasing, right continuous at 0 and with ∞ −→ ∞ ψ(0) = 0.

Assumption N: For all ε> 0, there exists δ > 0, ψ , ψ Ψ, and a continuous linear operator 1 2 ∈ S : T (λ ) H L , such that for all λ =(θ, η) (λ ), 0 ⊆ 7→ 2 ∈ Bδ 0 (i)

(f f ) /f S(λ λ ) < εψ ( λ λ H); k λ − λ0 λ0 − − 0 k 1 k − 0k (ii)

φ(λ) φ(λ ) ψ ( λ λ H); and | − 0 | ≤ 2 k − 0k (iii) S(λ λ ) inf || − 0 || > 0. λ∈Bδ(λ0) ψ ( λ λ H) 1 k − 0k

Assumption N(i) and N(ii) are mild smoothness conditions that often hold in applications. Condition N(iii) is a positive nonparametric generalized information condition. Then, I have the following

Theorem 9.2 Let Assumption N hold. Then, φ(λ) is locally identified at φ(λ0).

41 −1 Proof of Theorem 9.2: Assumptions N(i-ii) imply that if (fλ fλ ) f ǫ then we can || − 0 λ0 || ≤ find a positive constant C and 0 <ε

̟(ǫ) = sup φ(λ) φ(λ0) , −1 | − | λ∈Bδ(λ0):|| fλ−fλ f ||≤ǫ ( 0 ) λ0

sup ψ2 ( λ λ0 H) ≤ ǫ k − k λ∈Bδ (λ0):ψ1(kλ−λ0kH)≤ C−ε ǫ ψ ψ−1 ≤ 2 1 C ε   −  0 as ǫ 0. → → Thus, the Theorem follows from Lemma 9.3. 

9.3.1 A Counterexample

I provide a counterexample, building on that given in Chen et al. (2014, pg. 791), that shows that

regular identification is not equivalent to Iφ > 0 in general (and hence to Van der Vaart’s (1991) diffentiability condition). Let λ =(λ1,λ2, ...) be a sequence of real numbers. Let (p1,p2, ...) be ∞ probabilities, pj > 0, j=1 pj = 1. Let f(x) be a twice continuously differentiable function of a scalar x that is bounded with bounded second derivative. Suppose f(x) = 0 if and only if P 2 x 0, 1 and ∂f(0)/∂x = 1. Let m(λ)=(f(λ1), f(λ2), ...) also be a sequence with m(λ) = ∈{ } 1/4 k k ∞ p f 2(λ ). Then, for λ = ∞ p λ4 the mapping m is Frechet differentiable at j=1 j j k kΛ j=1 j j λ = 0 with derivative Sb = b, but λ = 0 is not locally identified (Chen et al. 2014). P0 0P  Consider the nonlinear functional ∞

φ(λ)= f(λj)pj. j=1 X This functional has a derivative at λ0 = 0 given by ∞ ˙ φ(b)= bjpj, j=1 X and by Cauchy-Schwarz ∞ 2 ˙ 2 φ(b) bj pj ≤ ! j=1 X = Sb 2 . k k

42 Hence, I 1 > 0. However, the functional is not identified, since φ(αk)=0= φ(0), where φ ≥ αk = (0, .., 0, 1, 1, 1...) has zeros in the first k positions and a one everywhere else.

43 References

Alan, S., Attanasio, O. and M. Browning (2009): “Estimating Euler Equations With Noisy Data: Two Exact GMM Estimators,” Journal of Applied Econometrics, 24, 309-324.

Alvarez, F., Borovickovˇ a,´ K. and R. Shimer (2016): “Decomposing Duration Depen- dence in a Stopping Time Model,” manuscript.

Altonji, J.G. and A. Siow (1987): “Testing the Response of Consumption to Income Changes with Noisy Panel Data,” Quaterly Journal of Economics, 102, 293-328.

Andrews, D.W.K (2011): “Examples of L2-Complete and Boundedly-Complete Distribu- tions,” Cowles Foundation for Research in Economics.

Andrews, D.W.K and M. A. Schafgans (1998): “Semiparametric Estimation of the In- tercept of a Sample Selection Model,” Review of Economic Studies, 65, 497-517.

Bajari, P., J. Hahn, H. Hong, and G. Ridder (2011): “A Note on Semiparametric Estimation of Finite Mixtures of Discrete Choice Models With Application to Game Theoretic Models,” International Economic Review, 52, 807-824.

Bekker, P. and Wansbeek, T. (2001): “Identification in Parametric Models,” in B. Balt- agi, ed., ‘Companion to Theoretical Econometrics’, Blackwell Companions to Contemporary Economics, Basil Blackwell, Oxford, U.K., chapter 7, pp. 144-161.

Begun, J. M., W. J. Hall, W. M. Huang, and J. A. Wellner (1983): “Information and Asymptotic Efficiency in Parametric-Nonparametric Models,” The Annals of Statistics, 11, 432-452.

Bickel, P. J., C. A. J. Klassen, Y. Ritov, and J. A. Wellner (1998): Efficient and Adaptive Estimation for Semiparametric Models. New York: Springer-Verlag.

Bonhomme, S. (2011): “Panel Data, Inverse Problems, and the Estimation of Policy Parame- ters,” Unpublished manuscript.

Bonhomme, S. (2012): “Functional Differencing,” Econometrica, 80, 1337-1385.

Carrasco, M., J. P. Florens, and E. Renault (2007): “Linear Inverse Problem in Structural Econometrics Estimation Based on Spectral Decomposition and Regularization,” in Handbook of Econometrics, vol. 6, ed. by J. J. Heckman and E. E. Leamer. Amsterdam: North-Holland, 5633–5751.

44 Carson, R.T. and W.M. Hanemann (2005): ”Contingent Valuation,” Handbook of Envi- ronmental Economics 2, Eds. K.-G. M¨aler and J.R. Vincent, 821-936.

Cattaneo, M. and Escanciano, J.C. (2017): “Regression Discontinuity Designs: Theory and Applications,” Volume 38 of Advances in Econometrics, Emerald Group Publishing.

Chamberlain, G. (1986): “Asymptotic Efficiency in Semi-Parametric Models with Censor- ing,” Journal of Econometrics, 34, 305-334.

Chamberlain, G. (1992): “Efficiency Bounds for Semiparametric Regression,” Econometrica, 60, 567-596.

Chamberlain, G. (2010): “Binary Response Models for Panel Data: Identification and Infor- mation,” Econometrica, 78,159-168.

Chen, X., V. Chernozhukov, S. Lee, and W. Newey (2014), “Identification in Semi- parametric and Nonparametric Conditional Moment Models,” Econometrica, 82, 785-809.

Chen, X. and Z. Liao (2014): “Sieve M-Inference of Irregular Parameters”, Journal of Econometrics, 182, 70-86.

Chen, X. and S. C. Ludvigson (2009): “Land of Addicts? An Empirical Investigation of Habit-Based Asset Pricing Models,” Journal of Applied Econometrics, 24, 1057-1093.

Chen, X., Pouzo, D. (2015) “Sieve quasi likelihood ratio inference on semi/nonparametric conditional moment models,” Econometrica 83, 1013-1079.

Chen, X., and M. Reiss (2011): “On Rate Optimallity for Ill-Posed Inverse Problems in Econometrics,” Econometric Theory, 27, 497–521.

Chen, X., and A. Santos (2018): “Overidentification in Regular Models,” Econometrica 86, 1771-1817.

Christensen, T.M (2017): “Nonparametric Stochastic Discount Factor Decomposition,” Econometrica, 85, 1501-1536.

Debnath, L and P. Mikusinski (2005): “Hilbert Spaces With Applications,” Elsevier Academy Press.

Dellavigna, S. and M.D. Paserman (2005): “Job Search and Impatience,” Journal of Labor Economics, 23, 527-588.

45 Donoho, D.L. and Liu, R.C. (1987): “Geometrizing Rates of Convergence, I.” Technical Report 137, Dept. Statistics, Univ. California, Berkeley.

Dufour, J.M. and Liang, X. (2014): “Necessary and Sufficient Conditions for Nonlinear Parametric Function Identification”, unpublished manuscript.

Dynam, K. (2000): “Habit Formation in Consumer Preferences: Evidence from Panel Data”, American Economic Review, 90, 391-402.

Escanciano, J.C. and S. Hoderlein, (2010): “Nonparametric Identification of Euler Equa- tions,” unpublished manuscript.

Escanciano, J.C., S. Hoderlein, A. Lewbel, O. Linton and S. Srisuma, (2015): “Nonparametric Euler Equation Identification and Estimation,” unpublished manuscript.

Fan, J. (1991): “On the Optimal Rates of Convergence for Nonparametric Deconvolution Problems,” The Annals of Statistics, 19 1257-1272.

Fisher, F. M. (1966), The Identification Problem in Econometrics, McGraw-Hill, New York.

Goh, C. (2017): “Rate-Optimal Estimation of the Intercept in a Semiparametric Sample- Selection Model,” working paper.

Graham, B. W., and J. L. Powell (2012): ”Identification and Estimation of Average Partial Effects in ’Irregular’ Correlated Random Coefficient Panel Data Models,” Econometrica, 80, 2105-2152.

Hahn, J. (1994): “The Efficiency Bound of the Mixed Proportional Hazard Model,” Review of Economic Studies, 61, 607-629.

Hansen, L. P. and J. A. Scheinkman (2009): “Long-Term Risk: An Operator Approach,” Econometrica, 77, 177-234.

Heckman, J. J. (1990), “Varieties of selection bias”, The American Economic Review 80, 313–318.

Heckman, J.J. and B. Singer (1984a): “The Identifiability of the Proportional Hazard Model”, Review of Economic Studies, 51,231-241.

Heckman, J.J. and B. Singer (1984b): “A Method for Minimizing the Impact of Distribu- tional Assumptions in Econometric Models for Duration Data,” Econometrica, 52, 271-320.

Hu, Y., and S. M. Schennach (2008): “Instrumental Variable Treatment of Nonclassical Measurement Error Models,” Econometrica, 76 (1), 195-216.

46 Hu, Y., and M. Shum (2012): “Nonparametric Identification of Dynamic Models With Un- observed State Variables,” Journal of Econometrics, 171, 32-44.

Hurwicz L. (1950): “Generalization of the Concept of Identification,” In in Dynamic Economic Models, ed. T.C. Koopmans. New York: Wiley.

Ishwaran, H. (1996): “Identifiability and Rates of Estimation for Scale Parameters in Location Mixture Models,” The Annals of Statistics 24 1560-1571.

Ishwaran, H. (1999): “Information in Semiparametric Mixtures of Exponential Families,” The Annals of Statistics 27, 159-177.

Khan, S. and D. Nekipelov (2016): “Information Structure and Statistical Information in Discrete Response Models,” unpublished manuscript.

Khan, S. and E. Tamer (2010): “Irregular Identification, Support Conditions, and Inverse Weight Estimation,” Econometrica, 6, 2021-2042.

Koopmans, T.C. (1949): “Identification Problems in Economic Model Construction,” Econo- metrica, 17, 125-144.

Koopmans, T. C., and O. Reirsol (1950): “The Identification of Structural Characteris- tics,” Annals of Mathematical Statistics, 21, 165-181.

Koˇsevnik, Y.A. and B. Y. Levit (1976): “On a Non-parametric Analogue of the Information Matrix,” Theory of Probability and its Applications, 21, 738-753.

Kress, R. (1999): Linear Integral Equations. Springer.

LeCam, L. (1973): “Convergence of Estimates Under Dimensionality Restrictions”, The Annals of Statistics, 1, 38-53.

Lewbel, A. (1997): “Semiparametric Estimation of Location and Other Discrete Choice Mo- ments,” Econometric Theory, 13, 32-51.

Lewbel, A. (2016): “The Identification Zoo – Meanings of Identification in Econometrics” Forthcoming in the Journal of Economic Literature.

Lewbel, A., Linton, O.B.B and D. McFadden (2011): “Estimating Features of a Distri- bution From Binomial Data,” Journal of Econometrics, 162, 170-188.

Lewbel, O. Linton and S. Srisuma, (2011): “Nonparametric Euler Equation Identification and Estimation,” Boston College Working Papers in Economics 757.

47 Luenberger, D. G. (1997): Optimization by Vector Space Methods. New York: John Wiley & Sons. 1969 edition.

MaCurdy, T.E (1981): “An Empirical Model of Labor Supply in a Life-Cycle Setting,” Journal of Political Economy, 89, 1059-1085

Manski, C. (1975): “Maximum Score Estimation of the Stochastic Utility Model of Choice,” Journal of Econometrics 3, 205-228.

Manski, C. (1987): “Semiparametric Analysis of Random Effects Linear Models From Binary Panel Data,” Econometrica, 55, 357-362.

Matzkin, R. L. (2007). Nonparametric Identification. In J.J. Heckman and E.E. Leamer (eds.), Handbook of Econometrics, Vol. 6b, pp. 5307-5368, Elsevier, New York.

Matzkin, R. L. (2013): “Nonparametric Identification in Structural Economic Models,” An- nual Review of Economics, 5.

Newey, W. K. (1990): “Semiparametric Efficiency Bounds,” Journal of Applied Econometrics, 5, 99-135.

Polyanin, A. D. and A. V. Manzhirov (2008): “Handbook of Integral Equations: Second Edition,” Chapman and Hall/CRC.

Ridder, G. and T. Woutersen (2003): “The Singularity of the Efficiency Bound of the Mixed Proportional Hazard Model,” Econometrica, 71, 1579-1589.

Rothenberg, R. J. (1971): “Identification in Parametric Models, ” Econometrica, 39(3), 577-591.

Rudin, W. (1973). Functional analysis. McGraw-Hill, New York.

Runkle, D.E. (1991): “Liquidity Constraints and the Permanent Income Hypothesis,” Journal of Monetary Economics, 27, 73-98.

Samwick, A.A. (2006): “Saving for Retirement: Understanding the Importance of Hetero- geneity,” Business Economics, 41, 21-27.

Santos, A. (2011), ‘Instrumental Variables Methods for Recovering Continuous Linear Func- tionals’, Journal of Econometrics 161, 129-146.

Sargan, J.D. (1983): “Identification and Lack of Identification,” Econometrica, 51, 1605-1633.

48 Severini, T. A., and G. Tripathi (2006): “Some Identification Issues in Nonparametric Linear Models with Endogenous Regressors,” Econometric Theorey, 22(2), 258–278.

Severini, T. A., and G. Tripathi (2012): “Efficency Bounds for Estimating Linear Func- tionals of Nonparametric Regression Models with Endogenous Regressors,” Journal of Econo- metrics, 170(2), 491-498.

Shapiro, M.D. (1984): “The Permanent Income Hypothesis and the Real Interest Rate, Eco- nomics Letters, 14, 93-100.

Stein, C. (1956): “Efficient Nonparametric Testing and Estimation,” in Proceedinigs of the Third Berkeley Symposium on Mathematical Statistics and Probability, 1. Berkeley: University of California Press.

van der Vaart, A. W. (1991): “On Differentiable Functionals,” The Annals of Statistics, 19, 178-204. van der Vaart, A. W. (1998). Asymptotic Statistics, vol. 3 of Cambridge Series in Statistical and Probabilistic Mathematics. Cambridge University Press, Cambridge.

Venti S.F and D.A. Wise (1998): “The Cause of Wealth Dispersion at Retirement: Choice or Chance? American Economic Review, 88, 185-191.

49