Skip to main content
Log in

A new bounded log-linear regression model

  • Published:
Metrika Aims and scope Submit manuscript

Abstract

In this paper we introduce a new regression model in which the response variable is bounded by two unknown parameters. A special case is a bounded alternative to the four parameter logistic model. The four parameter model which has unbounded responses is widely used, for instance, in bioassays, nutrition, genetics, calibration and agriculture. In reality, the responses are often bounded although the bounds may be unknown, and in that situation, our model reflects the data-generating mechanism better. Complications arise for the new model, however, because the likelihood function is unbounded, and the global maximizers are not consistent estimators of unknown parameters. Although the two sample extremes, the smallest and the largest observations, are consistent estimators for the two unknown boundaries, they have a slow convergence rate and are asymptotically biased. Improved estimators are developed by correcting for the asymptotic biases of the two sample extremes in the one sample case; but even these consistent estimators do not obtain the optimal convergence rate. To obtain efficient estimation, we suggest using the local maximizers of the likelihood function, i.e., the solution to the likelihood equations. We prove that, with probability approaching one as the sample size goes to infinity, there exists a solution to the likelihood equation that is consistent at the rate of the square root of the sample size and it is asymptotically normally distributed.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

References

  • Cheng RCH, Amin NAK (1983) Estimating parameters in continuous univariate distributions with a shifted origin. J R Stat Soc Ser B (Methodological) 45:394–403

    MATH  MathSciNet  Google Scholar 

  • Cheng RCH, Iles TC (1987) Corrected maximum likelihood in non-regular problems. J R Stat Soc Ser B (Methodological) 49:95–101

    MATH  MathSciNet  Google Scholar 

  • Cheng RCH, Traylor L (1995) Non-regular maximum likelihood problems. J R Stat Soc Ser B (Methodological) 57:3–44

    MATH  MathSciNet  Google Scholar 

  • Cohen AC (1951) Estimating parameters of logarithmic-normal distributions by maximum likelihood. J Am Stat Assoc 46:206–212

    Article  MATH  Google Scholar 

  • Cohen AC, Whitten BJ (1980) Estimation in the three-parameter lognormal distribution. J Am Stat Assoc 75:399–404

    Article  MATH  MathSciNet  Google Scholar 

  • DeLean A, Munson PJ, Rodbard D (1978) Simultaneous analysis of families of sigmoidal curves: application to bioassay, radioligand assay, and physiological dose-response curves. Am J Phys Endocrinol Metab 235:E97–E102

    Google Scholar 

  • Dragalin V, Hsuan F, Padmanabhan SK (2007) Adaptive designs for dose-finding studies based on sigmoid \(e_{\text{ max }}\) model. J Biopharm Stat 17:1051–1070

    Article  MathSciNet  Google Scholar 

  • Ernst AA, Nick TG, Weiss SJ, Houry D, Mills T (1997) Domestic violence in an inner-city ED. Ann Emerg Med 30:190–197

    Article  Google Scholar 

  • FDA (2010) Characterization and qualification of cell substrates and other biological materials used in the production of viral vaccines for infectious disease indications. Guidance for Industry U.S, Department of Health and Human Services, Food and Drug Administration, Center for Biologics Evaluation and Research

  • Feng F, Sales AP, Kepler TB (2011) A Bayesian approach for estimating calibration curves and unknown concentrations in immunoassays. Bioinformatics 27(5):707–712

    Article  Google Scholar 

  • Finke MD, DeFoliart GR, Benevenga NJ (1987) Use of a four-parameter logistic model to evaluate the protein quality of mixtures of mormon cricket meal and corn gluten meal in rats. J Nutr 117:1740–1750

    Google Scholar 

  • Finke MD, DeFoliart GR, Benevenga NJ (1989) Use of a four-parameter logistic model to evaluate the quality of the protein from three insect species when fed to rats. J Nutr 119:864–871

    Google Scholar 

  • Galambos J (1978) The asymptotic theory of extreme order statistics. Wiley, New York

    MATH  Google Scholar 

  • Giesbrecht F, Kempthorne O (1976) Maximum likelihood estimation in the three-parameter lognormal distribution. J R Stat Soc Ser B (Methodological) 38:257–264

    MATH  MathSciNet  Google Scholar 

  • Hall P (1982) On estimating the endpoint of a distribution. Ann Stat 10:556–568

    Article  MATH  Google Scholar 

  • Hall P, Wang JZ (2005) Bayesian likelihood methods for estimating the end point of a distribution. J R Stat Soc Ser B (Statistical Methodology) 67:717–729

    Article  MATH  MathSciNet  Google Scholar 

  • Harter HL, Moore AH (1966) Local-maximum-likelihood estimation of the parameters of three-parameter lognormal populations from complete and censored samples. J Am Stat Assoc 61:842–851

    Article  MathSciNet  Google Scholar 

  • Hill BM (1963) The three-parameter lognormal distribution and Bayesian analysis of a point-source epidemic. J Am Stat Assoc 58:72–84

    Article  MATH  Google Scholar 

  • Holford NH, Sheiner LB (1981) Understanding the dose-effect relationship: clinical application of pharmacokinetic-pharmacodynamic models. Clin Pharmacokinet 6:429–453

    Article  Google Scholar 

  • Kpamegan, E., Jani, D.: Anti-F IgG ELISA based assay data from a qualification study, personal communication, Novavax Inc (2013)

  • Leonov S, Miller S (2009) An adaptive optimal design for the \(E_{\text{ max }}\) model and its application in clinical trials. J Biopharm Stat 19:360–385

    Article  MathSciNet  Google Scholar 

  • MacDougall J (2006) Analysis of dose-response studies \({E}_{\text{ max }}\). In: Ting N (ed) Dose finding in drug development. Springer, New York, pp 127–145

    Chapter  Google Scholar 

  • Menon A, Bhandarkar S (2004) Predicting polymorphic transformation curves using a logistic equation. Int J Pharm 286:125–129

    Article  Google Scholar 

  • Nix B, Wild D (2001) Calibration curve-fitting. In: Wild D (ed) The immunoassay handbook. Nature Publishing Group, New York, pp 198–210

    Google Scholar 

  • R Core Team (2013) R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna

  • Rao CR (1973) Linear statistical inference and its applications, 2nd edn. Wiley, New York

    Book  MATH  Google Scholar 

  • Ratkowsky DA, Reedy TJ (1986) Choosing near-linear parameters in the four-parameter logistic model for radioligand and related assays. Biometrics 42:575–582

    Article  Google Scholar 

  • Sebaugh JL (2011) Guidelines for accurate EC50/IC50 estimation. Pharm Stat 10:128–134

    Article  Google Scholar 

  • Serfling RJ (1980) Approximation theorems of mathematical statistics. Wiley, New York

    Book  MATH  Google Scholar 

  • Smith RL (1985) Maximum likelihood estimation in a class of nonregular cases. Biometrika 72:67–90

    Article  MATH  MathSciNet  Google Scholar 

  • Smith RL (1994) Nonregular regression. Biometrika 81:173–183

    Article  MATH  MathSciNet  Google Scholar 

  • Stein CM (1981) Estimation of the mean of a multivariate normal distribution. Ann Stat 9:1135–1151

    Article  MATH  Google Scholar 

  • Triantafilis J, Laslett G, McBratney AB (2000) Calibrating an electromagnetic induction instrument to measure salinity in soil under irrigated cotton. Soil Sci Soc Am J 64:1009–1017

    Article  Google Scholar 

  • Varadhan R, Gilbert P (2009) BB: an R package for solving a large system of nonlinear equations and for optimizing a high-dimensional nonlinear objective function. J Stat Softw 32:1–26

    Google Scholar 

  • Vedenov D, Pesti GM (2008) A comparison of methods of fitting several models to nutritional response data. J Ani Sci 86:500–507

    Article  Google Scholar 

  • Vølund A (1978) Application of the four-parameter logistic model to bioassay: comparison with slope ratio and parallel line models. Biometrics 34:357–365

    Article  MATH  Google Scholar 

  • Weiss L, Wolfowitz J (1973) Maximum likelihood estimation of a translation parameter of atruncated distribution. Ann Stat 1:944–947

    Article  MATH  MathSciNet  Google Scholar 

  • Woodroofe M (1972) Maximum likelihood estimation of a translation parameter of a truncated distribution. Ann Math Stat 43:113–122

    Article  MATH  MathSciNet  Google Scholar 

  • Woodroofe M (1974) Maximum likelihood estimation of translationparameter of truncated distribution II. Ann Stat 2:474–488

    Article  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgments

We thank Dr. Valerii Fedorov for suggesting to us the importance of model (3).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nancy Flournoy.

Appendix: Technical details

Appendix: Technical details

1.1 Proof of Proposition 1

Proof

First, following the idea in Sect. 2.3 of Galambos (1978), for any \(t\),

$$\begin{aligned}&\lim _{n\rightarrow \infty } \text {Pr}\left[ \left\{ \frac{\log (B_0-Y)-\log (Y-A_0)-\mu _0}{\sigma _0}\right\} _{(n)}<r_n+s_nt\right] =e^{-e^{-t}}\\&\quad =\lim _{n\rightarrow \infty } \text {Pr}\left\{ \left( \frac{B_0-Y}{Y-A_0}\right) _{(n)}<e^{\mu _0+\sigma _0r_n+\sigma _0s_nt}\right\} \\&\quad =\lim _{n\rightarrow \infty } \text {Pr}\left\{ \frac{B_0-Y_{(1)}}{Y_{(1)}-A_0}<e^{\mu _0+\sigma _0r_n}+\sigma _0s_ne^{\mu _0+\sigma _0r_n}(1+v\sigma _0s_nt)t\right\} \!, \end{aligned}$$

where \(|v|\le 1\). Since \(v\sigma _0s_nt\rightarrow 0\) as \(n\rightarrow \infty \), it follows that

$$\begin{aligned} \lim _{n\rightarrow \infty } \text {Pr}\left( \frac{B_0-Y_{(1)}}{Y_{(1)}-A_0}<c_n+d_nt\right) =e^{-e^{-t}}, \end{aligned}$$

where \(c_n=e^{\mu _0+\sigma _0r_n}\) and \(d_n=\sigma _0s_ne^{\mu _0+\sigma _0r_n}\).

Then, for any \(t\ne 0\),

$$\begin{aligned} \text {Pr}\left( \frac{B_0-Y_{(1)}}{Y_{(1)}-A_0}<c_n+d_nt\right)&=\text {Pr}\left( \frac{Y_{(1)}-A_0}{B_0-A_0}>\frac{1}{1+c_n+d_nt}\right) \\&=\text {Pr}\left[ \frac{Y_{(1)}-A_0}{B_0-A_0}>\frac{c_n-1}{c_n^2}-\left\{ \frac{d_n}{c_n^2}-\frac{(1+d_nt)}{c_n^2(1+c_n+d_nt)t}\right\} t\right] \!. \end{aligned}$$

It can be shown that \([{(1+d_nt)}/\{c_n^2(1+c_n+d_nt)t\}]/({d_n}/{c_n^2})\rightarrow 0\) and \(({1}/{c_n^2})/({d_n}/{c_n^2})\) \(\rightarrow 0\) as \(n\rightarrow \infty \). So from Lemma 2.2.2 in Galambos (1978),

$$\begin{aligned} \lim _{n\rightarrow \infty } \text {Pr}\left\{ \frac{B_0-Y_{(1)}}{Y_{(1)}-A_0}<c_n+d_nt\right\} =\lim _{n\rightarrow \infty } \text {Pr}\left\{ \frac{Y_{(1)}-A_0}{B_0-A_0}>\frac{1}{c_n}-\frac{d_n}{c_n^2}t\right\} \!. \end{aligned}$$

When \(t=0\), the result can be verified by using the properties of the extreme order statistics of normal distribution directly. The second equation can be proved similarly. \(\square \)

1.2 Derivation of the Fisher information

The lemma below is useful in deriving the Fisher information.

Lemma 1

From Lemma 2 of Stein (1981), we obtain that, if \(E|h'(Z)|<\infty \) for a normal random variable \(Z\sim N(\mu ,\sigma ^2)\) and some differentiable function \(h\). Then

$$\begin{aligned} E\{(Z-\mu )h(Z)\}=\sigma ^2E\{ h'(Z)\}. \end{aligned}$$

The log-likelihood function of model (1) based on one observation \((\mathbf{x},Y)\) is

$$\begin{aligned} \ell (\varvec{\theta },\mathbf{x},Y)&= -\frac{\log {(2\pi )}}{2}-\log \sigma +\log (B-A)-\log (Y-A)\\&\,\quad -\log (B-Y)-\frac{\left\{ \log (B-Y)-\log (Y-A)-\mathbf{x}^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right\} ^2}{2\sigma ^2} \end{aligned}$$

for \(Y\in (A, B)\) and 0 otherwise. Let \(Z=\log (B-Y)-\log (Y-A)\). By direct calculation,

$$\begin{aligned} \frac{\partial \ell (\varvec{\theta },\mathbf{x},Y)}{\partial A}&=\frac{1}{Y-A+\frac{(Y-A)^2}{B-Y}}-\frac{\left\{ \log (B-Y)-\log (Y-A)-\mathbf{x}^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right\} }{\sigma ^2(Y-A)};\\ \frac{\partial ^2\ell (\varvec{\theta },\mathbf{x},Y)}{\partial A^2}&=\frac{1+2\frac{Y-A}{B-Y}}{\left\{ Y-A+\frac{(Y-A)^2}{B-Y}\right\} ^2}-\frac{\left\{ \log (B-Y)-\log (Y-A)-\mathbf{x}^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right\} }{\sigma ^2(Y-A)^2}\\&-\frac{1}{\sigma ^2(Y-A)^2}=\frac{e^{2Z}+2e^{Z}}{(B-A)^2}-\frac{(Z-\mathbf{x}^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta })(1+e^{Z})^2}{\sigma ^2(B-A)^2}-\frac{(1+e^{Z})^2}{\sigma ^2(B-A)^2}. \end{aligned}$$

Then from Lemma 1,

$$\begin{aligned} \begin{aligned} E\left\{ \frac{\partial ^2\ell (\varvec{\theta },\mathbf{x},Y)}{\partial A^2}\right\}&=-E\left\{ \frac{e^{2Z}}{(B-A)^2}+\frac{1+2e^{Z}+e^{2Z}}{\sigma ^2(B-A)^2}\right\} \\&=-\frac{e^{2\mathbf{x}^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }+2\sigma ^2}}{(B-A)^2}-\frac{1+2e^{\mathbf{x}^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }+\frac{\sigma ^2}{2}}+e^{2\mathbf{x}^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }+2\sigma ^2}}{\sigma ^2(B-A)^2}. \end{aligned} \end{aligned}$$

Other elements of the Fisher information can be derived similarly.

1.3 Proof of Theorems for the regression model

The proof of Theorem 1 begins with some lemmas.

Lemma 2

For constant sequences \(v_n\downarrow v\) and \(w_n\uparrow w\) as \(n\rightarrow \infty \), let \(\xi _{v_n}\in (v_{n+1},v_n)\) and \(\xi _{w_n}\in (w_n,w_{n+1})\). If a continuous function sequence \(f_n(\cdot )>0\), which is decreasing in \(n\), satisfies \(n^{1+\alpha }f_n(\xi _{v_n})\rightarrow 0\) and \(n^{1+\alpha }f_n(\xi _{w_n})\rightarrow 0\) for \(\alpha >0\) as \(n\rightarrow \infty \), then

$$\begin{aligned} \limsup _n\int \limits _{v_n}^{w_n}f_n(x)\mathrm{d}x<\infty . \end{aligned}$$

Proof

Let \(S_n=\int _{v_n}^{w_n}f_n(x)\mathrm{d}x\). Then

$$\begin{aligned} S_{n}-S_{n-1}&=\int \limits _{v_n}^{w_n}f_n(x)\mathrm{d}x-\int \limits _{v_{n-1}}^{w_{n-1}}f_{n-1}(x)\mathrm{d}x\\&\le (v_{n-1}-v_n)f_{n-1}(\xi _{v_{n-1}})+(w_n-w_{n-1})f_{n-1}(\xi _{w_{n-1}})\\&=\frac{(v_{n-1}-v_n)n^{1+\alpha }f_{n-1}(\xi _{v_{n-1}})+(w_n-w_{n-1})n^{1+\alpha }f_{n-1}(\xi _{w_{n-1}})}{n^{1+\alpha }}\\&=o\left( \frac{1}{n^{1+\alpha }}\right) \!. \end{aligned}$$

So \(\limsup _nS_n=\limsup _n\sum _{i=1}^n(S_n-S_{n-1})\) is finite.

Lemma 3

For any \(\alpha >0\), let \(\delta _n=n^{-\alpha }\). Then for any \(k_1\ge 0\) and \(k_2\ge 0\), there exists a constant \(M\) such that

$$\begin{aligned} \begin{aligned}&\lim _{n\rightarrow \infty }\text {Pr}\left\{ \frac{1}{n}\sum _{i=1}^{n}\frac{|\log (B-Y_i)|^{k_1}}{(Y_i-A)^{k_2}}<M\right\} =1,\\&\lim _{n\rightarrow \infty }\text {Pr}\left\{ \frac{1}{n}\sum _{i=1}^{n}\frac{|\log (Y_i-A)|^{k_1}}{(B-Y_i)^{k_2}}<M\right\} =1 \end{aligned} \end{aligned}$$
(10)

uniformly in \(A\) and \(B\) so long as \(|A-A_0|<\delta _n\) and \(|B-B_0|<\delta _n\).

Proof

We gives the details of proof for the first quantity in (10). The proof for the other one is similar.

$$\begin{aligned} \frac{\left| \log (B-Y_i)\right| ^{k_1}}{(Y_i-A)^{k_2}}&= \frac{|\log (B-B_0+B_0-Y_i)|^{k_1}}{(Y_i-A_0+A_0-A)^{k_2}}I{(Y_i-A_0>2\delta _n,B_0-Y_i>2\delta _n)}+o_p(1)\\&= \frac{|\log (B-B_0+B_0-Y_i)|^{k_1}}{(Y_i-A_0+A_0-A)^{k_2}}I{(Y_i-A_0>2\delta _n,1-2\delta _n>B_0\!-\!Y_i>2\delta _n)}\\&+\,\frac{|\log (B-B_0+B_0-Y_i)|^{k_1}}{(Y_i-A_0+A_0-A)^{k_2}}I{(Y_i-A_0>2\delta _n,B_0-Y_i>1)}+o_p(1)\\&< \frac{I{(Y_i-A_0>2\delta _n,1-2\delta _n>B_0-Y_i>2\delta _n)}}{(B_0-Y_i-\delta _n)^{k_1}(Y_i-A_0-\delta _n)^{k_2}}\\&+\,\frac{(B_0-Y_i+\delta _n)^{k_1}}{(Y_i-A_0-\delta _n)^{k_2}}I{(Y_i-A_0>2\delta _n,B_0-Y_i>1)}+o_p(1)\\&< \frac{I{(B_0-2\delta _n>Y_i>A_0+2\delta _n)}}{(B_0-Y_i-\delta _n)^{k_1}(Y_i-A_0-\delta _n)^{k_2}}\\&+\,\frac{(B_0-A_0+1)^{k_1}}{(Y_i-A_0-\delta _n)^{k_2}}I{\left( B_0-1>Y_i>A_0+2\delta _n\right) }+o_p(1)\\&= C_{in1}+C_{in2}+o_p(1). \end{aligned}$$

Note that

$$\begin{aligned}&(2\pi )^{\frac{1}{2}}E(C_{in1})=\frac{1}{\sigma _0}\int \limits _{A_0+2\delta _n}^{B_0-2\delta _n}\frac{1}{(B_0-y-\delta _n)^{k_1}(y-A_0-\delta _n)^{k_2}}\\&\qquad \times \frac{B_0-A_0}{(B_0-y)(y-A_0)}\exp \left[ -\frac{\left\{ \log \left( \frac{B_0-y}{y-A_0}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_0\right\} ^2}{2\sigma _0^2}\right] \mathrm{d}y\\&\quad \le \frac{1}{\sigma _0}\int \limits _{A_0+2\delta _n}^{B_0-2\delta _n}\frac{1}{(B_0-y-\delta _n)^{k_1+1}(y-A_0-\delta _n)^{k_2+1}}\\&\qquad \times \exp \left[ -\frac{\frac{1}{4}\left\{ \log \left( \frac{B_0-y}{y-A_0}\right) \right\} ^2-(\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_0)^2}{2\sigma _0^2}\right] \mathrm{d}y\\&\quad =\frac{\exp \left\{ \frac{(\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_0)^2}{2\sigma _0^2}\right\} }{\sigma _0}\int \limits _{A_0+2\delta _n}^{B_0-2\delta _n}\frac{1}{(B_0-y-\delta _n)^{k_1+1}(y-A_0-\delta _n)^{k_2+1}}\\&\qquad \times \exp \left[ -\frac{\left\{ \log \left( \frac{B_0-y}{y-A_0}\right) \right\} ^2}{8\sigma _0^2}\right] \mathrm{d}y. \end{aligned}$$

From Lemma 2, \(\limsup _nE(C_{in1})\) is bounded by a finite constant, say \(C_1\). Similarly, it can also be shown that \(\limsup _nE(C_{in2})\) is bounded by a finite constant, say \(C_2\). So using the formula \(X_n=E(X_n)+O_P\{{\text {var}(X_n)^{1/2}}\}\), we have

$$\begin{aligned} \frac{1}{n}\sum _{i=1}^{n}\frac{|\log (B-Y_i)|^{k_1}}{(Y_i-A)^{k_2}}=\frac{1}{n}\sum _{i=1}^nE(C_{in1})+\frac{1}{n}\sum _{i=1}^nE(C_{in2}) + O_P\left( n^{-\frac{1}{2}}\right) + o_P(1). \end{aligned}$$

Thus any \(M\) that is greater than \(C_1+C_2\) satisfys the requirement.

Lemma 4

If assumptions 1- 3 hold, then \(-n^{-1}\partial {^2\ell _n(\varvec{\theta })}/(\partial {\varvec{\theta }\partial \varvec{\theta }^{{ {\mathrm{\scriptscriptstyle T} }}}})\rightarrow \mathcal I (\varvec{\theta }_0)\) in probability uniformly over \(\Vert \varvec{\theta }-\varvec{\theta }_0\Vert <\delta _n\).

Proof

The first element of \({\partial ^2\ell _n(\varvec{\theta })}/({\partial \varvec{\theta }\partial \varvec{\theta }^{{ {\mathrm{\scriptscriptstyle T} }}}})\) is

$$\begin{aligned} \frac{\partial ^2\ell _n(\varvec{\theta })}{\partial A^2} =\sum _{i=1}^n\left[ \frac{1-\frac{1}{\sigma ^2}}{(Y_i-A)^2}-\frac{1}{(B-A)^2}-\frac{\left\{ \log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right\} }{\sigma ^2(Y_i-A)^2}\right] \!. \end{aligned}$$

So it is straightforward to get

$$\begin{aligned}&\frac{1}{n}\left| \frac{\partial ^2\ell _n(\varvec{\theta })}{\partial A^2}-\frac{\partial ^2\ell _n(\varvec{\theta }_0)}{\partial A^2}\right| \\&\quad \le \frac{1}{n}\sum _{i=1}^n\left| \frac{1}{(B-A)^2}-\frac{1}{(B_0-A_0)^2}\right| + \frac{1}{n}\sum _{i=1}^n\left| \frac{1+\frac{1}{\sigma ^2}}{(Y_i-A)^2}-\frac{1+\frac{1}{\sigma _0^2}}{(Y_i-A_0)^2}\right| \\&\quad \quad +\frac{1}{n}\sum _{i=1}^n\left| \frac{\left\{ \log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right\} }{\sigma ^2(Y_i-A)^2}-\frac{\left\{ \log \left( \frac{B_0-Y_i}{Y_i-A_0}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_0\right\} }{\sigma _0^2(Y_i-A_0)^2}\right| \\&\quad ={\varDelta }_1+{\varDelta }_2+{\varDelta }_3. \end{aligned}$$

\({\varDelta }_1\) goes to 0 as \(\delta _n\) goes to 0. By straightforward but tedious calculation, we obtain

$$\begin{aligned} {\varDelta }_3&\le \frac{1}{n}\sum _{i=1}^n\left| \log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right| \left| \frac{1}{\sigma ^2(Y_i-A)^2}-\frac{1}{\sigma _0^2(Y_i-A_0)^2}\right| \\&\quad \quad +\frac{1}{n}\sum _{i=1}^n\left| \frac{\log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }}{\sigma _0^2(Y_i-A_0)^2}-\frac{\log \left( \frac{B_0-Y_i}{Y_i-A_0}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_0}{\sigma _0^2(Y_i-A_0)^2}\right| \\&\le \left| \frac{1}{\sigma ^2}-\frac{1}{\sigma _0^2}\right| \times \frac{1}{n}\sum _{i=1}^n\frac{\left| \log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right| }{(Y_i-A)^2}\\&\quad \quad +\frac{1}{n}\sum _{i=1}^n\frac{\left| \log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right| }{\sigma _0^2}\times \left| \frac{1}{(Y_i-A)^2}-\frac{1}{(Y_i-A_0)^2}\right| \\&\quad \quad +\frac{1}{n}\sum _{i=1}^n\frac{\left| \log \left( \frac{B-Y_i}{B_0-Y_i}\right) \right| }{\sigma _0^2(Y_i-A_0)^2}+ \frac{1}{n}\sum _{i=1}^n\frac{\left| \log \left( \frac{Y_i-A}{Y_i-A_0}\right) \right| }{\sigma _0^2(Y_i-A_0)^2}+ \frac{1}{n}\sum _{i=1}^n\frac{\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}(\varvec{\beta }-\varvec{\beta }_0)}{\sigma _0^2(Y_i-A_0)^2}\\&\le \frac{1}{n}\sum _{i=1}^n\frac{\left| \log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right| }{(Y_i-A)^2}\times \left| \frac{1}{\sigma ^2}-\frac{1}{\sigma _0^2}\right| + \frac{4B|A-A_0|}{\sigma _0^2}\\&\qquad \times \frac{1}{n}\sum _{i=1}^n\frac{\left| \log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right| }{(Y_i-A)^2(Y_i-A_0)^2}+\frac{1}{n}\sum _{i=1}^n\frac{|B-B_0|}{\sigma _0^2(B^*-Y_i)(Y_i-A_0)^2}\\&\quad \quad + \frac{1}{n}\sum _{i=1}^n\frac{|A-A_0|}{\sigma _0^2(Y_i-A^*)(Y_i-A_0)^2}+ \frac{1}{n}\sum _{i=1}^n\frac{\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}(\varvec{\beta }-\varvec{\beta }_0)}{\sigma _0^2(Y_i-A_0)^2}\\&={\varDelta }_{3.1}+{\varDelta }_{3.2}+{\varDelta }_{3.3}+{\varDelta }_{3.4}+{\varDelta }_{3.5}, \end{aligned}$$

where \(A^*\) is between \(A\) and \(A_0\) and \(B^*\) is between \(B\) and \(B_0\). Now we look into each term in the last equation above.

$$\begin{aligned} {\varDelta }_{3.2}\le \frac{2B|A-A_0|}{\sigma _0^2} \times \frac{1}{n}\sum _{i=1}^n\left| \log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right| \left\{ \frac{1}{(Y_i-A)^4}+\frac{1}{(Y_i-A_0)^2}\right\} \!.\nonumber \\ \end{aligned}$$
(11)

The right hand side term in (11) goes to 0 in probability uniformly since the second factor is bound with probability tending to 1 by Lemma 3 and the boundedness of \(x_i\). Similarly, \({\varDelta }_{3.1}, {\varDelta }_{3.3}, {\varDelta }_{3.4}\) and \({\varDelta }_{3.5}\) can be shown to go to 0 in probability uniformly which implies \({\varDelta }_3\) goes to 0 in probability uniformly. Similarly but more easily, \({\varDelta }_1\) and \({\varDelta }_2\) can be showen to converge to 0 in probability uniformly, which implies \(n^{-1}\left| \partial {^2\ell _n(\varvec{\theta })}/\partial {A^2}-\partial {^2\ell _n(\varvec{\theta }_0)}/\partial {A^2}\right| \rightarrow 0\) in probability uniformly. By similiar arguments, other components of \({\partial ^2\ell _n(\varvec{\theta })}/({\partial \varvec{\theta }\partial \varvec{\theta }^{{ {\mathrm{\scriptscriptstyle T} }}}})\) can be shown to have the same property. This implys that \(-n^{-1}\partial {^2\ell _n(\varvec{\theta })}/(\partial \varvec{\theta }\partial \varvec{\theta }^{{ {\mathrm{\scriptscriptstyle T} }}})\rightarrow \mathcal I (\varvec{\theta }_0)\) in probability uniformly over \(\Vert \varvec{\theta }-\varvec{\theta }_0\Vert <\delta _n\). \(\square \)

The following lemma is the Lemma 5 of Smith (1985). We state it for integrity and skip the proof.

Lemma 5

Let \(h\) be a continuously differentiable real-valued function of \(p+1\) real variables and let \(H\) denote the gradient vector of \(h\). Suppose that the scalar product of \(\mathbf{u}\) and \(H(\mathbf{u})\) is negative whenever \(\Vert \mathbf{u}\Vert =1\). Then \(h\) has a local maximum, at which \(H=0\), for some \(\mathbf{u}\) with \(\Vert \mathbf{u}\Vert <1\).

Proof

(of Theorem 1) It suffices to show for any \(\epsilon \), there exists a constant \(c\) such that

$$\begin{aligned} \text {Pr}\left\{ \mathbf{u}^{{ {\mathrm{\scriptscriptstyle T} }}}\frac{\partial \ell _n\left( \varvec{\theta }_0+{n^{-{1}/{2}}c\mathbf{u}}\right) }{\partial \varvec{\theta }}<0\right\} >1-\epsilon \end{aligned}$$
(12)

for any vector \(\mathbf{u}\) such that \(\Vert \mathbf{u}\Vert =1\). Using Taylor’s expansion,

$$\begin{aligned} \frac{\partial \ell _n\left( \varvec{\theta }_0+{n^{-{1}/{2}}c\mathbf{u}}\right) }{\partial \varvec{\theta }}&=\frac{\partial \ell _n\left( \varvec{\theta }_0\right) }{\partial \varvec{\theta }}+cn^{-1/2}\frac{\partial ^2\ell _n\left( \varvec{\theta }_0+{n^{-1/2}c\mathbf{u}^*}\right) }{\partial \varvec{\theta }\partial \varvec{\theta }^{{ {\mathrm{\scriptscriptstyle T} }}}}\mathbf{u}\\&=\frac{\partial \ell _n\left( \varvec{\theta }_0\right) }{\partial \varvec{\theta }}-cn^{1/2}\mathcal I (\varvec{\theta }_0)\mathbf{u}+ n^{1/2}\epsilon _{n,\mathbf{u}}, \end{aligned}$$

where \(\mathbf{u}^*\) is a vector satisfying \(\Vert \mathbf{u}^*\Vert \le 1\) and, by Lemma 3, \(\epsilon _{n,\mathbf{u}}\rightarrow 0\) in probability uniformly over \(\Vert \mathbf{u}\Vert \le 1\) as \(n\rightarrow \infty \). It follows that

$$\begin{aligned} n^{-1/2}\mathbf{u}^{{ {\mathrm{\scriptscriptstyle T} }}}\frac{\partial \ell _n\left( \varvec{\theta }_0+n^{-1/2}\mathbf{u}\right) }{\partial \varvec{\theta }}=n^{-1/2}\mathbf{u}^{{ {\mathrm{\scriptscriptstyle T} }}}\frac{\partial \ell _n\left( \varvec{\theta }_0\right) }{\partial \varvec{\theta }}-c\mathbf{u}^{{ {\mathrm{\scriptscriptstyle T} }}}\mathcal I (\varvec{\theta }_0)\mathbf{u}+ \mathbf{u}^{{ {\mathrm{\scriptscriptstyle T} }}}\epsilon _{n,\mathbf{u}}. \end{aligned}$$
(13)

Note that \(n^{-1/2}\mathbf{u}^{{ {\mathrm{\scriptscriptstyle T} }}}\partial {\ell _n\left( \varvec{\theta }_0\right) }/\partial {\varvec{\theta }}\) is \(O_P(1)\). So the second term dominates the first term in (13) for large enough \(c\). This proves Eq. (12) and the result follows from Lemma 5.

Proof

(of Theorem 2, part 1) For any \(\varvec{\theta }_1\in S\), \(E\ell _n(\varvec{\theta }_1)\!<\!\infty \), so \(E[\ell _n(\varvec{\theta }_1)-\ell _n(\varvec{\theta }_0)]\!<\!0\) by Jensen’s inequality. This implies the existence of \(\xi _{\varvec{\theta }_1}\) such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\text {Pr}\left\{ \ell _n(\varvec{\theta }_1)-\ell _n(\varvec{\theta }_0)<-\xi _{\varvec{\theta }_1}\right\} =1. \end{aligned}$$

For \(\varvec{\theta }\) and \(\eta \) such that \(|\varvec{\theta }-\varvec{\theta }_1|<\eta <|\varvec{\theta }_1-\varvec{\theta }_0|-\delta \),

$$\begin{aligned}&|\ell _n(\varvec{\theta })-\ell _n(\varvec{\theta }_1)|\\&\quad \le |\log \sigma -\log \sigma _1| +\frac{1}{n}\sum _{i=1}^n\left| \log \left( \frac{1}{B-Y_i}+\frac{1}{Y_i-A}\right) -\log \left( \frac{1}{B_1-Y_i}+\frac{1}{Y_i-A_1}\right) \right| \\&\quad \quad +\frac{1}{n}\sum _{i=1}^n\left| \frac{\left\{ \log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right\} ^2}{\sigma ^2}-\frac{\left\{ \log \left( \frac{B_1-Y_i}{Y_i-A_1}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_1\right\} ^2}{\sigma _1^2}\right| \\&\quad ={\varDelta }_4+{\varDelta }_5+{\varDelta }_6. \end{aligned}$$

\({\varDelta }_4\) can be made smaller than \({\xi _{\varvec{\theta }_1}}/{4}\) by choosing \(\eta \) small enough. By the mean value theorem,

$$\begin{aligned} {\varDelta }_5=&\frac{1}{n}\sum _{i=1}^n\left| \frac{1}{B^*-Y_i}\frac{Y_i-A^*}{B^*-A^*}(B-B_1)+\frac{1}{Y_i-A^*}\frac{B^*-Y_i}{B^*-A^*}(A-A_1)\right| \\ \le&\frac{1}{n}\sum _{i=1}^n\left\{ \frac{B_0-A_1+\eta }{B_0-A_0}\frac{|B-B_1|}{B_0-Y_i}+\frac{B_1-A_0+\eta }{B_0-A_0}\frac{|A-A_1|}{Y_i-A_0}\right\} \!, \end{aligned}$$

for some \(A^*\) between \(A_0\) and \(A_1\) and \(B^*\) between \(B_0\) and \(B_1\). So \(E({\varDelta }_5)\) can be make arbitrarily small by choosing small enough \(\eta \), which implies

$$\begin{aligned} \lim _{n\rightarrow \infty }\text {Pr}\left( {\varDelta }_5<\frac{\xi _{\varvec{\theta }_1}}{4}\right) =1 \end{aligned}$$

for small enough \(\eta \).

$$\begin{aligned} {\varDelta }_6&\le \frac{1}{n}\sum _{i=1}^n\left| \frac{\left\{ \log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right\} ^2}{\sigma ^2}-\frac{\left\{ \log \left( \frac{B_1-Y_i}{Y_i-A_1}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_1\right\} ^2}{\sigma ^2}\right| \\&\quad +\frac{1}{n}\sum _{i=1}^n\left\{ \log \left( \frac{B_1-Y_i}{Y_i-A_1}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_1\right\} ^2\left| \frac{1}{\sigma ^2}-\frac{1}{\sigma _1^2}\right| \\&\le \frac{1}{n}\sum _{i=1}^n\left| \frac{\log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }+\log \left( \frac{B_1-Y_i}{Y_i-A_1}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_1}{\sigma ^2}\right| \\&\quad \times \left\{ \frac{|A-A_0|}{A_0-Y_i}+\frac{|B-B_0|}{B_0-Y_i}+|\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }-\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_1|\right\} \\&\quad +\frac{1}{n}\sum _{i=1}^n\left\{ \log \left( \frac{B_1-Y_i}{Y_i-A_1}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_1\right\} ^2\left| \frac{1}{\sigma ^2}-\frac{1}{\sigma _1^2}\right| . \end{aligned}$$

So, for small enough \(\eta \), we obtain

$$\begin{aligned} \lim _{n\rightarrow \infty }\text {Pr}\left( {\varDelta }_6<\frac{\xi _{\varvec{\theta }_1}}{4}\right) =1. \end{aligned}$$

Combining results for \({\varDelta }_4\), \({\varDelta }_5\) and \({\varDelta }_6\),

$$\begin{aligned} \lim _{n\rightarrow \infty }\text {Pr}\left\{ \sup _{|\varvec{\theta }-\varvec{\theta }_1|<\eta }\ell _n(\varvec{\theta })-\ell _n(\varvec{\theta }_0)<-\frac{\xi _{\varvec{\theta }_1}}{4}\right\} =1. \end{aligned}$$

For any compact set \(K, S_\delta \cap K\) can be covered by a finite number of neighborhoods of points in \(S_\delta \), so it follows that

$$\begin{aligned} \lim _{n\rightarrow \infty }\text {Pr}\left\{ \sup _{S_\delta \cap K}\ell _n(\varvec{\theta })-\ell _n(\varvec{\theta }_0)<-\xi _{m}\right\} =1. \end{aligned}$$

Proof

(of Theorem 2, part 2) First, if \(A_0\) and \(B_0\) are known, model (1) can be transformed to a linear model with normal random error with unknown mean and variance. It follows that

$$\begin{aligned} \lim _{n\rightarrow \infty }\text {Pr}\left\{ \sup _{\Vert \varvec{\beta }-\varvec{\beta }_0\Vert >\delta \, |\sigma -\sigma _0|>\delta }\ell _n(\varvec{\beta },\sigma ,A_0,B_0)-\ell _n(\varvec{\theta }_0)<-\xi \right\} =1. \end{aligned}$$
(14)

For \(\varvec{\beta }_1, \sigma _1, \eta \) and \((\varvec{\beta },\sigma ,A,B)\in T\) such that \((\varvec{\beta }_1,\sigma _1, A,B)\in T\), \(\Vert \varvec{\beta }-\varvec{\beta }_1\Vert <\eta \), \(|\sigma -\sigma _1|<\eta \) and \(\delta <\eta \),

$$\begin{aligned}&|\ell _n(\varvec{\beta },\sigma ,A,B)-\ell _n(\varvec{\beta }_1,\sigma _1,A_0,B_0)|\nonumber \\&\quad \le |\log \sigma -\log \sigma _1|+|\log (B-A)-\log (B_0-A_0)|\nonumber \\&\qquad +\frac{1}{n}\sum _{i=1}^n|\log (B-Y_i)-\log (B_0-Y_i)| +\frac{1}{n}\sum _{i=1}^n\left| \log (Y_i-A)-\log (Y_i-A_0)\right| \nonumber \\&\qquad +\frac{1}{2n}\sum _{i=1}^n\left| \frac{\left\{ \log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right\} ^2}{\sigma ^2}-\frac{\left\{ \log \left( \frac{B_0-Y_i}{Y_i-A_0}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_1\right\} ^2}{\sigma _1^2}\right| \nonumber \\&={\varDelta }_7+{\varDelta }_8+{\varDelta }_9+{\varDelta }_{10}+{\varDelta }_{11}. \end{aligned}$$
(15)

The terms \({\varDelta }_7\) and \({\varDelta }_8\) can be made smaller than \({\xi }/{8}\) by choosing \(\eta \) small enough. By the mean value theorem,

$$\begin{aligned} {\varDelta }_9=\frac{1}{n}\sum _{i=1}^n\left| \frac{B-B_0}{B^*-Y_i}\right| \le \frac{|B-B_0|}{n}\sum _{i=1}^n\frac{1}{\min (B,B_0)-Y_i} \end{aligned}$$

with probability tending to 1. If \(B\ge B_0\),

$$\begin{aligned} n^{-1}\sum _{i=1}^n\frac{1}{\left| \min (B,B_0)-Y_i\right| }\le n^{-1}\sum _{i=1}^n\frac{1}{(B_0-Y_i)}, \end{aligned}$$

and the right hand side of the upper inequality goes to the limit of

$$\begin{aligned} \frac{1+n^{-1}\sum _{i=1}^ne^{-\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }+\sigma ^2/2}}{(B_0-A_0)} \end{aligned}$$

in probability. If \(B_0-\delta _n<B<B_0\), Lemma 3 provides that there exists some constant \(M^*\) such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\text {Pr}\left( \frac{1}{n}\sum _{i=1}^n\frac{1}{\left| B-Y_i\right| }<M^*\right) =1 \end{aligned}$$

for small enough \(\eta \). This implies that for small enough \(\eta \),

$$\begin{aligned} \lim _{n\rightarrow \infty }\text {Pr}\left( {\varDelta }_9<\frac{\xi }{8}\right) =1. \end{aligned}$$
(16)

The same result can be found for \({\varDelta }_{10}\) using similar arguments.

$$\begin{aligned} {\varDelta }_{11}\le&\frac{1}{2n}\sum _{i=1}^n\left| \frac{\left\{ \log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }\right\} ^2}{\sigma ^2}-\frac{\left\{ \log \left( \frac{B_0-Y_i}{Y_i-A_0}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_1\right\} ^2}{\sigma ^2}\right| \\&+\frac{1}{2n}\sum _{i=1}^n\left\{ \log \left( \frac{B_0-Y_i}{Y_i-A_0}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_1\right\} ^2\left| \frac{1}{\sigma ^2}-\frac{1}{\sigma _1^2}\right| \\ \le&\frac{1}{2n}\sum _{i=1}^n\left| \frac{\log \left( \frac{B-Y_i}{Y_i-A}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }+\log \left( \frac{B_0-Y_i}{Y_i-A_0}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_1}{\sigma ^2}\right| \times |\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }-\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_1|\\&+\frac{1}{2n}\sum _{i=1}^n\left\{ \log \left( \frac{B_0-Y_i}{Y_i-A_0}\right) -\mathbf{x}_i^{{ {\mathrm{\scriptscriptstyle T} }}}\varvec{\beta }_1\right\} ^2\left| \frac{1}{\sigma ^2}-\frac{1}{\sigma _1^2}\right| . \end{aligned}$$

So we obtain, for small enough \(\eta \),

$$\begin{aligned} \lim _{n\rightarrow \infty }\text {Pr}\left( {\varDelta }_{11}<\frac{\xi }{8}\right) =1. \end{aligned}$$
(17)

Combining (14), (15), (16) and (17), we have

$$\begin{aligned} \lim _{n\rightarrow \infty }\text {Pr}\left\{ \sup \ell _n(a,b,\sigma ,A,B)-\ell _n(\varvec{\theta }_0)<-\frac{3\xi }{8}\right\} =1, \end{aligned}$$

where the supermum is taken over all \(\varvec{\theta }\) satisfying \((\varvec{\beta }_1,\sigma _1,A,B)\in T, \Vert \varvec{\beta }-\varvec{\beta }_1\Vert <\eta \) and \(|\sigma -\sigma _1|<\eta \) for fixed \(\varvec{\beta }_1\) and \(\sigma _1\). This result can be extended directly to any finite set of values of \(\varvec{\beta }_1\) and \(\sigma _1\), and then to any compact sets of values of \(\varvec{\beta }_1\) and \(\sigma _1\).

Proof

(of Theorem 3) By Taylor expansion,

$$\begin{aligned} 0=\frac{\partial \ell _n({\widehat{\varvec{\theta }}}_n)}{\partial \varvec{\theta }}=\frac{\partial \ell _n\left( \varvec{\theta }_0\right) }{\partial \varvec{\theta }}+\frac{\partial ^2\ell _n({\widehat{\varvec{\theta }}}^*)}{\partial \varvec{\theta }\partial \varvec{\theta }^{{ {\mathrm{\scriptscriptstyle T} }}}}({\widehat{\varvec{\theta }}}_n-\varvec{\theta }_0), \end{aligned}$$

where \({\widehat{\varvec{\theta }}}^*\) is between \(\varvec{\theta }_0\) and \({\widehat{\varvec{\theta }}}_n\). From Lemma 4, \(n^{-1}\partial {^2\ell _n({\widehat{\varvec{\theta }}}^*)}/(\partial {\varvec{\theta }\partial \varvec{\theta }^{{ {\mathrm{\scriptscriptstyle T} }}}})\rightarrow -\mathcal I (\varvec{\theta }_0)\) in probability. So

$$\begin{aligned} n^{1/2}({\widehat{\varvec{\theta }}}_n-\varvec{\theta }_0)=\left\{ \mathcal I (\varvec{\theta }_0)\right\} ^{-1}n^{-1/2}\frac{\partial \ell _n\left( \varvec{\theta }_0\right) }{\partial \varvec{\theta }}+o_P(1). \end{aligned}$$
(18)

Note \(n^{-1/2}\partial {\ell _n\left( \varvec{\theta }_0\right) }/\partial {\varvec{\theta }}=n^{-1/2}\sum _{i=1}^n\partial \ell (\varvec{\theta }_0,\mathbf{x}_i,Y_i)/\partial \varvec{\theta }\) is summation of independent random vectors and its variance converges to \(\mathcal I (\varvec{\theta }_0)\). Also we have for \(t>0\),

$$\begin{aligned}&\frac{1}{n}\sum _{i=1}^nE\left[ \left\| \frac{\partial \ell (\varvec{\theta }_0,\mathbf{x}_i,Y_i)}{\partial \varvec{\theta }}\right\| ^2I{\left\{ \left\| \frac{\partial \ell (\varvec{\theta }_0,\mathbf{x}_i,Y_i)}{\partial \varvec{\theta }}\right\| >n^{1/2}\epsilon \right\} }\right] \\&\quad \le \frac{1}{n}\frac{1}{(n^{1/2}\epsilon )^t}\sum _{i=1}^nE\left[ \left\| \frac{\partial \ell (\varvec{\theta }_0,\mathbf{x}_i,Y_i)}{\partial \varvec{\theta }}\right\| ^{2+t}I{\left\{ \left\| \frac{\partial \ell (\varvec{\theta }_0,\mathbf{x}_i,Y_i)}{\partial \varvec{\theta }}\right\| >n^{1/2}\epsilon \right\} }\right] \nonumber \\&\quad \le \frac{1}{n}\frac{1}{(n^{1/2}\epsilon )^t}\sum _{i=1}^nE\left[ \left\| \frac{\partial \ell (\varvec{\theta }_0,\mathbf{x}_i,Y_i)}{\partial \varvec{\theta }}\right\| ^{2+t}\right] \rightarrow 0 \quad \text { as }\quad n\rightarrow \infty . \end{aligned}$$

By the multivariate central limit theorem (cf. Rao 1973; Serfling 1980),

$$\begin{aligned} n^{-1/2}\frac{\partial \ell _n\left( \varvec{\theta }_0\right) }{\partial \varvec{\theta }}\rightarrow N\{0,\mathcal I (\varvec{\theta }_0)\} \end{aligned}$$
(19)

in distribution. Combining (18), (19) and applying Slutsky’s theorem, the result follows.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Wang, H., Flournoy, N. & Kpamegan, E. A new bounded log-linear regression model. Metrika 77, 695–720 (2014). https://doi.org/10.1007/s00184-013-0460-x

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00184-013-0460-x

Keywords

Navigation