Skip to main content
Log in

Identifiability of Asymmetric Circular and Cylindrical Distributions

  • Published:
Sankhya A Aims and scope Submit manuscript

Abstract

Identifiability of statistical models is a fundamental and essential condition that is required to prove the consistency of maximum likelihood estimators. The identifiability of the skew families of distributions on the circle and cylinder for estimating model parameters has not been fully investigated in the literature. In this paper, a new method combining the trigonometric moments and the simultaneous Diophantine approximation is proposed to prove the identifiability of asymmetric circular and cylindrical distributions. Using this method, we prove the identifiability of general sine-skewed circular distributions, including the sine-skewed von Mises and sine-skewed wrapped Cauchy distributions, and that of a Möbius transformed cardioid distribution, which can be regarded as asymmetric distributions on the unit circle. In addition, we prove the identifiability of two cylindrical distributions wherein both marginal distributions of a circular random variable are the sine-skewed wrapped Cauchy distribution, and conditional distributions of a random variable on the non-negative real line given the circular random variable are a Weibull distribution and a generalized Pareto-type distribution, respectively.

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.

Similar content being viewed by others

References

  • Abe, T. (2015). Discussion: on families of distributions with shape parameters. Int. Stat. Rev. 83, 193–197.

    Article  MathSciNet  Google Scholar 

  • Abe, T. and Ley, C. (2017). A tractable, parsimonious and flexible model for cylindrical data, with applications. Econ. Stat. 4, 91–104.

    MathSciNet  Google Scholar 

  • Abe, T. and Pewsey, A. (2011). Sine-skewed circular distributions. Statist. Papers 52, 683–707.

    Article  MathSciNet  MATH  Google Scholar 

  • Abramowitz, M. and Stegun, I. A. (1972). Handbook of mathematical functions with formulas, graphs, and mathematical tables. Dover, New York.

    MATH  Google Scholar 

  • Bhattacharya, A. and Bhattacharya, R. (2008). Nonparametric statistics on manifolds with applications to shape spaces. In Pushing the Limits of Contemporary Statistics: Contributions in Honor of Jayanta K. Ghosh, Inst. Math. Stat. (IMS) Collect., vol 3, Institute Mathematical Statistics, Beachwood, (282–301).

  • Chikuse, Y. (2003). Statistics on special manifolds, Lecture Notes in Statistics, vol 174. Springer, New York.

    Book  MATH  Google Scholar 

  • Dryden, I. L. and Kent, J. T. (2015). Geometry driven statistics, Wiley, Chichester.

  • Frühwirth-Schnatter, S. (2006). Finite mixture and Markov switching models. Springer Series in Statistics. Springer, New York.

    MATH  Google Scholar 

  • Holzmann, H., Munk, A. and Stratmann, B. (2004). Identifiability of finite mixtures–with applications to circular distributions. Sankhyā 66, 440–449.

    MathSciNet  MATH  Google Scholar 

  • Imoto, T., Shimizu, K. and Abe, T. (2019). A cylindrical distribution with heavy-tailed linear part. Jpn. J. Stat. Data Sci. 2, 129–154.

    Article  MathSciNet  MATH  Google Scholar 

  • Jones, M. C. and Pewsey, A. (2012). Inverse Batschelet distributions for circular data. Biometrics 68, 183–193.

    Article  MathSciNet  MATH  Google Scholar 

  • Kato, S. and Jones, M. C. (2010). A family of distributions on the circle with links to, and applications arising from, Möbius transformation. J. Am. Stat. Assoc. 105, 249–262.

    Article  MATH  Google Scholar 

  • Ley, C. and Verdebout, T. (2017a). Modern directional statistics. CRC Press, Boca Raton.

    Book  MATH  Google Scholar 

  • Ley, C. and Verdebout, T. (2017b). Skew-rotationally-symmetric distributions and related efficient inferential procedures. J. Multivariate Anal. 159, 67–81.

    Article  MathSciNet  MATH  Google Scholar 

  • Mardia, K. V. and Jupp, P. E. (2000). Directional statistics. Wiley, Chichester.

    MATH  Google Scholar 

  • Miyata, Y., Shiohama, T. and Abe, T. (2020). Estimation of finite mixture models of skew-symmetric circular distributions. Metrika 83, 895–922.

    Article  MathSciNet  MATH  Google Scholar 

  • Prakasa Rao, B. L. S. (1992). Identifiability in stochastic models. Academic Press, Boston.

    MATH  Google Scholar 

  • Rothenberg, T. J. (1971). Identification in parametric models. Econometrica 39, 577–591.

    Article  MathSciNet  MATH  Google Scholar 

  • Schmidt, W. M. (1991). Diophantine approximations and Diophantine equations. Lecture Notes in Mathematics, vol 1467. Springer, Berlin.

    Google Scholar 

  • Schmidt, W.M. (1996). Diophantine approximation, vol 785. Springer Science & Business Media.

  • Teicher, H. (1963). Identifiability of finite mixtures. Ann. Math. Stat.34, 1265–1269.

    Article  MathSciNet  MATH  Google Scholar 

  • Titterington, D. M., Smith, A F M and Makov, U E (1985). Statistical analysis of finite mixture distributions. Wiley, Chichester.

    MATH  Google Scholar 

  • Umbach, D and Jammalamadaka, S R (2009). Building asymmetry into circular distributions. Stat. Probab. Lett. 79, 659–663.

    Article  MathSciNet  MATH  Google Scholar 

  • Wang, M and Shimizu, K. (2012). On applying Möbius transformation to cardioid random variables. Stat. Methodol. 9, 604–614.

    Article  MathSciNet  MATH  Google Scholar 

  • Yakowitz, S. J. and Spragins, J. D. (1968). On the identifiability of finite mixtures. Ann. Math. Stat. 39, 209–214.

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We thank the editor and an anonymous referee for helpful comments. Yoichi Miyata was supported in part by JSPS KAKENHI Grant Number 19K11863 and the competitive research expenses of Takasaki City University of Economics. Takayuki Shiohama was supported in part by JSPS KAKENHI Grant Numbers 18K01706 and 22K11944 and Nanzan University Pache Research Subsidy I-A-2 for the 2021 and 2022 academic year. Toshihiro Abe was supported in part by JSPS KAKENHI Grant Numbers 19K11869 and 19KK0287.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yoichi Miyata.

Ethics declarations

Conflict of Interest

The authors declare that they have no conflict of interest.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix : Appendix A: Proofs

Appendix A: Proofs

Proof Proof of Theorem 2.1.

We prove this by contradiction. Assume conditions (a) and (b), and that there exist parameters γ1, γ2 such that γ1γ2 and F(x|γ1) = F(x|γ2). By these assumptions, it holds that ϕi(p|γ2)/ϕi(p|γ1) = 1 for every i and any p, which contradicts condition (b). □

Lemma A.1.

For any \(s\in \mathbb {N}\), any \(a_{1},...,a_{s}\in \left [\right .0,2\pi \left .\right )\) and any 𝜖 > 0, there exist infinitely many \(p\in \mathbb {N}\) such that |pai(mod 2π) | < 𝜖 (i = 1,...,s), where “ mod” indicates the modulo operation.

Proof.

Fix any \(s\in \mathbb {N}\) and any 𝜖 > 0. For any ai ∈ [0,2π) (i = 1,...,s), there exist constants ci ∈ [0,2) such that ai = ciπ (i = 1,...,s). From the simultaneous Diophantine approximation (e.g., see Theorem 1A on page 27 of Schmidt (1996)), for any natural number Q with Q ≥ 2, there exist integers q, p1,...,ps such that q ∈{1,...,Qs − 1}, and

$$ \left| c_{i}-\frac{p_{i}}{q}\right| \leq \frac{1}{qQ},\quad (i=1,...,s) $$

which leads to

$$ p\left| c_{i}\pi-\frac{p_{i}}{q}\pi\right| \leq \frac{p\pi}{qQ}, $$
(A.12)

where p is any natural number. For simplicity, we let \(c_{i}^{*}=p_{i}/q\). Then, choosing the number Q with 2πQ− 1/2 < 𝜖 and \(Q^{1/2}\in \mathbb {N}\), and putting p = 2qQ1/2 in Eq. A.12 yields

$$ p\left| c_{i}\pi-c_{i}^{*}\pi\right| \leq \frac{2\pi}{Q^{1/2}}<\epsilon \qquad (i=1,...,s). $$
(A.13)

In addition, we have \(pc_{i}^{*}\pi =2Q^{1/2}p_{i}\pi =0\) (mod 2π). Therefore, because \(pc_{i}\pi =(Q^{1/2}p_{i})(2\pi )+(pc_{i}\pi -pc_{i}^{*}\pi )\) and \(Q^{1/2}p_{i}\in \mathbb {Z}\), it follows from the inequality (A.13) that

$$ \begin{array}{@{}rcl@{}} |pc_{i}\pi \textrm{(mod 2$\pi$)}| &=&|pc_{i}\pi -pc_{i}^{*}\pi | <\epsilon , \end{array} $$

which completes the proof. Note that we can take infinitely many p = 2qQ1/2 by choosing Q successfully because there are infinitely many Q with 2πQ− 1/2 < 𝜖 and \(Q^{1/2}\in \mathbb {N}\), and q ≥ 1. □

Proof of Theorem 2.2.

For different parameters γ1 and γ2, we consider the following three steps: Step 1. ψ1ψ2, Step 2. ψ1 = ψ2, λ1λ2, Step 3. ψ1 = ψ2, λ1 = λ2, and μ1μ2. Then, under each step, we verify the conditions in Theorem 2.1.

Step 1. We consider two parameter vectors γ1 and γ2 with ψ1ψ2, and set ϕ1(p|γ) = ρp(γ)2 where ρp(γ) is defined in Eq. 4. Then, the ratio of ϕ1(p|γ1) to ϕ1(p|γ2) is

$$ \frac{\phi_{1}(p|\boldsymbol{\gamma}_{1})}{\phi_{1}(p|\boldsymbol{\gamma}_{2})}=\frac{\alpha_{0,p}(\boldsymbol{\psi}_{1} )^{2}+({\lambda_{1}^{2}}/4)\left\{ \alpha_{0,p-1}(\boldsymbol{\psi}_{1} )-\alpha_{0,p+1}(\boldsymbol{\psi}_{1} )\right\}^{2}}{\alpha_{0,p}(\boldsymbol{\psi}_{2} )^{2}+({\lambda_{2}^{2}}/4)\left\{ \alpha_{0,p-1}(\boldsymbol{\psi}_{2} )-\alpha_{0,p+1}(\boldsymbol{\psi}_{2} )\right\}^{2}}. $$
(A.14)

By condition (5), there exists a function ep(ψ) of integer p and positive constant M > 0 such that \(\sup _{p\in \mathbb {N}}|e_{p}(\boldsymbol {\psi } )|\leq M\) and

$$ \begin{array}{@{}rcl@{}} \frac{\alpha_{0,p-1}(\boldsymbol{\psi} )-\alpha_{0,p+1}(\boldsymbol{\psi} )}{\alpha_{0,p}(\boldsymbol{\psi} )}=e_{p}(\boldsymbol{\psi} )p^{c}, \end{array} $$

where the constant c is given in condition (iii) of Theorem 2.2.

Hence, Eq. A.14 is expressed as

$$ \begin{array}{@{}rcl@{}} \frac{\alpha_{0,p}(\boldsymbol{\psi}_{1} )^{2}\left\{ 1+({\lambda_{1}^{2}}/4)e_{p}(\boldsymbol{\psi}_{1})^{2}p^{2c} \right\}}{\alpha_{0,p}(\boldsymbol{\psi}_{2} )^{2}\left\{ 1+({\lambda_{2}^{2}}/4)e_{p}(\boldsymbol{\psi}_{2})^{2}p^{2c} \right\}}. \end{array} $$
(A.15)

Because the domain of ϕ1(p|γ) is \(S_{1}(\boldsymbol {\gamma })=\mathbb {Z}\), and the closure of S1(γ) becomes the extended rational number set \(\overline {\mathbb {Z}}\) under the metric d(x,y) described in Section 2. Therefore, \(p\in \mathbb {Z}\) is allowed to increase to infinity.

If condition \(\{ \alpha _{0,p}(\boldsymbol {\psi }_{1})/\alpha _{0,p}(\boldsymbol {\psi }_{2})\} p^{c}\rightarrow 0\) \((p\to \infty )\) in Eq. 6 holds, then the ratio (A.15) is bounded above by

$$ \begin{array}{@{}rcl@{}} \frac{\alpha_{0,p}(\boldsymbol{\psi}_{1} )^{2}}{\alpha_{0,p}(\boldsymbol{\psi}_{2} )^{2}}\left\{ 1+ \frac{M^{2}}{4}p^{2c}\right\} &=&\left( \frac{\alpha_{0,p}(\boldsymbol{\psi}_{1})}{\alpha_{0,p}(\boldsymbol{\psi}_{2} )}p^{c}\right)^{2}\left\{ p^{-2c}+ \frac{M^{2}}{4}\right\} \\ &\rightarrow& 0\quad (p\to\infty). \end{array} $$

On the other hand, if \(\frac {\alpha _{0,p}(\boldsymbol {\psi }_{1})}{\alpha _{0,p}(\boldsymbol {\psi }_{2})}p^{-c}\rightarrow \infty \) \((p\to \infty )\), then the ratio (A.15) is bounded below by

$$ \begin{array}{@{}rcl@{}} \frac{\alpha_{0,p}(\boldsymbol{\psi}_{1} )^{2}}{\alpha_{0,p}(\boldsymbol{\psi}_{2} )^{2}}\frac{1}{ 1+ (M^{2}/4)p^{2c}} &=&\left( \frac{\alpha_{0,p}(\boldsymbol{\psi}_{1})}{\alpha_{0,p}(\boldsymbol{\psi}_{2} )}p^{-c}\right)^{2} \frac{1}{ p^{-2c}+ (M^{2}/4)} \\ &\rightarrow& \infty \quad (p\to\infty). \end{array} $$

Accordingly, under assumption ψ1ψ2, the ratio \(\phi _{1}(p|\boldsymbol {\gamma }_{1})/\phi _{1}(p|\boldsymbol {\gamma }_{2}) \not \rightarrow 1\) \((p\to \infty )\).

Step 2. Next, we consider two parameter vectors γ1 and γ2 with ψ1 = ψ2 = ψ and λ1λ2, and set ϕ2(p|γ) = βp(γ), which is defined in Eq. 3. Then, by Lemma A.1, for possibly different parameters μ1 and μ2 in [0,2π), there exists a sequence \(\{ p_{n}\}_{n\in \mathbb {N}}\) with \(p_{n}\in \mathbb {N}\) such that \(\lim _{n\to \infty }p_{n}=\infty \) and \(\lim _{n\to \infty }|p_{n}\mu _{i} \pmod {2\pi }|=0\) (i = 1,2). By using the sequence \(\{ p_{n}\}_{n\in \mathbb {N}}\), it follows that

$$ \begin{array}{@{}rcl@{}} &&\left| \sin (p_{n}\mu_{i}) \frac{\alpha_{0,p_{n}}(\boldsymbol{\psi}_{i})}{\alpha_{0,p_{n}-1}(\boldsymbol{\psi}_{i} )-\alpha_{0,p_{n}+1}(\boldsymbol{\psi}_{i} )}\right|\\ &\leq& |\sin (p_{n}\mu_{i})|\frac{1}{\inf_{p\in\mathbb{N}}\left| \frac{\alpha_{0,p-1}(\boldsymbol{\psi}_{i} )-\alpha_{0,p+1}(\boldsymbol{\psi}_{i} )}{\alpha_{0,p}(\boldsymbol{\psi}_{i})}\right|} \\ &\leq& M|\sin (p_{n}\mu_{i})| \\ &\rightarrow& 0\quad (n\to\infty ). \end{array} $$
(A.16)

If λ2≠ 0, from Eq. A.16, the ratio of ϕ2(p|γ1) to ϕ2(p|γ2) with p replaced by pn is

$$ \begin{array}{@{}rcl@{}} \frac{\phi_{2}(p_{n}|\boldsymbol{\gamma}_{1})}{\phi_{2}(p_{n}|\boldsymbol{\gamma}_{2})}\!\!\! &&=\frac{\sin (p_{n}\mu_{1} )\alpha_{0,p_{n}}(\boldsymbol{\psi} )\!+\cos (p_{n}\mu_{1} )\lambda_{1} \{ \alpha_{0,p_{n}-1}(\boldsymbol{\psi} ) - \alpha_{0,p_{n}+1}(\boldsymbol{\psi} )\} /2 }{\sin (p_{n}\mu_{2} )\alpha_{0,p_{n}}(\boldsymbol{\psi} )\!+\cos (p_{n}\mu_{2} )\lambda_{2} \{ \alpha_{0,p_{n}-1}(\boldsymbol{\psi} ) - \alpha_{0,p_{n}+1}(\boldsymbol{\psi} )\} /2 } \\ &&=\frac{\sin (p_{n}\mu_{1})\frac{\alpha_{0,p_{n}}(\boldsymbol{\psi} )}{\alpha_{0,p_{n}-1}(\boldsymbol{\psi} )-\alpha_{0,p_{n}+1}(\boldsymbol{\psi} )}+\frac{\cos (p_{n}\mu_{1} )\lambda_{1}}{2} }{\sin (p_{n}\mu_{2} )\frac{\alpha_{0,p_{n}}(\boldsymbol{\psi} )}{\alpha_{0,p_{n}-1}(\boldsymbol{\psi} )-\alpha_{0,p_{n}+1}(\boldsymbol{\psi} )}+\frac{\cos (p_{n}\mu_{2} )\lambda_{2}}{2} } \\ && \rightarrow\frac{\lambda_{1}}{\lambda_{2}}\ne 1\qquad (n\to\infty ). \end{array} $$

If λ2 = 0, then \(|\phi _{2}(p_{n}|\boldsymbol {\gamma }_{1})/\phi _{2}(p_{n}|\boldsymbol {\gamma }_{2})|\to \infty \) as \(n\to \infty \).

Step 3. We consider two parameter vectors γ1 and γ2 with ψ1 = ψ2 = ψ, λ1 = λ2 = λ and μ1μ2. Let i be the imaginary unit defined as i2 = − 1, and set \(\phi _{3}(\boldsymbol {\gamma })=\alpha _{1}(\boldsymbol {\gamma })+\mathrm {i}\beta _{1}(\boldsymbol {\gamma })=\exp (\mathrm {i}\mu )\{ \alpha _{0,1}(\boldsymbol {\psi } )+\mathrm {i}\lambda M_{1}(\boldsymbol {\psi } )\}\) which is the polar form of the mean direction where M1(ψ) = {α0,0(ψ) − α0,2(ψ)}/2. Then, we have

$$ \begin{array}{@{}rcl@{}} &\phi_{3}(\boldsymbol{\gamma}_{1})-\phi_{3}(\boldsymbol{\gamma}_{2})=\left\{ \exp (\mathrm{i}\mu_{1})-\exp (\mathrm{i}\mu_{2}) \right\}\{ \alpha_{0,1}(\boldsymbol{\psi} )+\mathrm{i}\lambda M_{1}(\boldsymbol{\psi} )\} . \end{array} $$

Because |α0,1(ψ) + iλM1(ψ)|≥|α0,1(ψ)| > 0 and \(\exp (\mathrm {i}\mu _{1})-\exp (\mathrm {i}\mu _{2})\ne 0\), we have ϕ3(γ1)≠ϕ3(γ2), which completes the proof. □

Proof Proof of Proposition 2.1.

Condition (i) is obvious. It follows from \(\alpha _{0,p}(\rho )=E_{0,\rho }\{ {\cos \limits } (p{\varTheta } )\} =\rho ^{p}\) that

$$ \begin{array}{@{}rcl@{}} \frac{\alpha_{0,p-1}(\rho )-\alpha_{0,p+1}(\rho )}{\alpha_{0,p}(\rho )}=\frac{1}{\rho}-\rho . \end{array} $$

Thus, conditions (ii) and (iii) hold with c = 0, which completes the proof. □

To prove Proposition 2.2, we present a lemma on the modified Bessel function.

Lemma A.2.

The following results hold:

$$ \begin{array}{@{}rcl@{}} &&\frac{1}{p!}\left( \frac{\kappa}{2}\right)^{p}\leq I_{p}(\kappa )\leq \frac{1}{p!}\left( \frac{\kappa}{2}\right)^{p}\exp\left( \frac{\kappa^{2}}{4}\right) ,\textrm{ and }\\ &&\frac{A_{p}(\kappa )}{A_{p}(\kappa^{\prime})}=O\left( \left( \frac{\kappa}{\kappa^{\prime}}\right)^{p}\right) \textrm{ for }\kappa <\kappa^{\prime}\quad (p\to\infty ), \end{array} $$
(A.17)

where Ap(κ) = Ip(κ)/I0(κ).

Proof.

Using the expansion in Mardia and Jupp (2000, p.349), we have

$$ \begin{array}{@{}rcl@{}} I_{p}(\kappa )&=&\sum\limits_{r=0}^{\infty}\frac{1}{(p+r)!r!}\left( \frac{\kappa}{2}\right)^{2r+p} \\ &=&\frac{1}{p!}\left( \frac{\kappa}{2}\right)^{p}\left\{ 1+\frac{1}{(p+1)1!}\left( \frac{\kappa}{2}\right)^{2}+\frac{1}{(p+2)(p+1)2!}\left( \frac{\kappa}{2}\right)^{4}+{\cdots} \right\} \\ &\leq& \frac{1}{p!}\left( \frac{\kappa}{2}\right)^{p}\left\{ 1+\frac{1}{1!}\left( \frac{\kappa^{2}}{4}\right) +\frac{1}{2!}\left( \frac{\kappa^{2}}{4}\right)^{2}+\frac{1}{3!}\left( \frac{\kappa^{2}}{4}\right)^{3}+{\cdots} \right\} \\ &=&\frac{1}{p!}\left( \frac{\kappa}{2}\right)^{p}\exp\left( \frac{\kappa^{2}}{4}\right) , \end{array} $$
(A.18)

where the last equation holds from the Maclaurin expansion for ex. On the one hand, by Eq. A.18, we have the left side inequality in Eq. A.17. Next, we prove the latter. Let \(\kappa <\kappa ^{\prime }\). Because

$$ \frac{A_{p}(\kappa )}{A_{p}(\kappa^{\prime})}=\frac{I_{0}(\kappa^{\prime})I_{p}(\kappa)}{I_{0}(\kappa )I_{p}(\kappa^{\prime})}, $$

applying the inequality (A.17) to Ip(κ) and \(I_{p}(\kappa ^{\prime } )\) yields the result. □

Proof Proof of Proposition 2.2.

We verify conditions (i)–(iii) in Theorem 2.2 to demonstrate the identifiability of the SSvM distribution. Because the p th cosine-moment is given by \(\alpha _{0,p}:=E_{0,\kappa }\{ {\cos \limits } (p{\varTheta } )\} =I_{p}(\kappa )/I_{0}(\kappa )\), condition (i) holds. In addition, we have

$$ \begin{array}{@{}rcl@{}} \frac{\alpha_{0,p-1}(\kappa )-\alpha_{0,p+1}(\kappa )}{\alpha_{0,p}(\kappa )}&=&\frac{I_{p-1}(\kappa )-I_{p+1}(\kappa )}{I_{p}(\kappa )}. \end{array} $$
(A.19)

By using the result Iν− 1(z) − Iν+ 1(z) = (2ν/z)Iν(z), which is given in Abramowitz and Stegun (1972, p.376, 9.6.26), Eq. A.19 equals (2p)/κ. Therefore, conditions (ii) and Eq. 5 in (iii) hold with c = 1.

Finally, we verify Eq. 6 in condition (iii). It follows from Lemma A.2 that

$$ \begin{array}{@{}rcl@{}} \frac{\alpha_{0,p}(\kappa_{1})}{\alpha_{0,p}(\kappa_{2})}=\frac{I_{p}(\kappa_{1} )/I_{0}(\kappa_{1} )}{I_{p}(\kappa_{2})/I_{0}(\kappa_{2})}=O\left( \left( \frac{\kappa_{1}}{\kappa_{2}}\right)^{p}\right) . \end{array} $$

Thus, if κ1 > κ2, then \((\alpha _{0,p}(\kappa _{1})/\alpha _{0,p}(\kappa _{2}))p^{-c}\rightarrow \infty \) \((p\to \infty )\). By contrast, if κ1 < κ2, then \((\alpha _{0,p}(\kappa _{1})/\alpha _{0,p}(\kappa _{2}))p^{c}\rightarrow 0\) \((p\to \infty )\), which implies condition (iii). Hence, the result is proved. □

Proof of Proposition 2.3.

For different parameters γWS1 = (μ1,ρα1, \(\bar {\rho }_{1},\xi _{1})^{T}\) and \(\boldsymbol {\gamma }_{\textsc {WS2}}=(\mu _{2},\rho _{\alpha 2},\bar {\rho }_{2},\xi _{2})^{T}\), we consider the following four steps: Step 1: ρα1ρα2, Step 2: ρα1 = ρα2 and \(\bar {\rho }_{1}\ne \bar {\rho }_{2}\), Step 3: ρα1 = ρα2, \(\bar {\rho }_{1}=\bar {\rho }_{2}\), and ξ1ξ2. Step 4: ρα1 = ρα2, \(\bar {\rho }_{1}=\bar {\rho }_{2}\), ξ1 = ξ2, and μ1μ2.

First, we consider Step 1 with ρα1ρα2, and set ϕ1(p|γWS) = αp(γWS) which is the cosine moment. If ρα1 > ρα2, then ϕ1(p|γWS1)/ϕ1(p|γWS2) becomes

$$ \begin{array}{@{}rcl@{}} &&\frac{p\bar{\rho}_{1}\rho_{\alpha 1}^{p-1}(1-\rho_{\alpha 1}^{2})\cos (p\mu_{1} +\xi_{1} )+\rho_{\alpha 1}^{p}\cos (p\mu_{1})}{p\bar{\rho}_{2}\rho_{\alpha 2}^{p-1}(1-\rho_{\alpha 2}^{2})\cos (p\mu_{2} +\xi_{2} )+\rho_{\alpha 2}^{p}\cos (p\mu_{2})} \\ &=&\frac{p\bar{\rho}_{1}(\rho_{\alpha 1}/\rho_{\alpha 2})^{p-1}(1-\rho_{\alpha 1}^{2})\cos (p\mu_{1} +\xi_{1} )+(\rho_{\alpha 1}/\rho_{\alpha 2})^{p-1}\rho_{\alpha 1}\cos (p\mu_{1})/p}{\bar{\rho}_{2}(1-\rho_{\alpha 2}^{2})\cos (p\mu_{2} +\xi_{2} )+\rho_{\alpha 2}\cos (p\mu_{2})/p} \\ &\rightarrow& \infty \qquad (p\to \infty ). \end{array} $$

The case when ρα1 < ρα2 is proved similarly.

Next, we consider Step 2 with ρα1 = ρα2 = ρα and \(\bar {\rho }_{1}\ne \bar {\rho }_{2}\), and set ϕ2(p|γWS) to the squared p th mean resultant length, that is ϕ2(p|γWS) = ρp(γWS)2. When \(\bar {\rho }_{1}>\bar {\rho }_{2}\), then

$$ \begin{array}{@{}rcl@{}} \frac{\phi_{2} (p|\boldsymbol{\gamma}_{\textsc{WS1}})}{\phi_{2} (p|\boldsymbol{\gamma}_{\textsc{WS2}})}&=&\frac{p^{2}\bar{\rho}_{1}^{2}\rho_{\alpha}^{2(p-1)}(1-\rho_{\alpha}^{2})^{2}+\rho_{\alpha}^{2p}+2p\bar{\rho}_{1}\rho_{\alpha}^{2(p-1)+1}(1-\rho_{\alpha}^{2})\cos \xi_{1}}{p^{2}\bar{\rho}_{2}^{2}\rho_{\alpha}^{2(p-1)}(1-\rho_{\alpha}^{2})^{2}+\rho_{\alpha}^{2p}+2p\bar{\rho}_{2}\rho_{\alpha}^{2(p-1)+1}(1-\rho_{\alpha}^{2})\cos \xi_{2}} \\ &=&\frac{\bar{\rho}_{1}^{2}(1-\rho_{\alpha}^{2})^{2}+\rho_{\alpha}^{2}/p^{2}+2\bar{\rho}_{1}\rho_{\alpha}(1-\rho_{\alpha}^{2})(\cos \xi_{1})/p}{\bar{\rho}_{2}^{2}(1-\rho_{\alpha}^{2})^{2}+\rho_{\alpha}^{2}/p^{2}+2\bar{\rho}_{2}\rho_{\alpha}(1-\rho_{\alpha}^{2}) (\cos \xi_{2})/p} \\ &\rightarrow& \frac{\bar{\rho}_{1}^{2}}{\bar{\rho}_{2}^{2}}>1 \qquad (p\to\infty ). \end{array} $$

The case when \(\bar {\rho }_{1}<\bar {\rho }_{2}\) can also be verified similarly.

Next, we consider Step 3 with ρα1 = ρα2 = ρα, \(\bar {\rho }_{1}=\bar {\rho }_{2}=\bar {\rho }\), ξ1ξ2. ϕ3(p|γWS) is set to the characteristic function αp(γWS) + iβp(γWS). For simplicity of notation, let \(B_{p}=p\bar {\rho }\rho _{\alpha }^{p-1}(1-\rho _{\alpha }^{2})\). If ξ1ξ2, then we have

$$ \begin{array}{@{}rcl@{}} &&\frac{\alpha_{p}(\boldsymbol{\gamma}_{\textsc{WS1}})}{\alpha_{p}(\boldsymbol{\gamma}_{\textsc{WS2}})} \\ &=&\frac{\cos (p\mu_{1} + \xi_{1} ) + (\rho_{\alpha}^{p}/B_{p})\cos (p\mu_{1}) + \mathrm{i}\left\{ \sin (p\mu_{1} +\xi_{1} )+(\rho_{\alpha}^{p}/B_{p})\sin (p\mu_{1})\right\}}{\cos (p\mu_{2} + \xi_{2} ) + (\rho_{\alpha}^{p}/B_{p})\cos (p\mu_{2}) + \mathrm{i}\left\{ \sin (p\mu_{2} +\xi_{2} )+(\rho_{\alpha}^{p}/B_{p})\sin (p\mu_{2})\right\}}.\\ \end{array} $$
(A.20)

By Lemma A.1, there exists a sequence {pn} of natural numbers such that \(p_{n}\to \infty \) and |pμi(mod2π)|→ 0(i = 1,2) as \(n\to \infty \). Hence, replacing p with pn in Eq. A.20 and taking the limit leads to

$$ \begin{array}{@{}rcl@{}} \frac{\phi_{3} (p_{n}|\boldsymbol{\gamma}_{\textsc{WS1}})}{\phi_{3} (p_{n}|\boldsymbol{\gamma}_{\textsc{WS2}})} \rightarrow \frac{\cos (\xi_{1})+\mathrm{i}\sin (\xi_{1})}{\cos (\xi_{2})+\mathrm{i}\sin (\xi_{2})} \ne 1 ,\qquad (n\to \infty ), \end{array} $$

for which the condition is verified under Step 3.

Finally, we consider Step 4 with ρα1 = ρα2 = ρα, \(\bar {\rho }_{1}=\bar {\rho }_{2}=\bar {\rho }\), ξ1 = ξ2 = ξ and μ1μ2. Here, we set ϕ4(p|γWS) = ϕ3(p|γWS). Then,

$$ \begin{array}{@{}rcl@{}} \phi_{4} (p|\boldsymbol{\gamma}_{\textsc{WS}})\!\!&=&\!B_{p}\cos (p\mu +\xi )+\rho_{\alpha}^{p}\cos (p\mu )+\mathrm{i}\left\{ B_{p}\sin (p\mu +\xi )+\rho_{\alpha}^{p}\sin (p\mu ) \right\} \\ \!&=&\!B_{p}\exp \{ \mathrm{i}(p\mu +\xi )\} +\rho_{\alpha}^{p}\exp (\mathrm{i}p\mu ). \end{array} $$

Hence, we have

$$ \begin{array}{@{}rcl@{}} &&\phi_{4} (p|\boldsymbol{\gamma}_{\textsc{WS1}})-\phi_{4} (p|\boldsymbol{\gamma}_{\textsc{WS2}}) \\ &=&B_{p}\left\{ \exp (\mathrm{i}(p\mu_{1}+\xi ))-\exp (\mathrm{i}(p\mu_{2}+\xi ))\right\} +\rho_{\alpha}^{p}\left( \exp (\mathrm{i}p\mu_{1})-\exp (\mathrm{i}p\mu_{2})\right) \\ &=&\left( B_{p}\exp (\mathrm{i}\xi)+\rho_{\alpha}^{p}\right) \left\{ \exp (\mathrm{i}p\mu_{1})-\exp (\mathrm{i}p\mu_{2})\right\} . \end{array} $$

Because \(|B_{p}\exp (\mathrm {i}\xi )+\rho _{\alpha }^{p}|^{2}={B_{p}^{2}}+2B_{p}\rho _{\alpha }^{p}{\cos \limits } \xi +\rho _{\alpha }^{2p}\geq (B_{p}-\rho _{\alpha }^{p})^{2}\) and \(B_{p}-\rho _{\alpha }^{p}>0\) for some large p, we have ϕ4(p|γWS1) − ϕ4(p|γWS2)≠ 0 for some large p, which the condition is verified under Step 4. Therefore, the family of distributions (8) is identifiable under ΓWS. □

Proof of Proposition 2.4.

Because the marginal distribution of Θ is the SSWC, from Proposition 2.1 and Remark 2.2, it is identifiable with respect to the parameter vector (μ,κ,λ)T. Hence, we begin with the case of (μ1,κ1,λ1)T = (μ2,κ2,λ2)T(= (μ,κ,λ)T). Set \(g(\alpha ,\beta )=\beta (1-\tanh (\kappa )\cos \limits (\theta -\mu ))^{1/\alpha }\). It follows from the identifiability of Weibull distribution that fAL(x|𝜃;α1,β1,κ) = fAL(x|𝜃;α2,β2,κ) implies α1 = α2 and g(α1,β1) = g(α2,β2), which leads to β1 = β2. This means that for any (α1,β1)T≠(α2,β2)T, fAL(x|𝜃;α1,β1,κ)≠fAL(x|𝜃;α2,β2,κ), which completes the proof. □

Proof of Proposition 2.5.

Because the marginal distribution for Θ in Eq. 9 is the SSWC distribution, the identifiability with respect to the parameter vector (μ,κ,λ)T holds from Proposition 2.1 and Remark 2.2. The conditional distribution of the linear part of Eq. 9 is given in Eq. 11. We need to check the case with (μ1,κ1)T = (μ2,κ2)T = (μ,κ)T and δ1δ2. First, we assume δ2 > δ1. In this case, the ratio on the conditional distribution of the linear part becomes

$$ \begin{array}{@{}rcl@{}} &&\frac{f_{X|{\varTheta}}(x | \theta; \delta_{1}, \tau_{1}, \sigma_{1})}{f_{X|{\varTheta}}(x | \theta; \delta_{2}, \tau_{2}, \sigma_{2})}\\ &=& \frac{ \frac{1}{\sigma_{1} \delta_{1}} \!\left( \!\frac{x}{\sigma_{1}}\!\right)^{1/\delta_{1} -1} \!\left\{ 1 - \kappa \cos(\theta - \mu) \right\} \!\!\left[1 \!+ \frac{\tau_{1}}{\delta_{1}} \!\left( \frac{x}{\sigma_{1}}\right)^{1/\delta_{1}} \{1 - \kappa \cos(\theta - \mu) \}\right]^{-(\delta_{1}/\tau_{1} + 1)} }{ \frac{1}{\sigma_{2} \delta_{2}} \!\left( \!\frac{x}{\sigma_{2}}\!\right)^{1/\delta_{2} -1} \!\left\{ 1 - \kappa \cos(\theta - \mu) \right\} \!\!\left[1 \!+ \frac{\tau_{2}}{\delta_{2}} \!\left( \frac{x}{\sigma_{2}}\right)^{1/\delta_{2}} \{1 - \kappa \cos(\theta - \mu) \}\right]^{-(\delta_{2}/\tau_{2} + 1)} } \\ &=& C_{1} x^{1/\delta_{1} - 1/\delta_{2}} \frac{ \left[1 + \frac{\tau_{2}}{\delta_{2}} \left( \frac{x}{\sigma_{2}}\right)^{1/\delta_{2}} \{1 - \kappa \cos(\theta - \mu) \}\right]^{(\delta_{2}/\tau_{2} + 1)} } {\left[1 + \frac{\tau_{1}}{\delta_{1}} \left( \frac{x}{\sigma_{1}}\right)^{1/\delta_{1}} \{1 - \kappa \cos(\theta - \mu) \}\right]^{(\delta_{1}/\tau_{1} + 1)}} =: K_{1}, \end{array} $$

where \(C_{1} = \sigma _{2}^{1/\delta _{2}}\delta _{2} / (\sigma _{1}^{1/\delta _{1}}\delta _{1} )\). Since 1/δ1 − 1/δ2 > 0, we observe that K1 → 0 as x → 0. We can similarly observe that for the case with δ2 < δ1, \(K_{1} \to \infty \) as x → 0. For the next step, we consider the case with (μ1,κ1,δ1)T = (μ2,κ2,δ2)T = (μ,κ,δ)T and τ1τ2. To simplify expressions, we let 1/τ2 − 1/τ1 = Δτ and \(C_{2}(\theta )=1 - \kappa \cos \limits (\theta - \mu )\). From the expression K1 and the fact that (δ/τ2) + 1 = (δ/τ1) + 1 + Δτδ, we see that

$$ \begin{array}{@{}rcl@{}} && \frac{f_{X|{\varTheta}}(x | \theta; \delta_{1}, \tau_{1}, \sigma_{1})}{f_{X|{\varTheta}}(x | \theta; \delta_{2}, \tau_{2}, \sigma_{2} )} = C_{1} \frac{ \left[1 + \frac{\tau_{2}}{\delta} \left( \frac{x}{\sigma_{2}}\right)^{1/\delta} C_{2}(\theta)\right]^{(\delta/\tau_{2} + 1)} } {\left[1 + \frac{\tau_{1}}{\delta} \left( \frac{x}{\sigma_{1}}\right)^{1/\delta} C_{2}(\theta)\right]^{(\delta/\tau_{1} + 1)}} \\ &=& C_{1} \left\{ \frac{ 1 + \frac{\tau_{2}}{\delta} \left( \frac{x}{\sigma_{2}}\right)^{1/\delta}C_{2}(\theta) } { 1 + \frac{\tau_{1}}{\delta} \left( \frac{x}{\sigma_{1}}\right)^{1/\delta}C_{2}(\theta) } \right\}^{\delta/\tau_{1}+1} \left[ 1+ \frac{\tau_{1}}{\delta}\left( \frac{x}{\sigma_{2}} \right)^{1/\delta} C_{2}(\theta) \right]^{{\Delta}_{\tau}}\\ &=:& C_{1}\times K_{2} \times K_{3} . \end{array} $$

We see that K2 = O(1) as \(x\to \infty \). For τ2 < τ1 which indicates Δτ > 0, we have \(K_{3} \to \infty \) as \(x \to \infty \). On the contrary, when τ2 > τ1 which indicates Δτ < 0, we have K3 → 0 as \(x \to \infty \). For the final step, we consider the case with (μ1,κ1,δ1,τ1)T = (μ2,κ2,δ2,τ2)T = (μ,κ,δ,τ)T and σ1σ2. We observe that

$$ \begin{array}{@{}rcl@{}} \frac{f_{X|{\varTheta}}(x | \theta; \delta, \tau, \sigma_{1})}{f_{X|{\varTheta}}(x | \theta; \delta, \tau, \sigma_{2})} = \frac{\sigma_{2}^{1/\delta}\delta }{\sigma_{1}^{1/\delta}\delta } \frac{ \left[1 + \frac{\tau}{\delta} \left( \frac{x}{\sigma_{2}}\right)^{1/\delta} C_{2}(\theta)\right]^{(\delta/\tau + 1)} } {\left[1 + \frac{\tau}{\delta} \left( \frac{x}{\sigma_{1}}\right)^{1/\delta} C_{2}(\theta)\right]^{(\delta/\tau + 1)}} . \end{array} $$

Evaluating the ratio of the conditional distributions at x = 0, we have

$$ \begin{array}{@{}rcl@{}} \frac{f_{X|{\varTheta}}(0 | \theta; \delta, \tau, \sigma_{1})}{f_{X|{\varTheta}}(0| \theta; \delta, \tau, \sigma_{2} ) } = \frac{\sigma_{2}^{1/\delta}}{\sigma_{1}^{1/\delta}} \ne 1, \end{array} $$

which completes the proof of Proposition 2.5. □

Appendix B: Derivation of p th Cosine and Sine Moments

In this section, we describe how to represent p th cosine and sine moments in a circular model with location μ by using p th cosine and sine moments in the circular model with μ = 0. We consider the density Eq. 2 with μ = 0,

$$ \begin{array}{@{}rcl@{}} f^{*}(\theta |\boldsymbol{\gamma}^{*})&:=&f_{0}(\theta |\boldsymbol{\psi} )\left\{ 1+\lambda \sin (\theta )\right\} , \end{array} $$

where γ = (ψT,λ)T. In addition, we denote its p th cosine and sine moments as \(\alpha _{p}^{*}(\boldsymbol {\gamma }^{*}):=E^{*}\{{\cos \limits } (p{\varTheta })\}\) and \(\beta _{p}^{*}(\boldsymbol {\gamma }^{*}):=E^{*}\{{\sin \limits } (p{\varTheta })\}\), \((p\in \mathbb {Z})\). Because the density Eq. 2 is written as f(𝜃|γ) = f(𝜃μ|γ), the p th cosine moment in the density Eq. 2 is given by

$$ \begin{array}{@{}rcl@{}} \alpha_{p}(\boldsymbol{\gamma})&=&{\int}_{\mu-\pi}^{\mu+\pi}\cos (p\theta )f^{*}(\theta -\mu |\boldsymbol{\gamma}^{*}) d\theta . \end{array} $$
(A.21)

Letting 𝜃μ = η, Eq. A.21 becomes

$$ \begin{array}{@{}rcl@{}} &&{\int}_{-\pi}^{\pi}\cos (p(\mu +\eta ))f^{*}(\eta |\boldsymbol{\gamma}^{*})d\eta \\ &=&{\int}_{-\pi}^{\pi}\left\{ \cos (p\mu )\cos(p\eta )-\sin (p\mu )\sin (p\eta )\right\}f^{*}(\eta |\boldsymbol{\gamma}^{*}) d\eta \\ &=&\cos (p\mu )E^{*}\left\{ \cos (p\eta )\right\} -\sin (p\mu)E^{*}\left\{ \sin (p\eta )\right\} \\ &=&\cos (p\mu )\alpha_{p}^{*}(\boldsymbol{\gamma}^{*})-\sin (p\mu )\beta_{p}^{*}(\boldsymbol{\gamma}^{*}). \end{array} $$

Similarly, the p th sine moment is

$$ \begin{array}{@{}rcl@{}} \beta_{p}(\boldsymbol{\gamma})&=&{\int}_{-\pi}^{\pi}\sin (p(\mu +\eta ))f^{*}(\eta |\boldsymbol{\gamma}^{*})d\eta \\ &=&{\int}_{-\pi}^{\pi}\left\{ \sin (p\mu )\cos (p\eta )+\cos (p\mu )\sin (p\eta )\right\} f^{*}(\eta |\boldsymbol{\gamma}^{*})d\eta \\ &=&\sin (p\mu )E^{*}\left\{ \cos (p\eta )\right\} +\cos (p\mu)E^{*}\left\{ \sin (p\eta )\right\} \\ &=&\sin (p\mu )\alpha_{p}^{*}(\boldsymbol{\gamma}^{*}) +\cos (p\mu)\beta_{p}^{*}(\boldsymbol{\gamma}^{*}). \end{array} $$

This result shows that if we know the p th cosine and sine moments when μ = 0, we can automatically derive the p th cosine and sine moments of the circular model with location μ.

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Miyata, Y., Shiohama, T. & Abe, T. Identifiability of Asymmetric Circular and Cylindrical Distributions. Sankhya A 85, 1431–1451 (2023). https://doi.org/10.1007/s13171-022-00294-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13171-022-00294-3

Keywords and phrases

AMS (2000) subject classification

Navigation