Skip to main content
PLOS One logoLink to PLOS One
. 2025 Aug 22;20(8):e0329274. doi: 10.1371/journal.pone.0329274

Generalised linear regression GARMA model adopted in Denmark’s tourism industry

Hongxuan Yan 1, Xingyu Yan 2,3, Luoyi Sun 2,*
Editor: Mohamed R Abonazel4
PMCID: PMC12373212  PMID: 40845043

Abstract

This paper investigates the characteristics of seasonality in the tourism industry. The Gegenbauer long memory and seasonal features are clearly clarified in Denmark’s tourism data. By plotting ACF and periodogram graphs, the pattern of long memory is investigated. A generalised linear regression GARMA (GLRGARMA) model and a generalised linear regression SARMA (GLRSARMA) model with an innovative function of explanatory variables is proposed to capture data features. Furthermore, the generalised Poisson (GP) distribution with over- equal- and under-dispersion is adopted to improve model flexibility. Eight sub-models are implemented with the number of rented hotel rooms data set to explore the best-performed model structure. The Bayesian approach is adopted to implement in-sample fitting and out-of-sample forecast studies. Several model selection criteria are adopted to evaluate model performances. Overall, GLRGARMA model is the best model to handle the time series with Gegenbauer long memory feature, especially in the tourism area. The explanatory variable with the periodic sponge effect will dramatically enhance model performances.

Introduction

Tourism has become a vital economic pillar for nations globally. As consumers of goods and services, travelers drive demand that stimulates multiple industries, yielding employment opportunities, foreign currency inflows, and infrastructure development for host countries. Tourism stimulates the economic growth from different aspects, including directly boosting economic units from the tourism industry (housing, food, transportation, etc), creating new employment opportunities, bringing important revenues to the State budget in the form of taxes and fees and enhancing the developments of other sectors engaging in the accomplishment of the tourism product [1]. According to the statistical data from the WTO data for Tourism Sector including passenger Transportation Services (excluding freight), Travel Services, and the recreation portion of the Other Commercial Services sector [2], the proportion of the Tourism Sector is the sixth largest sector of the global economy and the largest Service Sector industry in the world [3]. Furthermore, the impacts of the development of tourism on industrial production have drawn great attention. [4] and [5] claimed that growing tourism will cause de-industralization, whereas [6] found that there is no evident negative impact of the development of tourism on the manufacturing industry in Thailand. Besides, tourism and hospitality increase the number of available jobs [7]. Consequently, tourism, a significant contributor to economic growth, needs to be investigated to reveal the dynamic mechanisms of tourism development through modeling and forecasting. The characteristics of development paths in tourism areas have also been underexplored.

Seasonality as a common characteristic has been widely found in the tourism industry [8]. Seasonality in tourism has traditionally been regarded as a major problem that needs to be overcome. [9] studied the characteristics of seasonality and developed a methodology to study this phenomenon.Recent evidence further highlights that seasonal variations in environmental conditions directly elevate pedestrian-accident risks in tourist areas by 18–24% [10]. [11] proposed quantitative solutions via financial portfolio theory to assist marketers in mitigating seasonal effects. [12] attempt to provide a rational framework for tourism seasonality by analysing the main characteristics of these challenges. Furthermore, in terms of the quantitative analysis method, [13] investigated the causalities of seasonality using a mixed effects panel data model for the main tourist destinations in the world. [14] proposes a new index to measure the seasonality in tourism by analysing the pattern of seasonal swings, including seasonal amplitude and similarity. This method focuses on the ordinal and cyclical structures of seasonal variations.The application of this methodology revealed a statistically significant association between seasonal variations and spatial distribution patterns across European countries. [15] found a seasonal phenomenon describing wane and wax shifting between the industry sector and the service sector. This phenomenon, named the periodic sponge effect, plays a vital role in economic development and addressing the unemployment issue. Moreover, they further propose a periodic sponge effect index that describes a reversed cyclical relationship between two time series. This index is a quantitative measure to define the strength of a periodic sponge effect accurately.

In terms of seasonal time series models in the tourism area, [16] studied ARIMA modeling seasonality in tourism forecasting with two settings, such as one for modeling stochastic nonstationary seasonality and another for a constant seasonality with three seasonal dummies. Moreover, the out-of-sample forecasting performances were evaluated. [17] argued that the simple deterministic model with seasonal dummy variables and AR(1) disturbances have better forecast performance. [18] compared the performance of various econometric time-series models in forecasting seasonal tourism demand and found that the methods of seasonality treatment, such as the pre-test for seasonal unit roots, affect the forecasting performance of the models. [19] forecasted tourism demand with ARMA-based methods. [20] adopted the seasonal ARIMA model to forecast the monthly outbound tourism departures. [21] investigated the performance of combination forecasts in international tourism demand. [22] proposed a SARIMA model with non-linear methods to forecast a seasonal tourist with a structural break in the data. [23] evaluated the performance of the Holt-Winters and Seasonal ARIMA models for forecasting foreign tourist arrivals in India. [24] proposed a multi-series structural time series method with one variable to predict seasonal tourism demand.

The long memory phenomenon has been widely studied in many areas, which motivates researchers to analyze this non-ignorable dependence between the present observation and all previous observations in a time series [25]. And the decay rate of this dependence can often be slower than exponential decay [26] provided a decent condition for a long memory stationary process using the autocorrelation function (ACF), denoted by ρ(j) for integers j, such that j=ρ(j)=. Furthermore, [27] and [28] proposed the autoregressive fractionally integrated moving average (ARFIMA) model by incorporating a fractional differencing operator of certain order d (0<d<1/2) with the classical autoregressive integrated moving average (ARIMA) model. To capture the real-world cyclical phenomena with long-range dependence, [29] and [30] adopted the seasonal autoregressive fractionally integrated moving-average (SARFIMA) model. Moreover, to describe the long memory with an oscillatory pattern in many fields, [28] extended the ARFIMA model to Gegenbauer ARMA (GARMA) model by introducing Gegenbauer polynomial. To overcome the difficulties of applying the classical time series models to model, forecast and analyze time series in the form of counts, [31] constructed generalised linear ARMA (GLARMA) model by modifying the linear predictor into a ARMA time series structure in a generalised linear regression model. With the prevalence of seasonal discrete time series in many areas such as biology, finance and engineering, [32] extended GLARMA to Gegenbauer GLARMA (GLGARMA) model by replacing the ARMA structure in the linear predictor to Gegenbauer ARMA structure. Moreover, [33] further developed GLGARMA model by incorporating the period effect component and cohort effect component from the mortality model structure. Compared with classical Lee-Carter models, the model risk and associated forecast errors can be dramatically reduced.

Contribution and structure

From a modeling perspective, our first contribution is to propose a generalised linear regression generalized autoregressive moving-average model (GLRGARMA) model and a generalised linear regression seasonal autoregressive integrated moving-average (GLRSARMA) model with an innovative function of explanatory variables in order to extend GLGARMA to incorporate relevant information for model fitting and forecast in the tourism area. Besides, the generalised Poisson (GP) distribution is adopted to accommodate over-, equal- and under-dispersion for certain tourism data. Model structures and properties of key components are precisely explained.

Our second contribution is to investigate the statistical features of the periodic sponge effect of tourism data. Especially, the pattern of long memory is examined. The analysis of the Hurst exponent, ACF plot and periodogram plot shows that Gegenbauer long memory features are presented in tourism data. Furthermore, the distinct characteristics between Gegenbauer long memory and seasonality are demonstrated to reveal that the GLRGARMA model is more suitable for modeling tourism data.

Our third contribution is the development of a Bayesian estimation framework for the proposed models, implemented via an efficient and user-friendly Rstan package. For the ML approach, the likelihood function is untractable because of involves very high dimensional integrals. Several monitors of convergence of posterior samples are discussed, such as the number of effective samples and R^ estimate. The criteria for modeling performance are also derived.

Our fourth contribution is to adopt both GLRGARMA and GLRSARMA to implement in-sample fitting and out-sample forecasts on Denmark’s economic data with an apparent periodic sponge effect. Overall, 8 nested sub-models are compared and evaluated using several well-known selection criteria. The models with Gegenbauer long memory components provide more accurate predictability than seasonal type models. Moreover, the explanatory variable with a stronger periodic sponge effect can dramatically enhance both in-sample fitting and out-sample forecast results.

The rest of the paper is structured as follows. The ’Periodic sponge effect’ section introduces and discusses this key economic phenomenon in Denmark’s tourism industry. Following this, the ’Generalised linear regression models with GARMA or SARMA latent processes’ section provides a foundational review of generalized linear models (GLMs) and long-memory time series frameworks for discrete data, followed by the formal introduction of the proposed GLRGARMA and GLRSARMA models, including their mean function specifications across eight distinct sub-models. The ’Bayesian inference’ section details the Bayesian estimation methodology for these models, alongside criteria for model selection and evaluation. In the ’Data analysis’ section, we demonstrate the practical utility of our framework through an empirical analysis of tourism data, reporting both in-sample fit metrics and out-of-sample forecasting performance. The ’Limitations and future work’ section discusses the constraints of the current study and potential directions for future research. Finally, the ’Conclusion’ section concludes with key insights and implications for future research.

Periodic sponge effect

Denmark is a renowned and attractive Scandinavian country for visitors. Like other famous tourist attractions, there exists an obvious seasonal pattern in tourism. [15] proposed that there exists an apparent wane and wax shifting phenomenon between the developments of tourism and industry. Both tourism and industry follow totally opposite moving directions. This phenomenon is defined as the periodic sponge effect. Fig 1 shows the time series trend of the number of rented hotel rooms (in black) and IPI (in blue) from Jan 2000 to Dec 2019 in Denmark. Overall, both tourism and industry exhibit strong seasonality. The trend of IPI is counter-cyclical to the seasonal period of tourism. Details of data sets can be found in Section Data analysis.

Fig 1. Time series plot for the number of rented hotel room (scaled in million) and IPI.

Fig 1

There are two main drivers for the existence of the periodic sponge effect. The first one is the seasonal characteristic in tourism. The second one is the flexicurity labour market with flexible employment policies in Denmark, because there also exists a periodic labour mobility pattern between tourism and industry. Other causes of this effect, like tight immigration policy, higher level of unemployment assistance, and low barriers to entry tourism labour market, are also be discussed from various aspects. The existence of periodic sponge effect not only helps to understand the dynamic of the economic developments but also enhances the economy’s capacity to withstand risks. In many other places, the service sector has stirred concern about the possible drawbacks caused by the excessively strong tourism industry. However, [15] proved that the existence of the periodic sponge effect will avoid the Dutch Disease, which means that the increment of tourism does not harm economic development. Moverover, after experiencing a global disaster, such as great recession or global pandemic, the service sector with high return, low industry threshold and asset-light operation can recover rapidly to help other sectors, beacuse the existence of a periodic sponge effect means there exists a broad, smooth, and efficient transmission mechanism between the service industry and other industries, enabling it to cope with disaster shocks through flexible economic pattern. Furthermore, to address the unemployment issue in a global disaster, the periodic sponge effect provides enough buffer areas in the labour market. It allows a certain amount of labour flow and solves a sudden increase in the unemployment rate. Hence, to study the periodic sponge effect is vital to understand the dynamic of the economic developments. It provides a guideline for policymakers to predict labour flow and economic activities. In this paper, We will conduct a quantitative study on this phenomenon, deeply exploring its mathematical characteristics by constructing statistical models. Additionally, we will use time-series models to investigate its patterns and forecast the changes and trends.

Generalised linear regression models with GARMA or SARMA latent processes

To model discrete time series with Gegenbauer long memory features, [32] extended the GLM to generalised linear generalized autoregressive moving-average model (GLGARMA) model, which combines the GLM and the GARMA time series model. The results from [32] show that generalised Poisson is consistently the best distribution choice compared with Poisson, NB and DP distributions. Furthermore, PD type models with higher flexibilities are often better than OD type models in both in-sample fitting and put-sample forecast. In terms of model structure, GLGARMA type model outperforms other types of models, including generalised autoregressive score (GAS) model [34], autoregressive conditional Poisson (ACP) model [35] and GLARFIMA model in both short and long memory data with or without obvious seasonal periodicity. In this study, since our models are extended from GLGARMA types model, based on the conclusions from [32], only PD type error term and GP distribution with Gegenbauer and seasonal periodic process are considered. Hence, generalised linear regression generalized autoregressive moving-average model (GLRGARMA) model and generalised linear regression seasonal autoregressive integrated moving-average (GLRSARMA) model are proposed to capture the periodic sponge effect in a time series.

Definition 1 ( GLRGARMA model) The GLGARMA model is a generalised state-space framework designed for discrete time series 𝐘1:T(Y1,Y2,,YT), integrating an observation equation and latent state dynamics. To ensure strictly positive intensity parameters in count-valued processes, a canonical log-link function governs the conditional mean structure. For a stationary discrete time series with observed data filtration 1:t1=σ(Y1,Y2,,Yt1) with t[1,T], a GLGARMA model with order (p,d,q) is defined by

Yt|1:t1,X1:t~GP(μt,ν),
Φ(B)ln(μt)=βg(Xt)+(12uB+B2)dΘ(B)εt
εt~i.i.dN(0,σ2)

where g(Xt)W are the function of explanatory variables representing the entire feature state at time t with regressors Xt,w for t[1,T] and w[1,W]. βW is a parameter vector that describes the relationship between Y and X. The dispersion parameter is defined as ν(1,1) for GP distribution. The Gegenbauer parameter |u|<1 controls the pattern of oscillation and the long memory parameter 0<d<1/2 determines the strength of long memory. B is the backshift operator, such that BYt=Yt1 and

Φ(B)=1ϕ1BϕpBp  and  Θ(B)=1+θ1B++θqBq,

are the autoregressive and moving-average characteristic polynomials, respectively, with no common roots.

Definition 2 ( GLRSARMA model) The GLSARIMA model with order (p,s,q) is given by

Yt|1:t1,X1:t~GP(μt,ν),
Φ(B)(1αBs)lnμt=βg(Xt)+Θ(B)(1γBs)εt
εt~i.i.dN(0,σ2)

where (1–Bs) is the standard integer seasonal difference operator with the integer seasonal period S that defines the frequency of the seasonal pattern.

In both GLRGARMA and GLRSARMA frameworks, explanatory variables interact with the latent processes via additive separation: βg(Xt) modulates the mean trend directly. It represents explanatory information at time t, characterising the strength of association between the explanatory variables and the response variable Yt. While GARMA/SARMA components focus on investigating fundamental oscillation features of data. These two components capture the long memory/seasonal dependencies of historical observations. Considering some special cases of βg(Xt), the GLRGARMA model and GLRSARMA model can be further divided into 8 different sub-models with their mean functions defined in Table 1.

Table 1. 8 different sub-models where model 1,2,3,4 are GLRGARMA models, model 5,6,7,8 are GLRSARMA models.

Model mean function
model 1 lnμt=β0+β1X1+(12uB+B2)dεt
model 2 lnμt=β0+β1X1+β2X2+(12uB+B2)dεt
model 3 lnμt=(12uB+B2)dX+εt
model 4 lnμt=(12uB+B2)dX+(12uB+B2)dεt
model 5 lnμt=β0+β1X1+αlnμts+(1γBs)εt
model 6 lnμt=β1X1+β2X2+lnμts+(1Bs)εt
model 7 lnμt=(1δBs)X+αlnμts+εt
model 8 lnμt=(1δBs)X+αlnμts+(1γBs)εt

Model 1 and model 5 are simple generalised linear regression models with a constant β0 and a single explanatory variable β1X1. These models can be regarded as baseline models for comparison. There are two explanatory components in model 2 and model 6 to evaluate the improvement of incorporating multi-source information. For model 3 and model 7, to test the cointegration between the periodic features (Gegenaber long memory and seasonal pattern) and explanatory variable, the periodic component is introduced to explanatory variable structure. Model 4 and model 8 are compositions of GLRGARMA/GLRSARMA model and explanatory variable with periodic component.

For the data distributions, GP distribution has the pmf, mean and variance given by

f(yt;μt,ν)=μt(1ν)[μt(1ν)+νyt]yt1eμt(1ν)νyt/yt!,μt>0, 1ν<1,
𝔼(Yt)=μtandVar(Yt)=μt(1ν)2,

respectively. Furthermore, the GP distribution is over-, under- and equi-dispersed when ν is greater than, less than and equal to 0, respectively. On the other hand, the pmf for the DP distribution is given by

f~(yt;μt,ν)=c(ν,μt)f(yt;μt,ν),μt>0, ν>0,

where the normalizing term c(ν,μt), given by [36], is

1c(ν,μt)=y=0f(yt;μt,ν)1+1ν12μtν(1+1μtν).

Because of the complicated structure of c(ν,μt), some properties of the DP distribution are difficult to derive. The unnormalised pmf, mean and variance are given by

f(yt;μt,ν)=(ν1/2eνμt)(eytytytyt!)(eμtyt)νyt,
𝔼(Yt)μtand Var(Yt)μtν,

respectively. The DP distribution is over-, under- and equi-dispersed when ν is less than, greater than and equal to 1 respectively.

Mean functions

The key component of GLRGARMA model in the mean functions μt is based on a long memory process structure. There are two typical types of long memory model structure, the autoregressive fractionally integrated moving average (ARFIMA) model class and the generalised form of ARFIMA called Gegenbauer autoregressive integrated moving average (GARMA) model. [27] and [28] extended the classical ARIMA model to the ARFIMA model, which describe a long memory stationary process with integrated order d(0,1/2). The Gegenbauer ARMA (GARMA) framework [28] extends the generalized ARFIMA specification to accommodate time series with oscillatory-damped autocorrelation structures. This generalization leverages Gegenbauer polynomials to inherently capture long-memory patterns, providing a natural parametrization for cyclical persistence in autocorrelation functions (ACFs).

For d(0,1/2), the ARFIMA model exhibit long memory features. The short memory ARMA model is a special case of ARFIMA model with d = 0 where the long memory operator (1−B)−2d = 1.

Definition 3 ( ARFIMA) Consider a stationary time series process with constant c, an ARFIMA model with order (p,d,q) is defined by

Φ(B)(μtc)=Θ(B)(1B)2dεt,  εt~i.i.dN(0,σε2),

where the long memory operator in ARFIMA model can be represented as

(1B)da=j=0Γ(j+da)Γ(j+1)Γ(da)Bj=j=0φjBj  with  da=2d.

For d(0,1/2) and u(1,1), the GARMA model exhibits a long memory with an oscillatory pattern. The ARFIMA model is the special case of GARMA model with u = 1 (see [27]) such that the factor (12uB+B2)d

Definition 4 ( GARMA) Consider a stationary time series process with constant c, a GARMA model with order (p,d,q) is defined by

Φ(B)(μtc)=Θ(B)(12uB+B2)dεtΘ(B)(j=0ψjεtj),  εt~i.i.dN(0,σε2),

The Gegenbauer long memory operator in GARMA model can be represented as

(12uB+B2)d=j=0ψjεtj,

and ψj denote the coefficients of the generating function for the Gegenbauer polynomials (12uB+B2)d [37]. These coefficients are formulated as

ψj=q=0[j/2](1)q(2u)j2qΓ(dq+j)q!(j2q)!Γ(d),

where [j/2] represents the integral part of j/2.

The coefficients ψj in Eq (4) are functionally dependent on d, which controls the strength of long memory and the Gegenbauer parameter u that controls the oscillation of ACF [38]. The coefficients ψj can be easily computed using the recursive formula:

ψj=2u(d1j+1)ψj1(2d1j+1)ψj2,

where the first three terms are ψ0=1, ψ1=2du and ψ2=d+2d(1+d)u2. Furthermore, [32] demonstrated that the bounds for the coefficients in the Gegenbauer polynomials ψj are the coefficients of ARFIMA φj

ψj|(2d)jj!=φj.

Gagenbauer fractional differences versus seasonal difference in both time domain and frequency domain

In time series settings, the terms of long memory refers to the strength of statistical dependence, extended temporal dependence or persistence between lagged observations in a time series. And the rate that such lagged dependency decreases should be slower than exponential decay, which is the main feature in the long memory structure time series [25]. The Wold representation introduced by [39] states that

Theorem 1. Any zero-mean nondeterministic covariance-stationary process Yst{1,2,3,,T} can be expressed as

Yt=ct+j=0ψjεtj=Ψ(B)εt+ct,  εt~WN(0,σ2), (1)

where εt and ψj are uniquely defined and satisfy ψ0=1,  j=0ψj2<, E(εt)=0, E(εt2)=σε2, E(εtεs)=0,t,s, the coefficients {ct;t} is a deterministic term with E(ct,εs)=0,t,s, WN stands for white noise.

Given a stationary time series process Y1:T(Y1,Y2,,YT) which admits Wold representation, with Y1:T({0})T, [26] defined a condition for a long memory stationary process in terms of the divergence of the autocorrelation function (ACF) for Yt and Yt + j at lag j, such that

limnj=nn|ρ(j)|  where  ρ(j)=Cov(Yt,Yt+j)𝕍ar(Yt)𝕍ar(Yt+j).

Parametric analysis of the autocorrelation function (ACF) in long-memory processes offers critical insights into their persistence characteristics, particularly through the interplay of fractional integration parameter d and Gegenbauer frequency parameter u. A notable case arises when u = −1 and 0<d<1/4 , where the ACF admits a closed-form representation:

ρ(j)=(1)jΓ(12d)Γ(j+2d)Γ(2d)Γ(j2d+1)~constant·(1)jj4d1, as  j.

For ARFIMA(0,d,0) with u = 1 and 0<d<1/4, it is given asymptotically by,

ρ(j)=Γ(12d)Γ(j+2d)Γ(2d)Γ(j2d+1)~constant·j4d1, as  j.

For the GARMA(0,d,0) process under the constraints |u|<1, 0<d<1/2, a closed-form expression for the autocorrelation function (ACF) remains analytically intractable [40]. However, its asymptotic behavior is characterized by:

ρ(j)~constant·j2d1sin(πdjλ0), as  j,

where λ0=cos1(u). The ACF plots of long-memory processes exhibit distinct patterns depending on parameters d and u.

According to the Szegö-Kolmogorov formula, the spectral density fs(λ) can be derived by taking the Fourier transform of autocovariance functions γθ(lj)=(Γθ)lj and IT(λk) can be derived by taking Discrete Fourier Transformation (𝔇(·)) of Yt{1,2,3,,T} where λk=2πkT for k=lj=1,,[T2], [·] represents the integer part and only half of frequencies are enough to demonstrate the features because of symmetry. The spectral density is given by

fs(λ)=12πγθ(k)exp(iλk)dk,  π<λ<π.

In practice, the periodogram IT(λ) is usually employed as an estimator of the spectral density fs(λ). [?] stated that IT(λ) is unbiased but inconsistent estimator of fs(λk) for a Gaussian white noise process Yt{1,2,3,,T}. The periodogram is given by

IT(λk)=12πT|𝔇(Yt{1,2,...,T})|2
=12πT|j=1TYjeijλk|2
=Yr2(λk)+Yi2(λk),
Yr(λk):=12πTj=1Tcos(jλk)Yj,
Yi(λk):=12πTj=1Tsin(jλk)Yj.

For the ARFIMA(p,d,q) model

Φ(B)(Ytc)=Θ(B)(1B)daεtΘ(B)(j=0φjεtj)  with  da=2d, (2)

The spectral density function can be expressed as follow:

fs(λ)=σ22π|1eiλ|2da|Θ(eiλ)|2|Φ(eiλ)|2
=σ22π(2sinλ2)4d|Θ(eiλ)|2|Φ(eiλ)|2.

Furthermore, [28] showed that as λ0,

fs(λ)=σ22π(2sinλ2)4d|Θ(eiλ)|2|Φ(eiλ)|2~σ22π|Θ(1)|2|Φ(1)|2λ4d~σ2λ4d,

for 0<λπ. Consequently, limλ0λ4dfs(λ) exists and is finite. For the special case of Φ(B)=Θ(B)=1, we have

fs(λ)=σ22π|1eiλ|4d=σ22π(2sinλ2)4d.

A useful generalised ARFIMA model called Gegenbauer ARMA (GARMA)(p,d,q) model proposed by [28] describes data showing slowly damping ACF with a cyclical pattern. The GARMA is given by

Φ(B)(Ytc)=Θ(B)(12uB+B2)dεtΘ(B)(j=0Ψjεtj). (3)

For the GARMA(p,d,q) process with long memory, the spectral density function was derived in [41] as follow:

fs(λ)=σ22π|12uexp(iλ)+exp(2iλ)|2d|Θ(eiλ)|2|Φ(eiλ)|2
=σ22π(4|sin(λ+λ02)sin(λλ02)|)2d|Θ(eiλ)|2|Φ(eiλ)|2,

where 0<d<1/2 and u=cos(λ0). Furthermore, it was demonstrated in [41] that the limiting behaviour of the spectral density function for the GARMA(p,d,q) model as λλ0 is

fs(λ)=σ22π|Θ(eiλ)|2|Φ(eiλ)|2(4|sin(λ+λ02)sin(λλ02)|)2d
~σ22π|Θ(λ0)|2|Φ(λ0)|2(4|sin(λ0)|)2d(λλ0)2d
~σ2(λλ0)2d. (4)

Hence, limλλ0(λλ0)2dfs(λ) exists and is finite. For the special case Φ(B)=Θ(B)=1, we have

fs(λ)=σ22π[4(cos(λ)cos(λ0))2]d=σ22π[4(cos(λ)u)2]d. (5)

The differences between the SARIMA model and GARMA model are demonstrated by [32]. They further clarified that there exists a clear distinct feature between seasonal oscillation and the oscillation that comes from a Gegenbauer long memory process. Their results further showed that the deseasonalisation cannot remove the oscillating behavior in the Gegenbauer long memory time series. To reveal the fundamental difference between Gagenbauer fractional differences and seasonal difference in tourism area, this study simulate two sets of data with oscillation period of 12 (1 year period for monthly data). For SARMA model, the period is set to 12 which is agree with our real world data. For GARMA data set, the long memory parameter is d = 0.49 and Gegenbauer parameter isu = 0.7, which is similar to the period of SARMA.

The first row in Fig 2 shows the time series plot simulated by SARMA (left panel) and GARMA (right panel). According to the time series, the tendencies of SARMA and GARMA look very similar. It is hard to distinguish the nature between these two series. The ACF plots (second row) and periodogram (last row) plots are provided to analyze the fundamental characteristics between SARMA and GARMA. The ACF plot for SARMA shows damped periodic peaks with an overall short memory pattern. For GARMA, the ACF plot shows a typical Gegenbauer long memory ACF. There exists an oscillated long memory pattern. The periodogram plot as a representation of ACF in frequency domain can easily reveal the differences between SARMA and GARMA. The peaks for SARMA model in the periodogram plot are allocated in several places. These peaks must be located in 0, π/2 and π because the number of peaks represents the period, which can be regarded that these peaks chop the region [0,π] into 12 pieces. For GARMA model, the location of the peak represents the period of Gegenbauer long memory process, which can be interpreted as λ=cos1(u) where λ is the location of the peak.

Fig 2. Simulation studies for SARMA and GARMA models.

Fig 2

Bayesian inference

In this study, we employ Bayesian inference to perform in-sample fitting and out-of-sample forecasting, leveraging Bayes’ theorem to integrate prior structural knowledge into state-space formulations. This approach circumvents the computational challenges of evaluating marginal likelihood functions in partially observed models (e.g., high-dimensional integration over latent variables) [32]. Furthermore, our framework generates posterior predictive distributions for probabilistic forecasting, enabling nuanced interpretation of model characteristics. For instance, credible intervals for all parameters can be directly derived from their posterior distributions.

Lety1:T=(y1,y2,,yT)denote a discrete-time series of non-negative integer observations(yt+{0})and ϑ*the parameter vector. Under the Bayesian paradigm, the posterior distribution of ϑ* conditioned on y1:T is defined as:

π(ϑ*|y1:T)=f(y1:T|ϑ*) π(ϑ*)f(y1:T|ϑ*) π(ϑ*) dϑ*f(y1:T|ϑ*)π(ϑ*),

which is proportional to the likelihood function f(y1:T|ϑ*)and the prior densities π(ϑ*), where priors are specified via empirical evidence or historical data. In the absence of prior knowledge, non-informative or reference priors can be adopted to preserve Bayesian objectivity.

Bayesian model

Let ϑ*=(ϑ,z) denote the vector of all model parameter ϑ=(β0,u,d,θj,ϕj,σ2,ν) and state parameter ε=ε1:T=(ε1,ε2,,εT), with each εt. For demonstrating purpose, both GLRGARMA model and GLRSARMA model with simple regression structure βXt, p = 0 for Φ(B) and q = 0 for Θ(B) are proposed as examples. Hence, for GLRGARMA model, the latent process is ln(μt)=β0+β(Xt)+(12uB+B2)dεt and the set of model parameter is ϑ=(β0,u,d,β,σ2,ν).

The priors π(ϑ) are defined as:

u~U(1,1),d~U(0,1/2),β0~N(0,σc2),
β~N(0,σϕ2),σ2~Γ(a,b)andν~U(1,1),

are adopted in which U(au,bu) denotes the uniform priors on the range (au,bu) for the long memory parameters u, d and ν and Γ(a,b) denotes the gamma prior with shape and scale parameters a and b respectively for the scale parameter σ2. The joint posterior distribution for the GLRGARMA(0,d,0) model with GP data distribution is

f(ϑ*|y1:T,x1:T)=f(y1:T|ε1:T,x1:T,ϑ)f(ε1:T|ϑ)π(ϑ)t=1T[exp(β0+β(Xt)+j=0ψjεtj)(1ν)[exp(β0+β(Xt)+j=0ψjεtj)(1ν)+ytν]yt1Γ(yt+1)expytνexp(β0+β(Xt)+j=0ψjεtj)(1ν)]×1σexp(εt22σ2)exp(β022σβ02)·(β22σβ2)·(σ2)a1e(σ2)bIu(1,1)Id(0,0.5)Iν(1,1),

where the hyperparameters are set to be σc2=σβ2=10, a = 3 and b = 1.

For GLRSARMA model, the latent process is (1Bs)lnμt=β0+βg(Xt)+(1Bs)εt and the set of model parameter is ϑ=(β0,β,σ2,ν). The priors π(ϑ) are defined as:

β0~N(0,σβ02),β~N(0,1),γ~N(0,1),α~N(0,1),σ2~Γ(a,b)andν~U(1,1),

The joint posterior distribution for the GLRSARMA(0,s,0) model with GP data distribution is

f(ϑ*|y1:T,x1:T)=f(y1:T|ε1:T,x1:T,ϑ)f(ε1:T|ϑ)π(ϑ)t=1T[exp(β0+β(Xt)+αlnμts+(1γBs)εtj)(1ν)Γ(yt+1)expytνexp(β0+β(Xt)+αlnμts+(1γBs)εtj)(1ν).×[exp(β0+β(Xt)+αlnμts+(1γBs)εtj)(1ν)+ytν]yt1]×1σexp(εt22σ2)exp(β022σβ02)·(β22σβ2)·(σ2)a1e(σ2)bIν(1,1).

Bayesian forecasting

A key strength of Bayesian forecasting lies in its capacity to generate full posterior predictive distributions for multi-step predictions. This framework implements m-step forecasts yT+1:T+m, through iterative one-step-ahead forecasting, where each prediction yT+s, s=1,,m is conditioned on the expanding information set s:T+s1 and covariates x1:T. The posterior predictive distribution for yT+s, s=1,,m is formally defined as:

f(yT+s|s:T+s1,xs:T+s)=f(yT+s|μT+s,ϑ,s:T+s1,xs:T+s)f(μT+s|μs:T+s1,ϑ,s:T+s1,xs:T+s)f(ϑ|s:T+s1,xs:T+s) dμs:T+sdϑ, (6)

And this integral is approximated by a Monte Carlo estimator using posterior samples:

f^(yT+s|s:T+s1,xs:T+s)=1Ll=1Lf(yT+s|μs:T+s(l),ϑs(l),s:T+s1,xs:T+s).

In this analysis, we employ L=90,000 post-burn-in iterations per MCMC chain for each information window s:T+s1 and xs:T+s. Here, μs:T+s(l) and ϑs(l)represent the l-th posterior draws of the latent states μs:T+s and parameters ϑ, respectively. Beyond posterior predictive distributions, Bayesian inference further yields point estimators and predictive credible intervals for forecasts. Empirical Bayes forecasting, which conditions on in-sample posterior point estimates (ϑ~ and μ~s:T+s) rather than marginalizing over parameter uncertainty, enhances computational efficiency [42]. Comparative analyses between frequentist and empirical Bayes forecasting methodologies are detailed in [43], with the latter omitting posterior integration steps in predictive distributions [32].

f^EB(yT+s|s:T+s1)=f(yT+s|μ~s:T+s,ϑs~,s:T+s1).

The point estimators ϑs~ and μ~s:T+s are typically derived from either the maximum a posteriori (MAP) estimate or the estimator minimizing the posterior expected loss. The MAP estimator generalizes the maximum likelihood (ML) principle under uninformative priors, as ϑ~s then coincides with the mode of the posterior distribution

ϑ~s,MAP=arg maxϑsfϑs(ϑs|s:T+s1).

Another Bayes estimator that minimises the posterior expected loss (PEL) is defined as

ϑ~s,PEL=argminϑs E(L(ϑs,ϑ~s|s:T+s1)),

where L(ϑs,ϑ~s|s:T+s1) is the loss function. One example is the commonly used minimum mean square error (MSE) estimator defined as

ϑ~s,MSE=argminϑs E([ϑsϑ~s]2|s:T+s1),

where ϑ~s,MSE corresponds to the posterior mean E(ϑs|s:T+s1)=ϑs¯. If the minimum absolute error (AE) estimator L(ϑs,ϑs~)=ϑsϑ~s is used, it gives ϑ~s,AE=ϑs,0.5 which is the posterior median.

Bayesian tool: Implementation with Rstan

The proposed Bayesian models are implemented via the Rstan package, which interfaces with the Stan probabilistic programming language (C++ backend) for efficient inference. Rstan employs Hamiltonian Monte Carlo (HMC) sampling [44,45], a Markov chain Monte Carlo (MCMC) variant [46] that supersedes conventional methods like random-walk Metropolis [46] and Gibbs sampling [47] in high-dimensional parameter spaces. HMC accelerates convergence by reparameterizing the sampling problem through Hamiltonian dynamics [48], leveraging gradient information to generate distant proposals. This mechanism circumvents the inefficient state-space exploration inherent to random-walk methods, which exhibit diffusive behaviour in complex posterior landscapes.

For the HMC sampler, to assess the dependence, precision and convergence of the posterior sample, three measures are reported in Rstan. The first measure is the number of effective samples which indicates dependence within a Monte Carlo sample. The second measure is the Monte Carlo standard error (MCSE)

MCSE=posterior standard deviationnumber of effective samples,

which reports the error of estimation for the posterior mean. To monitor the convergence for k > 2 chains of length 2n each, [49] proposed R^ which is defined as

R^=V^W·dfdf2,

where

V^=n1nW+k+1knB,W=i=1ksi2k,B=ni=1k(ϑi.ϑ..)2k1,df=2V^2V^ar(V^), (7)
V^ar(V^)=(n1n)21kV^ar(si2)+(k+1kn)22k1B2+2(k+1)(n1)kn2·nk[C^ov(si2,ϑi.2)2ϑ..C^ov(si2,ϑi.)], (8)

si is the within-chain variance and ϑij is the j-th parameter in chain i. If R^ is close to 1, the parameter ϑ has converged.

This analysis employs a single MCMC chain (k = 1), rendering between-chain variance B = 0 in Eqs (7) and (8). The chain executes 100,000 iterations, with the initial 10,000 iterations discarded as burn-in, yielding L=90,000 post-convergence draws (thinning interval = 1). Convergence diagnostics—including trace plots and R^ statistics—confirm all parameters satisfy R^ between 1.0000 and 1.0003, indicating stationarity and negligible between-chain variability. For both in-sample and out-of-sample analyses, effective sample sizes (ESS) range from 75,000 to 86,000 across parameters, reflecting moderate autocorrelation consistent with HMC’s sampling efficiency in high-dimensional spaces.

Model selection and forecast performance

The performance of each model was evaluated using the deviance information criterion (DIC) [50], a widely adopted Bayesian model selection metric. DIC balances model fit and complexity, extending classical information criteria to hierarchical models and addressing limitations in traditional approaches. As a generalized version of Akaike’s Information Criterion (AIC), DIC extends its applicability to models incorporating informative prior distributions, particularly hierarchical Bayesian frameworks. This adaptation addresses a critical limitation of AIC in handling models with parameter constraints imposed through prior specification. Since informative priors inherently restrict parameter freedom, conventional parameter counting methods become inadequate for AIC calculations. DIC resolves this ambiguity by introducing a probabilistic estimate of the model’s effective dimensionality through the concept of "effective number of parameters."

The DIC can be calculated using the equation

DIC=D¯+pD=2D¯D(ϑ¯x),

where the deviance is defined as D(ϑx)=2ln(f(yx|ϑx)), D¯=Eϑx|yx[2ln(f(yx|ϑx))] measures the model fit and the estimated number of parameters pD=D¯D(ϑ¯x) measures model complexity [51]. To compare models, calculate the DIC for each model and choose the model with the lowest DIC. A lower DIC suggests a better fit given the model complexity.

The forecast performance of m-step-ahead predictions y^x,t (obtained via posterior mean or median estimators) relative to observed values yx,t across T temporal intervals and gg demographic cohorts is quantitatively assessed through three fundamental error metrics: residuals rx,t=yx,ty^x,t percentage errors px,t=rx,tyx,t×100 and scaled errors ϵx,t as specified in Eq (11). Based on rx,t and px,t, three popular criteria, namely mean absolute error (MAE), root mean squared error (RMSE) and mean absolute percentage error (MAPE), are defined respectively below

MAE=1gx=1g[1mt=1m|rx,T+t|],RMSE=1gx=1g[1mt=1mrx,T+t2]
andMAPE=1gx=1g[1mt=1m|px,T+t|], (9)

While relative errors rx,t suffer from scale dependency, complicating cross-dataset comparisons, percentage errors px,t—though scale-invariant—exhibit heightened sensitivity to near-zero observations. To address these limitations, we introduce the Mean Absolute Scaled Error (MASE) as a fourth evaluation criterion, defined by:

MASE=1gx=1g[1mt=1m|ϵx,T+t|], (10)

making use of the scaled errors

ϵx,T+t=rx,T+t1m1t=2m|yx,T+tyx,T+t1|, (11)

proposed by [52]. Furthermore, this analytical framework can be extended to evaluate posterior estimates μ^x,t derived from mean or median estimators, where analogous error metrics—including residuals rx,ts=μx,tμ^x,t, percentage errors px,ts=rx,tsμx,t×100 and scaled errors

ϵx,ts=rx,ts1m1t=2m|μx,tμx,t1|,

can be systematically employed to establish parallel evaluation criteria for μ estimation, with MAE, RMSE, MAPE, and MASE directly computable using the unified computational formulations specified in Eqs (9) to (10).

Data analysis

In this study, several economic data sets of Denmark are analysed because of the special political-economic features and the existence of periodic sponge effect. The Gini coefficient is the lowest (latest OECD figures from 2012) and there are numerous strong local companies with great competitiveness [53]. The data sets are obtained from [54] which is the central authority on Danish statistics.The complete dataset used in this analysis is provided in S1 File. This reliable data source is a state institution under the Ministry of Economic Affairs and the Interior. To ensure methodological consistency, the analysis utilizes 2000-2019 data exclusively, as Denmark implemented revised seasonal adjustment models in December 2020, rendering pre-February 2021 datasets incomparable with post-recalibration series [54]. From the macroeconomic scope point of view, Denmark has rescheduled their economic development from a predominantly agricultural country to a modern industrial country. Currently, the agriculture sector only contributes less than 2% of the overall GDP. The industrial base and services contribute around 18% and 76%, respectively. [54]. For the service sector, tourism is the most important component, which can be treated as a crucial representative indicator. Moreover, [15] shows the existence of the periodic sponge effect between tourism and manufacturing and their findings reveal the dynamic mechanism of the economic developments in Denmark. Furthermore, the flexicurity labour market with flexible employment policies is another special characteristic in Denmark. Employers maintain a very high level of flexibility, which means they can hire and fire whenever they want. On the other hand, the unemployment compensation is relatively high to guarantee a stable living standard for unemployed persons [55].

[15] also claimed that the flexicurity labour market with flexible employment policies can be regarded as another powerful driver to cause the periodic sponge effect. Consequently, on a micro-scale with monthly data, number of rented hotel rooms (scaled in a million), power production, industrial production index (IPI) and unemployed rate are adopted to investigate the dynamic mechanism of tourism sector statistics. In order to avoid the complex composition of IPI, Power production as a single statistical indicator of industrial production activities is also applied. The Power production is a physical quantity indicator that measures the total electricity consumption in the industrial sector, focusing on reflecting energy input in the production process. Sourced from power authorities, it is highly real-time and sensitive to energy-intensive industries, often used for monitoring short-term economic activity and analyzing high-energy-consuming sectors. The IPI is a relative index that comprehensively reflects changes in industrial output scale, covering all industries (including manufacturing, mining, etc.). Calculated from enterprise statistical data with a lag in publication, it focuses on output results and capacity utilization efficiency, suitable for medium-to-long-term trend research and macroeconomic cycle analysis. The two are closely related but may deviate due to factors such as industrial structure adjustment or changes in energy efficiency. For example, when the proportion of low-energy-consuming industries increases, the growth rate of the IPI may exceed that of power consumption, whereas the opposite may occur if energy-intensive industries dominate.

Black curves in Fig 3 show a time series plot of rented hotel rooms (scaled in a million), power production, IPI and unemployed rate from January 2007 to June 2019. Grey lines are smoothed trends using a simple moving average approach. All of power production, IPI and unemployed rate show outstanding cyclical fluctuations with totally reversed period against number of rented hotel rooms. There exist strong periodic sponge effects in this group of data. The intensity of periodic sponge effect between number of rented hotel rooms and power production is strongest with an apparent reversed cyclical oscillation pattern. Moreover, the periodic sponge effect index between number of rented hotel rooms and power production is 0.89 which is the highest among these pairs [15].

Fig 3. Time series plot of number of rented hotel rooms (scaled in a million), power production, industrial production index (IPI) and unemployed rate in a monthly scale.

Fig 3

Empirical data analysis

The Hurst exponent H, also known as the index of long-range dependence, was proposed by [56]. It is a classical self-similarity parameter that measures the long memory feature in a time series [57]. Since it is robust with few assumptions about the underlying system, it has been widely applied to many fields [58]. A value of H in the range (12,1) indicates long memory in a time series, which means that a high value in the series will more likely be followed by another high value and such an effect is likely to maintain for a long period into the future. A value of H=12 can indicate a standard Brownian motion which is a short memory process. Furthermore, there exists a relationship between d and H, which is then given by d = H−0.5. Consequently, the estimator of the Hurst exponent H can approximate the long memory parameter d [32]. There exist various estimators of H, in this section, a well-known estimator called rescaled range analysis (R/S) is adopted to implement the following empirical studies.

[56] proposed the first Hurst exponent estimator using the rescaled range R/S analysis to measure the intensity of long-range dependence. Given a time series Yt{1,2,3,,T}, the sample mean and the standard deviation process are given by

YT=1Tj=1TYj  and  St=1t1j=1t(Xj)2, (12)

where the mean adjusted series Xt=YtYT. Then a cumulative sum series is given by Zt=j=1tXj and the cumulative range based on these sums is

Rt=Max(0,Z1,,Zt)Min(0,Z1,,Zt). (13)

An important proposition for the estimator of H was derived by [59].Consider a time series YtR and define St and Rt in Eqs (12) and (13) respectively, then   CR such that the following asymptotic property of the rescaled range R/S holds

[R/S](T)=1Tt=1TRt/St~CTH,  as  T.

In addition, for small sample size T, the rescaled range R/S can also be approximated by following equation [60]

[R/S](T)=T-1/2TΓ((T-1)/2)π(T/2)j=1T-1T-jj,  for  T340T-1/2T1Tπ/2j=1T-1T-jj,  for  T>340

where the T1/2T term was added by [61]. The H estimate can be obtained by a simple linear regression

logR/S(T)=logC+HlogT.

Hence, the definition for the estimator of H is given by the following equation

Definition 5 ( Estimator H^ by R/S) The estimator H^ based on the rescaled range R/S analysis is given by

H^R/S=T(t=1TlogR/S(t)logt)(t=1TlogR/S(t))(t=1Tlogt)T(t=1T(logt)2)(t=1Tlogt)2. (14)

The empirical confidence interval of H^ given in Eq (14) with sample size T = 2N [62] is

(0.5exp(7.33log(logN)+4.21),exp(7.20log(logN)+4.04)+0.5).

In this paper, the R package called pracma is adopted to estimate the value of H adopting the R/S analysis. For the data set of rented hotel room numbers in a month scale, the estimated Corrected R over S Hurst exponent is 0.908, which indicates that there exist a strong long memory in this data. Furthermore, according to the analytic plots for rented hotel room numbers in Fig 4, the class of long memory structure is a typical Gegenbauer long memory pattern with apparent oscillatory structure in ACF plot. For the periodogram plot, the peaks locate at non-zero position which aligns with the characteristics for Gegenbauer long memory type models.

Fig 4. Plot of ACF and periodogram for rented hotel room numbers.

Fig 4

Model fitting

In this section, Model 1 to Model 8 are adopted to fit rented hotel room numbers in a month. Industrial production index (IPI), power production (PG) and unemployment rate (UR) are incorporated to improve model feasibility. The in-sample fitting performances of seasonal component and Gegenbauer long memory component are compared in this study. To monitor the convergence of Bayesian approach, the values of R^ for each estimator are between 1.0000 and 1.0003 and the number of effective sample is always more than 80,000. Fig 5 is an example of a convergence test, which reports the MCMC sample path for several key parameters for Model 1. According to these plots, the model parameters are properly estimated.

Fig 5. Trace plots for part of Model 1 parameters.

Fig 5

The goodness of fit for the eight models is evaluated by using DIC to select the best fitting model for each data set. Table 2 reported the DIC values of these models. The performance of the models incorporating Gegenbauer long memory or seasonal component on error terms are significantly better than the models with these components on the explanatory variables X. This indicates that the periodic historical information provided by the explanatory variables X will cause negative impacts on modeling performance. As contrast, the error terms only keep essential characteristics of previous knowledge for modeling. Furthermore, the performance of the model with seasonal component is similar to the long memory model since the differences in DIC values between both type models are very small. Consequently, for in-sample fitting study, Model 1, Model 2, Model 5 and Model 6 outperform other models.

Table 2. DIC results.

Model 1 X= IPI X= UR X= PG
DIC 1440.54 1285.48 825.4
Model 2 X1 = IPI, X2=PG X1 = UR, X2=IPI X1 = UR, X2=PG
DIC 863.48 1325.06 794.77
Model 3 X = IPI X = UR X = PG
DIC 2510.72 2367.663 2115.049
Model 4 X = IPI X = UR X = PG
DIC 2516.81 2416.19 2291.37
Model 5 X = IPI X = UR X = PG
DIC 1327.78 1287.67 825.87
Model 6 X1 = IPI, X2=PG X1 = UR, X2=IPI X1 = UR, X2=PG
DIC 847.93 1318.26 820.23
Model 7 X = IPI X = UR X = PG
DIC 2508.97 2397.63 2187.52
Model 8 X = IPI X = UR X = PG
DIC 2607.18 2417.82 2128.53

The figures show the in-sample fit performance of selected best-performed models, which confirms that there is no significant evidence to claim long memory model superior than seasonal model in model fitting. Figs 6, 7 and 8 show the pairwise comparison of baseline model in in-sample fitting performance. The black dots with a dash line is the time series plot of observed data and the purple dots with dash line is the results generated from models. The grey area is the credible interval calculated from posterior samples. Both type models show reasonable and reliable fitting results. Furthermore, different information from explanatory variables will lead to distinct model performances. An explanatory variable with a significant periodic sponge effect, such as power production, will dramatically enhance model fitting results. Figs 9, 10 and 11 investigate the changes of model performances with an extra explanatory variable. The model fitting performance can be notably improved by introducing an explanatory variable with a stronger periodic sponge effect. Consequently, the periodic sponge effect plays a vital role in increasing model fitting accuracy.

Fig 6. In-sample fitting plot for model 1 and model 5 with X = IPI.

Fig 6

Fig 7. In-sample fitting plot for model 1 and model 5 with X = PG.

Fig 7

Fig 8. In-sample fitting plot for model 1 and model 5 with X = UR.

Fig 8

Fig 9. In-sample fitting plot for model 2 and model 6 with X1 = IPI and X2 = UR.

Fig 9

Fig 10. In-sample fitting plot for model 2 and model 6 with X1 = IPI and X2 = PG.

Fig 10

Fig 11. In-sample fitting plot for model 2 and model 6 with X1 = PG and X2 = UR.

Fig 11

Model forecast

In this section, we calculate one-step ahead forecast for m = 20 time points based on the posterior predictive distributions and the posterior sample size of L=90,000. Only Model 1, Model 2, Model 5 and Model 6 are adopted in out-of-sample forecasting study because these models show reasonable in-sample fitting performance. The forecasts y^t are given by the posterior mean or median. To evaluate the forecast performance, three types of measures, namely residuals rt=yty^t, percentage errors pt=rtyt×100 and scaled errors qt are adopted. Based on rt and pt, three popular criteria, namely the mean absolute error (MAE), root mean squared error (RMSE) and mean absolute percentage error (MAPE) are calculated in Table 3. Overall, Gegenbauer long memory models with smaller criteria values provide more accurate forecast results than seasonal models. Moreover, the forecast performance can be greatly improved by incorporating more explanatory variables.

Table 3. Comparison of models in forecasts with 95% Credible Intervals.

Model 1 X= IPI[CI] X= UR[CI] X= PG[CI]
MAE 15.43 [14.20, 16.80] 10.71 [9.80, 11.90] 8.08 [7.20, 9.10]
RMSE 16.75 [15.30, 18.40] 12.01 [10.90, 13.50] 9.59 [8.50, 10.80]
MAPE 0.16 [0.14, 0.18] 0.11 [0.10, 0.13] 0.09 [0.08, 0.10]
MASE 1.72 [1.58, 1.88] 1.34 [1.22, 1.48] 0.90 [0.82, 1.00]
Model 5 X = IPI[CI] X = UR[CI] X = PG[CI]
MAE 18.08 [16.50, 19.90] 12.01 [10.90, 13.40] 12.91 [11.70, 14.30]
RMSE 20.57 [18.70, 22.60] 13.16 [11.90, 14.70] 14.89 [13.50, 16.50]
MAPE 0.18 [0.16, 0.20] 0.12 [0.11, 0.14] 0.13 [0.11, 0.15]
MASE 2.02 [1.85, 2.20] 1.19 [1.08, 1.32] 1.44 [1.31, 1.59]
Model 2 X1 = IPI, X2=PG[CI] X1 = IPI, X2=UR[CI] X1 = UR, X2=PG[CI]
MAE 6.91 [6.20, 7.80] 9.91 [8.90, 11.10] 7.63 [6.80, 8.60]
RMSE 8.02 [7.10, 9.10] 11.27 [10.20, 12.60] 9.17 [8.20, 10.30]
MAPE 0.07 [0.06, 0.08] 0.10 [0.09, 0.11] 0.08 [0.07, 0.09]
MASE 0.77 [0.70, 0.85] 1.10 [1.00, 1.22] 0.85 [0.77, 0.94]
Model 6 X1 = IPI, X2=PG[CI] X1 = IPI, X2=UR[CI] X1 = UR, X2=PG
MAE 10.92 [9.80, 12.20] 9.92 [8.90, 11.10] 8.27 [7.40, 9.30]
RMSE 12.77 [11.50, 14.30] 11.54 [10.40, 12.90] 9.84 [8.80, 11.10]
MAPE 0.11 [0.10, 0.12] 0.10 [0.09, 0.11] 0.09 [0.08, 0.10]
MASE 1.22 [1.11, 1.35] 1.11 [1.01, 1.23] 0.92 [0.84, 1.02]

Note: Values represent point estimates with 95% credible intervals in brackets, calculated from 90,000 MCMC posterior samples.

Figures below show the forecast results, which agree with Table 3. Gegenbauer long memory type model should be the best choice in dealing with the number of rented hotel room data with Gegenbauer long memory features. Seasonal models cannot replace Gegenbauer long memory model since some fundamental features cannot be captured by a seasonal structure. Figs 12, 13 and 14 compare the predictability of the baseline model between Gegenbauer long memory type model and seasonal component type model. The black dots with dash line is the time series plot of observed data and the purple dots with dash line is the results forecasted from models. The grey area is the credible interval calculated from predictive posterior distribution. Similar to in-sample fitting study, the model incorporating the explanatory variable with a stronger periodic sponge effect produces more accurate forecasts. Figs 15, 16 and 17 describe the improvement of forecastability by adopting an extra explanatory variable with an intensive periodic sponge effect.

Fig 12. Out-of-sample forecasting plot for model 1 and model 5 with X = IPI.

Fig 12

Fig 13. Out-of-sample forecasting plot for model 1 and model 5 with X = PG.

Fig 13

Fig 14. Out-of-sample forecasting plot for model 1 and model 5 with X = UR.

Fig 14

Fig 15. Out-of-sample forecasting plot for model 2 and model 6 with X1 = IPI and X2 = UR.

Fig 15

Fig 16. Out-of-sample forecasting plot for model 2 and model 6 with X1 = IPI and X2 = PG.

Fig 16

Fig 17. Out-of-sample forecasting plot for model 2 and model 6 with X1 = PG and X2 = UR.

Fig 17

Limitations and future work

There are some limitations to our current study. Modeling assumptions, such as linear relationships between variables, fixed distributional forms (e.g., Generalized Poisson), and strict separation of long memory from seasonality, may oversimplify real-world complexities like nonlinear interactions, time-varying heteroscedasticity, and dynamic feature couplings. Estimation procedures relying on Bayesian inference via Rstan face computational challenges in high-dimensional spaces, suffer from subjective prior specifications, and lack robustness due to single-model-selection criteria (e.g., DIC), while struggling to handle high-dimensional explanatory variables. To address these gaps, future work could incorporate flexible, data-driven methods like gradient-boosted machines (GBM), LSTM/Transformer networks, or Bayesian neural networks (BNNs) to capture nonlinear patterns and long-range dependencies. Additionally, integrating high-dimensional data through automated feature engineering, dimensionality reduction (e.g., autoencoders), and causal inference techniques (e.g., Double Machine Learning) would enhance predictive power, while model ensembles and distributed computing could improve computational efficiency and scalability for real-world tourism forecasting.

Conclusion

This paper proposes a generalised linear regression structure with an innovative function of explanatory variables. Essential relevant information for modeling can be taken into consideration via explanatory variables. To capture the periodic oscillation features in a time series, both the seasonal component and Gegenbauer long memory component are incorporated to enhance model feasibility. Especially the fundamental differences between the seasonal component and Gegenabuer long memory component are distinguished. Moreover, the generalised Poisson (GP) distribution with over- equal- and under-dispersion is adopted to improve model flexibility. Furthermore, the existence of a periodic sponge effect among several key indices, including the number of rented hotel rooms, power production, IPI and unemployment rate, is discussed. The Gegenbauer long memory feature in the number of rented hotel rooms data is revealed. By plotting ACF and periodogram graphs, the long memory pattern is investigated. Overall, eight sub-models are implemented with the number of rented hotel rooms data to evaluate the model performance. The Bayesian approach is applied to implement in-sample fitting and out-of-sample forecast studies. Several model selection criteria are adopted to select the most feasible model. Overall, GLRGARMA model is evaluated to be the best model to handle the time series with Gegenbauer long memory feature, especially in the tourism area. The explanatory variable with a periodic sponge effect produces more accurate in-sample fitting results and out-sample forecasts.

Supporting information

S1 File. Denmark data.
pone.0329274.s001.csv (15.2KB, csv)

Data Availability

All relevant data are within the manuscript and its Supporting information files.

Funding Statement

This work was supported by the grant (3440012222519), Beijing Institute of Technology Research Fund Program for Young Scholars, awarded to L. Y. Sun. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

References

  • 1.Bunghez CL. The importance of tourism to a destination’s economy. J East Eur Res Bus Econ. 2016;1–9. [Google Scholar]
  • 2.World Trade Organization (WTO). 2008. http://www.wto.org/english/res_e/statis_e/its2007_e/its07_merch_trade_product_e.htm
  • 3.Lew AA. Tourism’s role in the global economy. Tour Geogr. 2011;13(1):148–51. [Google Scholar]
  • 4.Copeland BR. Tourism, welfare and de-industrialization in a small open economy. Economica. 1991:515–29. [Google Scholar]
  • 5.Ojaghlou M, et al. Tourism-led growth and risk of the Dutch disease: Dutch disease in Turkey. Int Bus Res. 2019;12(7):103–20. [Google Scholar]
  • 6.Kenell L. Dutch disease and tourism-the case of Thailand; 2008.
  • 7.Aynalem S, Birhanu K, Tesefay S. Employment opportunities and challenges in tourism and hospitality sectors. Journal of Tourism & Hospitality. 2016;5(6):1–5. [Google Scholar]
  • 8.Jolliffe L, Farnsworth R. Seasonality in tourism employment: human resource challenges. Int J Contemp Hosp Manag. 2003. [Google Scholar]
  • 9.Butler R. Seasonality in tourism: Issues and implications. Tourist Rev. 1998. [Google Scholar]
  • 10.Abdulrazaq MA, Fan WD. Temporal dynamics of pedestrian injury severity: A seasonally constrained random parameters approach. Int J Transp Sci Technol. 2024. doi: 10.1016/j.ijtst.2024.11.009 [DOI] [Google Scholar]
  • 11.Jang SS. Mitigating tourism seasonality: A quantitative approach. Ann Tour Res. 2004;31(4):819–36. [Google Scholar]
  • 12.Cannas R. An overview of tourism seasonality: Key concepts and policies. Almatour-J Tour Cult Territ Dev. 2012;3(5):40–58. [Google Scholar]
  • 13.Duro JA, Turrión-Prats J. Tourism seasonality worldwide. Tour Manag Perspect. 2019;31:38–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Ferrante M, Magno GLL, De Cantis S. Measuring tourism seasonality across European countries. Tour Manag. 2018;68:220–35. [Google Scholar]
  • 15.Yan H, Wang S. Poison or cure? A study on the periodic sponge effect in Denmark’s tourism industry; 2021. https://ssrn.com/abstract=3899761
  • 16.Kulendran N, Wong KK. Modeling seasonality in tourism forecasting. J Travel Res. 2005;44(2):163–70. [Google Scholar]
  • 17.Gil-Alana LA, Cunado J, Perez de Gracia F. Tourism in the Canary Islands: Forecasting using several seasonal time series models. J Forecast. 2008;27(7):621–36. [Google Scholar]
  • 18.Shen S, Li G, Song H. Effect of seasonality treatment on the forecasting performance of tourism demand models. Tour Econ. 2009;15(4):693–708. [Google Scholar]
  • 19.Chu FL. Forecasting tourism demand with ARMA-based methods. Tour Manag. 2009;30(5):740–51. [Google Scholar]
  • 20.Chang YW, Liao MY. A seasonal ARIMA model of tourism forecasting: The case of Taiwan. Asia Pacific J Tour Res. 2010;15(2):215–21. [Google Scholar]
  • 21.Shen S, Li G, Song H. Combination forecasts of international tourism demand. Ann Tour Res. 2011;38(1):72–89. doi: 10.1016/j.annals.2010.05.003 [DOI] [Google Scholar]
  • 22.Saayman A, Botha I. Non-linear models for tourism demand forecasting. Tour Econ. 2017;23(3):594–613. [Google Scholar]
  • 23.Mishra PK, Rout HB, Pradhan B. Seasonality in tourism and forecasting foreign tourist arrivals in India. Iran J Manag Stud. 2018;11(4):629–58. [Google Scholar]
  • 24.Chen JL, Li G, Wu DC, Shen S. Forecasting seasonal tourism demand using a multiseries structural time series method. J Travel Res. 2017;58(1):92–103. doi: 10.1177/0047287517737191 [DOI] [Google Scholar]
  • 25.Graves T, Gramacy RB, Watkins N, Franzke C. A brief history of long memory; 2014. doi: 10.48550/arXiv.1406.6018 [DOI] [Google Scholar]
  • 26.Beran J. Statistics for long-memory processes. CRC press. 1994.
  • 27.Granger CWJ, Joyeux R. An introduction to long-memory time series models and fractional differencing. J Time Ser Anal. 1980;1(1):15–29. [Google Scholar]
  • 28.Hosking JRM. Fractional differencing. Biometrika. 1981;68(1):165–76. [Google Scholar]
  • 29.Porter-Hudak S. An application of the seasonal fractionally differenced model to the monetary aggregates. J Am Stat Assoc. 1990;85(410):338–44. [Google Scholar]
  • 30.Hassler U. (Mis) specification of long memory in seasonal time series. J Time Ser Anal. 1994;15(1):19–30. [Google Scholar]
  • 31.Davis RA, Dunsmuir WTM, Wang Y. Modeling time series of count data. Stat Textb Monogr. 1999. p. 63–114.
  • 32.Yan H, Chan JS, Peters GW. Long memory models for financial time series of counts and evidence of systematic market participant trading behaviour patterns in futures on US treasuries; 2017. https://ssrn.com/abstract=2962341
  • 33.Yan H, Peters GW, Chan J. Mortality models incorporating long memory for life table estimation: A comprehensive analysis. Ann Actuar Sci. 2021:1–38. [Google Scholar]
  • 34.Creal D, Koopman SJ, Lucas A. Generalized autoregressive score models with applications. J Appl Econ. 2013;28(5):777–95. [Google Scholar]
  • 35.Groß-KlußMann A, Hautsch N. Predicting bid–ask spreads using long-memory autoregressive conditional Poisson models. J Forecast. 2013;32(8):724–42. [Google Scholar]
  • 36.Efron B. Double exponential families and their use in generalized linear regression. J Am Stat Assoc. 1986;81(395):709–21. doi: 10.1080/01621459.1986.10478327 [DOI] [Google Scholar]
  • 37.Stein EM, Weiss GL. Introduction to Fourier analysis on Euclidean spaces. Princeton University Press. 1971.
  • 38.Rainville ED. Special functions. New York. 1960.
  • 39.Wold H. A study in the analysis of stationary time series. Almqvist & Wiksell. 1938.
  • 40.Woodward WA, Cheng QC, Gray HL. A k-factor GARMA long-memory model. J Time Ser Anal. 1998;19(4):485–504. [Google Scholar]
  • 41.Gray HL, Zhang N, Woodward WA. On generalized fractional processes. J Time Ser Anal. 1989;10(3):233–57. doi: 10.1111/j.1467-9892.1989.tb00026.x [DOI] [Google Scholar]
  • 42.Robins H. Some thoughts on empirical bayes estimation. The Annals of Statistics. 1983;11(3):713–23. [Google Scholar]
  • 43.Knox T, Stock JH, Watson MW. Empirical bayes forecasts of one time series using many predictors. National Bureau of Economic Research; 2001. p. 269. Available from: http://www.nber.org/papers/t0269
  • 44.Duane S, Kennedy AD, Pendleton BJ, Roweth D. Hybrid monte carlo. Phys Lett B. 1987;195(2):216–22. [Google Scholar]
  • 45.Neal RM. An improved acceptance procedure for the hybrid Monte Carlo algorithm. J Computat Phys. 1994;111(1):194–203. [Google Scholar]
  • 46.Metropolis N, Rosenbluth AW, Rosenbluth MN, Teller AH, Teller E. Equation of state calculations by fast computing machines. J Chem Phys. 1953;21(6):1087–92. [Google Scholar]
  • 47.Geman S, Geman D. Stochastic relaxation, gibbs distributions, and the bayesian restoration of images. IEEE Trans Pattern Anal Mach Intell. 1984;6(6):721–41. doi: 10.1109/tpami.1984.4767596 [DOI] [PubMed] [Google Scholar]
  • 48.Neal RM. MCMC using Hamiltonian dynamics. Handbook of Markov Chain Monte Carlo; 2011. p. 113–62.
  • 49.Gelman A, Rubin DB. Inference from iterative simulation using multiple sequences. Stat Sci. 1992;7(4):457–72. [Google Scholar]
  • 50.Spiegelhalter DJ, Best NG, Carlin BP, Linde A. The deviance information criterion: 12 years on. J R Stat Soc: Ser B (Stat Methodol). 2014;76(3):485–93. [Google Scholar]
  • 51.Spiegelhalter DJ, Best NG, Carlin BP, Van Der Linde A. Bayesian measures of model complexity and fit. J R Stat Soc: Ser B (Stat Methodol). 2002;64(4):583–639. [Google Scholar]
  • 52.Hyndman RJ, Koehler AB. Another look at measures of forecast accuracy. Int J Forecast. 2006;22(4):679–88. [Google Scholar]
  • 53.Refslund B, Sørensen OH. Islands in the stream? The challenges and resilience of the Danish industrial relations model in a liberalising world. Ind Relat J. 2016;47(5–6):530–46. [Google Scholar]
  • 54.Statistics Denmark; 2020. https://www.dst.dk/en
  • 55.World Bank Country and Lending Groups. 2020. https://datahelpdesk.worldbank.org/knowledgebase/articles/906519#High_income
  • 56.Hurst HE. Long-term storage capacity of reservoirs. Trans Amer Soc Civil Eng. 1951;116:770–808. [Google Scholar]
  • 57.Millen S, Beard R. Estimation of the Hurst exponent for the Burdekin River using the Hurst-Mandelbrot rescaled range statistic. In: First Queensland Statistics Conference; 2003.
  • 58.Qian B, Rasheed K. Hurst exponent and financial market predictability. In: Proceedings of the 2nd IASTED international conference on financial engineering and applications; 2004. p. 203–9.
  • 59.Mandelbrot BB. Limit theorems on the self-normalized range for weakly and strongly dependent processes. Probab Theory Relat Fields. 1975;31(4):271–85. [Google Scholar]
  • 60.Annis AA, Lloyd EH. The expected value of the adjusted rescaled Hurst range of independent normal summands. Biometrika. 1976;63(1):111–6. [Google Scholar]
  • 61.Peters EE. Fractal market analysis: Applying chaos theory to investment and economics. John Wiley & Sons. 1994.
  • 62.Weron R. Estimating long-range dependence: Finite sample properties and confidence intervals. Phys A: Stat Mech Appl. 2002;312(1):285–99. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

S1 File. Denmark data.
pone.0329274.s001.csv (15.2KB, csv)

Data Availability Statement

All relevant data are within the manuscript and its Supporting information files.


Articles from PLOS One are provided here courtesy of PLOS

RESOURCES