Next Article in Journal
A Penny Saved Is a Penny Earned: How Executive Cognitive Flexibility Drives Performance Through Strategic Resource Reallocation
Previous Article in Journal
Cultural Ecosystem Services of Grassland Communities: A Case Study of Lubelskie Province
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring Precipitable Water Vapor (PWV) Variability and Subregional Declines in Eastern China

1
School of Geography and Environment, Jiangxi Normal University, Nanchang 330022, China
2
School of Geography, Archaeology and Irish Studies, University of Galway, H91CF50 Galway, Ireland
3
Key Laboratory of Natural Disaster Monitoring, Early Warning and Assessment of Jiangxi Province, Jiangxi Normal University, Nanchang 330022, China
4
Key Laboratory of Poyang Lake Wetland and Watershed Research, Ministry of Education, Jiangxi Normal University, Nanchang 330022, China
5
School of Geography and Information Engineering, China University of Geosciences, Wuhan 430078, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Sustainability 2025, 17(15), 6699; https://doi.org/10.3390/su17156699
Submission received: 11 June 2025 / Revised: 11 July 2025 / Accepted: 21 July 2025 / Published: 23 July 2025
(This article belongs to the Section Sustainability in Geographic Science)

Abstract

In recent years, China has experienced growing impacts from extreme weather events, emphasizing the importance of understanding regional atmospheric moisture dynamics, particularly Precipitable Water Vapor (PWV), to support sustainable environmental and urban planning. This study utilizes ten years (2013–2022) of Global Navigation Satellite System (GNSS) observations in typical cities in eastern China and proposes a comprehensive multiscale frequency-domain analysis framework that integrates the Fourier transform, Bayesian spectral estimation, and wavelet decomposition to extract the dominant PWV periodicities. Time-series analysis reveals an overall increasing trend in PWV across most regions, with notably declining trends in Beijing, Wuhan, and southern Taiwan, primarily attributed to groundwater depletion, rapid urban expansion, and ENSO-related anomalies, respectively. Frequency-domain results indicate distinct latitudinal and coastal–inland differences in the PWV periodicities. Inland stations (Beijing, Changchun, and Wuhan) display annual signals alongside weaker semi-annual components, while coastal stations (Shanghai, Kinmen County, Hong Kong, and Taiwan) mainly exhibit annual cycles. High-latitude stations show stronger seasonal and monthly fluctuations, mid-latitude stations present moderate-scale changes, and low-latitude regions display more diverse medium- and short-term fluctuations. In the short-term frequency domain, GNSS stations in most regions demonstrate significant PWV periodic variations over 0.5 days, 1 day, or both timescales, except for Changchun, where weak diurnal patterns are attributed to local topography and reduced solar radiation. Furthermore, ERA5-derived vertical temperature profiles are incorporated to reveal the thermodynamic mechanisms driving these variations, underscoring region-specific controls on surface evaporation and atmospheric moisture capacity. These findings offer novel insights into how human-induced environmental changes modulate the behavior of atmospheric water vapor.

1. Introduction

Water vapor plays a vital role in processes from local precipitation to global climate regulation [1,2]. Specifically, variations in the lower troposphere influence the frequency and severity of extreme weather events and are closely related to precipitation [3,4]. Precipitable Water Vapor (PWV) represents the total amount of atmospheric water vapor in the atmosphere [5] and effectively summarizes the distribution of water vapor throughout the entire atmospheric column. Therefore, PWV is fundamental to understanding and monitoring atmospheric moisture, offering valuable insights into sustainable climate management, water resource planning, and the assessment of weather- and climate-related risks.
PWV is usually retrieved using radiosondes and water vapor radiometers [6,7]. However, these strategies are constrained by their inadequacies. For instance, sounding stations, usually located 200–300 km apart, launch sounding balloons twice daily [8]. Consequently, radiosondes offer water vapor distribution data with limited horizontal and temporal resolutions. PWV can also be obtained from reanalysis products, which provide variables of global atmospheric circulation with excellent spatial integrity and continuity of data, using sophisticated data assimilation techniques [9,10]. However, the accuracy and quality of the PWV measurements are limited [11,12,13,14]. Since the seminal work of Bevis et al. [15], which introduced the concept of PWV retrieval using the data of Global Navigation Satellite Systems (GNSSs), the field of GNSS meteorology has undergone rapid advancement. The utilization of GNSS PWV data offers distinct advantages, including cost-effectiveness, enhanced precision, high temporal resolutions, and capacity to operate under diverse meteorological conditions [16,17].
In recent years, numerous researchers have studied precipitable water using time-series data. In the time domain, some researchers have utilized multiple data sources to retrieve PWV and found that most regions of the globe exhibit an increasing trend, while a few areas remain unchanged or show a decreasing trend [18,19,20,21]. Studies on the frequency-domain characteristics of atmospheric water vapor have highlighted distinct patterns across different climates and regions. Typical bimodal distributions of water vapor frequency have been found in monsoon regions, and lognormal distributions have been found in subtropical and temperate zones [22]. Some studies have extracted multiyear periods in the Greenland region using the Fast Fourier Transform (FFT) method based on GPS and ERA5 reanalysis PWV data [23]. Khutorova et al. [24] investigated the long-term oscillations in the time series of the surface partial pressure of water vapor for Russia’s European region and found that variations in timescales ranging from two to six years accounted for between 35 and 60 percent of the interannual variance. Jadala et al. [25] used an FFT and harmonic analysis to extract the amplitude and phase of diurnal, semi-diurnal, and 8 h oscillations at the tropical Hyderabad site.
Similarly, some researchers analyzed GNSS water vapor data from stations of the Crustal Movement Observation Network of China (CMONOC), and the results indicated the existence of annual and semi-annual periods and revealed notable regional disparities and seasonal fluctuations in water vapor frequency characteristics, further emphasizing the complexity of atmospheric moisture dynamics [26,27]. These findings further underscore the complexity of atmospheric moisture dynamics. However, most studies only briefly discuss the PWV trends and periodic behaviors, without a thorough investigation into the underlying causes. Moreover, the feature extraction methods employed for PWV analysis in these studies remain relatively limited, lacking integration for a comprehensive assessment.
Hu Huanyong’s foundational research historically revealed that eastern China holds an absolute advantage over the western regions in terms of both population and economics [28]. Moreover, eastern China encompasses a broader range of the country’s major climate zones, many of which have become focal areas for scientific investigation. In recent decades, this regional disparity has intensified, and the global increase in extreme precipitation events has posed greater risks to eastern China [29]. These factors underscore the urgent need for a comprehensive investigation of the PWV characteristics in this region.
In this study, the PWV time series over key regions of eastern China were derived from GNSS observations at stations of the International GNSS Service (IGS). Through frequency-domain analysis, we explored the multiscale temporal characteristics and dominant factors influencing PWV variability. These findings contribute to a deeper understanding of regional atmospheric water dynamics, supporting sustainable climate monitoring, environmental planning, and early warning systems for extreme weather events.
Furthermore, PWV contributes to achieving the 2030 Agenda goals. The long-term trend of PWV can be used as a proxy variable for changes in atmospheric humidity, complementing traditional temperature-based indicators and aiding in the prediction of urban heat island effects. Its cyclical characteristics directly reflect the frequency and intensity of extreme precipitation events, enabling early warning capabilities. These insights offer a new perspective for assessing urban hydrometeorological risks and can inform the optimization of urban drainage infrastructure and disaster-prevention strategies.

2. Research Data and Methods

2.1. Research Data

2.1.1. GNSS Data

The GNSS observation files used in this study were collected from 15 in situ stations of the IGS (Figure 1), an international organization founded in 1993 by the International Association of Geodesy (IAG), to provide services for GNSS applications. The data are stored in RINEX format, and all 15 stations are equipped with receivers that continuously track GNSS signals. The observational data can be downloaded from the IGS Data Center in Wuhan (http://www.igs.gnsswhu.cn/index.php, last access on 10 December 2023), and the study period extends from 1 January 2013 to 31 December 2022. The data used in this study have been quality-checked according to WMO standards, with non-compliant data excluded. Table 1 shows the locations of the 15 stations, temporal coverage, and number of valid points before and after exclusion in this study. The JFNG station has provided observations since 2014, whereas the CKSV, KMNM, NCKU, and WUH2 stations have provided observations since 2016.

2.1.2. Radiosonde PWV

Radiosondes are meteorological sounding instruments carried by balloons to measure atmospheric parameters, such as temperature, pressure, and humidity, at different altitudes. These measurements enable the accurate calculation of PWV using standard formulas [30]. Radiosonde PWV is commonly used as a reference value for the validation of satellite PWV products owing to its excellent accuracy [31,32]. In this study, radiosonde data from six sites, obtained from the University of Wyoming (http://weather.uwyo.edu/upperair/sounding.html, last access on 5 March 2024), were downloaded for the period 2013–2022 to evaluate the accuracy of the GNSS-derived PWV. These radiosonde sites are located within a 50 km radius of the GNSS stations. Figure 2 depicts the distribution of the six GNSS stations and neighboring radiosonde sites.

2.1.3. Temperature

GNSS stations lack meteorological measurements, while radiosonde stations provide such data but are sparse. Ground-based weather stations offer only surface-level observations. ERA5 is the fifth generation of the European Center for Medium-Range Weather Forecasts (ECMWF) atmospheric reanalysis dataset of global climate from January 1950 to the present, produced by the Copernicus Climate Change Service (C3S) [33]. In contrast, the ERA5 reanalysis data provide an adequate vertical temperature dataset across the study area. Therefore, the temperature dataset from ERA5 can be used to analyze vertical atmospheric temperature profiles. The temperatures used in this study are labeled as T2m (thermodynamic temperature at a height of 1.5–2 m) and Th (atmospheric temperature at 14 pressure levels from 1000 hPa to 600 hPa). The grid for latitude and longitude is 0.25° × 0.25°, and the time resolution is one hour. We then matched the latitude and longitude of the GNSS stations on a 0.25° grid to obtain atmospheric data from the GNSS stations.

2.2. GNSS PWV Retrieval

2.2.1. GNSS Data Solution Software

As is well known, there are numerous GNSS data processing software tools, among which GAMIT and Bernese are representative and widely used globally. GAMIT is designed for scientific use, supporting satellite orbit calculation, atmospheric delay correction, earth rotation parameter estimation, and other purposes. It provides millimeter-level accuracy and allows code customization. Bernese is another widely adopted tool that provides comprehensive functions for both precise point positioning (PPP) and network solutions [34]. Its automated processing capabilities and relatively user-friendly interface make it suitable for both research and engineering applications. However, its high cost may limit access to individual users.
In recent years, China has developed several high-precision GNSS processing tools. PRIDE and GREAT are two notable platforms that offer competitive performance in high-precision GNSS applications. PRIDE is an open-source software package with a user-friendly GUI for high-frequency GNSS data processing [35,36,37]. It offers centimeter-level accuracy and has strong dynamic tracking capabilities, making it particularly suitable for crustal movement monitoring, time–frequency transfer, and aerial photogrammetry. GREAT is a powerful tool for precise positioning, multi-source fusion navigation, and satellite precise orbit determination [38,39,40]. Its modular and open-source structure is suited to complex navigation scenarios and integrated Earth science research.
Considering the high precision, scientific applicability, and flexible customization of GAMIT 10.7, it was selected in this study for GNSS data processing. Its robust tropospheric delay modelling capabilities and millimeter-level accuracy make it well-suited for extracting reliable PWV time-series data in climate-related research. In this study, double-difference relative positioning strategies were adopted to process the GNSS observations within the research area. The processing model and detailed settings are presented in Table 2. The sampling rate of GNSS observations was 30 s. The final ephemeris of the IGS was used to estimate the satellite orbits. A cutoff angle of 10° was adopted to remove low-quality observations. The RELAX mode of GAMIT was employed, with data corrections applied using the VMF1 mapping function and tidal model [41]. Zenith Tropospheric Delay (ZTD) corrections were estimated hourly using the GAMIT default horizontal gradient.

2.2.2. GNSS PWV Retrieval Method

ZTD includes Zenith Hydrostatic Delay (ZHD) and Zenith Wet Delay (ZWD). ZHD can be accurately calculated using the Saastamoinen model [42]:
Z H D = 2.2768 ± 0.0024 P s / ( 1 0.00266 × cos 2 θ + 0.00028 × H )
where P s is the atmospheric pressure obtained from a ground-based weather station co-located at the GNSS station, θ is the latitude of the GNSS station, and H is the GNSS station height.
ZWD is obtained by subtracting ZHD from ZTD, and PWV can be derived using the following formula [43]:
P W V = ×   Z W D
where the conversion coefficient, Π, can be calculated using the Bevis formula:
= 10 6 / [ ( k 3 × T m 1 + k 2 ) × R V ]
where k 2 (22.1 ± 2.2 K/hPa) and k 3 (3.735 × 105 ± 0.012 × 105 K2/hPa) are the atmospheric refractivity constants; R V (4.613 × 106 erg/g.°) is the gas constant of water vapor; and T m is the weighted mean temperature, derived from the global barometric pressure and temperature (GPT) model of Boehm and Schuhti [44,45].
By plotting the GNSS PWV time series, we filtered and removed the PWV with values less than or equal to zero.

2.3. Bayesian Statistics

Bayesian statistics uses prior knowledge to probabilistically deal with model parameters and structures and provide a unified framework for resolving various forms of uncertainty [46]. It uses Bayesian model averaging (BMA) [47], which applies Bayesian integrated modelling techniques to combine numerous competing models and uncover complex nonlinear dynamics from time-series data. Although the model was originally developed for remote sensing [48,49], it can be applied to any time-series data that satisfy its assumptions [50]. In Bayesian statistics, time-series data are equal to the sum of trends, seasonality, and residuals:
Y t = T t + S t + e t ,   t = 1 ,   , n
where Y t ( t denotes time) is the time series, T t is the trend, S t is the seasonality, e t is the random variation of the variable, and n is the number of observations in the time series.
The seasonal component, S t , mentioned above is modelled as a piecewise harmonic model, whose specific form depends on the seasonal change points in the time series. For each time interval [ ξ k , ξ k + 1 ], the seasonal component can be expressed as in Formula (5). The trend component, T t , is modeled as a piecewise linear function, whose specific form depends on the trend change points in the time series. For each time interval [ τ j , τ j + 1 ], the trend component can be expressed as in Formula (6):
S t = l = 1 L k a k , l     sin 2 π l t P + b k , l     cos ( 2 π l t P ) ;   ξ k t ξ k + 1
T t = a j + b j t ;   τ j t τ j + 1
where P denotes the seasonal cycle (e.g., one year), L k is the harmonic order of the k th time period, a k , l and b k , l are the coefficients of the l th harmonic of the k th time period, ξ k is the position of the seasonal change point, a j and b j denote the linear coefficients of the j th time period, and τ j is the position of the trend change point.
The trend (trend probability) is rigorously evaluated using a Bayesian approach with Markov Chain Monte Carlo sampling and model averaging, ensuring robust uncertainty quantification. This method provides a probabilistic understanding of trend changes and indirectly assesses trend probabilities through the statistical analysis of posterior distribution samples. The general form of the posterior distribution is as follows:
p β M ,   σ 2 ,   M D p ( D | β M ,   σ 2 ,   M ) π ( β M ,   σ 2 ,   M )
where p denotes the probability density function; D is the observed data; β M and σ 2 are the model parameters and noise variance, respectively; and M is the model structure (including the number and location of trend and seasonal change points).
In this study, we used Bayesian statistics to obtain the trend/seasonality components and trend probabilities of the PWV time series to analyze the time-domain characteristics of PWV.

2.4. Wavelet Transform

Time series are often nonlinear and complex, and different results can be obtained by analyzing data at different timescales. The wavelet transform can be used for time–frequency analysis because of its multi-resolution property and ability to characterize the local features of a signal in the time and frequency domains [51,52]. The wavelet transform method is a process of shifting a basic wavelet function by τ and subsequently performing the inner product operation with the signal to be analyzed ( t ) at different scales μ . It can be expressed as follows:
W f a , τ = < f t ,   Ψ a , τ t > = 1 a R Ψ * ( t τ a ) d t
where f ( t ) denotes the original signal, W f a , τ describes the degree of similarity between a signal and a wavelet function Ψ a , τ ( t ) at a specific scale a and translation τ [53], and R is the set of real numbers. The duration of the wavelet at different scales broadens with increasing scale and decreases inversely with amplitude, a [54], while the shape of the wave remains constant.
This study employed the complex Morlet (Cmor) wavelet [55] to identify the main modes of variability and their temporal variations, as well as to examine the amplitude and phase of the oscillations. Moreover, the amplitude of the wavelet spectrum reflects the temporal variations in the relative contribution of the periodic components at different scales to the original signals. At any given moment, it allows for the estimation of the intensity of parameter variations across all timescales. The phase wavelet spectrum, on the other hand, indicates the timing of the maximum value of each mode. Owing to these capabilities, wavelet analysis has been widely employed to identify multi-timescale variations and reveal hidden features in complex time series, particularly in the geosciences [24,56]. Cmor is expressed as follows:
Ψ t = e i w 0 t e t 2 2 σ 2
where Ψ t denotes the value of the complex Morlet wavelet function at time t; i denotes the imaginary number; w 0 denotes the center frequency, which determines the principal frequency component of the wavelet function; and σ denotes the scale parameter, which controls the width of the wavelet function.

2.5. Fast Fourier Transform

The discrete Fourier transform (DFT) establishes a correspondence between the time-domain characteristics of the signal and the frequency-domain characteristics and realizes the mutual conversion of the signal in the time and frequency domains. The FFT, which was proposed by James W. Cooley in 1965, speeds up the computation process compared with the computationally intensive DFT [57]. The employment of the unit complex root as a rotational element not only adeptly captures energy fluctuations in multidimensional datasets, but also substantially diminishes the computational demands of the discrete Fourier transform. Furthermore, it elucidates the amplitude and phase attributes of continuous signals that remain imperceptible in the time domain in the frequency domain. The FFT is widely used in the frequency-domain analysis of various types of time series, and its core formula is as follows:
x k = 1 N j = 1 N X ( j ) W N ( j 1 ) ( k 1 ) , k = 1 , 2 , , N ;   j = 1 , 2 , , N
W N = e j 2 π N
where x k denotes the distribution of the signal in the frequency domain, X ( j ) denotes the distribution of the signal in the time domain, and W N is a rotation factor with periodicity and symmetry.
In this study, FFT was used to analyze the GNSS PWV time series in eastern China, and its oscillation characteristics were analyzed from a frequency-domain perspective.

2.6. Evaluation Indicators

Three evaluation metrics, namely, the correlation coefficient (R), Root Mean Square Error (RMSE), and mean bias (MB), were used to evaluate the precision of the GNSS PWV. The absolute values of the correlation coefficients are typically interpreted as very strong (0.8–1.0), strong (0.6–0.8), moderate (0.4–0.6), weak (0.2–0.4), or very weak or negligible (0.0–0.2) correlations. The RMSE measures the gap between the mean and true values of the test values. MB determines whether the test values are underestimated or overestimated. The formulas for the three evaluation metrics are as follows:
R = i = 1 N ( x x ¯ ) ( y y ¯ ) i = 1 N ( x x ¯ ) 2 i = 1 N ( y y ¯ ) 2
R M S E = 1 N i = 1 N ( x y ) 2
M B = 1 N i = 1 N ( x y )
where x is the test value, y is the validation value, x ¯ is the mean of the test value, y ¯ is the mean of the validation value, and N is the sample size.

3. Results and Discussion

3.1. Accuracy Assessment of GNSS PWV

Before exploring GNSS PWV, we used observations from radiosonde stations near the GNSS stations to evaluate the accuracy of GNSS PWV. Observations at radiosonde sites were performed twice daily. Therefore, only the GNSS PWV values at the corresponding times were adopted to align with the radiosonde PWV values.
Comparisons of the radiosonde and GNSS PWV time series at the six stations are shown in Figure 3. As can be seen from the subfigures, the regression line for each station nearly coincides with the one–one line. The correlation coefficients at the six stations range from 0.95 to 0.99, the RMSE values vary between 2.6 and 4.2 mm, and the MB values range from −1.2 to 0.5 mm. The results indicate that the GNSS PWV retrieval in this study has high accuracy, which is consistent with previous studies and meets the application requirements [58,59,60,61].

3.2. Spatial and Temporal Variations in GNSS PWV

3.2.1. Overall Characteristics of PWV in the Time Domain

Ten years of PWV time series were obtained by solving GNSS observations of each station from 1 January 2013 to 31 December 2022. Due to a significant number of missing observations, PWV datasets from two neighboring stations, CKSV and NCKU, were merged into a single time series, as were those from TCMS and TNML. The datasets from the BJNM, SHAO, and WUH2 stations were not considered in this section because of their short time spans. By utilizing Bayesian statistics, we extracted the time series of seasonality and trend components, as well as the trend probability of PWV, from the monthly average PWV data. The extracted components and trend probabilities of the GNSS PWV for each station are shown in Figure 4.
As can be seen from the first subplots of all stations, the monthly average GNSS PWV values reach their peaks during the summer months and drop to their valley values in winter. This pattern is primarily driven by the monsoon climate of eastern China, where winters are cold and dry and summers are warm and humid. As shown in Figure 4a,b, the BJFS and CHAN stations exhibit the highest peak values of approximately 40 mm, with a short peak duration. In contrast, as illustrated in Figure 4c–j, the stations in Wuhan, Hong Kong, Taiwan, and Kinmen County display higher peak values of approximately 60 mm, accompanied by a longer peak duration. Moreover, the valley values also exhibit obvious differences, as they are approximately 0–5 mm in Beijing and Changchun, 10–20 mm in Wuhan, and 20–30 mm in the other areas.
The inter-station disparities exhibit a statistically significant latitudinal dependence. This latitudinal gradient aligns with thermodynamic constraints: reduced solar irradiance reduces surface evaporation efficiency, and colder winter temperatures at higher latitudes suppress atmospheric moisture retention. Shanghai and Wuhan are located at lower latitudes than Beijing and Changchun, resulting in more sunlight. Meanwhile, Kinmen County, Hong Kong, and Taiwan, located closer to the equator, experience higher annual temperatures, which enhance the surface evaporation efficiency and increase the atmospheric moisture retention capacity, ultimately contributing to the observed PWV patterns.

3.2.2. Trend Variations in PWV Time Series

From the second subplots of Figure 4b,e,g,h,j, it is evident that the red envelopes are more pronounced, indicating that the PWV in Changchun, Hong Kong, northern Taiwan, and Kinmen County is likely to follow an increasing trend. This phenomenon has also been reported in other studies [27,62,63], indicating that the increasing trend may be associated with the long-term rise in temperature and human activities in recent years [62,64]. In contrast, Figure 4a,c,d,i show that the blue envelopes over Beijing, Wuhan, and southern Taiwan are prominent, suggesting that the PWV in these regions is likely to decrease.
To further explore the factors contributing to the declining trend in specific regions, we conducted an additional analysis from the perspective of natural conditions and atmospheric environments. For Beijing and Wuhan, two inland cities, we explored this from the perspective of variations in the origin of the water vapor. The China Land Cover Dataset (CLCD) for 2013 and 2022 was used to indicate land cover changes in Beijing and Wuhan. The CLCD categorizes land cover into nine classes, which we reclassified into two broad categories: impervious and non-impervious surfaces. Subsequently, a raster calculator was used to obtain the increase in impervious surfaces by 2022. Figure 5 presents the extent of impervious surfaces in 2013 and highlights the expansion of impervious areas in Wuhan and Beijing by 2022.
In Wuhan, which has experienced rapid urbanization over the past decade, the area of impervious surfaces in the city has increased by 26.83%. Rapid urbanization impacts the hydrological balance of a city [65], which can lead to urban dry and wet island effects [66]. This may reduce local evapotranspiration, decrease water vapor sources, and reduce atmospheric PWV. A previous study demonstrated that atmospheric humidity decreased with the rapid urbanization of Wuhan [67]. Simultaneously, Beijing has increased its impervious surface by 7.08% over the past ten years, resulting in lower humidity in the atmosphere [68,69]. Moreover, Beijing is located in a semi-arid region, and groundwater has been experiencing persistent drought over the past 10 years [70,71], leading to reduced groundwater evaporation through soil pores [72]. Therefore, increased urbanization and persistent groundwater droughts are likely to be the key drivers of the decreasing PWV trend in Beijing.
For tropical southern Taiwan, previous research has suggested that the decreasing trend may be attributed to changes in ENSO phases [73]. The trend of water vapor in the tropics during 1988–2003 may be related to interdecadal Pacific oscillation interdecadal variations from a warm period (1977–1998) to a cool period (1999–2003) [74]. Thus, we explored it from the ENSO perspective, utilizing the two most commonly used ENSO discriminant indices to characterize ENSO, namely, MEI and ONI (https://psl.noaa.gov/enso/, last access on 10 April 2024).
Figure 6 illustrates the MEI and ONI from 2013 to 2022 at a temporal resolution of one month. ENSO underwent several phase changes during the study period. A stronger El Niño event occurred during 2015–2016. However, in the following years, La Niña events occurred more frequently and lasted longer, especially during 2020–2022, when La Niña persisted for nearly two years. The La Niña phenomenon usually enhances easterly winds and upwelling, leading to a decrease in sea surface temperature [75], which reduces water vapor evaporation. Therefore, the PWV in southern Taiwan shows a downward trend.

3.3. Frequency-Domain Characterization of GNSS PWV

3.3.1. Annual/Semi-Annual Period of GNSS PWV

The time–frequency features of GNSS PWV from different areas were investigated using the Cmor wavelet. We downsampled the temporal resolution of the GNSS PWV time-series data to a daily scale and interpolated the missing values to obtain continuous datasets. To mitigate the boundary effects at both ends of the series, we extended the data on both sides and removed the extended wavelet transform coefficients, thereby generating wavelet coefficient plots of water vapor for each regional station.
Figure 7 shows the wavelet coefficient plots for BJFS, BJNM, CHAN, JFNG, WUH2, and WUHN, while those for SHAO, HKWL, HKWS, and KMNM are presented in Figure 8 and those for TWTF, TCMS, TNML, CKSV, and NCKU are shown in Figure 9.
The first column in Figure 7, Figure 8 and Figure 9 shows the contour maps of the modulus for the stations, where the horizontal and vertical coordinates indicate the observation time and timescale, respectively, which can determine whether there is a periodic change. The second column shows the wavelet variance of the wave energy distribution of the PWV time series at different scales for each station. The main period (timescale) was determined by calculating the wavelet variances. The third column shows the fluctuation in the change in the main periods.
According to column 1, there is a significant fluctuation at all stations in the vertical coordinate timescale of approximately 500–700 d. Moreover, for the BJFS, BJNM, CHAN, JFNG, WUH2, and WUHN stations, there is another fluctuation with a small amplitude and inconsistent fluctuation time in the timescale of approximately 200–300 d, whereas the other nine stations do not exhibit such a fluctuation. The exact first and second main periods for each station were identified by locating the amplitude peaks based on the variance trends shown in column 2. The main periods for all the stations are listed in Table 3. Except for the SHAO station, the first main periods at the other stations are concentrated around 550 d. Additionally, the second main periods at BJFS, BJNM, CHAN, JFNG, WUH2, and WUHN are nearly identical, at approximately 277–286 d.
The extracted first and second main periods of each station are further used to illustrate the wavelet coefficient fluctuations, as shown by the blue and red wavy lines in column 3, respectively. The first main period of each station exhibits fluctuations with a cycle of approximately one year. In the second main period at BJFS, BJNM, CHAN, JFNG, WUH2, and WUHN, a semi-annual (~6 month) cycle can be observed. However, its amplitude is relatively small, unstable, and occasionally variable compared to the annual cycle.
Therefore, it can be concluded that all stations across the seven regions exhibit a pronounced annual cycle with a significant amplitude. Additionally, inland areas, such as Beijing, Changchun, and Wuhan, also display semi-annual cycles of smaller amplitudes, which are absent in the coastal regions.
Based on the results of the annual/semi-annual period, it can be verified that the stations in coastal cities have only an annual period, whereas the stations in inland cities have both annual and semi-annual periods with low amplitudes. For coastal cities, water vapor is mainly dependent on the ocean, and the stabilizing and regulating effect of the ocean creates a continuous and smooth supply of water vapor, resulting in insignificant semi-annual cycle fluctuations. The ocean warms up more slowly than land in summer, creating low-pressure areas that may attract more water vapor inland; the opposite is true in winter. Moreover, periodic variations in monsoon activities (such as the East Asian monsoon) may also lead to phased increases or decreases in water vapor in inland regions during specific seasons (e.g., late spring to early summer and late autumn to early winter), resulting in semi-annual cycle fluctuations [76]. Therefore, sea–land thermal differences and monsoon systems may contribute to the phenomena where the semi-annual cycle is significant in inland cities but not in coastal cities.

3.3.2. Seasonal/Monthly Fluctuation Period of GNSS PWV

To identify the period on a smaller timescale, the time series of PWV in 2018 was selected for wavelet transformation to obtain wavelet coefficient maps of each regional station. Considering the continuity of the datasets, the BJNM, TCMS, and TNML stations were excluded from this part of the experiment. The wavelet coefficient plots of the stations in each region are shown in Figure 10 and Figure 11.
As shown in column 1 of Figure 10, there is a prominent wave peak and a secondary peak of smaller amplitude at BJFS and CHAN, corresponding to their first and second main periods, respectively. JFNG, WUH2, WUHN, and SHAO display three peaks of comparable amplitudes, suggesting more balanced periodic components. For HKSL, HKWS, KMNM, TWTF, CKSV, and NCKU, shown in column 1 of Figure 11, there are multiple large-amplitude and small peaks, indicating that the main periods are more complex in these areas. The main periods extracted for all the stations are listed in Table 4.
The wavelet coefficient fluctuations are plotted based on these identified dominant periods, as illustrated in column 3 of Figure 10 and Figure 11. At BJFS and CHAN, the first main period shows fluctuations with a seasonal cycle, whereas the second main period exhibits a shorter fluctuation of approximately one month. At JFNG, WUH2, WUHN, and SHAO, fluctuations with approximately two-month, forty-day, and one-month cycles can be identified. Several medium- and short-period fluctuations can be observed at stations with multiple complex dominant periods, such as HKSL, HKWS, KMNM, TWTF, CKSV, and NCKU.
In summary, cycles of seasonality and one month exist in Beijing and Changchun. Shanghai and Wuhan display two-month, forty-day and one-month cycles, while multiple medium and short periods are exhibited in Kinmen County, Hong Kong, and Taiwan. According to the findings for the seasonal/monthly period, the medium-term periodic characteristics of PWV demonstrate a clear latitude-dependent pattern.
At higher latitudes such as Beijing and Changchun, PWV cycles tend to be longer and more stable. This is largely due to the dominance of large-scale, stable atmospheric circulation systems such as the polar vortex and westerlies, which introduce regular seasonal variations [77,78]. In these regions, the relatively flat terrain also provides minimal interference with atmospheric motion, facilitating the formation of regular oscillations.
In contrast, mid- and low-latitude regions experience complex and irregular PWV variations. Mid-latitude areas, such as Shanghai and Wuhan, lie in transitional zones influenced by a mixture of weather systems, including the subtropical high, the Meiyu front, cold fronts, and typhoons. These interacting systems generate multiscale fluctuations, such as two-month plum rain cycles and forty-day typhoon intervals. In low-latitude regions such as Kinmen, Hong Kong, and Taiwan, atmospheric dynamics are dominated by tropical convection and short-lived systems, such as thunderstorms and tropical cyclones. Furthermore, water vapor transport is more stable and zonally directed at higher latitudes, while at lower latitudes it is dominated by disorganized local convection, further weakening identifiable cycles [79].

3.4. Short-Time Frequency-Domain Characterizations of GNSS PWV

To further investigate the short-term frequency-domain characteristics of PWV, time-series data from 2018 were selected and processed using the FFT method to extract the frequency-domain variations in PWV at each station. As shown in Figure 12, the vertical coordinates represent the amplitude and the horizontal coordinates represent the time spans from 0 to 2 d (0–48 h), which corresponds to a time frequency of 0.5~∞ cpd.
The stations located in the Shanghai and northern Taiwan areas have only a significant 1 d period (diurnal) in the frequency domain. In contrast, stations in Wuhan, Hong Kong, and Kinmen County display prominent period changes of 0.5 d (semi-diurnal). Stations in Beijing and southern Taiwan demonstrate both 0.5 d and 1 d periodic components, with the amplitude of the diurnal period being greater than that of the semi-diurnal period. Compared with these regions, the frequency-domain characteristics of PWV in the Changchun area exhibit minimal variation and lack any prominently amplified signals, reflecting more stable short-term fluctuations.
In summary, except for Changchun, GNSS stations in six typical cities display pronounced periodic variations in PWV at semi-diurnal, diurnal, or both frequencies. In contrast, Changchun exhibits insignificant variations in the frequency domain.
This may be because diurnal water vapor cycles are primarily driven by solar radiation. Nonetheless, other factors, such as temperature, local atmospheric circulation, vertical air motion, evapotranspiration, condensation, and precipitation, also modulate short-term variability [80]. Since solar radiation is generally stronger at lower latitudes, water vapor responds more sharply, with multiple fluctuation periods occurring within a single day.
Changchun, located at mid-to-high latitudes, has a temperate continental monsoon climate and receives less solar radiation. This limits evaporation and reduces atmospheric moisture. Although eastern and southern Changchun are not far from the ocean, moist summer winds are weakened by the Changbai Mountain. In winter, the open terrain of the Songliao Plain facilitates the dominance of cold, dry Siberian air masses. Together, these geographic and climatic factors limit the occurrence of short-term PWV periodicities in the Changchun region.

3.5. The Effect of Atmospheric Temperature on PWV

To further explore the main factors driving the periodic characteristics of PWV, experiments focusing on atmospheric temperature were conducted. Similarly, we selected the 10-year and 2018 time series to apply wavelet transforms to the temperature (T2m and Th) data of the representative stations. For Th, the stratified air temperature level with the highest correlation with the PWV at each station was selected based on the correlation analysis presented in Table 5.
As a result, the temperature at 600 hPa was selected for BJFS, CHAN, JFNG, and SHAO, whereas the temperatures at 850 hPa were chosen for HKSL, TWTF, CKSV, and KMNM in the wavelet transform analysis. The wavelet coefficient plots for the 10-year temperatures (T2m and Th) are presented in Figure 13 and Figure 14, respectively. As shown, all stations exhibit one main period of approximately 550 d, corresponding to an annual variation cycle. Notably, no significant semi-annual periodic components can be observed in the temperature data for any of the stations.
The wavelet coefficient plots of temperature (T2m and Th) in 2018 for the BJFS, CHAN, JFNG, and SHAO stations are presented in Figure 15, whereas those for the HKSL, TWTF, CKSV, and KMNM stations are shown in Figure 16. By comparing Figure 15 and Figure 16 with Figure 8 and Figure 9, it can be observed that the PWV characteristics of BJFS and CHAN align with the wavelet coefficient patterns of T2m, both exhibiting variability periods of approximately four months and one month. For the CKSV station, PWV characteristics on the timescale of 1–120 d is also more consistent with the wavelet coefficient patterns of T2m at 1–120 d. For the other stations, their PWV features on the timescale of 1–120 d conform to the wavelet coefficient patterns of Th at 1–120 d.
Temperature affects the atmospheric PWV in two ways. On the one hand, when the temperature near the surface rises, surface water evaporation accelerates, leading to an increase in PWV. On the other hand, when the relative humidity in the mid–lower troposphere (especially over oceanic regions) remains constant, the amount of water vapor increases with increasing air temperature. According to the Clausius–Clapeyron equation, the water vapor content can increase by 6–7% per 1 K increase in temperature [73,81]. As the temperature rises, water molecules move more rapidly, allowing more molecules to overcome the surface tension of the liquid and enter the atmosphere, causing the saturated water vapor pressure to rise. This also directly drives the actual water vapor pressure to increase in the same proportion to maintain the dynamic equilibrium between water vapor and liquid water.
The periodic characteristics of PWV in Beijing and Changchun are consistent with those of T2m, suggesting that PWV in these two regions is predominantly influenced by temperature-driven surface water evaporation. The characteristics of southern Taiwan within the timescale of 1–120 d are consistent with T2m, illustrating that the medium- and short-term periods of this region are also predominantly influenced by temperature-driven surface water evaporation. Conversely, in Wuhan, Shanghai, Hong Kong, northern Taiwan, and Kinmen County, the PWV variation conforms to the Th patterns of the strongest correlation coefficient at the pressure level in the timescale of 1–120 d. This indicates that, in these regions, the PWV is more affected by the second mechanism within this timescale.
The annual PWV period aligns with the temperature variation patterns, although the affecting mechanism remains unclear. For Beijing and Changchun, the seasonal and monthly periods are affected by temperature-accelerated surface water evaporation, and medium and short periods in southern Taiwan are also affected by this mechanism. However, for the other regions, except for the seasonal period, the medium and short periods are affected by temperatures which enhance the moisture retention capacity of the atmosphere, suggesting that seasonal periods may be affected by more complex weather systems. In summary, these findings indicate that temperature plays a crucial role in regulating annual, medium-, and short-term PWV variation periods across the major regions of eastern China.

4. Limitations and Prospects

In this study, we focused on major cities in eastern China to investigate PWV variability. While this focus was partly driven by the region’s economic development and high population density, it also imposed notable limitations. China is a vast country with huge differences in climate conditions, topography, and atmospheric circulation in different regions [82], all of which shape distinct PWV distribution characteristics, changing patterns, and influencing mechanisms. Therefore, the findings of this study may not be fully applicable to other regions not included in the study.
Moreover, PWV, a key parameter in the field of meteorology and hydrology, is affected by many complex factors, including atmospheric physical processes, geographic environmental features, and human activities. However, our study primarily considered the effect of temperature on PWV and only at a qualitative level. The absence of a quantitative assessment limits a deeper understanding of the temperature–PWV relationship in this region.
Future research should broaden the scope of this study to encompass all Chinese cities in the research field, comprehensively consider the mechanisms of multiple influencing factors on PWV, and deepen studies on the quantitative relationship between these factors and PWV. Notably, abrupt changes in PWV have been identified as precursors to rainfall events [83]. Previous studies have shown that PWV typically rises sharply within 0–12 h before precipitation onset, maintains high-level fluctuations during the event, and rapidly returns to baseline levels afterwards [84]. A deeper understanding of these dynamics will support more accurate characterization of PWV variability and provide a more reliable scientific basis for the fields of meteorological forecasting, water resource management, and climate change research.

5. Conclusions

In this study, GNSS observations from IGS stations between 2013 and 2022 were utilized to retrieve PWV data across representative regions in eastern China. By integrating long-term GNSS-derived PWV time series with multiple frequency-domain analysis methods (Bayesian statistics, wavelet transform, and FFT), the multiscale time–frequency characteristics of PWV were described in Changchun, Beijing, Shanghai, Wuhan, Taiwan, Hong Kong, and Kinmen County. Additionally, the vertical atmospheric temperature was uniquely incorporated to explore the driving mechanisms behind the regional PWV variability. The following findings can be drawn:
(1) Subregional declines in PWV can be identified in Wuhan, Beijing, and southern Taiwan—contrasting with the increasing trends in other regions. In Wuhan, rapid urbanization has driven the growth of impervious surfaces, resulting in the suppression of local atmospheric moisture levels. In Beijing, the decline is primarily driven by a combination of groundwater depletion and urbanization, both of which contribute to reduced local evapotranspiration. In southern Taiwan, this decline is primarily driven by ENSO-related climate anomalies, which disrupt regional moisture transport patterns.
(2) The PWV periodicities exhibit clear latitudinal and coastal–inland gradients. Inland stations (Beijing, Changchun, and Wuhan) show weaker annual and semi-annual signals, whereas coastal stations (Shanghai, Kinmen County, Hong Kong, and Taiwan) primarily exhibit annual cycles. Higher-latitude stations display stronger seasonal and monthly oscillations, mid-latitude stations present intermediate-scale variations, and low-latitude regions exhibit richer medium- and short-term fluctuations.
(3) Pronounced diurnal and semi-diurnal PWV oscillations can be observed in six typical cities, while Changchun exhibits minimal short-term fluctuations—likely due to its northern latitude and complex topographic influences.
(4) Temperature plays a key role in modulating PWV periodicities. The annual PWV cycle is closely aligned with temperature fluctuations. In Beijing, Changchun, and southern Taiwan, near-surface temperature drives groundwater evaporation, amplifying seasonal signals. In contrast, in other regions, the medium- and short-term PWV variations are influenced by temperature-driven increases in atmospheric moisture capacity, indicating more complex thermodynamic interactions.
Overall, this study offers solid data support and an analytical framework for discovering the geographic and climatic driving mechanisms underlying water vapor variations. These findings provide valuable insights for enhancing weather and climate modelling, sustainable urban planning, and regional drought and flood early warning.

Author Contributions

Conceptualization, T.Z. and M.H.; methodology, S.H., J.X. and L.Z.; formal analysis, J.X., T.Z., W.Z. and Y.X.; writing—original draft preparation, J.X. and T.Z.; writing—review and editing, M.H., S.H., W.Z. and T.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (42301469, 42201438); the Natural Science Foundation of Jiangxi, China (20224BAB202039); the Program of China Scholarship Council (202308360254); the Science and Technology Research Project of Jiangxi Provincial Department of Education (GJJ2200328, GJJ2200327); and the Youth Program of Major Discipline Academic and Technical Leaders Training Program of Jiangxi Talents Supporting Project (20232BCJ23086).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The IGS observational data that supported this study were collected from the IGS Data Center at Wuhan University and are openly available at http://www.igs.gnsswhu.cn/index.php (accessed on 10 December 2023). Raw radiosonde data were obtained from the University of Wyoming, Russia, at http://weather.uwyo.edu/upperair/sounding.html (accessed on 5 March 2024). Temperature reanalysis data were provided by the European Centre for Medium-Range Weather Forecasts (ECMWF) Fifth Generation Atmospheric Reanalysis dataset for Global Climate, available from https://cds.climate.copernicus.eu/datasets (accessed on 20 March 2024).

Acknowledgments

The authors would like to thank the IGS, the University of Wyoming, and the ECMWF for providing the datasets used in this study. We appreciate the constructive suggestions and comments from the editorial team and anonymous reviewers.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
PWVPrecipitable Water Vapor
GNSSGlobal Navigation Satellite System
IGSInternational GNSS Service
FFTFast Fourier Transform
ERA5ECMWF Reanalysis v5
ZHDZenith Hydrostatic Delay
ZTDZenith Tropospheric Delay
ZWDZenith Wet Delay

References

  1. Charlesworth, E.; Plöger, F.; Birner, T.; Baikhadzhaev, R.; Abalos, M.; Abraham, N.L.; Akiyoshi, H.; Bekki, S.; Dennison, F.; Jöckel, P.; et al. Stratospheric water vapor affecting atmospheric circulation. Nat. Commun. 2023, 14, 3925. [Google Scholar] [CrossRef] [PubMed]
  2. Zhu, H.; Chen, K.; Hu, S.; Liu, J.; Shi, H.; Wei, G.; Chai, H.; Li, J.; Wang, T. Using the Global Navigation Satellite System and Precipitation Data to Establish the Propagation Characteristics of Meteorological and Hydrological Drought in Yunnan, China. Water Resour. Res. 2023, 59, e2022WR033126. [Google Scholar] [CrossRef]
  3. Zhang, J.; Chen, H.; Fang, X.; Yin, Z.; Hu, R. Warming-induced hydrothermal anomaly over the Earth’s three Poles amplifies concurrent extremes in 2022. NPJ Clim. Atmos. Sci. 2024, 7, 8. [Google Scholar] [CrossRef]
  4. Zhou, X.; Li, Y.; Xiao, C.; Chen, W.; Mei, M.; Wang, G. High-impact Extreme Weather and Climate Events in China: Summer 2024 Overview. Adv. Atmos. Sci. 2025, 42, 1064–1076. [Google Scholar] [CrossRef]
  5. Liu, X.; Wang, Y.; Huang, J.; Yu, T.; Jiang, N.; Yang, J.; Zhan, W. Assessment and calibration of FY-4A AGRI total precipitable water products based on CMONOC. Atmos. Res. 2022, 271, 106096. [Google Scholar] [CrossRef]
  6. Jiang, P.; Liu, R.; Huo, Y.; Wu, Y.; Ye, S.; Wang, S.; Mu, X.; Zhu, L. Retrieving the Atmospheric Water Vapor Profile Combining FY-4A/GIIRS and Ground-Based GNSS PWV in Hong Kong Region. IEEE Trans. Geosci. Remote Sens. 2025, 63, 1–9. [Google Scholar] [CrossRef]
  7. Wang, J.; Liu, Z. Improving GNSS PPP accuracy through WVR PWV augmentation. J. Geod. 2019, 93, 1685–1705. [Google Scholar] [CrossRef]
  8. Sun, Q.; Vihma, T.; Jonassen, M.O.; Zhang, Z. Impact of Assimilation of Radiosonde and UAV Observations from the Southern Ocean in the Polar WRF Model. Adv. Atmos. Sci. 2020, 37, 441–454. [Google Scholar] [CrossRef]
  9. Dessler, A.E.; Davis, S.M. Trends in tropospheric humidity from reanalysis systems. J. Geophys. Res.-Atmos. 2010, 115, D19127. [Google Scholar] [CrossRef]
  10. Li, W.; Yao, Y.; Zhang, L.; Peng, W.; Du, Z.; Zuo, Y.; Wang, W. Evaluating global precipitable water vapor products from four public reanalysis using radiosonde data. J. Geod. Geodyn. 2025, in press. [Google Scholar] [CrossRef]
  11. Mo, Z.; Zeng, Z.; Huang, L.; Liu, L.; Huang, L.; Zhou, L.; Ren, C.; He, H. Investigation of Antarctic Precipitable Water Vapor Variability and Trend from 18 Year (2001 to 2018) Data of Four Reanalyses Based on Radiosonde and GNSS Observations. Remote Sens. 2021, 13, 3901. [Google Scholar] [CrossRef]
  12. Ssenyunzi, R.C.; Oruru, B.; D’ujanga, F.M.; Realini, E.; Barindelli, S.; Tagliaferro, G.; von Engeln, A.; van de Giesen, N. Performance of ERA5 data in retrieving Precipitable Water Vapour over East African tropical region. Adv. Space Res. 2020, 65, 1877–1893. [Google Scholar] [CrossRef]
  13. Wang, S.; Xu, T.; Nie, W.; Jiang, C.; Yang, Y.; Fang, Z.; Li, M.; Zhang, Z. Evaluation of Precipitable Water Vapor from Five Reanalysis Products with Ground-Based GNSS Observations. Remote Sens. 2020, 12, 1817. [Google Scholar] [CrossRef]
  14. Zhao, T.; Fu, C.; Ke, Z.; Guo, W. Global Atmosphere Reanalysis Datasets: Current Status and Recent Advances. Adv. Earth Sci. 2010, 25, 241–254. [Google Scholar]
  15. Bevis, M.; Businger, S.; Herring, T.A.; Rocken, C.; Anthes, R.A.; Ware, R.H. GPS meteorology: Remote sensing of atmospheric water vapor using the global positioning system. J. Geophys. Res.-Atmos. 1992, 97, 15787–15801. [Google Scholar] [CrossRef]
  16. Dong, C.; Yu, F.; Zhang, W.; Fang, L.; Wei, K.; Lou, Y.; Ou, S. An Improved Rainfall Forecasting Method with GNSS-PWV Three Factor Threshold. Geomat. Inf. Sci. Wuhan Univ. 2025, 50, 866–873. [Google Scholar] [CrossRef]
  17. Wang, Z.; Chai, H.; Zheng, N.; Ming, L.; Chen, P. Research on filling missing GNSS precipitable water vapor time series data using the PSORF model combined with reanalysis datasets. Meas. Sci. Technol. 2025, 36, 016014. [Google Scholar] [CrossRef]
  18. Borger, C.; Beirle, S.; Wagner, T. Analysis of global trends of total column water vapour from multiple years of OMI observations. Atmos. Chem. Phys. 2022, 22, 10603–10621. [Google Scholar] [CrossRef]
  19. Kim, S.; Sharma, A.; Wasko, C.; Nathan, R. Linking Total Precipitable Water to Precipitation Extremes Globally. Earth Future 2022, 10, e2021EF002473. [Google Scholar] [CrossRef]
  20. Parracho, A.C.; Bock, O.; Bastin, S. Global IWV trends and variability in atmospheric reanalyses and GPS observations. Atmos. Chem. Phys. 2018, 18, 16213–16237. [Google Scholar] [CrossRef]
  21. Wan, N.; Lin, X.; Pielke, R.A., Sr.; Zeng, X.; Nelson, A.M. Global total precipitable water variations and trends over the period 1958–2021. Hydrol. Earth Syst. Sci. 2024, 28, 2123–2137. [Google Scholar] [CrossRef]
  22. Foster, J.; Bevis, M.; Raymond, W. Precipitable water and the lognormal distribution. J. Geophys. Res.-Atmos. 2006, 111, D15102. [Google Scholar] [CrossRef]
  23. Liu, Y.; Zhang, B.; Yao, Y.; Zhao, Q.; Xu, C.; Yan, X.; Zhang, L. Revealing the spatiotemporal patterns of water vapor and its link to North Atlantic Oscillation over Greenland using GPS and ERA5 data. Sci. Total Environ. 2024, 918, 170596. [Google Scholar] [CrossRef] [PubMed]
  24. Khutorova, O.G.; Khutorov, V.E.; Teptin, G.M. Tropospheric Water Vapor Long-term Periodicities and ENSO Relation in European Territory of Russia. Earth Space Sci. 2019, 6, 2480–2486. [Google Scholar] [CrossRef]
  25. Jadala, N.B.; Sridhar, M.; Dashora, N.; Dutta, G. Annual, seasonal and diurnal variations of integrated water vapor using GPS observations over Hyderabad, a tropical station. Adv. Space Res. 2020, 65, 529–540. [Google Scholar] [CrossRef]
  26. Liu, Y.; Wang, Y.; Ding, K.; Liu, X.; Zhan, W. Short Term Frequency Domain Characteristics of GNSS PWV Based on CMONOC. J. Geod. Geodyn. 2021, 41, 1118–1122. [Google Scholar]
  27. Yang, F.; Gong, X.; Li, Z.; Wang, Y.; Song, S.; Wang, H.; Chen, R. Spatiotemporal distribution and impact factors of GNSS-PWV in China based on climate region. Adv. Space Res. 2024, 73, 4187–4201. [Google Scholar] [CrossRef]
  28. Wang, F.; Liu, C.; Xu, Y. Analyzing Population Density Disparity in China with GIS-automated Regionalization: The Hu Line Revisited. Chin. Geogr. Sci. 2019, 29, 541–552. [Google Scholar] [CrossRef]
  29. Pielke, R., Jr.; Burgess, M.G.; Ritchie, J. Plausible 2005–2050 emissions scenarios project between 2 °C and 3 °C of warming by 2100. Environ. Res. Lett. 2022, 17, 024027. [Google Scholar] [CrossRef]
  30. Ferreira, A.P.; Gimeno, L. Determining precipitable water vapour from upper-air temperature, pressure and geopotential height. Q. J. R. Meteorol. Soc. 2024, 150, 484–522. [Google Scholar] [CrossRef]
  31. Gong, S.; Fiifi Hagan, D.; Lu, J.; Wang, G. Validation on MERSI/FY-3A precipitable water vapor product. Adv. Space Res. 2018, 61, 413–425. [Google Scholar] [CrossRef]
  32. Liu, H.; Tang, S.; Zhang, S.; Hu, J. Evaluation of MODIS water vapour products over China using radiosonde data. Int. J. Remote Sens. 2015, 36, 680–690. [Google Scholar] [CrossRef]
  33. Hersbach, H.; Bell, B.; Berrisford, P.; Hirahara, S.; Horányi, A.; Muñoz-Sabater, J.; Nicolas, J.; Peubey, C.; Radu, R.; Schepers, D.; et al. The ERA5 global reanalysis. Q. J. R. Meteorol. Soc. 2020, 146, 1999–2049. [Google Scholar] [CrossRef]
  34. Glaser, S.; Fritsche, M.; Sośnica, K.; Rodríguez-Solano, C.J.; Wang, K.; Dach, R.; Hugentobler, U.; Rothacher, M.; Dietrich, R. A consistent combination of GNSS and SLR with minimum constraints. J. Geod. 2015, 89, 1165–1180. [Google Scholar] [CrossRef]
  35. Geng, J.; Wen, Q.; Zhang, Q.; Li, G.; Zhang, K. GNSS observable-specific phase biases for all-frequency PPP ambiguity resolution. J. Geod. 2022, 96, 11. [Google Scholar] [CrossRef]
  36. Geng, J.; Zhang, Q.; Li, G.; Liu, J.; Liu, D. Observable-specific phase biases of Wuhan multi-GNSS experiment analysis center’s rapid satellite products. Satell. Navig. 2022, 3, 23. [Google Scholar] [CrossRef]
  37. Zeng, J.; Geng, J.; Li, G.; Tang, W. Improving cycle slip detection in ambiguity-fixed precise point positioning for kinematic LEO orbit determination. GPS Solut. 2024, 28, 135. [Google Scholar] [CrossRef]
  38. Li, X.; Han, X.; Li, X.; Liu, G.; Feng, G.; Wang, B.; Zheng, H. GREAT-UPD: An open-source software for uncalibrated phase delay estimation based on multi-GNSS and multi-frequency observations. GPS Solut. 2021, 25, 66. [Google Scholar] [CrossRef]
  39. Li, X.; Yuan, L.; Li, X.; Huang, J.; Shen, Z.; Tan, Y. GREAT-PVT: An open-source software for multi-frequency and multi-GNSS PPP-AR and RTK. GPS Solut. 2025, 29, 157. [Google Scholar] [CrossRef]
  40. Li, X.; Zheng, H.; Li, X.; Yuan, Y.; Wu, J.; Han, X. Open-source software for multi-GNSS inter-frequency clock bias estimation. GPS Solut. 2023, 27, 84. [Google Scholar] [CrossRef]
  41. Xie, S.; Zhang, P.; Wang, X. Influence of Dynamic Mapping Functions on GPS Baseline Quality. J. Geod. Geodyn. 2017, 37, 192–195. [Google Scholar]
  42. Saastamoinen, J. Atmospheric Correction for the Troposphere and Stratosphere in Radio Ranging Satellites. In The Use of Artificial Satellites for Geodesy; American Geophysical Union: Washington, DC, USA, 1972; Volume 15, pp. 247–251. [Google Scholar]
  43. Bevis, M.; Businger, S.; Chiswell, S.; Herring, T.A.; Anthes, R.A.; Rocken, C.; Ware, R.H. GPS Meteorology: Mapping Zenith Wet Delays onto Precipitable Water. J. Appl. Meteorol. Climatol. 1994, 33, 379–386. [Google Scholar] [CrossRef]
  44. Boehm, J.; Heinkelmann, R.; Schuh, H. Short Note: A global model of pressure and temperature for geodetic applications. J. Geod. 2007, 81, 679–683. [Google Scholar] [CrossRef]
  45. Sun, J.; Fan, S.; Zang, J.; Zhou, C.; Liu, Y. Accuracy comparison and analysis for three GPT models. Sci. Surv. Mapp. 2018, 43, 63–67,75. [Google Scholar]
  46. Zhou, T.; Popescu, S.C.; Lawing, A.M.; Eriksson, M.; Strimbu, B.M.; Bürkner, P.C. Bayesian and Classical Machine Learning Methods: A Comparison for Tree Species Classification with LiDAR Waveform Signatures. Remote Sens. 2017, 10, 39. [Google Scholar] [CrossRef]
  47. Zhao, K.; Valle, D.; Popescu, S.; Zhang, X.; Mallick, B. Hyperspectral remote sensing of plant biochemistry using Bayesian model averaging with variable and band selection. Remote Sens. Environ. 2013, 132, 102–119. [Google Scholar] [CrossRef]
  48. Li, J.; Li, Z.-L.; Wu, H.; You, N. Trend, seasonality, and abrupt change detection method for land surface temperature time-series analysis: Evaluation and improvement. Remote Sens. Environ. 2022, 280, 113222. [Google Scholar] [CrossRef]
  49. Zhao, K.; Wulder, M.A.; Hu, T.; Bright, R.; Wu, Q.; Qin, H.; Li, Y.; Toman, E.; Mallick, B.; Zhang, X.; et al. Detecting change-point, trend, and seasonality in satellite time series data to track abrupt changes and nonlinear dynamics: A Bayesian ensemble algorithm. Remote Sens. Environ. 2019, 232, 111181. [Google Scholar] [CrossRef]
  50. White, J.H.R.; Walsh, J.E.; Thoman, R.L., Jr. Using Bayesian statistics to detect trends in Alaskan precipitation. Int. J. Climatol. 2021, 41, 2045–2059. [Google Scholar] [CrossRef]
  51. Morlet, J.; Arens, G.; Fourgeau, E.; Giard, D. Wave propagation and sampling theory; Part I, Complex signal and scattering in multilayered media. Geophysics 1982, 47, 203–221. [Google Scholar] [CrossRef]
  52. Morlet, J.; Arens, G.; Fourgeau, E.; Giard, D. Wave propagation and sampling theory; Part II, Sampling theory and complex waves. Geophysics 1982, 47, 222–236. [Google Scholar] [CrossRef]
  53. Li, L.; Song, Y.; Zhou, J.L. Preliminary Exploration of GNSS Meteorological Elements Using Wavelet Transform for Rainstorm Prediction. J. Geod. Geodyn. 2020, 40, 225–230. [Google Scholar]
  54. Wang, Y.; Liu, B.; Liu, Y. Correlation Analysis of GPS PWV and Meteorological Elements Based on Wavelet Transform. J. Geod. Geodyn. 2017, 37, 721–725. [Google Scholar]
  55. Grossmann, A.; Morlet, J. Decomposition of Hardy Functions into Square Integrable Wavelets of Constant Shape. SIAM J. Math. Anal. 1984, 15, 723–736. [Google Scholar] [CrossRef]
  56. Kumar, P.; Foufoula-Georgiou, E. Wavelet analysis for geophysical applications. Rev. Geophys. 1997, 35, 385–412. [Google Scholar] [CrossRef]
  57. Cooley, J.W.; Tukey, J.W. An Algorithm for the Machine Calculation of Complex Fourier Series. Math. Comput. 1965, 19, 297–301. [Google Scholar] [CrossRef]
  58. Shi, H.; Zhang, R.; Nie, Z.; Li, Y.; Chen, Z.; Wang, T. Research on variety characteristics of mainland China troposphere based on CMONOC. J. Geod. Geodyn. 2018, 9, 411–417. [Google Scholar] [CrossRef]
  59. Wang, J.; Zhang, L. Climate applications of a global, 2-hourly atmospheric precipitable water dataset derived from IGS tropospheric products. J. Geod. 2009, 83, 209–217. [Google Scholar] [CrossRef]
  60. Zhao, Q.; Yao, Y.; Yao, W.; Zhang, S. GNSS-derived PWV and comparison with radiosonde and ECMWF ERA-Interim data over mainland China. J. Atmos. Sol.-Terr. Phys. 2019, 182, 85–92. [Google Scholar] [CrossRef]
  61. Zhao, Y.; Zhou, S.; Wang, S.; Sun, J.; San, X. The Error Analysis for the Remote Sensing of Water Vapor Data by Ground Based GPS in Tengchong, Yunnan Province. J. Geosci. Environ. Prot. 2019, 7, 231–245. [Google Scholar] [CrossRef]
  62. Shangguan, S.; Lin, H.; Wei, Y.; Tang, C. Spatiotemporal Modes Characteristics and SARIMA Prediction of Total Column Water Vapor over China during 2002–2022 Based on AIRS Dataset. Atmosphere 2022, 13, 885. [Google Scholar] [CrossRef]
  63. Wu, M.; Jin, S.; Li, Z.; Cao, Y.; Ping, F.; Tang, X. High-Precision GNSS PWV and Its Variation Characteristics in China Based on Individual Station Meteorological Data. Remote Sens. 2021, 13, 1296. [Google Scholar] [CrossRef]
  64. Zhang, J.; Zhao, T.; Dai, A.; Zhang, W. Detection and Attribution of Atmospheric Precipitable Water Changes since the 1970s over China. Sci. Rep. 2019, 9, 17609. [Google Scholar] [CrossRef] [PubMed]
  65. Willett, K.M.; Gillett, N.P.; Jones, P.D.; Thorne, P.W. Attribution of observed surface humidity changes to human influence. Nature 2007, 449, 710–712. [Google Scholar] [CrossRef] [PubMed]
  66. Luo, Z.; Liu, J.; Zhang, Y.; Zhou, J.; Yu, Y.; Jia, R. Spatiotemporal characteristics of urban dry/wet islands in China following rapid urbanization. J. Hydrol. 2021, 601, 126618. [Google Scholar] [CrossRef]
  67. Liu, S. Assessment of Regional Climate Effects of Urbanization around Subtropical City Wuhan in Summer Using Numerical Modeling. Atmosphere 2024, 15, 185. [Google Scholar] [CrossRef]
  68. Li, X.; Fan, W.; Wang, L.; Luo, M.; Yao, R.; Wang, S.; Wang, L. Effect of urban expansion on atmospheric humidity in Beijing-Tianjin-Hebei urban agglomeration. Sci. Total Environ. 2021, 759, 144305. [Google Scholar] [CrossRef] [PubMed]
  69. Luan, Q.; Cao, Q.; Huang, L.; Liu, Y.; Wang, F. Identification of the Urban Dry Islands Effect in Beijing: Evidence from Satellite and Ground Observations. Remote Sens. 2022, 14, 809. [Google Scholar] [CrossRef]
  70. Chen, Y.; Zhang, Y.; Tian, J.; Tang, Z.; Wang, L.; Yang, X. Understanding the Propagation of Meteorological Drought to Groundwater Drought: A Case Study of the North China Plain. Water 2024, 16, 501. [Google Scholar] [CrossRef]
  71. Liu, Q.; Zhang, X.; Xu, Y.; Li, C.; Zhang, X.; Wang, X. Characteristics of groundwater drought and its correlation with meteorological and agricultural drought over the North China Plain based on GRACE. Ecol. Indic. 2024, 161, 111925. [Google Scholar] [CrossRef]
  72. Sprenger, M.; Leistert, H.; Gimbel, K.; Weiler, M. Illuminating hydrological processes at the soil-vegetation-atmosphere interface with water stable isotopes. Rev. Geophys. 2016, 54, 674–704. [Google Scholar] [CrossRef]
  73. Trenberth, K.E.; Fasullo, J.; Smith, L. Trends and variability in column-integrated atmospheric water vapor. Clim. Dyn. 2005, 24, 741–758. [Google Scholar] [CrossRef]
  74. Dong, B.; Dai, A. The influence of the Interdecadal Pacific Oscillation on Temperature and Precipitation over the Globe. Clim. Dyn. 2015, 45, 2667–2681. [Google Scholar] [CrossRef]
  75. Feng, L.; Zhang, R.-H.; Yu, B.; Han, X. Roles of Wind Stress and Subsurface Cold Water in the Second-Year Cooling of the 2017/18 La Niña Event. Adv. Atmos. Sci. 2020, 37, 847–860. [Google Scholar] [CrossRef]
  76. Sherwood, S.C.; Roca, R.; Weckwerth, T.M.; Andronova, N.G. Tropospheric water vapor, convection, and climate. Rev. Geophys. 2010, 48, RG2001. [Google Scholar] [CrossRef]
  77. Nakamura, M.; Miyama, T. Impacts of the Oyashio Temperature Front on the Regional Climate. J. Clim. 2014, 27, 7861–7873. [Google Scholar] [CrossRef]
  78. Xie, Y.; Wei, F.; Chen, G.; Zhang, T.; Hu, L. Analysis of the 2008 heavy snowfall over South China using GPS PWV measurements from the Tibetan Plateau. Ann. Geophys. 2010, 28, 1369–1376. [Google Scholar] [CrossRef]
  79. Kim, Y.-J.; Jee, J.-B.; Lim, B. Investigating the Influence of Water Vapor on Heavy Rainfall Events in the Southern Korean Peninsula. Remote Sens. 2023, 15, 340. [Google Scholar] [CrossRef]
  80. Wang, X.; Zhang, Q.; Zhang, S. Periodic Oscillation Analysis of GPS Water Vapor Time Series Using Combined Algorithm Based on EMD and WD. Geomat. Inf. Sci. Wuhan Univ. 2018, 43, 620–628. [Google Scholar] [CrossRef]
  81. O’Gorman, P.A.; Muller, C.J. How closely do changes in surface and column water vapor follow Clausius–Clapeyron scaling in climate change simulations? Environ. Res. Lett. 2010, 5, 025207. [Google Scholar] [CrossRef]
  82. Shi, P.; Sun, S.; Wang, M.; Li, N.; Wang, J.A.; Jin, Y.; Gu, X.; Yin, W. Climate change regionalization in China (1961–2010). Sci. China Earth Sci. 2014, 57, 2676–2689. [Google Scholar] [CrossRef]
  83. Xian, T.; Su, K.; Zhang, J.; Hu, H.; Wang, H. Precipitable Water Vapor Retrieval Based on GNSS Data and Its Application in Extreme Rainfall. Remote Sens. 2025, 17, 2301. [Google Scholar] [CrossRef]
  84. Zhao, Q.; Liu, Y.; Yao, W.; Yao, Y. Hourly Rainfall Forecast Model Using Supervised Learning Algorithm. IEEE Trans. Geosci. Remote Sens. 2022, 60, 1–9. [Google Scholar] [CrossRef]
Figure 1. Distribution of (a) all GNSS stations and (b) GNSS stations in southeastern China.
Figure 1. Distribution of (a) all GNSS stations and (b) GNSS stations in southeastern China.
Sustainability 17 06699 g001
Figure 2. Locations of the selected GNSS stations and radiosonde sites. In each city, the radiosonde site is located within 50 km of the corresponding GNSS station.
Figure 2. Locations of the selected GNSS stations and radiosonde sites. In each city, the radiosonde site is located within 50 km of the corresponding GNSS station.
Sustainability 17 06699 g002
Figure 3. Comparison of GNSS PWV and radiosonde PWV time series at the six stations. The black and red lines represent the one–one lines and linear regression lines, respectively.
Figure 3. Comparison of GNSS PWV and radiosonde PWV time series at the six stations. The black and red lines represent the one–one lines and linear regression lines, respectively.
Sustainability 17 06699 g003
Figure 4. Time-series plots of the PWV seasonality, trend components, and trend probabilities for each area. The first subplot in each area shows the seasonality and trend components of the PWV. The second subplot displays the PWV trend probability, where the upper red, middle green, and lower blue envelopes indicate positive, zero, and negative trends, respectively.
Figure 4. Time-series plots of the PWV seasonality, trend components, and trend probabilities for each area. The first subplot in each area shows the seasonality and trend components of the PWV. The second subplot displays the PWV trend probability, where the upper red, middle green, and lower blue envelopes indicate positive, zero, and negative trends, respectively.
Sustainability 17 06699 g004
Figure 5. Impervious surface expansion in Beijing and Wuhan from 2013 to 2022.
Figure 5. Impervious surface expansion in Beijing and Wuhan from 2013 to 2022.
Sustainability 17 06699 g005
Figure 6. Time-series plots of MEI and ONI from 2013 to 2022. The red lines represent ±0.5. When the ONI and MEI are ≥0.5 for five consecutive months or more, an El Niño event is determined, and if the ONI and MEI are ≤−0.5 for five consecutive months or more, a La Niña event is determined.
Figure 6. Time-series plots of MEI and ONI from 2013 to 2022. The red lines represent ±0.5. When the ONI and MEI are ≥0.5 for five consecutive months or more, an El Niño event is determined, and if the ONI and MEI are ≤−0.5 for five consecutive months or more, a La Niña event is determined.
Sustainability 17 06699 g006
Figure 7. Wavelet coefficient plots of 10-year PWV time series. Rows 1–6 correspond to the BJFS, BJNM, CHAN, JFNG, WUH2, and WUHN stations, respectively. Columns 1–3 represent the contour maps of the wavelet coefficient modulus, wavelet variance, and fluctuation chart of the real part of the wavelet coefficients for each station, respectively.
Figure 7. Wavelet coefficient plots of 10-year PWV time series. Rows 1–6 correspond to the BJFS, BJNM, CHAN, JFNG, WUH2, and WUHN stations, respectively. Columns 1–3 represent the contour maps of the wavelet coefficient modulus, wavelet variance, and fluctuation chart of the real part of the wavelet coefficients for each station, respectively.
Sustainability 17 06699 g007
Figure 8. Wavelet coefficient plots of 10-year PWV time series. Rows 1–4 correspond to the SHAO, HKWL, HKWS, and KMNM stations, respectively. Columns 1–3 represent the contour maps of the wavelet coefficient modulus, wavelet variance, and fluctuation chart of the real part of the wavelet coefficients for each station, respectively.
Figure 8. Wavelet coefficient plots of 10-year PWV time series. Rows 1–4 correspond to the SHAO, HKWL, HKWS, and KMNM stations, respectively. Columns 1–3 represent the contour maps of the wavelet coefficient modulus, wavelet variance, and fluctuation chart of the real part of the wavelet coefficients for each station, respectively.
Sustainability 17 06699 g008
Figure 9. Wavelet coefficient plots of 10-year PWV time series. Rows 1–5 correspond to the TWTF, TCMS, TNML, CKSV, and NCKU stations, respectively. Columns 1–3 represent the contour maps of the wavelet coefficient modulus, wavelet variance, and fluctuation chart of the real part of the wavelet coefficients for each station, respectively.
Figure 9. Wavelet coefficient plots of 10-year PWV time series. Rows 1–5 correspond to the TWTF, TCMS, TNML, CKSV, and NCKU stations, respectively. Columns 1–3 represent the contour maps of the wavelet coefficient modulus, wavelet variance, and fluctuation chart of the real part of the wavelet coefficients for each station, respectively.
Sustainability 17 06699 g009
Figure 10. Wavelet coefficient plots of PWV time series in 2018. Rows 1 and 2 correspond to the BJFS, CHAN, JFNG, WUH2, WUHN, and SHAO stations, respectively. Columns 1–3 represent the contour maps of the wavelet coefficient modulus, wavelet variance, and fluctuation chart of the real part of the wavelet coefficients for each station, respectively.
Figure 10. Wavelet coefficient plots of PWV time series in 2018. Rows 1 and 2 correspond to the BJFS, CHAN, JFNG, WUH2, WUHN, and SHAO stations, respectively. Columns 1–3 represent the contour maps of the wavelet coefficient modulus, wavelet variance, and fluctuation chart of the real part of the wavelet coefficients for each station, respectively.
Sustainability 17 06699 g010
Figure 11. Wavelet coefficient plots of PWV time series in 2018. Rows 1–6 correspond to the HKSL, HKWS, KMNM, TWTF, CKSV, and NCKU stations, respectively. Columns 1–3 represent the contour maps of the wavelet coefficient modulus, wavelet variance, and fluctuation chart of the real part of the wavelet coefficients for each station, respectively.
Figure 11. Wavelet coefficient plots of PWV time series in 2018. Rows 1–6 correspond to the HKSL, HKWS, KMNM, TWTF, CKSV, and NCKU stations, respectively. Columns 1–3 represent the contour maps of the wavelet coefficient modulus, wavelet variance, and fluctuation chart of the real part of the wavelet coefficients for each station, respectively.
Sustainability 17 06699 g011
Figure 12. The frequency-domain changes of PWV in different areas in 2018.
Figure 12. The frequency-domain changes of PWV in different areas in 2018.
Sustainability 17 06699 g012
Figure 13. Wavelet coefficient plots of 10-year air temperature time series. Rows 1–4 correspond to the BJFS, CHAN, JFNG, and SHAO stations, respectively. Columns 1 and 2 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of T2m for each station, respectively. Columns 3 and 4 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of Th at a height of 600 hPa for each station.
Figure 13. Wavelet coefficient plots of 10-year air temperature time series. Rows 1–4 correspond to the BJFS, CHAN, JFNG, and SHAO stations, respectively. Columns 1 and 2 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of T2m for each station, respectively. Columns 3 and 4 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of Th at a height of 600 hPa for each station.
Sustainability 17 06699 g013
Figure 14. Wavelet coefficient plots of 10-year air temperature time series. Rows 1 to 4 correspond to HKSL, TWTF, CKSV, and KMNM stations, respectively. Columns 1 and 2 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of T2m for each station, respectively. Columns 3 and 4 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of Th at a height of 850 hPa for each station.
Figure 14. Wavelet coefficient plots of 10-year air temperature time series. Rows 1 to 4 correspond to HKSL, TWTF, CKSV, and KMNM stations, respectively. Columns 1 and 2 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of T2m for each station, respectively. Columns 3 and 4 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of Th at a height of 850 hPa for each station.
Sustainability 17 06699 g014
Figure 15. Wavelet coefficient plots of air temperature time series for 2018. Rows 1–4 correspond to the BJFS, CHAN, JFNG, and SHAO stations, respectively. Columns 1 and 2 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of T2m for each station, respectively. Columns 3 and 4 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of Th at a height of 600 hPa for each station.
Figure 15. Wavelet coefficient plots of air temperature time series for 2018. Rows 1–4 correspond to the BJFS, CHAN, JFNG, and SHAO stations, respectively. Columns 1 and 2 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of T2m for each station, respectively. Columns 3 and 4 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of Th at a height of 600 hPa for each station.
Sustainability 17 06699 g015
Figure 16. Wavelet coefficient plots of air temperature time series for 2018. Rows 1 to 4 correspond to HKSL, TWTF, CKSV, and KMNM stations, respectively. Columns 1 and 2 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of T2m for each station, respectively. Columns 3 and 4 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of Th at a height of 850 hPa for each station.
Figure 16. Wavelet coefficient plots of air temperature time series for 2018. Rows 1 to 4 correspond to HKSL, TWTF, CKSV, and KMNM stations, respectively. Columns 1 and 2 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of T2m for each station, respectively. Columns 3 and 4 present the wavelet variance and fluctuation chart of the real part of the wavelet coefficients of Th at a height of 850 hPa for each station.
Sustainability 17 06699 g016
Table 1. Statistics of GNSS station areas, temporal coverage, and number of valid data points before and after exclusion.
Table 1. Statistics of GNSS station areas, temporal coverage, and number of valid data points before and after exclusion.
StationsAreasTemporal CoverageData Points
Before Exclusion
Data Points
After Exclusion
BJFSBeijing2013-01-01–2022-12-3183,92683,829
BJNMBeijing2013-01-01–2020-07-0853,86553,772
CHANChangchun2013-01-01–2022-12-3183,54482,914
CKSVSouthern Taiwan2016-10-16–2022-12-3150,90050,900
HKSLHong Kong2013-01-01–2022-12-3185,18985,189
HKWSHong Kong2013-01-01–2022-12-3185,28085,280
JFNGWuhan2014-03-26–2022-12-3175,02675,024
KMNMKinmen County2016-10-16–2022-12-3150,68150,681
NCKUSouthern Taiwan2016-04-05–2022-05-1947,56047,560
SHAOShanghai2013-01-01–2021-07-0844,48944,479
TCMSNorthern Taiwan2013-01-01–2021-04-2448,59548,595
TNMLNorthern Taiwan2013-01-01–2022-12-3145,96645,966
TWTFNorthern Taiwan2013-01-01–2022-12-3181,81081,810
WUH2Wuhan2016-08-31–2022-12-3150,69350,690
WUHNWuhan2013-01-04–2022-12-3172,87472,867
Table 2. Data processing strategies in GAMIT.
Table 2. Data processing strategies in GAMIT.
Processing Key PointsSpecific Models and Parameters
Double-difference relative positioningCoordinate parametersDouble-difference network adjustment
Receiver clock offsetDifferential technique
ObservationsType of observationGPS observation data
Observation periods2013-01-01–2022-12-31
Sampling interval30 s
Correction modelsCutoff height angle10°
Phase entanglementCorrection
Tidal loadCorrection of Earth tides, polar tides, and ocean tides
Relativistic effectCorrection
ZTD prior modelVMF1+Saastamoinen
Projection function VMF 1   ( 2 ° × 2.5 °)
Parameter estimation strategiesSatellite orbitIGS Final Orbital Precision Ephemeris
Ambiguity solutionLeast-Squares Ambiguity Decorrelation Adjustment
Ambiguity parameterFloating point solution
Table 3. Main periods and fluctuation periods in the main periods of GNSS stations.
Table 3. Main periods and fluctuation periods in the main periods of GNSS stations.
StationsFirst Main Period/dFirst Fluctuation Period/dSecond Main Period/dSecond Fluctuation Period/d
BJFS552363278182
BJNM562370286184
CHAN552361277183
JFNG549365277182
WUH2542368280185
WUHN542364281180
SHAO611386//
HKSL553364//
HKWS553364//
KMNM544366//
TWTF551362//
TCMS558366//
TNML569365//
CKSV542364//
NCKU539366//
Table 4. Main periods and fluctuation periods of GNSS stations in 2018.
Table 4. Main periods and fluctuation periods of GNSS stations in 2018.
StationsMain Period/dFluctuation Period/d
BJFS160/43105/31
CHAN176/44113/26
JFNG94/59/3962/38/25
WUH297/59/3964/40/26
WUHN94/60/3964/40/26
SHAO97/60/4056/40/25
HKSL184/136/86/41/18121/90/56/27/11
HKWS184/136/86/42/18122/92/57/27/12
KMNM186/133/91/41/18125/93/58/27/12
TWTF132/90/41/1893/58/27/12
CKSV194/88/43/18 121/56/27/12
NCKU184/88/43/18121/56/28/12
Table 5. Correlation statistics of vertical atmospheric temperature (T2m and Th) and PWV at each station.
Table 5. Correlation statistics of vertical atmospheric temperature (T2m and Th) and PWV at each station.
HeightBJFSCHANJFNGSHAOHKSLTWTFCKSVKMNM
2 m0.710.760.710.760.740.730.710.78
1000 hPa0.700.770.700.760.730.750.740.77
975 hPa0.690.770.690.750.730.740.740.75
950 hPa0.690.770.700.750.750.740.740.75
925 hPa0.690.770.720.760.780.740.750.76
900 hPa0.700.780.730.770.80.740.760.77
875 hPa0.710.780.750.770.810.740.760.78
850 hPa0.720.790.760.780.810.740.760.78
825 hPa0.740.790.770.780.80.750.750.78
800 hPa0.750.800.780.780.790.760.740.77
775 hPa0.760.800.790.790.770.760.730.77
750 hPa0.770.810.800.790.750.750.720.76
700 hPa0.790.810.810.80.680.740.680.73
650 hPa0.80.810.820.810.590.700.600.67
600 hPa0.810.820.820.810.480.650.500.60
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, T.; Xiong, J.; Hu, S.; Zhao, W.; Huang, M.; Zhang, L.; Xia, Y. Exploring Precipitable Water Vapor (PWV) Variability and Subregional Declines in Eastern China. Sustainability 2025, 17, 6699. https://doi.org/10.3390/su17156699

AMA Style

Zhang T, Xiong J, Hu S, Zhao W, Huang M, Zhang L, Xia Y. Exploring Precipitable Water Vapor (PWV) Variability and Subregional Declines in Eastern China. Sustainability. 2025; 17(15):6699. https://doi.org/10.3390/su17156699

Chicago/Turabian Style

Zhang, Taixin, Jiayu Xiong, Shunqiang Hu, Wenjie Zhao, Min Huang, Li Zhang, and Yu Xia. 2025. "Exploring Precipitable Water Vapor (PWV) Variability and Subregional Declines in Eastern China" Sustainability 17, no. 15: 6699. https://doi.org/10.3390/su17156699

APA Style

Zhang, T., Xiong, J., Hu, S., Zhao, W., Huang, M., Zhang, L., & Xia, Y. (2025). Exploring Precipitable Water Vapor (PWV) Variability and Subregional Declines in Eastern China. Sustainability, 17(15), 6699. https://doi.org/10.3390/su17156699

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop