Next Article in Journal
Land Regulation and Local Service Provision: Can Economic Growth and Environmental Protection Be Achieved Simultaneously?
Previous Article in Journal
Photovoltaic Solar Farms Site Selection through “Policy Constraints–Construction Suitability”: A Case Study of Qilian County, Qinghai
Previous Article in Special Issue
Research on Improving the Accuracy of SIF Data in Estimating Gross Primary Productivity in Arid Regions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Spatio-Temporal Diversification of per Capita Carbon Emissions in China: 2000–2020

1
School of Economics and Management, Inner Mongolia University, Hohhot 010020, China
2
Institute of Geographic Sciences and Natural Resources Research, Chinese Academy of Sciences, Beijing 100101, China
3
Department of Geography, School of Geography and Information Engineering, China University of Geosciences, Wuhan 430078, China
4
College of Public Administration, Nanjing Agricultural University, Nanjing 210095, China
5
School of Economics and Management, Beijing University of Technology, Beijing 100124, China
*
Authors to whom correspondence should be addressed.
Land 2024, 13(9), 1421; https://doi.org/10.3390/land13091421
Submission received: 25 July 2024 / Revised: 23 August 2024 / Accepted: 28 August 2024 / Published: 3 September 2024
(This article belongs to the Special Issue Land-Based Greenhouse Gas Mitigation for Carbon Neutrality)

Abstract

:
Exploring the low-carbon transition in China can offer profound guidance for governments to develop relevant environmental policies and regulations within the context of the 2060 carbon neutrality target. Previous studies have extensively explored the promotion of low-carbon development in China, yet no studies have completely explained the mechanisms of the low-carbon transition in China from the perspective of per capita carbon emissions (PCEs). Based on the statistics and carbon emissions data of 367 prefecture level cities in China from 2000 to 2020, this study employed markov chain, kernel density analysis, hotspots analysis, and spatial regression models to reveal the spatiotemporal distribution patterns, future trends, and driving factors of PCEs in China. The results showed that China’s PCEs in 2000, 2010, and 2020 were 0.72 ton/persons, 1.72 ton/persons, and 1.91 ton/persons, respectively, exhibiting a continuous upward trend, with evident regional heterogeneity. PCEs in northern China and the eastern coastal region were higher than those of southern China and the central and southwestern regions. The PCEs in China showed obvious spatial clustering, with hot spots mainly concentrated in Inner Mongolia and Xinjiang, while cold spots were mainly in some provinces in southern China. The transition of PCEs in China exhibited a strong stability and a ‘club convergence’ phenomenon. A regression analysis revealed that the urbanization level and latitude had negative effects on PCEs, while the regional economic development level, average elevation, average slope, and longitude showed positive effects on PCEs. These findings have important implications for the promotion of the low-carbon transition and the effective achievement of the “dual carbon” goal.

1. Introduction

The phenomenon of global warming, coupled with the melting of glaciers and a series of extreme weather events, has a significant impact on human survival and social activities, thereby posing ongoing challenges to social development [1,2]. The acceleration of CO2 emissions has been unequivocally demonstrated as the dominant contributing factor to global warming [3,4]. Climate change has been induced by the explosive growth of greenhouse gases, primarily carbon emissions (CEs). How to tackle climate issues is the biggest environmental challenge in the world [5]. According to Climate Watch (2024) [6], China was the leading contributor to greenhouse gas emissions in 2020, accounting for 27%. In the ranking, the United States accounted for 11%, India accounted for 7%, while the European Union accounted for 6% of the emissions. The achievement of the Paris Agreement in 2015 has accelerated the global shift towards a low-carbon economy, thereby prompting significant transformations across various aspects of the world economy and society [7]. The Intergovernmental Panel on Climate Change reports underscore the importance of integrating renewable energy, enhancing energy efficiency, and fostering sustainable urban development. Initiatives such as the Global Carbon Project and the International Energy Agency’s sustainable development scenarios provide pathways and policy recommendations for achieving low-carbon transformation. As the most rapidly developing country in the world, China has become the largest emitter of CEs globally. In 2022, China’s CO2 emissions reached 12.1 billion tons, accounting for 30% of the world, placing the country under severe pressure to reduce CEs. Although China ranks first in total CEs, the United States has the highest per capita carbon emissions (PCEs). Reducing CO2 emissions has become a common goal for all countries [8]. In response, China is making efforts to carry forward low-carbon transition to reduce CO2 emissions.
The evolution of CEs has been the subject of the existing research. Examining CEs from a per capita perspective is particularly important for understanding the unequal distribution. China exhibits significant differences across regions in terms of natural conditions, economic development, and population density. By approaching CEs from a per capita perspective, we can achieve a more precise measurement of regional contributions and responsibilities in CEs, reflecting the intensity and efficiency of CEs in each region. Hickel (2020) highlighted that China, despite its significant contribution to cumulative emissions, has four times the population of the United States [9]. If this difference is taken into account, the United States should be more responsible for CEs than China. Matthews (2016) quantified climate debts by the principle of atmospheric commons from a per capita perspective between 1960 (or 1990) and 2013 and determined whether countries’ PCEs are above or below the global average [10]. It was observed that if a country’s PCEs exceed the global average, it can be considered to be in debt. Conversely, if a country’s PCEs are below the global average, it can be considered to be in credit. Therefore, exploring CEs from a per capita perspective is based on fairness and equity. Thus, it is imperative to elaborate the spatio-temporal patterns and driving mechanism of PCEs in China, which would provide significantly insights for understanding the low-carbon transition and formulating effective action plans to achieve “dual carbon” targets.
The reduction in CEs has emerged as a key area of interest and investigation. The primary focuses of current research concerning CEs are on three aspects: CEs measurement [11,12,13], spatial distribution patterns and discrepancy [14,15], and driving mechanisms [16,17,18]. The degree of CEs output has been widely measured in many fields, such as the energy [19,20], manufacturing [21], service [22], tourism [23], and transportation industries [24]. The study of CEs has been progressively integrated with that of land use [25], thereby significantly enhancing the depth and breadth of research into CEs. In terms of scale, this research has included countries [26,27], urban agglomerations [28], river basins [29], provinces [30], and counties [31]. Moreover, there is a growing trend in researching spatial patterns and heterogeneity related to CEs. Scholars employed quantitative methods to quantify spatial autocorrelation and discrepancies, such as Moran’s I, the Theil index, and Dagum’s Gini index [32,33]. Moran’s I is extensively applied to quantitatively describe spatial autocorrelation [8,34]. For example, Chen et al. (2022) confirmed that building CEs had a positive spatial autocorrelation [35]. Wang et al. (2022) argued that the distribution of PCEs demonstrated self-reinforcing agglomeration in spatial distribution [36]. Furthermore, the spatio-temporal patterns and disparities were assessed using the Theil index and Dagum’s Gini coefficient. For example, Ma et al. (2022) employed Dagum Gini coefficient to assess the PCEs of commercial buildings in different regions, revealing that intra-regional disparities were smaller than inter-regional disparities [37].
The investigation of the driving factors contributing to CEs has been a topic of considerable interest and debate, drawing upon both theoretical deduction and empirical evidence. Existing academic studies have discussed the impacts of influencing factors, such as financial development [38], economic development [39,40], foreign direct investment [41], urbanization level [42], high-speed railways [43], and environmental policies [44]. The findings of previous studies indicated that factors such as economic development, population growth, and traffic congestion were positively associated with CEs. In addition, economic growth may also induce the unintended consequence of environmental pollution, thus resulting in higher CEs [45]. In the early stage, the conventional economic model resulted in a considerable amount of CEs. The process of urbanization has been identified as a significant driver of CEs [46]. As a consequence of urbanization, the rising demand for clothing, food, shelter, and transportation leads to different lifestyles, consequently increasing CEs. From this perspective, the process of urbanization inevitably increases CEs [47].
Meanwhile, when analyzing which factor exerts the greatest influence on CEs, a variety of analytical techniques are available for this purpose, including structural decomposition analysis (SDA) [48], Stochastic Impacts by Regression on Population, Affluence and Technology (STIRPAT), the logarithmic mean Divisia index (LMDI), and spatial econometric methods. For instance, Xu et al. (2021) took Guangdong province in China as the research object, and revealed that the factors of consumption structure, per capita consumption, and population caused the largest increase in CEs [48]. Yu et al. (2023) used the STIRPAT method to analyze the factors of household CO2 emissions and identified that the total population, household size, unemployment rate, and urbanization level were the most significant factors [49]. Using a spatial econometric model, Liu et al. (2023) revealed that rapid economic growth and traditional industrial structure transform had promotion effects on CEs, while government expenditure, population clustering, and scientific innovation had inhibitory effects on CEs [50].
By the way of conclusion, previous research has provided a comprehensive examination of CEs, establishing a robust theoretical and empirical foundation. However, the previous research is not without limitations. Few studies have completely explained the mechanisms of the low-carbon transition in China from the perspective of PCEs. Furthermore, the study of how geographical factors influence PCEs has received only sporadic attention to date. To bridge the research gaps, this study aimed to analyze the spatiotemporal diversification of PCEs across different regions in China from 2000 to 2020, identify the trends and spatial autocorrelation in PCEs, assess the impact of economic and geographical factors on PCE levels in various regions, and provide insights into effective strategies for regional low-carbon transitions. So, we aimed to construct a framework to characterize the spatio-temporal patterns of PCEs, thereby expanding the depth and scope of research on PCEs. Secondly, we aimed to analyze the dynamic evolution of PCEs, recognizing that the CEs varied significantly across regions due to different influencing factors. Therefore, it is essential to explore this spatial heterogeneity, which can be visualized through spatial patterns. Finally, we attempted to explore the driving factors from both socio-economic and natural perspectives. While previous studies have identified various driving factors affecting CEs, there has been a tendency to focus on socio-economic factors, with relatively little attention paid to the role of natural and geographical factors. However, there is evidence that these factors exert a significant influence on CEs. Therefore, we integrated the effects of the urbanization rate, gross domestic product (GDP) density, average slope, average digital elevation model (DEM), latitude, and longitude factors to explore the effects on CEs. Revealing their driving mechanisms will help us to uncover the primary factor affecting China’s low-carbon transition, thereby contributing to achieving the goals of carbon peaking and neutrality.

2. Materials and Methods

2.1. Study Area

This study analysed 367 units at the prefecture level and above in mainland China (including a few county-level administrative districts; data for Hong Kong, Macao, and Taiwan were not yet available), including 4 municipalities directly under the central government, 293 prefectural-level cities, 7 districts, 30 autonomous prefectures, 3 leagues, and 30 provincial-level direct-administration units, for a total of 367 computational units (Figure 1). To scientifically reflect the socio-economic development of different regions in China, this study area was divided into four major economic subregions: Northeast, Central, Western, and Eastern China. The northeastern region mainly included the municipal units of Liaoning, Jilin, and Heilongjiang provinces; the eastern region includes the municipal units of Beijing, Tianjin, Hebei, Shanghai, Jiangsu, Zhejiang, Fujian, Shandong, Guangdong, and Hainan provinces; the central region includes the municipal units of Shanxi, Anhui, Jiangxi, Henan, Hubei, and Hunan provinces; and the rest is the western region.

2.2. Data Sources

This study selected 367 cities in China as the research sample, covering 30 provinces, municipalities, and autonomous regions. The CE data were derived from the Center for Global Environmental Research (https://db.cger.nies.go.jp/dataset/ODIAC/, accessed on 2 August 2024). The population data were mainly derived from the main data bulletins of the fifth, sixth, and seventh national censuses in 2000, 2010, and 2020, respectively. The land use data, DEM, and slope were derived from EPS data platform (https://www.epsnet.com.cn/index.html#/Index, accessed on 2 August 2024) and Data Center for Resources and Environmental Sciences (https://www.resdc.cn/, accessed on 2 August 2024). The GDP data were derived from the Statistical Yearbooks. In addition, for areas within the Xinjiang Uygur Autonomous Region where the seventh census data was unavailable, this study supplemented the missing values using linear regression.

2.3. Per Capita Carbon Emissions

In this context, PCEs represent the ratio of CEs to population [51,52,53]. The calculation equations were as follows:
P C E = C E s / P o p u l a t i o n
P C E C = C E s t 2 P o p u l a t i o n t 2 C E s t 1 P o p u l a t i o n t 1
where PCEC represents the change in PCEs; CEst1 and CEst2 represent CEs of a unit at time t1 and t2, and Populationt1 and Populationt2 represent the population of a unit at time t1 and t2.

2.4. Markov Chain

The markov chain provides a mechanism for explaining how the probability distribution of transitions from one state to another eventuates [54]. Due to its steady-state analysis, the markov chain can be utilized to predict the long-term patterns and trends of geographic phenomena with a high degree of accuracy. In this study, we used markov chain to predict the trend of PCEs in each region, and the probabilistic transfer model constructed by the markov chain explaining the PCE conversion state in each region, reflecting the future interconversion of the values of different magnitudes. In general, the state type of PCEs at moment t is represented by a 1 × k state probability vector as Et = [E1,t, E2,t, …, Ek,t], and the transition process of PCE state types can be represented by a k × k Markov transition matrix M. Based on the similar counts for each type of county, PCEs at the county level were categorized into four types using quartiles (0.25/0.5/0.75), labeled as k = 1, 2, 3, and 4, respectively.

2.5. Kernel Density Analysis

As one kind of nonparametric estimation, kernel density estimation was characterized by relative smoothness and unbiasedness, which can accurately describe the distribution of random variables [55,56]. The calculation equations were as follows:
f ( x ) = 1 n h i = 1 n K D i D h
K ( t ) = 1 2 π e t 2 2
where K(x) is the kernel function; Di is the independent and equally distributed observed values; D is mean value; n is the number of observed values; and h is the bandwidth. This study used Gaussian kernel function to effectively explore the changing law of PCEs in various regions so as to reflect the heterogeneity and the dynamic trend of PCEs in each region.

2.6. Hotspots Analysis

This study intended to use the Moran’s I index to reveal the spatial autocorrelation characteristics of PCEs. To further reflect the spatial agglomeration of PCEs and changes in PCEs, the Getis-Ord Gi* index was used to measure the statistically significant hot and cold spots of PCEs [35]. The Getis-Ord Gi* index can be used for measuring the spatial clustering, which can test whether variables exhibit the feature of high-value cluster or low-value cluster, thereby identifying the locations of hot and cold spots, and spatial outliers [57,58].

2.7. Spatial Regression

Spatial autocorrelation is a common phenomenon among geographical elements, and non-spatial models may overlook the spatial dependence between factors, which may lead to potential biases in identifying underlying influencing factors. To illustrate, an element is influenced not only by the level of the element in the study unit, but also by the level of the element in adjacent or distant units. Based on this, this study intended to use the least squares method (OLS), the spatial lag model (SLM), and spatial error model (SEM) to reveal the mechanism of China’s low-carbon transition [59]. The cross-sectional benchmark model was expressed as Equation (5):
P C E i = α 0 + α 1 L U I i + α 2 G D P D i + α 3 A S L O i + α 4 A D E M i + α 5 L A T i + α 6 L O N i + ε i
where i represents a city, and LUI is the development and utilization intensity of land [60,61], which represents the land urbanization level. The function, intensity, and efficiency of CEs are influenced by important driving forces such as land urbanization and associated human activities [62]. GDPD is GDP density, representing economic growth, which is an important factor. GDP promotes PCEs, while this effect weakens with the growth of GDP [63,64]. ASLO represents average slope, and ADEM represents average DEM. These two factors represent the topographic elements [20,51]. The ASLO and ADEM factors exert notable effects on CEs by influencing urban expansion [65,66]. In this study, we also added longitude (LON) and latitude (LAT) factors to reveal the rules of CEs in terms of latitude and longitude.
SLM was expressed as shown in Equation (6):
P C E i = ρ i = 1 n W i j × P C E i + β X i + ε i
where ρ is the spatial autoregressive coefficient, Wij is the spatial weight matrix, β is a vector of coefficients Xi, and Xi represents the set of independent variables.
SEM was expressed as shown in Equation (7):
P C E i = β X i + ε i ,   ε i = λ j = 1 n W i j ε i + φ i
where λ is the spatial autocorrelation coefficient of the error term.
By combining multiple methods, this study aimed to evaluate the influencing factors to provide a reference for integrated regional green development (see Figure 2).

3. Results

3.1. PCEs in China

China’s PCEs in 2000, 2010, and 2020 were 0.72 ton/persons, 1.72 ton/persons, and 1.91 ton/persons, respectively. These figures demonstrated a continuous increasing trend, with notable regional variations in PCEs. The PCEs were the highest in the northeast and the lowest was in the central region with values of 0.53 ton/persons in 2000, 1.23 ton/persons in 2010, and 1.43 ton/persons in 2020. Additionally, the PCEs southeast of Hu-line were was lower than those in the northwest. Specifically, in 2000, 2010, and 2020 the PCEs southeast of Hu-line were 0.72 ton/persons, 1.63 ton/persons, and 1.81 ton/persons, respectively, while that northwest of Hu-line were 1.09 ton/person, 2.81 ton/persons, and 3.06 ton/persons, respectively (Figure 3). From the perspective of spatial distribution, the PCEs in southern China were obviously lower than those in northern China. Also, the eastern coastal areas had higher PCEs than central and southwest China. Seeing the change from 2010 to 2020, the PCEs in different regions exhibited an increasing tendency (Figure 4).

3.2. Spatiotemporal Dynamics of PCEs

The markov transfer probability matrix was employed to predict future PCEs with a time span of 10 years (Table 1). In overall spatial terms, PCEs were expected to exhibit stable growth over the next 10 years, with PCEs showing a tendency to converge to higher values. Specifically, this tendency was particularly pronounced in the northeast, followed by the central region. It was also observed that, with the exception of the northeast and the east, the shift probabilities on the main diagonal were higher than those off the diagonal, indicating relatively stable PCE levels. The probability of shifting from high to higher values in the northeast and eastern regions suggested that higher growth rates in PCEs exist in these regions.

3.3. Kernel Density

The changes in the study elements over the course of the study can be encapsulated in the contour plots of the kernel density of the normalized PCEs. The contour lines should remain close to the 45° diagonal if no drastic changes have occurred. Figure 5a,b,d showed the kernel density of PCEs over a 10-year time span for northeast, eastern, and central China, respectively, and it was found that there were shifts to higher or lower values of PCEs in these regions, with a more pronounced shift towards higher values in eastern and central China. Figure 5c shows the trend of the shift in western China over a 10-year time span, revealing that the PCEs shifted to higher values, while in some regions with higher values, the PCEs shifted to lower values. The changes in PCEs in China, as shown in Figure 5e, indicated that some peaks of the kernel density contours were basically located near 45°, showing that there was some degree of stability in PCEs, as well as shifts from low to high and from high to low.

3.4. Change in PCEs in China

From 2000 to 2010, the number of cities with a reduction in PCEs accounted for 0.544%, and only a small proportion of prefecture-level cities achieved a decreases in PCEs during that period. Correspondingly, a large number of prefecture-level cities had an evident increase in PCEs. For example, the proportion of prefecture-level cities with PCEs increasing by more than 1 ton/persons was 40.871%. However, the number of cities with a decrease in PCEs from 2010 to 2020 was 10.354%, which was higher than that of the previous decade. It is noteworthy that the proportion of cities with an increase in PCEs greater than 1 ton/persons was 1.907%, which was evidently lower than that in the previous decade. Most cities (29.700%) had PCEs increases of less than 1 ton/persons, indicating that China has undergone a drastic transition in terms of PCEs, and there has been remarkable progress towards the low-carbon transition (Figure 6).
To gain further insight into the distribution of PCEs, we calculated the Moran’s I indexes of PCEs, which were 0.258, 0.316, and 0.316 in 2000, 2010, and 2020, respectively, and the p-value was significant at the level of 0.001, obviously showing a significant spatial agglomeration pattern in PCEs. The specific spatial distribution characteristics can be obtained from the hot spot distribution map. We found that the hot spots were mainly distributed in Inner Mongolia and Xinjiang, while the cold spots were primarily distributed in some provinces of southern China (e.g., Sichuan, Chongqing, Guizhou, Hubei, Hunan, Jiangxi, and Hainan). From 2000 to 2020, it is evident that the PCE hot spots gradually expanded eastward into Inner Mongolia and northeastern China (Figure 7).
To further reveal the hotspots of changes in PCEs, we calculated the spatial autocorrelation index of PCE changes during 2000–2010 and 2010–2020. The spatial autocorrelation indexes of PCE changes during 2000–2010 and 2010–2020 were 0.354 and 0.179, respectively, and the p-value was significant at the level of 0.001. The hot spot analysis results demonstrated that the hot spots of PCE changes during 2000–2010 were distributed in Inner Mongolia, Northern Xinjiang, and Northern Qinghai Province. The cold spots of PCE changes were distributed in southern China, showing a similar pattern to those of the PCEs. The cold spots of PCE changes in China during 2010–2020 were distributed in the west of Xinjiang and the Pearl River Delta, while the hot spots were distributed in northeast China (Figure 8).

3.5. Driving Mechanism of Low-Carbon Transition

To reveal the driving mechanism of the low-carbon transition in China, the OLS, SEM, and SLM were employed in this study. This study first carried out OLS model (Table 2) to diagnose. The coefficients of LUI in 2000, 2010, and 2020 were −3.239, −8.440, and −7.027, respectively, which indicated that the LUI had a significant negative impact on PCEs. The coefficients of GDPD in 2000, 2010, and 2020 were 26.842, 44.587, and 4.659, which indicated that GDPD had a significant positive impact on PCEs. Higher GDPD typically indicates more economic activity and a greater population density, thus leading to higher energy consumption and CEs. The coefficients of ASLO in 2000, 2010, and 2020 were −1.209, −2.707, and −2.993, respectively, indicating that the ASLO had a significant negative impact on PCEs. Steeper slopes may limit large-scale agricultural or construction activities, potentially leading to lower levels of energy consumption and CEs. The coefficients of ADEM in 2000, 2010, and 2020 were −2.076, −3.205, and −3.131, illustrating that the ADEM had a negative impact on PCEs. Regions with higher DEM, characterized by mountainous and hilly landscapes, were less suitable for extensive agricultural and industrial operations, leading to a reduction in energy consumption and CEs. The coefficients of LAT in 2000, 2010, and 2020 were 0.068, 0.152, and 0.156, respectively, indicating that the LAT had a significant positive impact on PCEs. Higher latitudes often have colder climates, which may lead to increased energy consumption for heating, thus increasing PCEs. The coefficients of LON in 2000, 2010, and 2020 were −0.043, −0.066, and −0.056, respectively, indicating that the LON had a negative impact on PCEs. This is largely because that the eastern regions were the most economically developed areas in China, boasting advanced technological levels that were conducive to reducing CEs. The results showed that the Moran’s I indexes of PCEs in 2000, 2010, and 2020 were 3.225, 4.291, and 4.341, respectively, and the p-value was significant at the level of 0.05. Therefore, the results obtained using the OLS regression model cannot fully and scientifically explain the relationship between them. When considering spatial lag and spatial error terms, the SLM and SEM performed better.
The coefficient of LUI was negative, indicating that the urbanization promotes low-carbon transition (Table 3). The coefficient of the GDPD was positive, but it was significant only in the SEM and SLM models of 2010. The coefficient of the ASLO was negative, but it was significant only in the SEM model of 2010 and the SLM and SEM models of 2020. The coefficient of the ADEM was negative in all models, indicating that PCE decreases with increasing altitude. PCEs in China increased significantly with rising latitude, while PCEs decreased significantly with rising longitude. In addition, PCEs were influenced not only by their own unit factors but also by those of adjacent units, as indicated by the spatial lag terms being significant at the 0.001 level. An average 1% increase in PCEs in surrounding areas in 2000, 2010, and 2020 would result in increases of 0.238%, 0.290%, and 0.304% increases in PCEs in their own units. In addition, the spatial error terms were also statistically significant at the 0.001 level, indicating that PCE changes were not only influenced by the aforementioned factors, but also by other factors.
The OLS regression analysis results (Table 4) indicated that the Moran’s I indexes of the 2000, 2010, and 2020 settlements in the eastern region were 4.070, 2.709, and 3.959, respectively, with a significant p-value at the 0.01 level, which indicated that the model’s settlements in the eastern region had a strong spatial autocorrelation, so the spatial lag and spatial error terms were considered, and the SLM and SEM were used to improve the goodness of fit (Table 5 and Table 6). The SLM and SEM regression results analysis in Table 5 and Table 6 indicated that the main factors affecting PCEs varied across the four regions of China. In the western region, LUI and GDPD were the main factors significantly impact on PCEs. LUI exhibited a significant negative effect, while GDPD had a significant positive effect. The reason was that urbanization often led to a shift from high-carbon-emission industries and agriculture to low-energy-consuming service and high-tech industries, thereby resulting in reduced PCEs. However, economic growth often lead to increased energy consumption and production, which can result in higher CEs. In the northeastern region, LUI and ASLO were the main factors, which both had significant negative effects on PCEs. This can be attributed to the abundant forest resources in northeastern China. The steep areas with dense vegetation act as carbon sinks, absorbing CO2 from the atmosphere and thereby reducing the overall carbon footprint. In the central region, LUI and GDPD were the main factors which had significant effect on PCEs. LUI had a significant negative effect, while GDPD had a significant positive effect. In the eastern region, GDPD and LON were the main factors, which both had significant positive effects on PCEs.

4. Discussion

4.1. Interpretation of Findings

Theoretical and empirical research on PCEs made a significant contribution to the promotion of China’s low-carbon transition. In this research, we analyzed the various spatiotemporal patterns and projected future state of PCEs. Our findings revealed an upward trend in China’s PCEs, with northern China’s PCEs being significantly higher than those of southern China, and the east coastal region’s PCEs exceeding that of the central and southwestern regions. With economic growth, an increasing number of individuals have migrated to urban areas, which has subsequently resulted in a rise in CEs [67]. Northern China is currently the largest energy base, and energy consumption and greenhouse gas emissions have been rising sharply due to the rapid development of energy-intensive and polluting industries. For instance, Xue et al. (2011) confirmed that fossil energy production regions produce significantly more CEs than other regions [68]. Wang et al. (2022) demonstrated that the PCE level was positively correlated with the geographical and scale distribution of cities [69]. Compared to other regions, the eastern region has higher PCEs and CEs, demonstrating a connection between the region’s economic level and PCE level.
In the markov transition probability matrix, the highest probability was found in the H-type, reaching 97.5%, with an average stability probability of 75.27%, indicating that China’s PCEs were stable and exhibited a “club convergence” phenomenon. At the same time, the low-level (0.720) and high-level (0.975) convergence probabilities were greater than the medium–low-level (0.655) and medium–high-level (0.661) convergence probabilities, indicating that low- and high-level regions tended to stabilize at their own levels, while medium–low and medium–high levels tended to transfer to low and high levels. From the perspective of probability values, the probabilities of downward transfer were 32.7% and 32.2%, while those of upward transfer were 1.10% and 1.60%, indicating that the types with lower and higher intensities showed a good trend of downward transfer. The result obtained was consistent with the findings of Cui et al. (2022), Liu et al. (2023), and Wang et al. (2019) [16,50,52]. A dynamical procedure underlying the transition matrix showed that a discernible “Matthew effect” was observed in PCEs, whereby cities with both low and high emissions exhibited tendencies to maintain their initial state throughout the transfer process. This illustrated that the region may become trapped in a development path locked into specific trajectories, making it challenging to swiftly reduce PCEs through technological means.
By conducting a comprehensive review on China’s CEs, we can clearly identify the differences between this study and previous research. These differences primarily focus on the following three main aspects. (1) Regional level. Previous studies may focus on the disparities in CEs across different regions. This study analyzed 367 municipal units at the prefecture level. There were presently many studies in the literature on different regions. Jin et al. (2024) [70] took the urban agglomeration of the Yangtze River Economic Belt’s as the study areas of CEs. Bei et al. (2024) [71] conducted the study of CEs at Wuhan. Yan et al. (2023) [72] considered 30 provincial units (provinces, cities, and autonomous regions) in mainland China as their basic research units. (2) Research methods. This study used markov chain, kernel density, and spatial economic regression analyses. Feng et al. (2024) [73] and Chen and Bi (2022) [74] used Geodetector models to empirically analyze the driving mechanisms of CEs. Li et al. (2024) [75] used the LMDI model to investigate each province’s CEs drivers. Chen et al. (2023) [76] used kernel density estimation, a spatial autocorrelation analysis, and the spatial-temporal LMDI model to explore the spatiotemporal patterns and driving mechanism of CEs. (3) Driving factors. This study utilized a spatial economic model to explore the driving factors of CEs. Wei et al. (2021) [77] verified that GDP was the dominant factor affecting CEs efficiency. Wang et al. (2023) [78] revealed that technology and population size played important roles in CEs reductions. Jiang et al. (2024) [79] deduced that the scale of urban construction presented different promoting or inhibiting effects in different regions. Zhang et al. (2024) [80] demonstrated that technological progress was the main factor influencing CEs.

4.2. Driving Mechanisms of Low-Carbon Transition

Based on the results of the OLS, SLM, and SEM, we revealed the driving mechanism of China’s low-carbon transition and explored how socioeconomic and geographical factors that have significantly influenced low-carbon transition. The LUI was proven to have a significantly negative impact on PCEs, which indicated that the increase in urbanization promoted the low-carbon transition, which was inconsistent with previous studies showing increasing CEs due to urbanization [81]. It is a widely accepted hypothesis that the LUI can increase PCEs. On the contrary, some studies have pointed out that urbanization can indeed lead to a reduction in CEs [82,83]. The reason for this was that urbanization promoted the transformation of industrial structures and thus contributes to CE efficiency. The coefficients of the LUI affecting PCEs were −3.239, −8.440, and −7.027 in 2000, 2010, and 2020, which showed that urbanization was more crucial to the reduction in PCEs in 2010 and 2020 than in 2000. In the early stage, due to the problem of blindly pursuing quantity rather than quality under the traditional urbanization and economic model, the reduction in CEs has been inhibited [84]. However, as urbanization levels increased and shifted towards equity, a green economy, and efficiency as core objectives have contributed to the reduction in CEs. Moreover, several studies have corroborated the assertion that urbanization can result in a reduction in CEs, albeit with notable spatial heterogeneity. For instance, Li et al. (2023) [85] revealed that a higher level of urbanization reduced the CEs intensity and explored the significant implications of urbanization on the low-carbon transition by promoting green technological innovation. GDPD was also proven to be positively correlated with CEs. This is because economic growth was usually accompanied by the development of industrialization and manufacturing, which often required more energy, especially fossil fuels, thereby releasing large amounts of CEs [86]. Many scholars have found that economic growth inevitably promoted CEs, leading to global warming and glacier melting, which posed serious challenges to sustainable economic development. For example, Guo and Fang (2023) [87] revealed that CEs was positively correlated with economic growth but present a fluctuating trend. Similarly, Li et al. (2022) [88] concluded that there was a positive effect on GDP per capita and total CO2. The difference of this study was investigating the impact of GDPD on PCEs, with GDP being categorized by land area.
The regression results also demonstrated that the low-carbon transition was driven by the joint action of socio-economic systems and natural systems, rather than being solely influenced by either socio-economic factors or geographical factors. This finding underscores the comprehensiveness of the driving mechanism and emphasized the importance of formulating region-specific policies. The European Union emissions trading system (EU ETS) was launched in 2005, which has demonstrated significant success in reducing greenhouse gas emissions across 31 countries, accounting for over 40% of the EU’s total greenhouse gas emissions. The EU ETS has led to substantial emission reductions while increasing the regulated firms’ revenues and fixed assets, as verified by Dechezleprêtre et al. (2023) [89]. The Regional Greenhouse Gas Initiative (RGGI) in the United States was officially implemented in 2009 and established a cap-and-trade program for CEs in the power sector. The RGGI not only achieved significant emission reductions but also generated economic benefits by reinvesting auction proceeds into energy efficiency and renewable energy projects. However, the RGGI has caused significant emissions reductions within regulated states and emissions leakages in nearby unregulated states (Yan, 2021) [90]. In future research, we would like to explore this driving mechanism by taking China’s environmental policy into consideration.

4.3. Policy Implications of Low-Carbon Transition

4.3.1. Promoting the Low-Carbon Transition

A significant proportion of prefecture-level cities exhibited a noticeable increase in PCEs. China’s CEs have remained high level due to the decision that was made for its economic development mode. As rising CEs contributed to global warming, China, as the biggest developing country, is enacting a low-carbon transition and has made remarkable achievements in curbing the growth of CEs. New-type urbanization is the strategic fulcrum of future development, which is crucial for promoting the low-carbon transition. New urbanization has the potential to establish a favorable policy environment and offer significant scientific and technological support, which can facilitate the low-carbon transition. Throughout the progress of new urbanization [91,92], it is of paramount importance to underscore the significance of enhancing the quality of urbanization [93].

4.3.2. Focusing on the Spatial Agglomeration Effects of PCE

PCEs showed a relatively obvious distribution pattern of spatial agglomeration, and the degree of this continued to strengthen. The PCEs of each city were susceptible to influences from neighboring cities. Therefore, it is crucial for each city to not only prioritize its own PCEs but also to collaborate with neighboring cities. By establishing regional cooperation mechanisms, cities can share information and technology related energy and environmental protection, jointly implement emission reduction measures, as well as collectively assume the responsibility of promoting coordinated economic development and environmental protection. Considering the differences across regions, for provinces such as Chongqing, Sichuan, Guizhou, Anhui, and Jiangxi, which are experiencing rapid economic growth, it is recommended to optimize their industrial structure, support low-energy consuming and clean industries, and promote the shift from high-speed to high-quality economic development. For resource-rich provinces such as Inner Mongolia and Xinjiang, which have resource-endowed advantages, it is suggested that these governments appropriately increase their carbon allowances and set higher thresholds for their high-energy-consuming enterprises, thereby promoting the introduction of high-energy-efficiency technologies from the east to the west.

4.3.3. Coordinating Economic Growth and CEs

In terms of driving factors, economic growth acts as a “double-edged” sword, not only increasing CEs but also providing momentum for the low-carbon transition. The contradiction between the increase in GDP and CEs still exists. However, with the improvement of quality, economic progress can stimulate technological innovation and accelerate emission reductions. Enhancing the quality of economic growth quality would inevitably result in a reduction in CEs and the promotion of the low-carbon transition. Economic growth should leverage the advantages of agglomeration effects and scale effects.

4.3.4. Formulating Differential Strategy

Based on the different regions of CEs and their spatial-temporal characteristics, we proposed a targeted strategy to reduce PCEs. At the national level, the focus should be on improving the efficiency of fossil energy use and replacing fossil energy with non-fossil energy sources, such as developing hydropower, wind power, and solar power. Additionally, carbon sinks can be increased by using trees or other plants to absorb the carbon dioxide emitted into the atmosphere. At the regional level, for the eastern and central regions, the priority should be to strengthen green technology upgrades, leverage their advantages in low-carbon technology research and development, and actively explore innovative systems and mechanisms for low-carbon economic development. For the northeastern regions, it is essential to promote the transformation of resource-depleted cities and the adjustment of old industrial bases in conjunction with the development of a low-carbon economy. Considering the reality that the northeastern region has more per capita arable land and forestry resources, agriculture and forestry should play a significant role in the low-carbon economic development in this region. For the western region, the focus should be on reducing CEs in resource development, distinctive industry development, and infrastructure construction. In terms of increasing carbon sinks, efforts should be made to strengthen carbon sink construction in ecological functional zones, integrating the construction of carbon sinks in the western region with the establishment of a national ecological security barrier.

5. Conclusions

This study aimed to investigate the distribution patterns, future trends, and driving factors of PCEs in 367 cities in China. The results revealed that the PCEs presented a continuous upward trend, with obvious differences across regions. Therefore, CEs reduction policies should be formulated based on a per capita perspective and regional differences. The PCEs also showed obvious spatial clustering, with hot spots mainly concentrated in Inner Mongolia and Xinjiang, while cold spots were mainly distributed in some provinces in southern China. Meanwhile, China’s low-carbon transition has achieved noteworthy outcomes, with the majority of prefecture-level cities (29.700%) having PCEs of less than 1 ton/persons. The overall PCE level in China remained stable despite the dynamic changes observed over the ten-year period. In particular, the shift of the PCEs of eastern and central China to higher values was obvious. The regression analysis indicated that China’s PCEs undergone changes due to a confluence of factors. The LUI and latitude had negative effects on PCEs in prefecture-level cities, and regional economic development, elevation, slope, and longitude had positive effects. The results can serve as a basis for developing CEs policies, which are of strategic importance for promoting sustainable development and formulating rational CE reduction policies.
However, this study still has several limitations. First, we acknowledge that it is insufficient to use data from 2000, 2010, and 2020 to explore the distribution patterns, future trends, and driving factors of China’s PCEs. This is because the PCE calculations were based on data extracted from the bulletins of the fifth, sixth, and seventh national censuses, which covered a ten-year period. In future studies, we will use annual data to supplement or verify the findings of the studied ten-year period. Second, although this study verified the importance of natural geographical factors and socioeconomic factors in influencing the low-carbon transition, due to data availability limitations, other driving forces, such as the digital economy and trading policies, were not taking into account. In the future, we will consider more driving factors to analyze the mechanism.

Author Contributions

Conceptualization, X.Z. and Y.Z.; Methodology, X.Z. and Y.Z.; Software, X.Z., Y.Z. and W.C.; Resources, Y.Z. and W.C.; Data curation, X.Z. and G.Z.; Writing—original draft, X.Z. and Y.Z.; Writing—review and editing, S.P. and F.D.; Supervision, X.Z. and F.D. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Innovative Research Group Project of the National Natural Science Foundation of China (Grant number: 42121001). This study was also supported by the Natural Science Foundation of China (Grant numbers: 72364027; 42371258; 72164030). This study was supported in part by the China Postdoctoral Science Foundation (Grant number: 2023M733466), the Social Science Foundation of China (Grant number: 22VMZ013), the Natural Science Foundation of Inner Mongolia (Grant No. 2023QN07008), and the Social Science Foundation of Inner Mongolia (Grant No. 2024EY51).

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bakhtyar, B.; Ibrahim, Y.; Alghoul, M.A.; Aziz, N.; Fudholi, A.; Sopian, K. Estimating the CO2 abatement cost: Substitute price of avoiding CO2 emission (SPAE) by renewable energy’s feed in tariff in selected countries. Renew. Sustain. Energy Rev. 2014, 35, 205–210. [Google Scholar] [CrossRef]
  2. Gökmenoğlu, K.; Taspinar, N. The relationship between CO2 emissions, energy consumption, economic growth and FDI: The case of Turkey. J. Int. Trade. Econ. Dev. 2016, 25, 706–723. [Google Scholar] [CrossRef]
  3. Allen, M.R.; Frame, D.J.; Huntingford, C.; Jones, C.D.; Meinshausen, N. Warming caused by cumulative carbon emissions towards the trillionth tonne. Nature 2009, 458, 1163–1166. [Google Scholar] [CrossRef] [PubMed]
  4. Höhne, N.; den Elzen, M.; Rogelj, J.; Metz, B.; Fransen, T.; Kuramochi, T.; Olhoff, A.; Alcamo, J.; Winkler, H.; Fu, S.; et al. Emissions: World has four times the work or one-third of the time. Nature 2020, 579, 25–28. [Google Scholar] [CrossRef] [PubMed]
  5. Meinshausen, M.; Meinshausen, N.; Hare, W.; Raper, S.C.; Frieler, K.; Knutti, R.; Frame, D.J.; Allen, M.R. Greenhouse-gas emission targets for limiting global warming to 2 C′. Nature 2009, 458, 1158–1162. [Google Scholar] [CrossRef]
  6. Climate Watch. Historical GHG Emissions. 2024. Available online: https://www.climatewatchdata.org/ghg-emissions (accessed on 26 August 2024).
  7. Söderholm, P.; Hildingsson, R.; Johansson, B.; Khan, J.; Wilhelmsson, F. Governing the transition to low-carbon futures: A critical survey of energy scenarios for 2050. Futures 2011, 43, 1105–1116. [Google Scholar] [CrossRef]
  8. Wang, S.B.; Liu, H.M.; Pu, H.X.; Yang, H. Spatial disparity and hierarchical cluster analysis of final energy consumption in China. Energy 2020, 197, 117195. [Google Scholar] [CrossRef]
  9. Hickel, J. Quantifying national responsibility for climate break down:an equality-based attribution approach for carbon dioxide emissions in excess of the planetary boundary. Lancet Planet. Health 2020, 4, e399–e404. [Google Scholar] [CrossRef]
  10. Matthews, H.D. Quantifying historical carbon and climate debts among nations. Nat. Clim. Chang. 2016, 6, 60–64. [Google Scholar] [CrossRef]
  11. Adriana, D.B.; Michela, G.; Nicolò, S.; Baccelli, O.; Croci, E.; Molteni, T. Impact of circular measures to reduce urban CO2 emissions: An analysis of four case studies through a production- and consumption-based emission accounting method. J. Clean. Prod. 2022, 380, 134932. [Google Scholar]
  12. Ali, U.; Guo, Q.B.; Zhanar, N.; Nurgazina, Z.; Sharif, A.; Kartal, M.T.; Depren, S.K.; Khan, A. Heterogeneous impact of industrialization, foreign direct investments, and technological innovation on carbon emissions intensity: Evidence from Kingdom of Saudi Arabia. Appl. Energy 2023, 336, 120804. [Google Scholar] [CrossRef]
  13. Wang, Z.; Zeng, Y.; Wang, X.; Gu, T.; Chen, W. Impact of urban expansion on carbon emissions in the urban agglomerations of Yellow River Basin, China. Land 2024, 13, 651. [Google Scholar] [CrossRef]
  14. Shi, Y.; Han, B.; Han, L.; Wei, Z. Uncovering the national and regional household carbon emissions in China using temporal and spatial decomposition analysis models. J. Clean. Prod. 2019, 232, 966–979. [Google Scholar] [CrossRef]
  15. Huo, T.; Xu, L.; Feng, W.; Cai, W.; Liu, B. Dynamic scenario simulations of carbon emission peak in China’s city-scaleurban residential building sector through 2050. Sci. Total Environ. 2021, 159, 112612. [Google Scholar]
  16. Wang, S.J.; Huang, Y.Y. Spatial spillover effect and driving forces of carbon emission intensity at city level in China. Acta Geogr. Sin. 2019, 74, 1131–1148. (In Chinese) [Google Scholar] [CrossRef]
  17. Li, S.L.; Wang, Z.Z. The effects of agricultural technology progress on agricultural carbon emission and carbon sink in China. Agriculture 2023, 13, 793. [Google Scholar] [CrossRef]
  18. Chen, C.H.; Luo, Y.Q.; Zhou, H.; Huang, J.B. Understanding the driving factors and finding the pathway to mitigating carbon emissions in China’s Yangtze River Delta region. Energy 2023, 278, 127897. [Google Scholar] [CrossRef]
  19. Wang, S.; Xie, Z.; Wu, R.; Feng, K. How does urbanization affect the carbon intensity of human well-being? A global assessment. Appl. Energy 2022, 312, 118798. [Google Scholar] [CrossRef]
  20. Zhou, Y.Q.; Zou, S.; Duan, W.L.; Chen, Y.N.; Takara, K.R.; Di, Y.F. Analysis of energy carbon emissions from agroecosystems in Tarim River Basin, China: A pathway to achieve carbon neutrality. Appl. Energy 2022, 325, 119842. [Google Scholar] [CrossRef]
  21. Huang, Y.M.; Zhang, Y.N. Digitalization, positioning in global value chain and carbon emissions embodied in exports: Evidence from global manufacturing production-based emissions. Ecol. Econ. 2023, 205, 107674. [Google Scholar] [CrossRef]
  22. Wang, S.J.; Fang, C.L.; Wang, Y. Spatiotemporal variations of energy-related CO2 emissions in China and its influencing factors: An empirical analysis based on provincial panel data. Renew. Sustain. Energy Rev. 2016, 55, 505–515. [Google Scholar] [CrossRef]
  23. Ehigiamusoe, K.U.; Shahbaz, M.; Vo, X.V. How Does Globalization Influence the Impact of Tourism on Carbon Emissions and Ecological Footprint? Evidence from African Countries. J. Travel. Res. 2023, 62, 1010–1032. [Google Scholar] [CrossRef]
  24. Jan, S.; Liu, Y.; Stefan, S. Carbon emissions from European land transportation: A comprehensive analysis. Transp. Res. Part D Transp. Environ. 2023, 121, 103851. [Google Scholar]
  25. Zhang, C.Y.; Zhang, H.T.; Chen, M.N.; Fang, R.Y.; Yao, Y.; Zhang, Q.P.; Wang, Q. Spatial-temporal characteristics of carbon emissions from land use change in Yellow River Delta region. China Ecol. Indic. 2022, 136, 108623. [Google Scholar] [CrossRef]
  26. Dogan, E.; Turkekul, B. CO2 emissions, real output, energy consumption, trade, urbanization and financial development: Testing the EKC hypothesis for the USA. Environ. Sci. Pollut. Res. 2016, 23, 1203–1213. [Google Scholar] [CrossRef]
  27. Erdogan, S. Dynamic nexus between technological innovation and building sector carbon emissions in the BRICS Countries. J. Environ. Manag. 2021, 293, 112780. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, X.L.; Di, Q.B.; Jia, W.H.; Hou, Z.W. Spatial correlation network of pollution and carbon emission reductions coupled with high-quality economic development in three Chinese urban agglomerations. Sustain. Cities Soc. 2023, 94, 104552. [Google Scholar] [CrossRef]
  29. Wu, H.; Yang, Y.; Li, W. Analysis of spatiotemporal evolution characteristics and peak forecast of provincial carbon emissions under the dual carbon goal: Considering nine provinces in the Yellow River basin of China as an example. Atmos. Pollut. Res. 2023, 14, 101828. [Google Scholar] [CrossRef]
  30. Fan, J.J.; Wang, J.L.; Qiu, J.X.; Li, N. Stage effects of energy consumption and carbon emissions in the process of urbanization: Evidence from 30 provinces in China. Energy 2023, 276, 127655. [Google Scholar] [CrossRef]
  31. Liu, Q.F.; Song, J.P.; Dai, T.Q.; Shi, A.; Xu, J.H.; Wang, E.R. Spatio-temporal dynamic evolution of carbon emission intensity and the effectiveness of carbon emission reduction at county level based on nighttime light data. J. Clean. Prod. 2022, 362, 132301. [Google Scholar] [CrossRef]
  32. Liu, H.; Lei, J. The impacts of urbanization on Chinese households’ energy consumption: An energy input-output analysis. J. Renew. Sustain. Energy 2018, 10, 015903. [Google Scholar] [CrossRef]
  33. Wang, S.J.; Liu, X.P. China’s city-level energy-related CO2 emissions: Spatiotemporal patterns and driving forces. Appl. Energy 2017, 200, 204–214. [Google Scholar] [CrossRef]
  34. Zhang, Y.J.; Hao, J.F.; Song, J. The CO2 emission efficiency, reduction potential and spatial clustering in China’s industry: Evidence from the regional level. Appl. Energy 2016, 174, 213–223. [Google Scholar] [CrossRef]
  35. Chen, W.X.; Chi, G.Q. Ecosystem services trade-offs and synergies in China, 2000–2015. Int. J. Environ. Sci. Technol. 2022, 20, 3221–3236. [Google Scholar] [CrossRef]
  36. Chancel, L. Global carbon inequality over 1990–2019. Nat. Sustain. 2022, 5, 931–938. [Google Scholar] [CrossRef]
  37. Ma, T.; Liu, Y.S.; Yang, M. Spatial-temporal heterogeneity for commercial building carbon emissions in China: Based the dagum gini coefficient. Sustainability 2022, 14, 5243. [Google Scholar] [CrossRef]
  38. Tao, M.M.; Sheng, M.S.; Wen, L. How does financial development influence carbon emission intensity in the OECD countries: Some insights from the information and communication technology perspective. J. Environ. Manag. 2023, 335, 117553. [Google Scholar] [CrossRef]
  39. Jebli, M.B.; Kahia, M. The interdependence between CO2 emissions, economic growth, renewable and non-renewable energies, and service development: Evidence from 65 countries. Clim. Chang. 2020, 162, 193–212. [Google Scholar] [CrossRef]
  40. Akram, R.; Umar, M.; Gan, X.L.; Chen, F. Dynamic linkages between energy efficiency, renewable energy along with economic growth and carbon emission. A case of MINT countries an asymmetric analysis. Energy Rep. 2022, 8, 2119–2130. [Google Scholar] [CrossRef]
  41. Nasir, M.A.; Huynh, T.L.D.; Tram, H.T.X. Role of financial development, economic growth & foreign direct investment in driving climate change: A case of emerging ASEAN. J. Environ. Manag. 2019, 242, 131–141. [Google Scholar]
  42. Chen, W.X.; Wang, G.Z.; Xu, N.; Ji, M.; Zeng, J. Promoting or inhibiting? New-type urbanization and urban carbon emissions efficiency in China. Cities 2023, 140, 104429. [Google Scholar] [CrossRef]
  43. Tang, Z.P.; Wang, L.H.; Wu, W. The impact of high-speed rail on urban carbon emissions: Evidence from the Yangtze River Delta. J. Transp. Geogr. 2023, 110, 103641. [Google Scholar] [CrossRef]
  44. Wang, K.K.; Su, X.W.; Wang, S.H. How does the energy-consuming rights trading policy affect China’s carbon emission intensity? Energy 2023, 276, 127579. [Google Scholar] [CrossRef]
  45. Li, R.; Wang, Q.; Liu, Y.; Jiang, R. Per-capita carbon emissions in 147 countries: The effect of economic, energy, social, and trade structural changes. Sustain. Prod. Consum. 2021, 27, 1149–1164. [Google Scholar] [CrossRef]
  46. Yu, M.; Meng, B.; Li, R. Analysis of China’s urban household indirect carbon emissions drivers under the background of population aging. Struct. Chang. Econ. Dynam. 2022, 60, 114–125. [Google Scholar] [CrossRef]
  47. Ahmed, M.; Shuai, C.; Ahmed, M. Influencing factors of carbon emissions andtheir trends in China and India: A machine learning method. Environ. Sci. Pollut. Res. 2022, 29, 48424–48437. [Google Scholar] [CrossRef]
  48. Xu, W.H.; Xie, Y.L.; Xia, D.H.; Ji, L.; Huang, G.H. A multi-sectoral decomposition and decoupling analysis of carbon emissions in Guangdong province, China. J. Environ. Manag. 2021, 298, 113485. [Google Scholar] [CrossRef] [PubMed]
  49. Yu, Y.; Dai, Y.Q.; Xu, L.Y.; Zheng, H.Z.; Wu, W.H.; Chen, L. A multi-level characteristic analysis of urban agglomeration energy-related carbon emission: A case study of the Pearl River Delta. Energy 2023, 263, 125651. [Google Scholar] [CrossRef]
  50. Liu, X.J.; Jin, X.B.; Luo, X.L.; Zhou, Y.K. Quantifying the spatiotemporal dynamics and impact factors of China’s county-level carbon emissions using ESTDA and spatial econometric models. J. Clean. Prod. 2023, 410, 137203. [Google Scholar] [CrossRef]
  51. Chen, W.X.; Zeng, J.; Li, N. Change in land-use structure due to urbanisation in China. J. Clean. Prod. 2021, 321, 128986. [Google Scholar] [CrossRef]
  52. Cui, Y.; Khan, S.U.; Deng, Y.; Zhao, M. Spatiotemporal heterogeneity, convergence and its impact factors: Perspective of carbon emission intensity and carbon emission per capita considering carbon sink effect. Environ. Impact Assess. Rev. 2022, 92, 106699. [Google Scholar] [CrossRef]
  53. Wang, M.; Feng, C. Tracking the inequalities of global per capita carbon emissions from perspectives of technological and economic gaps. J. Environ. Manag. 2022, 315, 115144. [Google Scholar] [CrossRef]
  54. Wu, S.; Zhang, K. Influence of urbanization and foreign direct investment on carbon emission efficiency: Evidence from urban clusters in the Yangtze River Economic Belt. Sustainability 2021, 13, 2722. [Google Scholar] [CrossRef]
  55. Wang, X.; Wang, C.X.; Shi, J.L. Evaluation of urban resilience based on Service-Connectivity-Environment (SCE) model: A case study of Jinan city, China. Int. J. Disaster Risk Reduct. 2023, 95, 103828. [Google Scholar] [CrossRef]
  56. Yang, L.Y.; Fang, C.L.; Chen, W.X.; Zeng, J. Urban-rural land structural conflicts in China: A land use transition perspective. Habitat. Int. 2023, 138, 102877. [Google Scholar] [CrossRef]
  57. Getis, A.; Aldstadt, J. Constructing the Spatial Weights Matrix Using a Local Statistic. Geogr. Anal. 2004, 36, 90–104. [Google Scholar] [CrossRef]
  58. Getis, A.; Griffith, D.A. Comparative spatial filtering in regression analysis. Geogr. Anal. 2002, 34, 130–140. [Google Scholar] [CrossRef]
  59. Chen, W.X.; Wang, G.Z.; Zeng, J. Impact of urbanization on ecosystem health in Chinese urban agglomerations. Environ. Impact Assess. Rev. 2023, 98, 106964. [Google Scholar] [CrossRef]
  60. Felipe-Lucia, M.R.; Soliveres, S.; Penone, C.; Fischer, M.; Ammer, C.; Boch, S.; Boeddinghaus, R.S.; Bonkowski, M.; Buscot, F.; Fiore-Donno, A.M.; et al. Land-use intensity alters networks between biodiversity, ecosystem functions, and services. Proc. Natl. Acad. Sci. USA 2020, 117, 28140–28149. [Google Scholar] [CrossRef]
  61. Xu, F.; Chi, G.Q. Spatiotemporal variations of land use intensity and its driving forces in China, 2000–2010. Reg. Environ. Chang. 2019, 19, 2583–2596. [Google Scholar] [CrossRef]
  62. Zhou, Y.; Chen, M.X.; Tang, Z.P.; Mei, Z. Urbanization, land use change, and carbon emissions: Quantitative assessments for city-level carbon emissions in Beijing-Tianjin-Hebei region, Sustain. Cities Soc. 2021, 66, 102701. [Google Scholar] [CrossRef]
  63. Wang, W.Z.; Liu, L.C.; Liao, H.; Wei, Y.M. Impacts of urbanization on carbon emissions: An empirical analysis from OECD countries. Energy Policy 2021, 151, 112171. [Google Scholar] [CrossRef]
  64. Zhang, H.M.; Xu, L.; Zhou, P.; Zhu, X.D. Coordination between economic growth and carbon emissions: Evidence from 178 cities in China. Econ. Anal. Policy 2024, 81, 164–180. [Google Scholar] [CrossRef]
  65. Braimoh, A.K.; Onishi, T. Spatial determinants of urban land use change in Lagos, Nigeria. Land Use Policy 2007, 24, 502–515. [Google Scholar] [CrossRef]
  66. Liu, X.; Xu, H.Z.; Zhang, M. The effects of urban expansion on carbon emissions: Based on the spatial interaction and transmission mechanism. J. Clean. Prod. 2024, 434, 140019. [Google Scholar] [CrossRef]
  67. Li, K.; Lin, B.Q. Impacts of urbanization and industrialization on energy consumption/CO2 emissions: Does the level of development matter? Renew. Sustain. Energy Rev. 2015, 52, 1107–1122. [Google Scholar] [CrossRef]
  68. Xue, B.; Li, C.; Liu, Z.; Geng, Y.; Xi, F. Analysis on CO2 emission and urbanization at global level during 1970–2007. Clim. Chang. Res. 2011, 7, 423–427. [Google Scholar]
  69. Wang, S.J.; Wang, Z.H.; Fang, C.L. Evolutionary characteristics and drivers of carbon emission performance of Chinese cities. Sci. Sin. Terrae 2022, 52, 1613–1626. [Google Scholar]
  70. Jin, Y.Z.; Zhang, K.R.; Li, D.Y.; Wang, S.Y.; Liu, W.Y. Analysis of the spatial–temporal evolution and driving factors of carbon emission efficiency in the Yangtze River economic Belt. Ecol. Indic. 2024, 165, 112092. [Google Scholar] [CrossRef]
  71. Bei, L.; Yang, W.; Wang, B.; Gao, Y.W.; Wang, A.N.; Lu, T.F.; Liu, H.T.; Sun, L.S. Characteristics of residents’ carbon emission and driving factors for carbon peaking: A case study in Wuhan, China. Energy Sustain. Dev. 2024, 81, 101471. [Google Scholar] [CrossRef]
  72. Yan, D.S.; Liu, C.G.; Li, P.X. Effect of carbon emissions and the driving mechanism of economic growth target setting: An empirical study of provincial data in China. J. Clean. Prod. 2023, 415, 137721. [Google Scholar] [CrossRef]
  73. Feng, X.H.; Wang, S.S.; Li, Y.; Yang, J.Y.; Lei, K.G.; Yuan, W.K. Spatial heterogeneity and driving mechanisms of carbon emissions in urban expansion areas: A research framework coupled with patterns and functions. Land Use Policy 2024, 143, 107209. [Google Scholar] [CrossRef]
  74. Chen, C.Y.; Bi, L.L. Study on spatio-temporal changes and driving factors of carbon emissions at the building operation stage-A case study of China. Build. Environ. 2022, 219, 109147. [Google Scholar] [CrossRef]
  75. Li, S.Y.; Yao, L.L.; Zhang, Y.C.; Zhao, Y.X.; Sun, L. How do driving factors of carbon emissions and scenario forecasting differ across provinces in China? Investigation and analysis. Environ. Sustain. Indic. 2024, 22, 100390. [Google Scholar]
  76. Chen, H.D.; Du, Q.X.; Huo, T.F.; Liu, P.R.; Cai, W.G.; Liu, B.S. Spatiotemporal patterns and driving mechanism of carbon emissions in China’s urban residential building sector. Energy 2023, 263, 126102. [Google Scholar] [CrossRef]
  77. Wei, W.; Zhang, X.; Zhou, L.; Xie, B.; Zhou, J.; Li, C. How does spatiotemporal variations and impact factors in CO2 emissions differ across cities in China? Investigation on grid scale and geographic detection method. J. Clean. Prod. 2021, 321, 128933. [Google Scholar] [CrossRef]
  78. Wang, Z.; Shao, H.-A.-O. Spatiotemporal differences in and influencing factors of urban carbon emission efficiency in China’s Yangtze River Economic Belt. Environ. Sci. Pollut. Res. 2023, 30, 121713–121733. [Google Scholar] [CrossRef] [PubMed]
  79. Jiang, H.; Yin, J.; Wei, D.; Luo, X.; Ding, Y.; Xia, R. Industrial carbon emission efficiency prediction and carbon emission reduction strategies based on multi-objective particle swarm optimization-backpropagation: A perspective from regional clustering. Sci. Total Environ. 2024, 906, 167692. [Google Scholar] [CrossRef] [PubMed]
  80. Zhang, N.; Sun, F.; Hu, Y. Carbon emission efficiency of land use in urban agglomerations of Yangtze River Economic Belt, China: Based on three-stage SBM-DEA model. Ecol. Indic. 2024, 160, 111922. [Google Scholar] [CrossRef]
  81. Xu, J.H.; Li, Y.Y.; Hu, F.; Wang, L.; Wang, K.; Ma, W.H.; Ruan, N.; Jiang, W.Z. Spatio-temporal variation of carbon emission intensity and spatial heterogeneity of influencing factors in the Yangtze River Delta. Atmosphere 2023, 14, 163. [Google Scholar] [CrossRef]
  82. Adams, S.; Nsiah, C. Reducing carbon dioxide emissions: Does renewable energy matter? Sci. Total Environ. 2019, 693, 133288. [Google Scholar] [CrossRef] [PubMed]
  83. Kwakwa, P.A.; Adzawla, W.; Alhassan, H.; Oteng-Abayie, E.F. The effects of urbanization, ICT, fertilizer usage, and foreign direct investment on carbon dioxide emissions in Ghana. Environ. Sci. Pollut. Res. Int. 2023, 30, 23982–23996. [Google Scholar] [CrossRef] [PubMed]
  84. Pan, J.N.; Huang, J.T.; Chiang, T.F. Empirical study of the local government deficit, land finance and real estate markets in China. China Econ. Rev. 2015, 32, 57–67. [Google Scholar] [CrossRef]
  85. Li, X.Q.; Zheng, Z.J.; Shi, D.Q. New urbanization and carbon emissions intensity reduction: Mechanisms and spatial spillover effects. Sci. Total Environ. 2023, 905, 167172. [Google Scholar] [CrossRef] [PubMed]
  86. Abraham, D.; Huseyin, O.; Mehdi, S. The effect of GDP, renewable energy and total energy supply on carbon emissions in the EU 27: New evidence from panel GMM. Environ. Sci. Pollut. Res. 2023, 30, 28206–28216. [Google Scholar]
  87. Guo, X.M.; Fang, C.L. How does urbanization affect energy carbon emissions under the background of carbon neutrality? J. Environ. Manag. 2023, 327, 103851. [Google Scholar]
  88. Li, W.Y.; Ji, Z.S.; Dong, F.G. Spatio-temporal evolution relationships between provincial CO2 emissions and driving factors using geographically and temporally weighted regression model. Sustain. Cities Soc. 2022, 81, 103836. [Google Scholar] [CrossRef]
  89. Dechezleprêtre, A.; Nachtigall, D.; Venmans, F. The joint impact of the European Union emissions trading system on carbon emissions and economic performance. J. Environ. Econ. Manag. 2023, 118, 102758. [Google Scholar] [CrossRef]
  90. Yan, J.C. The impact of climate policy on fossil fuel consumption: Evidence from the Regional Greenhouse Gas Initiative (RGGI). Energy Econ. 2021, 100, 105333. [Google Scholar] [CrossRef]
  91. Zhang, W.S.; Xu, Y.; Lei, J.; Streets, D.G.; Wang, C. Direct and spillover effects of new-type urbanization on CO2 emissions from central heating sector and EKC analyses: Evidence from 144 cities in China. Resour. Conserv. Recy. 2023, 192, 106913. [Google Scholar] [CrossRef]
  92. Zhang, W.S.; Xu, Y.; Streets, D.G.; Wang, C. Can new-type urbanization realize low-carbon development? A spatiotemporal heterogeneous analysis in 288 cities and 18 urban agglomerations in China. J. Clean. Prod. 2023, 420, 138426. [Google Scholar] [CrossRef]
  93. Guo, X.M.; Fang, C.L.; Mu, X.F.; Chen, D. Coupling and coordination analysis of urbanization and ecosystem service value in Beijing-Tianjin-Hebei urban agglomeration. Ecol. Indic. 2022, 137, 108782. [Google Scholar]
Figure 1. Study area.
Figure 1. Study area.
Land 13 01421 g001
Figure 2. Research design.
Figure 2. Research design.
Land 13 01421 g002
Figure 3. PCEs in different regions in 2000, 2010, and 2020.
Figure 3. PCEs in different regions in 2000, 2010, and 2020.
Land 13 01421 g003
Figure 4. PCEs in China in 2000, 2010, and 2020.
Figure 4. PCEs in China in 2000, 2010, and 2020.
Land 13 01421 g004
Figure 5. Kernel density contour map of PCEs.
Figure 5. Kernel density contour map of PCEs.
Land 13 01421 g005
Figure 6. Change in PCEs in China during 2000–2020.
Figure 6. Change in PCEs in China during 2000–2020.
Land 13 01421 g006
Figure 7. Hotspots of PCEs in China during 2000–2020.
Figure 7. Hotspots of PCEs in China during 2000–2020.
Land 13 01421 g007
Figure 8. Hotspots of PCE changes in China during 2000–2020.
Figure 8. Hotspots of PCE changes in China during 2000–2020.
Land 13 01421 g008
Table 1. Markov transfer probability matrix of PCEs.
Table 1. Markov transfer probability matrix of PCEs.
VariablesSpatial ExtentTime SpanTypeLMLMHHNumber
PCEChina10L0.7200.2740.0040211
ML0.0110.6550.3270.005180
MH00.0160.6610.322183
H000.0250.975160
Eastern China10L0.7240.2410.034058
ML0.0350.4820.4280.05356
MH00.0600.4600.48050
H000.1000.90040
Central China10L0.7140.2850049
ML00.5530.446047
MH00.0260.7360.23638
H000140
Western China10L0.5810.3830.034086
ML00.6860.313067
MH00.0140.7200.26468
H000.0460.95365
Northeastern China10L0.5450.4090.045022
ML00.4730.526019
MH000.3500.65020
H000111
Notes: L stands for low-level type; ML stands for medium–low-level type; MH stands for medium–high-level type; and H stands for high-level type.
Table 2. Regression results of OLS.
Table 2. Regression results of OLS.
Variables200020102020
LUI−3.239 ***−8.440 ***−7.027 ***
(0.950)(1.818)(1.816)
GDPD26.842 *44.587 ***4.659 *
(13.565)(11.510)(2.115)
ASLO−1.209 *−2.707 *−2.993 **
(0.578)(1.109)(1.127)
ADEM−2.076 ***−3.205 **−3.131 **
(0.574)(1.120)(1.143)
LAT0.068 ***0.152 ***0.156 ***
(0.010)(0.020)(0.020)
LON−0.043 ***−0.066 ***−0.056 ***
(0.008)(0.016)(0.016)
Constant5.883 ***10.269 ***8.721 ***
(1.029)(2.033)(2.088)
Moran’s I (error)3.225 **4.291 ***4.341 ***
LM (lag)8.744 **16.081 ***16.967 ***
Robust LM (lag)1.9462.5223.339
LM (error)7.324 **13.865 ***14.408 ***
Robust LM (error)0.5260.3050.779
LM (SARMA)9.271 **16.387 ***17.747 ***
Measures of fit
Log likelihood−607.080−846.958−853.786
AIC1228.1601707.9201721.570
SC1255.5001735.2501748.910
R-Squared0.2480.2720.262
N367367367
Notes: *** p ≤ 0.001, ** p ≤ 0.01, * p ≤ 0.05. Standard deviation is shown in parentheses.
Table 3. Regression results of SLM and SEM.
Table 3. Regression results of SLM and SEM.
VariablesSLMSEMSLMSEMSLMSEM
200020102020
LUI−2.470 **−2.468 *−6.279 ***−6.749 **−5.061 **−5.034 **
(0.937)(1.109)(1.795)(2.163)(1.771)(2.153)
GDPD21.54719.39934.852 **38.262 **3.4283.013
(13.206)(13.780)(11.219)(12.964)(2.032)(2.162)
ASLO−0.988−1.200−2.105−2.639 *−2.299 *−2.818 *
(0.565)(0.656)(1.075)(1.288)(1.090)(1.319)
ADEM−1.858 **−2.316 ***−2.814 **−3.76 **−2.716 *−3.641 *
(0.574)(0.685)(1.098)(1.387)(1.116)(1.434)
LAT0.055 ***0.072 ***0.114 ***0.162 ***0.115 ***0.167 ***
(0.011)(0.013)(0.022)(0.027)(0.022)(0.027)
LON−0.038 ***−0.051 ***−0.058 ***−0.085 ***−0.050 **−0.075 ***
(0.009)(0.010)(0.016)(0.021)(0.016)(0.022)
Constant5.085 ***6.259 ***8.599 ***11.161 ***7.322 ***9.366 ***
(1.067)(1.297)(2.062)(2.708)(2.075)(2.826)
Spatial lag term0.238 *** 0.290 *** 0.304 ***
(0.070)(0.068)(0.067)
Spatial error term 0.257 *** 0.310 *** 0.326 ***
(0.071)(0.069)(0.068)
Measures of fit
Log likelihood−602.400−602.588−839.089−839.433−845.333−845.680
AIC1220.8001219.181694.1801692.8701706.6701705.360
SC1252.0401246.5101725.4201720.2101737.9101732.700
R-Squared0.2750.2760.3150.3150.3090.309
N367367367367367367
Notes: *** p ≤ 0.001, ** p ≤ 0.01, * p ≤ 0.05. Standard deviation is shown in parentheses.
Table 4. Regression results of OLS for different regions of China.
Table 4. Regression results of OLS for different regions of China.
VariablesWestern ChinaNortheastern ChinaCentral ChinaEastern China
200020102020200020102020200020102020200020102020
LUI−1.592−10.968 *−10.676 *−0.667−6.731−5.800−3.502 *−10.198 **−9.015 *−2.423 ***−5.624−1.775
(2.550)(4.397)(4.456)(1.738)(4.465)(5.667)(1.398)(3.276)(3.681)(1.284)(2.963)(3.003)
GDPD50.300361.097 **90.605 **90.51566.9902.06553.92091.123 ***15.105 *17.80024.656 ***0.988
(228.985)(115.654)(30.547)(49.639)(52.383)(29.286)(37.676)(24.616)(7.083)(4.536)(5.562)(0.945)
ASLO−1.234−1.811−2.133−1.606−6.625−5.410−1.699 *−4.247 *−3.269−0.275 *−2.146−2.965
(1.099)(2.028)(2.008)(1.724)(4.490)(5.302)(0.820)(1.864)(2.140)(0.720)(1.593)(1.797)
ADEM−1.954 *−3.098−3.2352.1070.443−2.9972.0215.1963.328−3.360−3.2370.309
(0.991)(1.871)(1.879)(3.246)(8.574)(9.434)(1.667)(3.762)(4.264)(1.699)(3.875)(4.229)
LAT−0.038−0.0170.0020.0360.135 *0.1400.1130.123 *0.134 *−0.0250.0060.061
(0.020)(0.035)(0.035)(0.024)(0.062)(0.072)(0.026)(0.058)(0.067)(0.014)(0.033)(0.036)
LON0.095 ***0.213 ***0.228 ***0.0370.002−0.0140.113 ***0.221 ***0.279 ***0.037 ***0.072 *0.025
(0.029)(0.052)(0.052)(0.038)(0.100)(0.115)(0.029)(0.069)(0.078)(0.011)(0.028)(0.027)
Constant3.5683.7871.644−4.647−10.126−9.6270.028−13.254−16.710*4.350 **2.614−4.425
(2.702)(5.116)(5.141)(3.249)(8.346)(9.460)(3.079)(6.893)(7.909)(1.632)(3.801)(3.825)
Moran’s I (error)1.9151.6761.711−1.055−1.000−1.4041.7062.182 * 1.6882.709 **3.959 ***
LM (lag)1.3311.5671.5043.3793.790 *3.3531.3330.9601.19112.149 ***5.5038.524 ***
Robust LM (lag)0.1091.3240.7500.3180.7910.5292.5910.0282.0441.9112.6540.263
LM (error)1.5001.0021.0973.0653.0994.139 *0.5451.2530.52910.240 ***3.5259.558 ***
Robust LM (error)0.2780.7590.3430.0040.1001.3151.8030.3211.3810.0010.6751.298
LM (SARMA)1.6092.3261.8473.3843.8904.6693.1361.2822.57312.151 **6.179 *9.822 **
Measures of fit
Log likelihood−296.557−383.134−382.904−12.723−46.081−52.108−46.992−117.917−129.776−32.805−115.618−126.148
AIC607.113780.268779.80839.447106.162118.218107.985249.833273.55279.611245.235266.296
SC627.853801.008800.54850.531117.247129.302125.246267.095290.81497.986263.61284.671
R-Squared0.2570.3190.3260.2940.3150.2640.5080.5040.4390.2350.3490.250
N143143143363636878787102102102
Notes: *** p ≤ 0.001, ** p ≤ 0.01, * p ≤ 0.05. Standard deviation is shown in parentheses.
Table 5. Regression results of SLM for different regions of China.
Table 5. Regression results of SLM for different regions of China.
VariablesWestern ChinaNortheastern ChinaCentral ChinaEastern China
200020102020200020102020200020102020200020102020
LUI−0.874−9.350 *−9.289 *−0.583−6.453−4.865−3.156 *−9.456 **−8.175 *−2.007−4.197−1.361
(2.467)(4.279)(4.325)(1.341)(3.406)(4.404)(1.366)(3.259)(3.606)(1.136)(2.756)(2.721)
GDPD 23.714335.386 **87.693 **49.49732.640−12.27652.69187.954 ***14.178 *13.715 ***18.992 ***0.501
(221.392)(111.95)(29.534)(38.662)(39.884)(22.700)(35.743)(24.093)(6.747)(4.015)(5.279)(0.855)
ASLO−1.171−1.675−1.966−2.343−7.222 *−6.313−1.424−3.755 *−2.729−0.201−1.484−2.227
(1.066)(1.971)(1.951)(1.336)(3.443)(4.119)(0.796)(1.815)(2.065)(0.639)(1.484)(1.643)
ADEM−1.790−2.802−2.8851.702−2.903−4.4491.5614.3492.405−2.803−2.8150.440
(0.977)(1.833)(1.840)(2.518)(6.543)(7.315)(1.596)(3.620)(4.071)(1.499)(3.585)(3.827)
LAT−0.037−0.024−0.0050.061 ***0.200 ***0.218 ***0.0240.1080.115−0.019−0.0020.039
(0.019)(0.034)(0.034)(0.019)(0.048)(0.058)(0.024)(0.057)(0.065)(0.013)(0.031)(0.033)
LON0.082 **0.179 ***0.191 ***0.0380.0230.02040.096 **0.192 **0.241 **0.027 **0.054 *0.015
(0.030)(0.056)(0.057)(0.031)(0.082)(0.094)(0.032)(0.073)(0.083)(0.010)(0.027)(0.025)
Constant3.4084.2652.362−6.997 **−17.122 *−19.265 *−3.247−11.318−14.1623.301 *2.720−2.592
(2.625)(4.949)(4.970)(2.572)(6.741)(7.809)(2.976)(6.820)(7.832)(1.452)(3.514)(3.490)
Spatial lag term0.1630.1600.160−0.743 ***−0.763 ***−0.733 ***0.1890.1600.1820.435 ***0.316 **0.387 ***
(0.111)(0.109)(0.109)(0.211)(0.203)(0.212)(0.146)(0.143)(0.147)(0.099)(0.111)(0.109)
Measures of fit
Log likelihood−295.739−382.252−382.037−9.469−42.470−48.971−46.287−117.397−129.135−26.245−112.564−121.633
AIC607.478780.503780.07534.938100.942113.942108.575250.794274.2768.491241.128259.266
SC631.181804.206803.77747.606113.61126.61128.302270.521293.99789.490262.128280.265
R-Squared0.2700.3310.3380.4780.5070.4510.5190.5130.4510.3610.4030.340
N143143143363636878787102102102
Notes: *** p ≤ 0.001, ** p ≤ 0.01, * p ≤ 0.05. Standard deviation is shown in parentheses.
Table 6. Regression results of SEM for different regions of China.
Table 6. Regression results of SEM for different regions of China.
VariablesWestern ChinaNortheastern ChinaCentral ChinaEastern China
200020102020200020102020200020102020200020102020
LUI−0.275−9.211 *−8.918 *−0.142−5.562−3.579−3.511 *−10.718 ***−9.135 *−2.045−4.122−2.234
(2.653)(4.527)(4.544)(1.191)(3.115)(3.765)(1.403)(3.360)(3.692)(1.258)(2.999)(2.962)
GDPD −15.50927.583 **85.35 **34.575−3.413−21.06053.43597.235 ***14.347 *11.67318.752 ***0.305
(222.797)(113.220)(29.620)(35.537)(32.089)(20.643)(36.688)(24.908)(6.892)(3.804)(5.803)(0.842)
ASLO−1.410−2.156−2.372−2.748 *−7.250 *−6.033−1.587−3.885−3.069−0.805−2.083−3.670
(1.177)(2.128)(2.117)(1.119)(3.008)(3.620)(0.829)(1.934)(2.164)(0.769)(1.715)(1.964)
ADEM−2.078−3.158−3.2212.559−2.851−3.8301.7454.2372.683−2.527−2.6210.790
(1.110)(2.033)(2.045)(1.960)(5.208)(5.436)(1.672)(3.863)(4.286)(1.535)(3.686)(3.882)
LAT−0.046 *−0.031 ***−0.0100.052 ***0.150 ***0.166 ***0.0270.1220.133−0.0120.0150.071
(0.023)(0.039)(0.040)(0.011)(0.031)(0.035)(0.027)(0.066)(0.071)(0.022)(0.043)(0.051)
LON0.099 **0.2130.230 ***0.001−0.075−0.0670.117 ***0.238 ***0.289 ***0.0270.0590.023
(0.033)(0.058)(0.058)(0.024)(0.064)(0.072)(0.030)(0.071)(0.079)(0.014)(0.032)(0.033)
Constant3.7764.4452.082−5.238 **−8.728−11.849 *−3.917−13.374−16.717 *2.9411.047−5.023
(3.166)(5.760)(5.800)(1.801)(4.683)(4.956)(3.246)(7.724)(8.344)(2.384)(4.728)(5.387)
Spatial error term0.2070.1640.168−0.923 ***−0.810 ***−0.957 ***0.1340.2130.1330.513 ***0.336 **0.449 ***
(0.110)(0.112)(0.112)(0.199)(0.214)(0.193)(0.156)(0.149)(0.156)(0.096)(0.115)(0.104)
Measures of fit
Log likelihood−295.458−382.430−382.149−7.776−40.236−47.226−46.686−117.175−129.477−25.921−113.195−120.578
AIC604.917778.861778.29929.55394.472108.452107.374248.351272.95465.843240.391255.156
SC625.657799.601799.03940.637105.557119.537124.635265.612290.21684.218258.766273.531
R-Squared0.2750.3300.3370.5550.5550.5450.5130.5170.4450.3800.3970.364
N143143143363636878787102102102
Notes: *** p ≤ 0.001, ** p ≤ 0.01, * p ≤ 0.05. Standard deviation is shown in parentheses.
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, X.; Zeng, Y.; Chen, W.; Pan, S.; Du, F.; Zong, G. Spatio-Temporal Diversification of per Capita Carbon Emissions in China: 2000–2020. Land 2024, 13, 1421. https://doi.org/10.3390/land13091421

AMA Style

Zhang X, Zeng Y, Chen W, Pan S, Du F, Zong G. Spatio-Temporal Diversification of per Capita Carbon Emissions in China: 2000–2020. Land. 2024; 13(9):1421. https://doi.org/10.3390/land13091421

Chicago/Turabian Style

Zhang, Xuewei, Yi Zeng, Wanxu Chen, Sipei Pan, Fenglian Du, and Gang Zong. 2024. "Spatio-Temporal Diversification of per Capita Carbon Emissions in China: 2000–2020" Land 13, no. 9: 1421. https://doi.org/10.3390/land13091421

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