Next Article in Journal
Parental Educational Anxiety during Children’s Transition to Primary School in China
Next Article in Special Issue
Spatial Spillover Effects of Resource Misallocation on the Green Total Factor Productivity in Chinese Agriculture
Previous Article in Journal
Measuring Family Quality of Life: Scoping Review of the Available Scales and Future Directions
Previous Article in Special Issue
Air Pollution and Corporate Green Financial Constraints: Evidence from China’s Listed Companies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Evaluating the Spatiotemporal Characteristics of Agricultural Eco-Efficiency Alongside China’s Carbon Neutrality Targets

1
College of Economics and Management, Nanjing Forestry University, Nanjing 210037, China
2
School of Economics and Management, Chuzhou University, Chuzhou 239000, China
3
Science and Technology Department, Nanjing Forestry University, Nanjing 210037, China
4
School of Urban Economics and Management, Beijing University of Civil Engineering and Architecture, Beijing 100044, China
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2022, 19(23), 15478; https://doi.org/10.3390/ijerph192315478
Submission received: 27 October 2022 / Revised: 21 November 2022 / Accepted: 21 November 2022 / Published: 22 November 2022
(This article belongs to the Special Issue Agricultural Green Transformation and Sustainable Development)

Abstract

:
Agriculture has the dual effect of contributing to both carbon emissions and sequestration, and thus plays a critical role in mitigating global climate change and achieving carbon neutrality. Agricultural eco-efficiency (AEE) is an important measurement through which we can assess the efforts toward reduced emissions and increased sequestration. The purpose of this study was to understand the relationship between China’s target of carbon neutrality and AEE through an evaluative model, so as to improve AEE and ultimately achieve sustainable agricultural development. The Super-SBM model scientifically measures the AEE based on provincial panel data collected between 2000 and 2020. We selected kernel density function and spatial distribution to explore the spatial and temporal evolutionary trends, and used a Tobit model to identify the drivers of AEE. The research shows that (1) China’s agricultural system functions as a net carbon sink, with all provinces’ agricultural carbon sequestration levels recorded as higher than their carbon emissions from 2000 to 2020. (2) Despite sequestration levels, the level of AEE in China is not high enough, and the average efficiency level from 2000 to 2020 is 0.7726, showing an overall trend where AEE decreased at first and then increased. (3) The AEE of each province is clearly polarized; there are obvious core–periphery characteristics and spatial distribution of clustered contiguous areas. Central provinces generally have lower efficiency, eastern and northeastern provinces have higher efficiency, and northeastern provinces always remain in the high-efficiency group. (4) Influencing factors show that urbanization, upgrading of industrial structure, financial support for agriculture, and mechanization have a significant positive impact on AEE. These findings have important implications for the promotion of the low-carbon green development of Chinese agriculture.

1. Introduction

Climate change is a global concern as it causes extreme weather and destroys natural resources such as glaciers, which negatively impacts economic and social development. Increasing CO2 emissions are considered to be an important factor contributing to global warming [1,2,3]. According to the data from ITA, China will emit more than 11.9 billion tons of CO2 in 2021, accounting for 33% of total global emissions [4]. As an active participant and supporter of global climate governance, China actively implements green development strategies and is committed to energy conservation, emission reduction, and environmental protection [5,6,7]. With these goals in mind, the Chinese government announced that it would strive to reach peak carbon emissions by 2030 and achieve carbon neutrality by 2060, which is a significant contribution toward achieving the global carbon neutrality target [2]. China’s agricultural carbon emissions account for 17% of the country’s carbon emissions; comparatively, with the global average measured at 11% [8]. This makes China the world’s largest emitter of carbon from agriculture, and further demonstrates why the goal of carbon neutrality cannot be achieved without the full engagement of the agricultural sector [9].
The agricultural sector not only emits greenhouse gases, but also functions as a carbon sequestration sink [10,11,12]. Greenhouse gas emissions in agricultural production mostly emanate from practices such as tillage, irrigation, use of agricultural materials, and the cultivation of rice [13,14], and 13% of global CO2, 44% of CH2, and 82% of N2O emissions come from agricultural systems [3]. On the other hand, specific agricultural practices and spaces can absorb large amounts of CO2 through photosynthesis, such as forestry, crop growth, and pastured land, which all have a significant carbon sequestration capacity [15,16]. China has a large population and the area of arable land per capita is small. The quality of this land is not high, and agricultural growth has long relied on the excessive use of chemicals [17,18,19], which has severely overstretched the available arable land and produced large amounts of carbon emissions. These issues demonstrate how serious the challenges are for the agricultural eco-system [20,21]. However, the proper use of cropland can effectively reduce carbon emissions and increase carbon sequestration [22], and studies have shown that Chinese agriculture is a huge carbon sequestration system with significant potential to promote carbon neutrality [23].
The Chinese government places great importance on promoting the agricultural reduction in emissions and increasing sequestration. The Opinions on the Complete and Accurate Implementation of the New Development Concept and the Good Work of Carbon Peaks and Carbon Neutrality in October 2021 proposed accelerating the development of green agriculture, promoting the sequestration of carbon, and enhancing ecological agricultural carbon sequestration. Likewise, China’s No.1 Central Document 2022 proposed to research and apply agricultural technologies to reduce carbon, increase carbon sinks, and explore research into the value of carbon sink products. Therefore, in the context of carbon neutrality, exploring the balance between economic growth and low-carbon agricultural development through green and sustainable development is an effective means in China. An effective action towards sustainable development requires a scientific measurement of current AEE and an analysis of spatial and temporal differences and driving factors, in order to explore appropriate paths for improving AEE.
AEE under the carbon neutral target is an effective tool to reflect the effectiveness of green and low-carbon agricultural development, and this refers to the practice of generating as much agricultural output with as little environmental pollution as possible in order to achieve concurrent economic and environmental benefits [24,25,26]. At present, many scholars have conducted research on how to measure AEE, and at the core of this research is the selection of indicators and the construction of models [27,28]. When selecting indicators, undesirable outputs should be included in the research framework. Scholars usually consider agricultural pollution emissions to be undesirable outputs, including pollutants such as greenhouse gas emissions, chemical oxygen demand, and so on [29,30,31], and agricultural carbon emissions are adopted by most scholars [32]. Desired outputs are mainly measured using agricultural added value or crop yields. Model construction mainly includes the SFA model, the DEA model, and their derivative models, such as the SBM model, the Super-SBM model, and the directional distance function (DDF) model [33,34,35,36]. Since SFA models can only use one output and have difficulty taking undesirable outputs into account [37], non-radial, non-angular SBM models are widely used in agroecological efficiency measurements, because they have the ability to handle undesirable outputs [38,39]. Existing research further explores the influencing factors of AEE [38,40].
Currently, existing studies have achieved useful results, but there are some limitations. Firstly, most studies have taken into account the undesired output factor of carbon emissions in agricultural production, but agriculture also has the attribute of carbon sequestration, and there is a lack of systematic studies that integrate agricultural carbon emissions and sequestration into the same analytical framework. Secondly, most existing studies are confined to the measurement of AEE and lack in-depth analysis of the contributing spatial and temporal characteristics and influencing factors of eco-efficiency, making it difficult to accurately explore effective countermeasures to improve AEE. To address these issues, the Super-SBM model is used in this research, which includes carbon emissions and sequestration, to conduct an empirical study of AEE throughout China. On the basis of scientific measurements of AEE, we further examine the spatial and temporal characteristics and influencing factors of AEE in China, and attempt to expand the current literature in the following ways: (1) Effectively measure agricultural carbon emissions (material inputs, rice fields, and soils) and carbon sequestration (crops) per province from 2000 to 2020 and explore the dynamic evolutionary characteristics. (2) Construct a Super-SBM model including carbon emissions and carbon sequestration to measure AEE in China, recording AEE by province and by region (eastern, northeastern, central, and western China), and exploring spatial and temporal evolutionary trends using kernel density functions and spatial distributions. (3) Analyze the influencing factors of AEE in China with the help of a panel Tobit regression model, and screen effective paths to improve AEE.
The rest of the study is arranged as follows. Section 2 covers the data and methods used in this paper. Section 3 analyzes agricultural carbon emissions and sequestration, the spatiotemporal characteristics of AEE, and the influencing factors of AEE. Section 4 discusses the conclusions and policy implications.

2. Materials and Methods

2.1. Accounting for Agricultural Carbon Emissions and Sequestration

Based on current research [3,41,42], this paper examines agricultural carbon emissions from three perspectives, namely agricultural materials, rice fields, and soil (Table 1). Specifically, (1) the use of agricultural chemicals, including fertilizers, pesticides, agricultural films, agricultural diesel fuel, and irrigation activities that consume electricity, will produce carbon emissions (Appendix A for coefficients); (2) rice field CH4 emissions include data from rice growth periods because, due to different hydrothermal conditions in different parts of China, CH4 emissions during the growth cycle may vary. Taking into account the soil, climate, and hydrological conditions of rice cultivation, the emission coefficients of CH4 for early-season rice, mid-season rice, and late-season rice in different provinces were determined and used to calculate the CH4 emissions from rice fields (Appendix B for details of the coefficients); (3) during crop cultivation, soil layer destruction causes N2O emissions. We took measurements of N2O emissions specifically from rice, spring wheat, winter wheat, soybeans, corn, and vegetables (Appendix C for details of the coefficients).
Agricultural carbon sequestration mainly measures the carbon uptake by crops, that is, the carbon dioxide absorbed by crops through photosynthesis during their lifecycle, which is one of the important sources of carbon sequestration [45]. Based on relevant studies, the total agricultural carbon sequestration in this study is calculated using the following formula:
C s = j = 1 n C s i = j = 1 n C j × Y j × 1 W j / H j
where C s is the total carbon sequestration and C s j is the carbon sequestration of j . For the same crop, C j is the carbon content rate, Y j is the economic yield, W j is the moisture factor, and H j is the economic factor. Details of the parameters are shown in Appendix D. According to prior research of [46], the effect of 1 t CH4 and N2O is equivalent to 25 t CO2 (6.8182 t C) and 298 t CO2 (81.2727 t C), which are all converted to C emissions later.

2.2. Methodology Specification

2.2.1. Measuring AEE: Super-SBM Model

During agricultural production, alongside any expected economic benefits, there are factor inputs that accompany undesirable outputs, but it is vital to limit these outputs [38]. As traditional DEA models suffer from input factor “crowding” or “slack”, they tend to produce biased results and are unable to deal with undesirable output indicators [47]. The SBM model is capable of considering unexpected outputs and can effectively compensate for the shortcomings of traditional DEA models. However, the SBM model has an issue consistent with the traditional DEA model, in that it is also unable to distinguish between decision units (DMUs) that are both efficient at 1. On this basis, the Super-SBM model is used in this paper to measure AEE, which is thus able to deal with undesirable output, and further compare and differentiate efficient DMUs that are on the frontier [48]. The model is constructed as follows:
ρ = min 1 m i = 1 m x ¯ x i k 1 s 1 + s 2 ( r = 1 s 1 y ¯ g y r k g + t = 1 s 2 y ¯ b y t k b )
s . t . x ¯ j = 1 , j k n λ j x i j , y ¯ g j = 1 , j k n λ j y r j g , y ¯ b j = 1 , j k n λ j y t j b x ¯ x k , y ¯ g y k g , y ¯ b y k b , λ j 0 i = 1 , 2 , , m ; j = 1 , 2 , , n r = 1 , 2 , , s 1 , t = 1 , 2 , , s 2
where ρ is the value of AEE, n is the number of DMUs, m is the number of inputs, r 1 is the number of desirable outputs, and r 2 is the number of undesirable outputs. Vectors x , y g , and y b represent inputs, desirable outputs, and undesirable outputs. When ρ 1 , the AEE of the target decision unit is relatively effective; when ρ < 1 , the AEE of the target decision unit has not reached efficiency, and there is redundancy or a shortage of inputs or outputs.

2.2.2. Inspect the Dynamic Evolution Characteristics: Kernel Density Estimation

As a kind of non-parametric probability density estimation, kernel density estimation is a common method for the study of disequilibrium distribution, which can describe the distribution pattern of random variables through continuous density curves using kernels as weights [49]. The changes in the distribution pattern of the kernel density curve, changes in kurtosis, and changes in the location of the curve can be analyzed to reveal the dynamic evolution characteristics of AEE. In this paper, we choose the Gaussian kernel, which is commonly used in existing studies, and the calculation formula is as follows:
f ^ h ( x ) = 1 n i = 1 n K h x x i = 1 n h i = 1 n K x x i h
where x 1 , x 2 , , x i is an independent distribution of n sample points, K x is a random kernel function, and the magnitude of the bandwidth h value affects the smoothness of the kernel density curve distribution.

2.2.3. Verifying the Influencing Factors: Tobit Model

The AEE measured in this paper using the Super-SBM model is non-negative truncated data, representing a restricted dependent variable. Therefore, using OLS estimation would lead to biased results [50]. To solve this issue, the Tobit model proposed by the American economist Tobin in 1958 is a suitable choice [51], so we used the Tobit model to construct an econometric model of the factors influencing AEE, which is calculated as follows:
y i t = β T x i t + ε i t β T x i t + ε i t > 0 0 o t h e r w i s e
where y i t is the AEE of the i province in the t year; x i t is the explanatory variable, which refers to the factors influencing AEE; β T is the regression coefficient of the explanatory variable; and ε i t is a random error term subject to N 0 , σ 2 .

2.3. Data

2.3.1. Data Description

The study area of this paper covers 30 provinces in China, excluding Hong Kong, Macao, Taiwan, and Tibet, given the availability of data. The study provinces are divided into four regions according to the Chinese statistical partitioning criteria (Figure 1). This paper uses the data of 30 provinces from 2000 to 2020, and the required data include an account of agricultural carbon emissions and sequestration, inputs and outputs of AEE, and its influencing factors. Data were obtained from the China Statistical Yearbook, China Agricultural Yearbook, China Rural Statistical Yearbook, China Population and Employment Statistical Yearbook, China Agricultural Products Import and Export Monthly Statistical Report, and provincial and municipal statistical yearbooks, and any missing data were filled in by the interpolation method.

2.3.2. Evaluation Indicators

Agriculture covers a wide range of practices, including crop farming, forestry, animal husbandry, fish farming, and sideline industries. Because each subsector varies greatly between geographic regions, agriculture in this paper is limited to the category of the plantation industry in an effort to elevate the relevance and specificity of the study. The measure of AEE under the carbon neutrality target includes inputs, desirable outputs, and undesirable output. Five input indicators, two desirable output indicators, and one undesirable output indicator are constructed by referring to the studies of related scholars [3,30,52,53]. The input indicators include labor, land, machinery, fertilizer, and irrigation, where agricultural employment = regional primary industry employment x (value added in agriculture/value added in the primary industry). Desirable outputs are expressed in terms of gross agricultural output and agricultural carbon sequestration, and the output value is converted into comparable data. The undesirable output choices agricultural carbon emissions, which are measured by agricultural materials, rice fields, and soil carbon emissions. The indicators and descriptive statistics are shown in Table 2.

2.3.3. Influencing Factors on AEE

Another focus of this study was to explore the influencing factors of AEE under the carbon neutrality target. The goal of the study is to provide information that will promote low-carbon sustainable development in agriculture, and to enable agriculture to play a more prominent role in promoting the carbon neutrality target. Based on the usefulness and validity of the data, we constructed the influencing factors driving AEE from the perspectives of economic foundations, production conditions, agricultural support policies, and technological innovation [53,54,55,56,57]. (1) Economic conditions were measured using the urbanization rate (URBAN) and the industrial structure upgrading index (ISU). (2) Production conditions were measured using the agricultural cultivation structure (ACS), the degree of agricultural disaster (DISA), and the multiple crop index (MCI). (3) Agricultural support policies were measured via the annual financial support for agriculture (FSFA). (4) The mechanization level (MECH) measured from the perspective of technological innovation. Details of each measurement are shown in Table 3.

3. Results and Analysis

3.1. Analysis of Agricultural Carbon Emissions and Sequestration in China

Using the above calculation list of agricultural carbon emissions and sequestration, the emissions and sequestration per hectare of farmland were estimated for 30 provinces in China; the results are shown in Figure 2.
The agricultural carbon sequestration across all provinces in China from 2000 to 2020 is higher than the carbon emissions, indicating that agricultural systems across all provinces belong to the net carbon sink [58], with an average net carbon sequestration of 3.754 t/hm2. This finding is consistent with the results of existing studies [3]. Specifically, the national average agricultural carbon sequestration is 5.585 t/hm2, among which 11 provinces, including Guangxi (11.778 t/hm2), Henan (8.878 t/hm2), Shandong (8.820 t/hm2), Jiangsu (7.645 t/hm2), and Anhui (7.543 t/hm2), have an average agricultural carbon sequestration higher than the national level. These provinces are all large agricultural provinces, with crop cultivation areas ranking among the top in China, such as sugar cane in Guangxi, wheat in Henan, vegetables in Shandong, and rice in Jiangsu. The agricultural carbon sequestration of Gansu, Qinghai, Ningxia, and Shaanxi are at the bottom of the list, all below 0.3 t/hm2.
As a comparison to carbon sequestration, the national average carbon emission intensity is 1.831 t/hm2, of which 0.923 t/hm2 is attributed to agricultural materials, 0.750 t/hm2 emanates from paddy fields, and 0.158 t/hm2 emanates from soils. This demonstrates that agricultural materials are the main contributor to carbon emissions, especially fertilizer, diesel, and agricultural films. The intensity of agricultural carbon emissions in most provinces is distributed between 1 and 4 t/hm2. Eight provinces, including Jiangxi (4.233 t/hm2), Shanghai (4.074 t/hm2), and Fujian (3.740 t/hm2), contribute the greatest amount to carbon emissions, with measurements all greater than 3 t/hm2. Contrastingly, 10 provinces, including Qinghai (0.365 t/hm2), Inner Mongolia (0.457 t/hm2), and Gansu (0.527 t/hm2), ranked lowest with less than 1 t/hm2 of recorded carbon emissions. The sources of carbon emissions vary greatly by province, with carbon emissions from rice paddies exceeding 60% in Jiangxi and Hunan, and agricultural carbon emissions accounting for over 90% in Qinghai and Xinjiang, while soil carbon emissions are low in all provinces, with the highest, Shanxi, accounting for only 21.7%. The provinces with high agricultural carbon emissions are mainly located in the central region where there are differences in agricultural operations and resource use, and the transition to low-carbon agriculture may require a differentiated approach.

3.2. Evaluation and Analysis of AEE in China

The Super-SBM model for undesirable outputs was applied, based on MATLAB2020b software, to evaluate the AEE per province from 2000 to 2020 under the carbon neutrality target. The results are shown in Table 4 and Figure 3.

3.2.1. Overall evolution of the AEE

In Figure 3, we can observe that China’s AEE declined but was followed by an increase between 2000 and 2020. The national average value of AEE in 2000 was 0.8645, followed by an oscillating decline, reaching the lowest value of 0.7125 in 2008, and then slowly rising to 0.8718 in 2020, with an overall upward U-shaped distribution. In terms of regional differences, the AEE value of the northeast region (1.0778) maintained its lead, ranking far above the national average, which has already reached an effective state, and showing a slow upward trend from 2000 to 2020 with an increase of 6.9% overall. In the eastern region, the AEE value presents a U-shaped change, with slowly declining rates from 2000 to 2008, which then fluctuate upward, reaching an efficiency value of 1.0171 in 2020. In the western region, AEE values fluctuated greatly from 2000 to 2009, and then much like the eastern region, shifted toward a trend of first decreasing and then increasing, with an average efficiency value of 0.7096 from 2000 to 2020. The central region has the lowest mean value of AEE at 0.5713. Overall, there are clear regional differences in AEE values; the efficiency of the northeast region is the highest, followed by the eastern and western, and the central region is the lowest.

3.2.2. Inter-Provincial Variation in AEE

Ten provinces, accounting for 33.3% of all provinces, have mean AEE values above the national average from 2000 to 2020: Jilin, Heilongjiang, and Liaoning in the northeast; Hainan, Beijing, Shanghai, Guangdong, and Jiangsu in the east; and Guangxi and Xinjiang in the west. Among them, eight provinces, including Hainan and Beijing, have efficiency averages above 1 and have reached the DEA effective status in the vast majority of years. Ningxia, Shanxi, Anhui, Hunan, Hubei, Gansu, Shaanxi, and Jiangxi have relatively low AEE, with mean values below 0.6, and all belong to the central and western regions, with Ningxia having the lowest efficiency value, at a mean value of 0.4628, which is only 39.2% of the mean value of the highest-ranking province, Hainan, and 59.9% of the mean value of the national average. It can be seen that from the perspective of geographic zones, the AEE of some provinces in the central and western regions is not high enough. In the future, emphasis in these regions should be placed on the integrated development of resource allocation and low-carbon transformation, the establishment of a synergistic development mechanism through exerting the demonstration effect and catch-up effect, and the improvement of the overall level of AEE.

3.3. Spatiotemporal Characteristics of AEE in China

3.3.1. Kernel Density Estimation of AEE

Using Stata16 software, the Kernel density function was applied to estimate the AEE values under the carbon neutrality target in 2000, 2005, 2010, 2015, and 2020, and a kernel density curve was drawn (Figure 4). The investigation of the dynamic evolution characteristics yielded the following observations. (1) From the position of the center of gravity of the kernel density curve, which shifts left and then right from 2000 to 2020, it is clear that China’s inter-provincial AEE under the carbon neutrality target first declines and then increases. (2) From the shape of the kernel density curves, where all five curves show the coexistence of the main peak and the secondary peak, we see the suggestion of a significant pattern of polarization in AEE. The height of the main peak gradually increases from 2000 to 2015, and the gap between provinces gradually widens, while the gap between the peaks decreases and flattens out in 2020; this indicates that the AEE gap among provinces narrowed and the degree of polarization was reduced. (3) From the trailing edge of the kernel density curve, it can be noted that the left and right sides of the corresponding curves in 2000 and 2020 are similar, while the trailing edge on the right side of the corresponding curves in 2005, 2010, and 2015 is longer than the trailing edge on the left side. This shift shows a clustering of low values of AEE under the carbon neutrality target in this period. This demonstrates that the AEE of different provinces in China has different patterns in different periods, with different dynamic evolutionary characteristics such as development level and polarization degree.

3.3.2. Spatial Distribution Patterns of AEE

To visualize the spatial characteristics of AEE per province, this paper uses ArcGIS 10.8 software to create a visual map of AEE at five time points and an average point, which is shown in Figure 5.
Figure 5 demonstrates that during the study period, China’s AEE first decreased, then increased, and had obvious spatial distribution characteristics of clustering and contiguity. In 2000, AEE was relatively high, with 15 provinces having efficiency values greater than 1, accounting for 50%. These provinces were mainly concentrated in the northeast, but also included most of the eastern and some parts of the western regions, whereas low efficiency was concentrated in the central region, which again shows high efficiency in the north and south and low efficiency in the central region in terms of spatial distribution. The years 2005, 2010, and 2015 saw a decline in AEE, with the number of efficient provinces decreasing. The number of provinces with AEE over 1 fluctuated to nine, eight, and nine again, respectively, while the number of provinces with AEE less than 0.7 likewise fluctuated but showed an increase, shifting between 17, 16, and 18, respectively, with a particularly marked decline in the central and western regions, showing a spatial distribution characterized by the clustering of low values. In 2020, AEE improved considerably, with only one province, Ningxia, having an efficiency value below 0.55, and with the number of provinces with AEE above 1 rising to 14. These regions are mainly clustered in the northeast and eastern regions, with a spatial distribution characterized by a concentration of high values and a gradual narrowing of the gap between regions.
As a whole, the spatial distribution of AEE in China has obvious core–periphery characteristics, whether from the perspective of five time points or the average efficiency distribution map from 2000 to 2020. As established, the efficiency of the central provinces is generally lower, while the eastern and northeastern provinces have higher efficiency rates, with the northeastern provinces always in the high-efficiency group. With its flat terrain, fertile resources, good lighting, and relatively high level of scale and intensification, the northeastern region is more conducive to promoting a low-carbon transition in agriculture and achieving a balance between carbon neutrality and agricultural production. The eastern region is economically developed to achieve more efficient and low-carbon operations while ensuring agricultural output via advanced technological tools and a strong policy environment. The western and central regions, however, have a relatively poor economic foundation, technological support has yet to be strengthened, and agricultural production needs to be transformed from extensive to refined.

3.4. Analysis on the Influencing Factors of AEE in China

When using the Tobit model, we can determine whether fixed or random effects should be used by testing for the presence of individual effects. The individual and random errors in the model are small, the variance ratios ρ are above 0.5, and the individual effect variances are large and all pass the LR test, strongly rejecting the original hypothesis. The Wald test passes the 1% significance test and the model worked well, so it is reasonable to use the random effects panel Tobit model.
According to Table 5, the coefficients of all variables pass the 1% or 5% significance level test. URBAN, ISU, FSFA, and MECH have positive effects, while ACS, DISA, and MCI have negative effects.
This summary more specifically means the following:
(1)
In relation to a region’s economic foundation, the urbanization rate positively affects AEE at the 1% significance level. This is mainly due to the fact that although urbanization brings about the loss of arable land and labor migration, it also induces a transition to more efficient specialization in agriculture as it increases the scarcity of inputs. Furthermore, agricultural productivity increases due to a rise in technological progress and the transformation of the industrial structure, which is brought about by urbanization [59]. The coefficient of ISU is 0.154, which significantly and positively affects AEE. This is because the industrial structure is optimized and upgraded, so the cluster effect and specialization effect gradually emerge, which not only reduces agricultural production costs and increases the added value of products, but also brings huge structural and scale dividends, which further help improve AEE [53].
(2)
In terms of production conditions, cropping structure negatively affects AEE at a significance level of 5%, which is consistent with the findings of [38], and rejects the assertion that cash crops increase the burden on the environment, and that it is feasible to moderately adjust agricultural cropping structure on the basis of ensuring food security. Natural disasters not only affect agricultural acreage and reduce crop yields and agricultural output, but also drive ecological degradation, which has a significant negative impact on AEE gains. The replanting coefficient, which reflects the intensity of cultivation of arable land, can increase agricultural value added, but can also bring about an increase in the number of tillage and chemical inputs. The replanting coefficient also requires an increase in input intensity, thereby bringing about an increase in undesired agricultural output, which likewise demonstrates the unsustainability of long-term and high-intensity cultivation [60].
(3)
As far as agricultural support policies are concerned, the financial support from the government for agriculture comes mainly in the form of investment and subsidies, which can support the construction of agricultural infrastructure and improve the input structure of agricultural production, thus enhancing AEE.
(4)
Regarding technological innovation, the widespread use of agricultural machinery can enhance production technology [61], resulting in improved productivity and production efficiency. However, increased machinery brings an increase in the use of petrochemical resources, so it is necessary to preferentially use low-carbon, green, and efficient agricultural machinery first.

4. Discussion and Policy Implications

Based on the measurement of agricultural carbon emissions and sequestration, this paper has explored the spatiotemporal characteristics of AEE as well as its influencing factors. The results from this study suggest, compared with rice fields and soil, that agricultural materials generate relatively large carbon emissions, which is consistent with [3]. Therefore, agricultural emission reduction and efficiency enhancement need to pay special attention to the control of agricultural material input. China’s AEE has obvious core–periphery characteristics and shows obvious regional differences. This means that a one-size-fits-all policy is no longer applicable, and it is necessary to balance the distribution of resources in different regions while improving the expected output. The main objectives are to reduce the external inputs, increase agricultural output value and carbon sink level, and reduce agricultural carbon emissions.
On the basis of the findings above, we make the following policy recommendations:
(1)
Deploy differentiated initiatives to reduce emissions and increase sequestration in agriculture. Specifically, work to improve the utilization rate of agricultural inputs [62], particularly in provinces such as Qinghai and Xinjiang, where the share of carbon emissions from agricultural inputs is relatively high. This can be achieved by encouraging the application of organic fertilizers and soil testing fertilizers, while promoting the resourceful use of straw. While the agricultural system works largely as a carbon sink, the agricultural cultivation structure should be further optimized to strengthen the carbon sequestration role of crops.
(2)
Assess the AEE under the carbon neutrality targets in each region, and change the behavior cease of pursuing high efficiency while ignoring environmental constraints. Agricultural carbon emissions and carbon sequestration must be central to the future research framework and focus on the balance between economy and environment. Furthermore, policies to promote AEE should be formulated in accordance with local conditions with consideration of regional differences in resource endowments, industrial structures, and economic bases. Agricultural carbon emission constraints should be made a government planning target, and any agricultural subsidy policies oriented towards green and low-carbon development should be constructed to cultivate and promote green agricultural technologies, so as to achieve a win-win situation for both environmental protection and effective allocation of scientific and technological resources.
(3)
The current problems of low AEE and regional imbalances require the development of cooperation plans for inter-regional collaboration, which must be formulated to balance the distribution of poorer and wealthier regional resources, to strengthen the supervision and management of resource elements, and to improve the allocation performance of various types of resources for AEE. The central and western regions have the opportunity to make large strides in the promotion of AEE by increasing investment in scientific research and by strengthening collaborative innovation [63]. The eastern and northeastern regions should continue to improve the level of resource allocation, increase research and development around core technologies, and play a leading role in the achievement of balanced and integrated AEE practices through active exchange and cooperation.
There are some shortcomings to this research. Firstly, our study only estimates agricultural carbon emissions and carbon sequestration from the narrow perspective of plantation agriculture, and lacks measurements that cover the broader scope of agricultural practices. Secondly, the coefficients for agricultural carbon accounting were obtained from the list published by the Chinese government and academic literature, but more accurate measurements could be taken, so there are some uncertainties around the current coefficients, thus affecting the reliability of the results to a certain extent. Thirdly, due to data availability constraints, this study lacks small-scale studies at the municipal or county level, and in the future, consideration will be given to a wider range of smaller-scale research units to improve accuracy.

5. Conclusions

The strategic goal of carbon neutrality has placed higher demands on the green and low-carbon development of agriculture in China. In this paper, we incorporated agricultural carbon emissions and sequestration into the model to build an AEE measurement model under the carbon neutrality target. On this basis, the Super-SBM model was employed to measure the AEE per province in order to assess the efforts toward reduced emissions and increased sequestration. Furthermore, we analyzed the spatial and temporal characteristics of AEE, and used the Tobit model to investigate the factors influencing AEE. The results of this study are as follows:
(1)
China’s agricultural system functions as a net carbon sink, with the agricultural carbon sequestration of all provinces from 2000 to 2020 measuring at higher rates than the carbon emissions. The national average carbon sequestration is 5.585 t/hm2 and the average net carbon sequestration is 3.754 t/hm2. Considering the national average carbon emission intensity of 1.831 t/hm2, including 0.923 t/hm2 for agricultural materials, 0.750 t/hm2 for paddy fields, and 0.158 t/hm2 for soils, it is clear that the use of agricultural materials is the main source of carbon emissions from agriculture.
(2)
From 2000 to 2020, the national average AEE was not high enough, with an average value of 0.7726, showing a trend of decreasing and then increasing, and there is still much room for improvement. In terms of spatial distribution, China’s AEE has obvious core–periphery characteristics and shows a clustered and contiguous spatial distribution, with central provinces generally having lower efficiency, eastern and northeastern provinces having higher efficiency, and northeastern provinces always in the high-efficiency group.
(3)
As for the influencing factors, urbanization, upgrading of industrial structure, financial support for agriculture, and mechanization can significantly contribute to AEE, with urbanization and financial support for agriculture having a greater degree of influence. In contrast, agricultural cultivation structure, agricultural disaster, and replanting have a negative impact on the AEE.

Author Contributions

Conceptualization, C.C. and Q.G.; methodology, C.C., J.L. and Y.Q.; software, C.C. and C.G.; validation, C.C. and J.L.; formal analysis, Q.G. and C.G.; writing—original draft preparation, C.C., J.L. and Y.Q.; writing—review and editing, C.C., J.L. and Q.G. 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 (71703077; 72203093) and the Scientific Research Startup Fund of Chuzhou University (2020qd21).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be available from the corresponding author on request.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Agricultural Carbon Emission Sources and Coefficients

Agricultural MaterialCoefficient
Fertilizer0.8956 kg C∙kg−1
Pesticide4.9341 kg C∙kg−1
Mulch5.1800 kg C∙kg−1
Diesel0.5927 kg C∙kg−1
Irrigation266.4800 kg C∙hm−2

Appendix B. CH4 Emission Coefficients of Different Rice Varieties in Chinese Provinces (Units: g∙m−2)

ProvinceEarly-Season RiceLate-Season RiceMid-Season RiceProvinceEarly-Season RiceLate-Season RiceMid-Season Rice
Beijing0013.23Henan0.000.0017.85
Tianjin0011.34Hubei17.513958.17
Hebei0015.33Hunan14.7134.156.28
Shanxi006.62Guangdong15.0551.657.02
Inner Mongolia008.93Guangxi12.4149.147.78
Liaoning009.24Hainan13.4349.452.29
Jilin005.57Chongqing6.5518.525.73
Heilongjiang008.31Sichuan6.5518.525.73
Shanghai12.4127.553.87Guizhou5.12122.05
Jiangsu16.0727.653.55Yunnan2.387.67.25
Zhejiang14.3734.557.96Shaanxi0012.51
Anhui16.7527.651.24Gansu006.83
Fujian7.7452.643.47Qinghai000.00
Jiangxi15.4745.865.42Ningxia007.35
Shandong0.000.0021.00Xinjiang0010.50

Appendix C. N2O Emission Coefficients of Soil From All Varieties of Crops (units: kg∙hm−2)

CropN2O Emission Coefficients
Rice0.24
Spring Wheat0.40
Winter wheat2.05
Soybeans0.77
Maize2.532
Vegetables4.21

Appendix D. Economic Coefficient and Carbon Sequestration Rate of Main Crops in China

CropEconomic CoefficientMoisture Content/%Sequestration RateCropEconomic CoefficientMoisture Content/%Sequestration Rate
Rice0.489120.414Yams0.667700.423
Wheat0.434120.485Sugar cane0.750500.450
Core0.438130.471Beet0.667750.407
Beans0.425130.450Vegetables0.830900.450
Rapeseed0.271100.450Melons0.700900.450
Peanut0.556100.450Tobacco0.830850.450
Sunflower0.300100.450Other crops0.400120.450
Cotton0.10080.450

References

  1. Ghommem, M.; Hajj, M.R.; Puri, I.K. Influence of natural and anthropogenic carbon dioxide sequestration on global warming. Ecol. Model. 2012, 235, 1–7. [Google Scholar] [CrossRef]
  2. Zhang, A.; Deng, R. Spatial-temporal evolution and influencing factors of net carbon sink efficiency in Chinese cities under the background of carbon neutrality. J. Clean. Prod. 2022, 365, 132547. [Google Scholar] [CrossRef]
  3. Wu, H.; Sipilainen, T.; He, Y.; Huang, H.; Luo, L.; Chen, W.; Meng, Y. Performance of cropland low-carbon use in China: Measurement, spatiotemporal characteristics, and driving factors. Sci. Total Environ. 2021, 800, 149552. [Google Scholar] [CrossRef] [PubMed]
  4. IEA. Global Energy Review: CO2 Emissions in 2021; IEA: Paris, France, 2021. [Google Scholar]
  5. Mi, Z.; Wei, Y.; Wang, B.; Meng, J.; Liu, Z.; Shan, Y.; Liu, J.; Guan, D. Socioeconomic impact assessment of China’s CO2 emissions peak prior to 2030. J. Clean. Prod. 2017, 142, 2227–2236. [Google Scholar] [CrossRef] [Green Version]
  6. Omri, A. Technological innovation and sustainable development: Does the stage of development matter? Environ. Impact Asses. 2020, 83, 106398. [Google Scholar] [CrossRef]
  7. Yu, Z.; Khan, S.; Ponce, P.; Jabbour, A.; Jabbour, C. Factors affecting carbon emissions in emerging economies in the context of a green recovery: Implications for sustainable development goals. Technol. Forecast. Soc. 2022, 176, 121417. [Google Scholar] [CrossRef]
  8. Huang, X.; Xu, X.; Wang, Q.; Zhang, L.; Gao, X.; Chen, L. Assessment of Agricultural Carbon Emissions and Their Spatiotemporal Changes in China, 1997–2016. Int. J. Environ. Res. Public Health 2019, 16, 3105. [Google Scholar] [CrossRef] [Green Version]
  9. Li, B.; Wang, C.; Zhang, J. Dynamic evolution and spatial spillover of China’s agricultural net carbon sink. China Popul. Resour. Environ. 2019, 29, 68–76. [Google Scholar]
  10. Huang, M.; Zeng, L.; Liu, C.; Li, X.; Wang, H. Research on the Eco-Efficiency of Rice Production and Its Improvement Path: A Case Study from China. Int. J. Environ. Res. Public Health 2022, 19, 8645. [Google Scholar] [CrossRef]
  11. Tian, Y.; Zhang, J.; Wu, X.; Cheng, L. Research on dynamic change and regional differences of China’s planting industry carbon sink surplus. J. Nat. Resour. 2015, 30, 1885–1895. [Google Scholar]
  12. Smith, P. Carbon sequestration in croplands: The potential in Europe and the global context. Eur. J. Agron. 2004, 20, 229–236. [Google Scholar] [CrossRef]
  13. Liu, W.; Zhang, G.; Wang, X.; Lu, F.; Ouyang, Z. Carbon footprint of main crop production in China: Magnitude, spatial-temporal pattern and attribution. Sci. Total Environ. 2018, 645, 1296–1308. [Google Scholar] [CrossRef] [PubMed]
  14. Linquist, B.; Adviento-Borbe, M.; Pittelkow, C.; van Kessel, C.; van Groenigen, K. Fertilizer management practices and greenhouse gas emissions from rice systems: A quantitative review and analysis. Field Crops Res. 2012, 135, 10–21. [Google Scholar] [CrossRef]
  15. Begum, K.; Kuhnert, M.; Yeluripati, J.; Ogle, S.; Parton, W.; Kader, M.; Smith, P. Model Based Regional Estimates of Soil Organic Carbon Sequestration and Greenhouse Gas Mitigation Potentials from Rice Croplands in Bangladesh. Land 2018, 7, 82. [Google Scholar] [CrossRef] [Green Version]
  16. Jiang, Z.; Yin, S.; Zhang, X.; Li, C.; Shen, G.; Zhou, P.; Liu, C. Research and Development of a DNDC Online Model for Farmland Carbon Sequestration and GHG Emissions Mitigation in China. Int. J. Environ. Res. Public Health 2017, 14, 1493. [Google Scholar] [CrossRef] [Green Version]
  17. Su, Y.; He, S.; Wang, K.; Shahtahmassebi, A.; Zhang LZhang, J.; Zhang, M.; Gan, M. Quantifying the sustainability of three types of agricultural production in China: An emergy analysis with the integration of environmental pollution. J. Clean. Prod. 2020, 252, 119650. [Google Scholar] [CrossRef]
  18. Rao, J.; Ji, X.; Ouyang, W.; Zhao, X.; Lai, X. Dilemma Analysis of China Agricultural Non-point Source Pollution Based on Peasants’ Household Surveys. In Proceedings of the 18th Biennial ISEM Conference on Ecological Modelling for Global Change and Coupled Human and Natural Systems, Beijing, China, 20–23 September 2011; Beijing Normal University: Beijing, China, 2012. [Google Scholar]
  19. Fan, M.; Shen, J.; Yuan, L.; Jiang, R.; Chen, X.; Davies, W.; Zhang, F. Improving crop productivity and resource use efficiency to ensure food security and environmental quality in China. J. Exp. Bot. 2012, 63, 13–24. [Google Scholar] [CrossRef] [PubMed]
  20. Wang, J.; Lin, Y.; Glendinning, A.; Xu, Y. Land-use changes and land policies evolution in China’s urbanization processes. Land Use Policy 2018, 75, 375–387. [Google Scholar] [CrossRef]
  21. Shi, H.; Xu, H.; Gao, W.; Zhang, J.; Chang, M. The impact of energy poverty on agricultural productivity: The case of China. Energy Policy 2022, 167, 113020. [Google Scholar] [CrossRef]
  22. Snyder, C.; Bruulsema, T.; Jensen, T.; Fixen, P. Review of greenhouse gas emissions from crop production systems and fertilizer management effects. Agric. Ecosyst. Environ. 2009, 133, 247–266. [Google Scholar] [CrossRef]
  23. Yang, Q.; Si, X.; Wang, J. The measurement and its distribution dynamic evolution of grain production efficiency in China under the goal of reducing pollution emissions and increasing carbon sink. J. Nat. Resour. 2022, 37, 600–615. [Google Scholar] [CrossRef]
  24. Pan, D.; Ying, R. Agricultural eco-efficiency evaluation in China based on SBM model. Acta Ecol. Sin. 2013, 33, 3837–3845. [Google Scholar] [CrossRef]
  25. Maia, R.; Silva, C.; Costa, E. Eco-efficiency assessment in the agricultural sector: The Monte Novo irrigation perimeter, Portugal. J. Clean. Prod. 2016, 138, 217–228. [Google Scholar] [CrossRef] [Green Version]
  26. Nsiah, C.; Fayissa, B. Trends in Agricultural Production Efficiency and their Implications for Food Security in Sub-Saharan African Countries. Afr. Dev. Rev. 2019, 31, 28–42. [Google Scholar] [CrossRef] [Green Version]
  27. Hou, X.; Liu, J.; Zhang, D.; Zhao, M.; Xia, C. Impact of urbanization on the eco-efficiency of cultivated land utilization: A case study on the Yangtze River Economic Belt, China. J. Clean. Prod. 2019, 238, 117916. [Google Scholar] [CrossRef]
  28. Yang, B.; Zhang, Z.; Wu, H. Detection and attribution of changes in agricultural eco-efficiency within rapid urbanized areas: A case study in the Urban agglomeration in the middle Reaches of Yangtze River, China. Ecol. Indic. 2022, 144, 109533. [Google Scholar] [CrossRef]
  29. Yang, B.; Wang, Z.; Zou, L.; Zou, L.; Zhang, H. Exploring the eco-efficiency of cultivated land utilization and its influencing factors in China’s Yangtze River Economic Belt, 2001–2018. J. Environ. Manag. 2021, 294, 112939. [Google Scholar] [CrossRef] [PubMed]
  30. Yang, H.; Wang, X.; Bin, P. Agriculture carbon-emission reduction and changing factors behind agricultural eco-efficiency growth in China. J. Clean. Prod. 2022, 334, 130193. [Google Scholar] [CrossRef]
  31. Hou, M.; Xi, Z.; Zhao, S. Evaluating the Heterogeneity Effect of Fertilizer Use Intensity on Agricultural Eco-Efficiency in China: Evidence from a Panel Quantile Regression Model. Int. J. Environ. Res. Public Health 2022, 19, 6612. [Google Scholar] [CrossRef] [PubMed]
  32. Chen, Y.; Miao, J.; Zhu, Z. Measuring green total factor productivity of China’s agricultural sector: A three-stage SBM-DEA model with non-point source pollution and CO2 emissions. J. Clean. Prod. 2021, 318, 128543. [Google Scholar] [CrossRef]
  33. Todorovic, M.; Mehmeti, A.; Scardigno, A. Eco-efficiency of agricultural water systems: Methodological approach and assessment at meso-level scale. J. Environ. Manag. 2016, 165, 62–71. [Google Scholar] [CrossRef] [PubMed]
  34. Angulo-Meza, L.; Gonzalez-Araya, M.; Iriarte, A.; Rebolledo-Leiva, R.; de Mello, J.C.S. A multiobjective DEA model to assess the eco-efficiency of agricultural practices within the CF plus DEA method. Comput. Electron. Agric. 2019, 161, 151–161. [Google Scholar] [CrossRef]
  35. Deng, X.; Gibson, J. Improving eco-efficiency for the sustainable agricultural production: A case study in Shandong, China. Technol. Forecast. Soc. 2019, 144, 394–400. [Google Scholar] [CrossRef]
  36. Quiroga, S.; Suarez, C.; Fernandez-Haddad, Z.; Philippidis, G. Levelling the playing field for European Union agriculture: Does the Common Agricultural Policy impact homogeneously on farm productivity and efficiency? Land Use Policy 2017, 68, 179–188. [Google Scholar] [CrossRef]
  37. Jin, G.; Li, Z.; Deng, X.; Yang, J.; Chen, D.; Li, W. An analysis of spatiotemporal patterns in Chinese agricultural productivity between 2004 and 2014. Ecol. Indic. 2019, 105, 591–600. [Google Scholar] [CrossRef]
  38. Liu, Y.; Zou, L.; Wang, Y. Spatial-temporal characteristics and influencing factors of agricultural eco-efficiency in China in recent 40 years. Land Use Policy 2020, 97, 104794. [Google Scholar] [CrossRef]
  39. Guo, Y.; Tong, L.; Mei, L. Spatiotemporal characteristics and influencing factors of agricultural eco-efficiency in Jilin agricultural production zone from a low carbon perspective. Environ. Sci. Pollut. R. 2022, 29, 29854–29869. [Google Scholar] [CrossRef] [PubMed]
  40. Zhao, Q.; Bao, H.; Zhang, Z. Off-farm employment and agricultural land use efficiency in China. Land Use Policy 2021, 101, 105097. [Google Scholar] [CrossRef]
  41. Li, B.; Zhang, J.; Li, H. Research on spatial-temporal characteristics and affecting factors decomposition of agricultural carbon emission in China. China Popul. Resour. Environ. 2011, 21, 80–86. [Google Scholar]
  42. Tian, Y.; Zhang, J.; He, Y. Research on Spatial-Temporal Characteristics and Driving Factor of Agricultural Carbon Emissions in China. J. Integr. Agric. 2014, 13, 1393–1403. [Google Scholar] [CrossRef] [Green Version]
  43. Commission PNDaR. Guidelines for the Preparation of Provincial Greenhouse Gas Inventories (Trial); PRC National Development and Reform Commission: Beijing, China, 2011. [Google Scholar]
  44. Min, J.; Hu, H. Calculation of greenhouse gases emission from agricultural production in China. China Popul. Resour. Environ. 2012, 22, 21–27. [Google Scholar]
  45. Chen, L.; Xue, L.; Xue, Y. Spatial-temporal characteristics of China’s agricultural net carbon sink. J. Nat. Resour. 2016, 31, 596–607. [Google Scholar]
  46. IPCC. Working Group I: The Physical Science Basis; IPCC Fourth Assessment Report Climate Change; IPCC: Geneva, Switzerland, 2007. [Google Scholar]
  47. Tone, K. A slacks-based measure of efficiency in data envelopment analysis. Eur. J. Oper. Res. 2001, 130, 498–509. [Google Scholar] [CrossRef] [Green Version]
  48. Tone, K. A Slacks-Based Measure of Super-Efficiency in Data Envelopment Analysis. Eur. J. Oper. Res. 2002, 143, 32–41. [Google Scholar] [CrossRef] [Green Version]
  49. Rosenblatt, M. Remarks on Some Nonparametric Estimates of a Density Function. Ann. Math. Stat. 1956, 27, 832–837. [Google Scholar] [CrossRef]
  50. Zhang, J.; Zeng, W.; Shi, H. Regional environmental efficiency in China: Analysis based on a regional slack-based measure with environmental undesirable outputs. Ecol. Indic. 2016, 71, 218–228. [Google Scholar] [CrossRef]
  51. Tobin, J. Estimation of Relationships for Limited Dependent Variables. Econometrica 1958, 26, 24–36. [Google Scholar] [CrossRef] [Green Version]
  52. Huang, X.; Feng, C.; Qin, J.; Wang, X.; Zhang, T. Measuring China’s agricultural green total factor productivity and its drivers during 1998–2019. Sci. Total Environ. 2022, 829, 154477. [Google Scholar] [CrossRef] [PubMed]
  53. Gao, Q.; Cheng, C.; Sun, G.; Li, J. The Impact of Digital Inclusive Finance on Agricultural Green Total Factor Productivity: Evidence From China. Front. Ecol. Evol. 2022, 10, 905644. [Google Scholar] [CrossRef]
  54. Wang, M.; Li, Y.; Liao, G. Research on the Impact of Green Technology Innovation on Energy Total Factor Productivity, Based on Provincial Data of China. Front. Environ. Sci. 2021, 9, 710931. [Google Scholar] [CrossRef]
  55. Luo, X.; Ao, X.; Zhang, Z.; Wan, Q.; Liu, X. Spatiotemporal variations of cultivated land use efficiency in the Yangtze River Economic Belt based on carbon emission constraints. J. Geogr. Sci. 2020, 30, 535–552. [Google Scholar] [CrossRef]
  56. Wu, H.; Huang, H.; Chen, W.; Meng, Y. Estimation and spatiotemporal analysis of the carbon-emission efficiency of crop production in China. J. Clean. Prod. 2022, 371, 133516. [Google Scholar] [CrossRef]
  57. Chen, J.; Wang, L.; Li, Y. Research on the impact of multi-dimensional urbanization on China’s carbon emissions under the background of COP21. J. Environ. Manag. 2020, 273, 111123. [Google Scholar] [CrossRef]
  58. Yang, G.; Chen, Y. China’s agriculture carbon sink estimation and its coupling analysis with agricultural economy development. China Popul. Resour. Environ. 2016, 26, 171–176. [Google Scholar]
  59. Han, J.; Qu, J.; Maraseni, T.; Xu, L.; Zeng, J.; Li, H. A critical assessment of provincial-level variation in agricultural GHG emissions in China. J. Environ. Manag. 2021, 296, 113190. [Google Scholar] [CrossRef] [PubMed]
  60. Yin, Y.; Hou, X.; Liu, J.; Zhou, X.; Zhang, D. Detection and attribution of changes in cultivated land use ecological efficiency: A case study on Yangtze River Economic Belt, China. Ecol. Indic. 2022, 137, 108753. [Google Scholar] [CrossRef]
  61. Xue, C.; Shi, X.; Zhou, H. Influence path of agricultural mechanization on total factor productivity growth in planting industry. J. Agrotech. Econ. 2020, 10, 87–102. [Google Scholar]
  62. Huang, W.; Jiang, L. Efficiency performance of fertilizer use in arable agricultural production in China. China Agric. Econ. Rev. 2019, 11, 52–69. [Google Scholar] [CrossRef]
  63. Chen, J.; Wang, L.; Li, Y. Natural resources, urbanization and regional innovation capabilities. Resour. Policy 2020, 66, 101643. [Google Scholar] [CrossRef]
Figure 1. Study areas.
Figure 1. Study areas.
Ijerph 19 15478 g001
Figure 2. Agricultural carbon emissions and sequestration in China from 2000 to 2020 (t/hm2).
Figure 2. Agricultural carbon emissions and sequestration in China from 2000 to 2020 (t/hm2).
Ijerph 19 15478 g002
Figure 3. Evolution of AEE in China and four regions from 2000 to 2020.
Figure 3. Evolution of AEE in China and four regions from 2000 to 2020.
Ijerph 19 15478 g003
Figure 4. Kernel density of AEE in major years.
Figure 4. Kernel density of AEE in major years.
Ijerph 19 15478 g004
Figure 5. Spatial distribution pattern of AEE in China.
Figure 5. Spatial distribution pattern of AEE in China.
Ijerph 19 15478 g005
Table 1. Agricultural carbon emissions and sequestration.
Table 1. Agricultural carbon emissions and sequestration.
Carbon EffectCategoryCauseFactorData RequiredReference
Carbon EmissionsAgricultural materialsThe production, application, and decomposition of fertilizers lead to carbon emissions.FertilizerConsumption of fertilizer[41]
The production, application, and decomposition of pesticides lead to carbon emissions.PesticideConsumption of pesticide
The production, application, and decomposition of mulches lead to carbon emissions.Agricultural filmThe amount of agricultural film used
The consumption of diesel by machinery leads to carbon emissions.DieselDiesel oil used in agriculture
The fossil fuels consumed for generating electricity in irrigation result in carbon emissions indirectly.IrrigationEffective irrigation area
Rice fieldsMethanogens in rice fields utilize organic matter from the roots of rice plants to form methane.Rice fieldPlanting area of early rice, medium rice, late rice[43,44]
SoilSoil surface releases carbon when planting crops.SoilYields of rice, winter wheat, spring wheat, soybeans, maize, vegetables[42]
Carbon Sequestration Crop sequestrationCrops absorb carbon dioxide through photosynthesis.CropYield of various crops, such as rice, wheat, maize, pulses, vegetables[42]
Table 2. Input and output variables for the measure of AEE.
Table 2. Input and output variables for the measure of AEE.
TypeVariableExplanationUnits
Input indicatorsLaborThe number of agricultural practitioners104 person
LandTotal sown areas of crops103 ha
MachineryTotal power of agricultural machinery104 kW
FertilizerApplication quantity of chemical fertilizer104 t
IrrigationEffective irrigation area103 ha
Output indicatorsDesirable outputActual output value of agriculture108 CNY
Agricultural carbon sequestration104 t
Undesirable outputAgricultural carbon emissions104 t
Table 3. Influencing factors of AEE growth.
Table 3. Influencing factors of AEE growth.
VariablesDescriptionMeanStd. Dev.MinMax
URBANUrbanization rate of resident population0.5140.1550.1960.896
ISU1 × Primary industrial added value/GDP + 2 × Secondary industrial added value/GDP+ tertiary industrial added value/GDP2.3360.1342.0692.834
ACSRatio of sown area of grain crops to total sown area of crops0.6580.1320.3540.971
DISARatio of disaster area to sown area0.2310.1620.0000.936
MCIRatio of total sown area of crops to total area of cultivated land1.4240.5070.4882.848
FSFARatio of agricultural financial expenditure to total financial expenditure0.0970.0360.0100.204
MECHRatio of total power of agricultural machinery to output of planting industry4.0742.1261.08311.781
Table 4. AEE in 30 Chinese provinces in major years.
Table 4. AEE in 30 Chinese provinces in major years.
Province20002005201020152020Average
Beijing1.16561.16261.17241.15481.02931.1442
Tianjin0.76080.64330.65310.69941.07190.7599
Hebei0.68120.57960.64790.63580.84510.6443
Shanxi0.50090.45780.48660.46030.58110.4871
Inner Mongolia0.76370.62240.6390.66510.77520.6568
Liaoning1.01901.01681.01731.05231.07371.0322
Jilin1.06771.10621.06151.11651.14521.1060
Heilongjiang1.07361.02631.13631.14311.15951.0953
Shanghai1.13171.10711.21101.08421.01321.0978
Jiangsu0.80370.70340.80121.01270.86450.8142
Zhejiang1.01200.64160.63770.64381.11250.7507
Anhui0.55370.46380.54290.53580.59030.5091
Fujian0.61440.62750.68130.69091.04170.7126
Jiangxi0.69550.55010.56570.61610.70260.5866
Shandong1.03060.68680.75210.74961.00150.7709
Henan1.01060.68340.76720.73580.87910.7602
Hubei0.80980.57500.52960.53190.56490.5514
Hunan0.66100.54660.55440.51340.58060.5335
Guangdong1.01500.82630.76440.78891.04050.8375
Guangxi1.07271.12261.17251.08751.12151.1223
Hainan1.22931.17771.18731.15401.15081.1792
Chongqing0.75020.69830.68760.67010.67910.6528
Sichuan1.01720.71790.70890.64610.71550.6920
Guizhou1.04500.68360.4980.49030.65600.6098
Yunnan1.07680.72220.70480.64260.77370.7610
Shaanxi0.61350.61300.61940.57650.62860.5821
Gansu0.61960.54990.53160.53000.69400.5703
Qinghai0.49011.01950.49560.42171.06500.6287
Ningxia0.50070.42380.47750.46000.53000.4628
Xinjiang1.14861.04131.09261.07931.06391.0674
Eastern0.94440.81560.85080.86141.01710.8711
Northeastern1.05341.04981.07171.10401.12611.0778
Central0.70530.54610.57440.56560.64980.5713
Western0.82710.74680.69340.66080.79110.7096
Average0.86450.75990.75990.75300.87170.7726
Table 5. Influencing factors of AEE.
Table 5. Influencing factors of AEE.
VariableCoefficientStandard ErrorZ-StatisticProbability
URBAN0.6480.03319.850.000 ***
ISU0.1540.0354.360.000 ***
ACS−0.0720.030−2.390.017 **
DISA−0.0260.012−2.150.031 **
MCI−0.0250.006−4.480.000 ***
FSFA0.5400.0816.650.000 ***
MECH0.0040.0022.390.017 **
Notes: **, *** denote statistical significance at 5% and 1%, respectively.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cheng, C.; Li, J.; Qiu, Y.; Gao, C.; Gao, Q. Evaluating the Spatiotemporal Characteristics of Agricultural Eco-Efficiency Alongside China’s Carbon Neutrality Targets. Int. J. Environ. Res. Public Health 2022, 19, 15478. https://doi.org/10.3390/ijerph192315478

AMA Style

Cheng C, Li J, Qiu Y, Gao C, Gao Q. Evaluating the Spatiotemporal Characteristics of Agricultural Eco-Efficiency Alongside China’s Carbon Neutrality Targets. International Journal of Environmental Research and Public Health. 2022; 19(23):15478. https://doi.org/10.3390/ijerph192315478

Chicago/Turabian Style

Cheng, Changming, Jieqiong Li, Yuqing Qiu, Chunfeng Gao, and Qiang Gao. 2022. "Evaluating the Spatiotemporal Characteristics of Agricultural Eco-Efficiency Alongside China’s Carbon Neutrality Targets" International Journal of Environmental Research and Public Health 19, no. 23: 15478. https://doi.org/10.3390/ijerph192315478

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