Next Article in Journal
Suspended Matter and Hydrocarbons Fluxes in the Kara and Laptev Seas
Previous Article in Journal
Depth-Dependent Concentrations of E. coli in Agricultural Irrigation Ponds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nitrate Contamination in Groundwater: Evaluating the Effects of Demographic Aging and Depopulation in an Island with Intensive Citrus Cultivation

1
Graduate School of Integrated Arts and Sciences, Hiroshima University, 1-7-1, Kagamiyama, Higashi-Hiroshima 739-8521, Japan
2
Graduate School of Advanced Science and Engineering, Hiroshima University, 1-7-1, Kagamiyama, Higashi-Hiroshima 739-8521, Japan
3
Fukuoka Labor Bureau, 2-11-1, Hakataekihigashi, Hakata-ku, Fukuoka 812-0006, Japan
4
Graduate School of Humanities and Social Sciences, Hiroshima University, 1-7-1, Kagamiyama, Higashi-Hiroshima 739-8521, Japan
*
Author to whom correspondence should be addressed.
Water 2022, 14(14), 2277; https://doi.org/10.3390/w14142277
Submission received: 12 June 2022 / Revised: 16 July 2022 / Accepted: 18 July 2022 / Published: 21 July 2022
(This article belongs to the Section Water Quality and Contamination)

Abstract

:
Despite rapid population aging and depopulation in Japan, groundwater nitrate contamination still poses serious environmental problems. One of the main factors contributing to elevated nitrate levels in Japanese groundwater sources is agricultural intensification, frequently because of increased fertilizer use. We investigated the impact of population aging and depopulation on groundwater nitrate contamination on a western Japanese island that has extensive citrus farming. In comparison to Ocho village, where the average age of farmers is 73 years, Kubi village’s farmers are slightly older on average, at 76 years, and agricultural land has decreased by 46% over the past ten years, from 2005 to 2015. Ocho had 830 residents, which was twice as many as Kubi. In comparison to Ocho (4.8 mg/L), Kubi village had higher average NO3−N concentrations (6.6 mg/L). NO3−N contamination sources in Kubi and Ocho were determined using stable nitrogen isotopes and a Bayesian isotope mixing model. These source distributions were more strongly associated with social and land use factors. It was believed that the older farmers in Kubi employed a significant quantity of chemical fertilizers, which caused inefficient nitrogen uptake by plants, leading to increased leaching and more groundwater contamination than in Ocho.

1. Introduction

In recent years, demographic trends have raised concerns with respect to the effects of population growth on water resources. Despite rapid global population growth, population densities are declining in some regions and the proportion of aged people in most developed countries is rising. However, Japan has experienced this demographic transition faster than most Western and Asian countries, with an increasing number of citizens over the age of 65 [1]. However, despite these changes in its demographics, it still faces serious environmental problems related to groundwater pollution, with nitrates being one of the more serious groundwater contaminants [2,3,4].
Agriculture-related nitrate contamination of groundwater is a widespread problem in many countries, and numerous studies have demonstrated the correlation between poor agricultural practices and high nitrate concentrations in groundwater [5,6,7]. Groundwater contamination occurs when nitrate (NO3) input into the soil exceeds the nitrate consumption capacity of plants, resulting in nitrogen loss to the environment [8,9]. In Japan, use of chemical fertilizers has steadily increased since the 1960s [9]. As a result, the effectiveness of nitrogenous fertilizers has diminished, and the nitrogen remaining in the soil after absorption by crops has leached into the groundwater in the form of nitrates. Demographic aging is a major socioeconomic problem in many regions of Japan [1] and has led to the increase in usage of chemical fertilizers to maximize production. Previously, high NO3 levels have been found in the groundwater in Japanese agricultural islands [1,10,11], which was associated with intensive cultivation methods. The widespread use of fertilizers by farmers was believed to be the primary source of groundwater contamination; however, agricultural farmland has varied sources of contamination input. To effectively manage and control NO3 pollution, it is important to identify its sources [12]. Sources of NO3 in agricultural fields can be classified into four groups: chemical fertilizers, nitrogen present in soil, atmospheric deposition of nitrogen, and manure and sewage [13,14,15,16,17,18,19,20]. Thus, examination of the different sources of NO3 in groundwater is crucial, especially for implementation of groundwater conservation measures [21].
In the past, conventional methods for determining the sources of groundwater NO3 relied on the association between types of land use and variations in groundwater chemical composition [22]. Based on this method, most nitrate sources were identified with a high level of uncertainty. Therefore, a technique has been proposed for identifying NO3 sources using N and O isotopes [23,24,25]. Different NO3 sources are characterized by their isotopic signatures, which provide a means of tracing point and nonpoint sources of contamination in ground and surface waters. However, a concern arose in relation to overlapping ranges of the δ15N-NO3, which might lead to identification of inaccurate nitrate sources, thus compromising the usefulness of these isotopes alone. As a result, it was determined that δ18O-NO3 could provide additional information for distinguishing nitrate from chemical fertilizers and atmospheric deposition because of its distinctive values compared to those of the overlapping N isotopes [25]. The isotopic compositions of N and O vary among different nitrate sources. Analysis of these isotopic compositions along with the NO3 content is an effective indicator for tracing the source of groundwater nitrate [15]. The typical ranges of δ15N-NO3 and δ18O-NO3 for various sources of NO3 pollution have been previously reported, following which the double isotope technique has been used to identify the sources of NO3 pollution in surface and groundwater [13,25,26,27,28,29]. Using these dual isotopes along with conventional tracing methods yields accurate results [30]. Overall, analysis of the isotopic composition of nitrates in groundwater can help in recognizing their sources such as synthetic or organic fertilizers, human waste or septic system effluents, and soil organic matter.
The Bayesian isotope mixing model has been used to calculate the proportionate contributions of the main nitrate sources in groundwater [16,31,32,33,34,35]. Previous studies have successfully used the δ15N-NO3 and δ18O-NO3 and the Bayesian isotopic mixing model to calculate the proportional contributions of the main sources of nitrates in groundwater samples [12,16,36]. The results have revealed that by using a Bayesian “fingerprint” for assessment of pollutants, the stable isotope analysis in R (SIAR) model could predict the proportional contributions of the probable NO3 sources. Nitrogen and oxygen isotopes can be used in conjunction with traditional methods to identify the sources of nitrates, and the SIAR model can be used to quantify the contributions of these sources of nitrate contamination in the groundwater to protect the ecosystem in intensive agricultural areas [12,35]. Nitrate contamination of groundwater and surface water in Osakishimojima Island is a cause for concern, as a previous study [11] has linked the elevated nitrate concentrations in water sources to the nitrogenous fertilizers used in citrus farms. Sometimes, the nitrate levels have been observed to exceed the World Health Organization’s [37] recommended guideline for maximum NO3 (50 mg/L) in drinking water or Japan’s recommended guideline for maximum NO3−N concentration (10 mg/L) in water supplies. Hence, exploring the spatial variation of NO3 and its sources in two study sites with different extents of areas under cultivation could provide insights into the sources contributing to NO3 pollution on this island, which would help in the effective management of groundwater sources, particularly considering that the island has a declining and aging population.
The aims of this investigation were: (1) to demonstrate how the dual isotope approach (δ15N-NO3 and δ18O-NO3) can be used to determine the dominant sources of nitrates in the shallow groundwater of a small island and (2) to evaluate the influence of social aspects such as population aging and depopulation on NO3 concentrations in groundwater.

2. Description of Study Site

This study was conducted in Osakishimojima (34°10′24″ N, 132°50′3″ E), a small island located in the Seto Inland Sea in Hiroshima Prefecture, western Japan (Figure 1), approximately 20 km WSW of Kure City. Groundwater samples from two neighboring villages and watersheds, Ocho in the eastern part of the island and Kubi in the northern part, were analyzed. The population sizes of Kubi and Ocho were approximately 413 and 830, respectively, in 2021 [38], and the average age of farmers was 76 and 73, respectively [39]. Osakishimojima has a warm temperate climate with all four seasons characterized by cold winters and hot summers. The average annual temperature is 15.6 C° and the average annual precipitation is approximately 1100 mm.
The land use of Osakishimojima Island includes areas allocated to citrus orchards, pine forests, residential areas, paddy weeds, and water bodies, and the proportions of these areas were 56.69, 36.46, 5.56, 0.82, and 0.49%, respectively [11]. Land use data were obtained from the Ministry of Land, Infrastructure, Transport, and Tourism of Japan [40] with a spatial resolution of 100 m, and detailed classification of land use was carried out using Google Earth to suit the purposes of this research. The agricultural land area in Ocho and Kubi was approximately 1.6 and 0.6 km2, respectively, which has decreased by 37 and 46%, respectively, in 10 years [40]. Further, citrus orchards and residential areas in the lowlands of Ocho and Kubi comprised 35.2 and 56.7% and 45.4 and 43.3%, respectively (Figure 2). The residential areas in these regions mainly consisted of scattered villages with low urbanization levels and substandard sewage treatment facilities. In addition, Ocho village has higher building density and larger surface area than Kubi village (low housing density, random distribution). General information about the farmers’ choice of fertilizers and agricultural practices in each village was determined based on oral discussions with the farmers and the employees of the agricultural companies in these areas; however, no formal interviews were conducted at the study sites during the survey.
Hydrogeological data for the study sites were unavailable; thus, only the geological outcrops were examined. The main geological features of Osakishimojima Island are Late Cretaceous non-alkaline felsic volcanic rocks and rhyolites (Figure S1). They are found on mountain slopes, and Late Pleistocene to Holocene marine and non-marine formations cover the basement rocks in the lowland.

3. Materials and Method

3.1. Sample Collection

Fifty-two water samples were collected in the study area in October 2021, which included 32 sites comprising 2 springs and 30 shallow wells in the Ocho area and 20 sites in the Kubi area consisting of 1 observation borehole, 2 rivers, and 17 shallow wells (Figure 1). The sampling location, depth, and utilization information for each sampling site were determined from GPS coordinates. All wells chosen for groundwater sampling were commonly used for domestic and/or agricultural purposes; these wells had a mean depth of 10.8 m and 4.84 m in the Kubi and Ocho villages, respectively. The groundwater pH, temperature, electrical conductivity (EC), dissolved oxygen (DO), and oxidation-reduction potential (ORP) were measured on-site using a portable multiparameter water quality meter (pH/COND Meter D-54, HACH LD0101, ORP Meter RM-20P). All samples were collected from a 0.5–1 m depth below the water level in the wells using a high-density polyethylene sampler and filtered through a 0.2 μm membrane into one 250 mL, three 10 mL, and one 100 mL high-density polyethylene sampling bottles. After filtration, the samples were stored at 4 °C and subsequently transported to the laboratory, where they were frozen at −4 °C until further hydrochemical and isotopic analyses. The concentrations of the major cations (Na+, K+, Mg2+, and Ca2+) and anions (Cl, NO3, and SO42−) were analyzed by ion chromatography with conductivity detection on ICS-2100 (Dionex integrion, Thermo Fisher Scientific, Waltham, MA, USA). Analytical precisions for cations were set at <0.2% for Na+ and <0.1% for Ca2+, Mg2+, and K+, while for anions, they were <0.1% for NO3 and Cl and <0.2% for SO42−. HCO3 was analyzed within 12 h of sample collection by acid titration using 0.02 N H2SO4, which was added to the sample until it reached the endpoint at pH 4.5. Dissolved organic carbon (DOC) was determined using a total organic carbon analyzer (Total Organic Carbon analyzer, SHIMADZU, Hiroshima University, Japan). The analyses of major ions and DOC were conducted at the Biogeochemistry Laboratory, Hiroshima University, Japan.
Isotope analyses were performed at the Institute of Geosciences, Yamanachi University, using a stable isotopic ratio mass spectrometer (MAT 253, Thermo Fisher Scientific, Waltham, MA, USA). Chemical reduction (cadmium and azide reduction) was used to quantify δ15NNO3 and δ18ONO3, respectively [41,42]. Briefly, to eliminate nitrite, the samples were treated with a 2.5 mmol/L sulfamic acid solution. The NO3 in the samples was then quantitatively converted to gaseous nitrous oxide (N2O) using a two-step reduction process. First, NO3 was reduced to nitrite with spongy cadmium, then nitrite was converted to nitrous oxide with an azide/acetic acid buffer, and finally, the δ15N and δ18O of N2O were determined. The stable isotope ratios of δ15N and δ18O were given in delta value (δ) and per mil (‰) notation in comparison to international standards as follows [13]:
δ s a m p l e = R s a m p l e / R s t a n d a r d 1 × 1000
where Rsample and Rstandard are the 15N/14N and 18O/16O ratios of the samples and standards for δ15N and δ18O, respectively. The δ15N values were reported relative to N2 in atmospheric air (AIR) and the δ18O values relative to Vienna Standard Mean Ocean Water (VSMOW). The precision values for δ15N and δ18O were ±0.3‰ and ±0.4‰, respectively. Two representative samples of organic and chemical fertilizer were collected from local farmers in Kubi village in January 2022 for analysis as end-member samples. The solid samples were transported to the laboratory for air-drying and then fully ground for chemical and isotopic analyses. The solid samples were converted to N2 by high-temperature redox reactions, and δ15N-N2 was measured using an isotopic ratio mass spectrometer. The precision for δ15N was ±0.3‰.

3.2. Source Apportionment Calculation

The Stable Isotope Analysis in R (SIAR) software program was used to develop the mixing model. The system model used in SIAR can be expressed by defining a set of N mixture measurements on J isotopes with K source contributors, and the mixing model can be expressed as follows [25]:
X i j = k = 1 k P k P j k + C j k + ε j k S j k ~   N   μ j k , ω j k 2 C j k ~   N   λ j k , τ j k 2 ε j k ~   N   O , j 2
where Xij is the isotope value j of mixture i, in which i = 1, 2, 3, …, I and j = 1, 2, 3, …, J; Sjk is the source value k of isotope j (k = 1, 2, 3, …, K) and is normally distributed with a mean value μjk and a standard deviation ωjk; Pk is the proportion of source k, which was estimated using the SIAR model; Cjk is the fractionation factor for isotope j of source k and is normally distributed with a mean value λjk; τjk is the standard deviation; and εij is the residual error representing the additional unquantified variation between individual mixtures and is normally distributed with a mean value of 0 and a standard deviation σj. Details of this model have been described in previous studies [8,43].
A Bayesian model was applied to determine the potential NO3 contribution from each of four sources for Kubi and Ocho villages: δ15N-NO3 in manure and sewage (M&S), chemical fertilizer (CF), and organic fertilizer (OF); and soil nitrogen (SN) × (k = 4). The end-member isotopic compositions for the three sources were used (j = 3) in the model setup. SN and M&S samples were not analyzed in the study area, so the SN values [25,33,34,35] were compared with the N present in the CF samples, and because similar values were obtained, they were considered as a single source (CF_SN) of N in this study area. The M&S values were obtained from the published literature and used in the model as an additional source. We considered M&S to be a suspension of water and solid waste from humans and animals transported by sewers and OF exclusively to be a mixture of chicken litter and sawdust. Although discrete samples of individual 15NNO3 sources in the study area were measured, the limited number obtained may induce a biased model outcome. Alternatively, the mean values of δ15N-NO3 isotopes (μjk) and associated SDs (ωjk) of different M&S were acquired by consulting previously published data [25,33,34,35], whereas CF and OF were obtained from the Kubi area.

3.3. Statistical Method

Statistical analysis was performed to determine the NO3−N relationship with DOC and the major ions using a scattered plot, which was demonstrated by the coefficient of determination (R2). In addition to correlation analysis, principal component analysis (PCA) was used to analyze groups of ions based on similarities or differences in concentrations. SPSS® ver. 23.2 (SPSS Inc., Chicago, IL, USA) was used to analyze the groundwater chemical data. PCA can explain most of the variability in a dataset with fewer variables than the original dataset, so parameters with similar loadings in the same principal component (PC) group may have similar contributory sources or geochemical behaviors [44,45]. PCA was used to identify the primary groundwater contamination indicators using the Kaiser varimax rotation method, which only accepts PCs with eigenvalues greater than one. An inverse distance weighting method (IDW) was also performed using ArcMap 10.6 to show the spatial distribution of NO3−N and δ15N-NO3 in Kubi and Ocho groundwaters.

4. Results and Discussion

4.1. General Hydrochemistry

The hydrochemical compositions of Kubi and Ocho are summarized in Table 1. The results showed that the pH of groundwater was near neutral in Kubi (6.4 to 7.5, mean pH 6.8). The pH values of the river water sample were neutral to alkaline (7.9 to 8.6) and higher than the groundwater pH values, which may be due to the strong evaporation of the surface water. In Ocho village, groundwater was slightly acidic to alkaline with pH values ranging from 6.2 to 8.5 with a mean of 6.7. All samples had moderate EC values varying from 106.3 to 472 μS/cm, (mean 306.7 μS/cm) in Kubi and 164.9 to 479 μS/cm (mean 308.5 μS/cm) in Ocho. The concentrations of total dissolved solids (TDSs) were estimated from the EC values using the equation mentioned by Evangelou [46] and are shown in Table 1. The DOC levels of groundwater in Kubi ranged from 2.5 to 13.33 mg/L (mean 5.1 mg/L), while those in Ocho ranged from 1.6 to 7.0 mg/L (mean 3.5 mg/L). The DO ranged from 1.1 mg/L to 8.6 mg/L with a mean of 4.9 in the wells in Kubi, while those in Ocho ranged from 0.8 mg/L to 9.1 mg/L with a mean value of 3.4 mg/L. All but one sample from Kubi had DO values above 2 mg/L, whereas in Ocho, nine samples had values less than 2 mg/L. Additionally, observed DO levels were higher in naturally occurring springs (mean 9.2 mg/L in Kubi and 8.6 mg/L in Ocho). Hence, groundwater samples in Ocho had the potential to undergo denitrification processes as opposed to those in Kubi.
Calcium was the most dominant cation in both study areas and ranged from 7.8–38.1 mg/L in groundwater samples and from 9.8 mg/L−35.5 mg/L in surface water samples in Kubi while in the Ocho area, concentrations ranged from 0.5 mg/L–4.6 mg/L in surface water and from 8.8 mg/L−50.8 mg/L in groundwater samples. Magnesium concentrations ranged from 1.6 mg/L−15.3 mg/L in groundwater samples and 4.5–14.4 mg/L in surface water in Kubi whereas in Ocho village, these concentrations where much lower, varying from 4.9 mg/L–22.2 mg/L in groundwater samples (Table 2). Calcium and magnesium can contaminate groundwater through ion exchange when minerals dissolve in rocks or when fertilizers run off and leach into the groundwater. According to Earle [47], soils containing CaCO3 and CaSO4 have high levels of Ca2+ in the groundwater because of leaching, and Keesari et al. [48] suggested that agricultural runoff and sewage disposal may lead to high Mg2+ levels in the groundwater. The most abundant anion detected in the water samples from both villages was HCO3. In the Kubi area, concentrations of HCO3 in the river samples ranged from 50.6–61.2 mg/L with a mean value of 55.9 mg/L, while HCO3 concentrations in wells ranged from 14.2–121.6 mg/L with a mean value of 45.3 mg/L (Table 1). Bicarbonate concentrations in Ocho ranged from 4.5 to 64.2 mg/L in the wells and from 4 to 9.25 mg/L in the spring samples. The dissolved species in the samples obtained from both villages (arranged in sequence from highest to lowest average concentrations) were Ca2+ > Na+ > Mg2+ > K+ cations and HCO3 > SO42− > NO3 > Cl anions. The chemical compositions of the water samples from Kubi and Ocho villages were examined and classified using trilinear plots (Figure 3).
The two groundwater facies in the Kubi area were Ca−HCO3 and Ca−SO4 water types, comprising 55 and 45% of the water samples, respectively. Meanwhile, the trilinear analysis for Ocho village suggested that 58.7% of the water samples were composed of Ca−HCO3 water type, while 41.3% of the water samples were plotted in the zones of CaMg−SO4 or CaMg−Cl water types. Groundwater samples from both areas showed high percentages of freshwater signatures in the Ca−HCO3 zone; however, they showed susceptibility to anthropogenic contamination as they were placed between Ca−HCO3 and CaMg−SO4 or CaMg−Cl types in the trilinear plots.

4.2. Spatial Variations in the NO3−N Concentration

When the nitrate nitrogen (NO3−N) concentration exceeds the threshold value of 3 mg/L, it is deemed to be affected by human factors [24,50]. NO3−N concentrations were calculated by converting NO3 to NO3−N using the equation documented by Kimbi et al. [11]. A mean value of 6.6 mg/L and a median value of 6.5 mg/L were observed in Kubi, with NO3−N concentrations ranging from 0.5−13 mg/L (Table S1). Approximately 33.3% of the groundwater samples had NO3−N concentrations below 5 mg/L, 47.7% of the samples had concentrations ranging between 5−10 mg/L, and 19% of the samples had concentrations above 10 mg/L. Meanwhile, NO3−N concentrations in Ocho varied from 0.7−15.1 mg/L with mean and median values of 4.8 and 4.5 mg/L, respectively. These concentrations were much lower than those of Kubi. In Ocho village, over 62.5% of the groundwater samples had NO3−N concentrations less than 5 mg/L, 31.3% of the samples had concentrations between 5 and 10 mg/L, and 6.2% of the samples had NO3−N concentrations above 10 mg/L. As shown in Figure 4b, most of the samples showing higher concentrations of NO3−N were collected from sites located upstream of the lowland area, which was directly below the upland citrus cultivation zone.
Leaching of chemical fertilizers into the groundwater, which are subsequently transported downstream, could be responsible for the elevated NO3−N concentrations in Kubi. In contrast, Ocho village had lower NO3−N concentrations, with some sampling points showing relatively high NO3−N concentrations, including upstream, midstream, and downstream sampling points (Figure 4c). Groundwater contamination was likely caused by organic fertilizer inputs or leakages from the septic systems. The NO3−N concentration was observed to be between 3 and 8 mg/L at well depth between 4 and 6 m (Figure 5).
A pattern could be observed in Kubi, where groundwater NO3−N concentrations gradually increased from deep to shallow water depths, but in some samples, high concentrations were observed in the samples collected from deep water layers. These results indicate variations in NO3−N residence time in the groundwater system. The topography and hydrogeological properties of the aquifer might equally be influencing the distribution of groundwater contaminants [51]; however, this relationship could not be confirmed in this study because of unavailability of data. In Ocho, NO3−N concentrations did not change significantly with depth, and results for most samples were plotted at 5 m depth.

4.3. Relationship between NO3−N and Water Chemical Parameters

A correlation matrix table is presented in Table S2 for all samples. The relatively high concentration of NO3−N in the Kubi groundwater is consistent with the significantly elevated concentrations of Mg2+, Ca2+, and SO42− (Figure 6). There was a weak positive association among NO3−N, DOC, and Cl in both Kubi and Ocho villages. In the absence of a geological chloride source, nitrate and chloride are frequently utilized as indicators of anthropogenic groundwater contamination because of their positive correlations. While chloride is a conservative tracer that does not normally undergo biological transformation in groundwater, the denitrification process can reduce the nitrate concentration in groundwater [52,53]. Thus, the presence of this process in the groundwater samples from Kubi and Ocho might have reduced the degree of correlation between the parameters shown in Figure 5. Furthermore, the positive relationship between NO3−N and SO42− could be due to the use of chemical fertilizers such as (NH4)2SO42− [50]. A positive relationship has been suggested among NO3, Mg2+, and Ca2+ [34], indicating that they are derived from chemical fertilizers, such as CaCO3 and MgCO3. In Kubi, concentrations of NO3−N showed strong significant positive correlations with SO42− (R2 = 0.76, p < 0.01) as opposed to Ocho, which presented a weak significant positive relationship (R2 = 0.45, p < 0.01). These results are corroborated by the findings of earlier studies [54,55], which have reported correlations between SO42− and NO3 concentrations in groundwater.
More chemical fertilizers may be used in Kubi since the chemical fertilizer derived component was higher in Kubi than Ocho. These results are corroborated by the findings of earlier studies [54,55], which have reported correlations between SO42− and NO3 concentrations in groundwater, indicating that locations with high SO42− concentrations also have high NO3 concentrations. Another study [56] found that the application of CaCO3 fertilizer generated a positive relationship between NO3 and Ca2+. Similarly, in Kubi village (R2 = 0.63, p < 0.01), there was a considerable positive correlation between NO3−N and Ca2+ compared to that in Ocho (R2 = 0.22, p < 0.01), confirming inputs from chemical fertilizers. K+ showed no correlation with NO3−N in both study sites.

4.4. Principal Component Analysis Results

A three-factor model controlled the groundwater chemistry in Kubi, and the cumulative variance of the three PCs was 73.2% (Table 2). In Kubi, PC1 explained 40.9% of the total variation, with positive loadings of six variables in the order Mg2+ > SO42− > NO3−N > Na+ > TDS > Ca2+ (Table 2). A strong positive loading of Mg2+ and SO42− and a moderate positive loading of Ca2+ and TDS suggests a high level of mineralization of groundwater from natural sources, whereas a strong positive loading of NO3−N indicates anthropogenic pollution because of agricultural activities, such as fertilizer and manure usage and land use [57]. Several studies [44,47,58] have linked domestic sewage and agricultural fertilizers to the presence of Na+, Mg2+, K+, SO42, NO3, and Cl in groundwater. Thus, PC1 was denoted by the infiltration of agricultural fertilizers and domestic sewage. PC2, accounting for 19.5% of the total variance, showed a moderate positive loading for DO and a moderate negative loading for K+. PC3 explained 13% of the total variance, with a moderate loading of pH. The groundwater chemistry of Ocho was also studied using a three-factor model, and the cumulative variance explained by the three PCs was 67.8% (Table 2). PC1, accounting for 38.2% of the total variance, showed strong loadings for SO42− and NO3−N and moderate loadings for Ca2+, Mg2+, and Na+.
Although these variables can be interpreted as anthropogenic contamination from agricultural fields, these loadings were lower in comparison to the loadings of PC1 in the Kubi area, possibly because of differences in the rates of leaching and residence times of contaminants in the groundwater systems. PC2 explained 18% of the total variance with moderate positive loadings of three variables (HCO3, Cl, and TDS) and a negative loading of DO. We believe that the dissolution of carbonate rocks due to water–rock interactions was responsible for the loading of HCO3, while domestic sewage was the source of Cl. The third PC, which accounted for 11.7% of the total variance in Ocho village, had moderate positive and negative loadings for pH and K+, respectively. The results of the PC analysis indicate that natural processes involving mineral components and anthropogenic contamination because of fertilizers and domestic sewage are the main processes affecting the groundwater chemistry in Kubi and Ocho.

4.5. Distribution Characteristics of Stable Isotopes

Ten samples from Kubi (one river and nine well samples) and nine well samples from Ocho were analyzed for δ15N-NO3 and δ18O-NO3 (Table S3). Samples were selected primarily on the bases of nitrate levels and spatial distribution to represent each study site. The δ15N-NO3 values of Osakishimojima Island showed high variability in the Kubi and Ocho villages (Figure 3). In Kubi village, the δ15N-NO3 values of groundwater samples ranged from 3.7‰ to 13.3‰, with a mean value of 6.7‰, and the δ18O-NO3 values ranged from −1.0‰ to 2.8‰, with a mean value of 1.5‰. Similarly, in Ocho village, the δ15N-NO3 values ranged from 4.9‰ to 26.0‰, with a mean value of 11.4‰, and the δ18O-NO3 values ranged from 0.1‰ to 14.9‰, with a mean value of 5.3‰ (Table S3). A relatively high value of δ15N-NO3 (26‰) in sample OG6 may suggest other controlling processes in the catchment area, such as denitrification; however, since the sample size was small, it was excluded from further analysis to avoid bias estimation, especially in the SIAR model. The mean values of δ15N-NO3 and δ18O-NO3 in Kubi village were significantly lower than those in Ocho village. This was contrary to the spatial distribution of NO3−N concentrations (Figure 3). The reason for this is that more chemical fertilizers with lower isotopic values and faster leaching rate were used for citrus cultivation in Kubi than in Ocho. In general, the leaching rate is determined by the degree of water solubility of a fertilizer in combination with the amount of soil water at any time, soil type, porosity, and compaction [59]. Kubi and Ocho have the same soil type and soil properties, but due to the differences in their choice of fertilizer, variations in NO3−N concentrations may arise. Highly soluble fertilizers dissolve quite rapidly into the soil and since plants can only absorb a certain amount of nutrition at a time, a greater percentage of the fertilizer simply leaches through the soil into the groundwater. The nitrate sources were identified from the cross-plot of δ15N-NO3 and δ18O-NO3 (Figure 7). Most of the data points were plotted in the range of synthetic fertilizer and soil nitrogen in Kubi village, while in Ocho village, points were plotted mainly in the manure and sewage domain. This suggests that the major inputs of NO3 are from synthetic fertilizer in Kubi, whereas in Ocho, they are mainly from organic fertilizer. The groundwater chemistry in Figure 5 indicates that more chemical fertilizer was used in Kubi than in Ocho, which explains the significant differences in the isotopic values between the sites. Additionally, higher δ15N-NO3 concentrations may have been derived from sewage systems, whose end-member samples were not measured in this study.
Ocho has a higher housing density than Kubi (Figure 2), and since the area is a rural depopulating and aging community, the sewage system was poorly managed, and contamination of the groundwater system could have occurred because of leakages through worn-out pipes. According to the Kure City Water Works and Sewerage Bureau [60], Kubi has a local wastewater treatment plant, whereas Ocho depends solely on septic treatment tanks. Considering the housing density in Ocho, sewer discharge into the groundwater is more likely. The difference in the age structure and population size influences farming practices, such as the method of fertilizer and pesticide application and the application schedule, which are important processes for quality crop yields. Chemical fertilizers are water soluble which makes them readily available for plant uptake than organic fertilizers; thus, aged farmers prefer to use chemical fertilizers that are more cost-effective, less time consuming, and more convenient for maximum crop yield. However, because the nutrient requirements for plant growth, productivity, and fruit quality fluctuate seasonally, the split fertilization method is commonly utilized in citrus farming [61]. While this strategy is acceptable and reduces inorganic N leaching while boosting N use efficiency [62], it necessitates the utilization of an additional labor force, which is unsuitable for Kubi and Ocho as the population is both declining and aging [37,38] with a limited crop production labor force. In studies conducted in China by [62,63], the association between fertilization application methods, population aging, and restricted labor force has been observed. However, the fact that Kubi farmers tend to use more fertilizer may possibly be because of their lower population size when compared to Ocho [38].
Additionally, the proportion of area under cultivation may influence the quantity of fertilizer applied, subsequently contributing to elevated NO3−N. As observed in Figure 2, Kubi has approximately 40.2% of its total area occupied by orchard fields, which is twice the coverage area than that of Ocho village, and even though at the catchment scale, Ocho has almost thrice the agricultural land than Kubi [40], the pathways of nitrate in the two study sites may be different. Nevertheless, the overlapping δ15N-NO3. and δ18O-NO3 values of the different sources hinder the quantitative estimation of their contributions. Furthermore, a previous study reported related investigations and concluded that the link among δ15N, δ18O, and NO3 is a complex web in which all environmental settings should be considered to accurately distinguish between different biogeochemical processes [16]. Hence, biological and geochemical processes may occur simultaneously in Ocho village, resulting in a shift in δ15N and δ18O levels.

4.6. Source Apportionment of NO3

The proportional contributions of NO3 sources in groundwater were identified using the SIAR model. Two isotopes (J = 2) (δ15N-NO3 and δ18O-NO3) and three potential sources (K = 3) (OF, CF_SN, and M&S) were used to estimate the contributions of NO3 sources in Kubi and Ocho village. The δ15N-NO3 and δ18O-NO3 of groundwater samples (Table S3) and sources (both mean and SD in Table 3) were input into the SIAR model.
Sewage was not measured in the study area; however, considering the residential area in one of the study sites, we applied the literature value for the model, while atmospheric deposition was not considered. Furthermore, the fractionation factor Cjk was set to zero because we did not investigate to determine its value. The results of source apportionment are shown in Figure 8.
The source contributions showed significant spatial variation under similar cropping systems with different land use proportions. The dominant source of NO3 pollution in Kubi village was CF_SN, which contributed 60.1%, followed by OF, which contributed 30.1%, and M&S, which contributed 9.8%. An increase in the fertilizer source contribution in the research area could also be due to the irrigation system. According to another study conducted by [29], higher yields from irrigated crops are generally associated with higher nitrogen fertilizer applications, increasing the risk of N loss to the environment. For Ocho village, the proportional contribution decreased in the following order: M&S (10.9%) < CF_SN (38.4%) < OF (50.7%). The results of NO3 source apportionment with SIAR in Ocho corresponded with our field investigation and the land use distribution shown in Figure 2. Residential areas in Ocho are more concentrated than those in Kubi and have a low urbanization level due to depopulation. Contamination from old and poorly maintained drainage systems is highly probable. According to the Kure City Water Works and Sewerage Bureau [60], 70% of the wastewater generated in Kubi is treated in a local wastewater treatment plant, while the remaining 30% is treated in septic tanks, whereas in Ocho, all the waste is treated by septic tanks. The differences between the waste management systems may explain the differences in the NO3−N concentrations in each area.
In summary, chemical fertilizers are mainly used in Kubi and organic fertilizers are used in Ocho (Figure 7). Due to differences in the wastewater treatment systems, Ocho has higher values of sewage and manure contribution than Kubi [60], but its impact is not as high as that of organic fertilizer input; therefore, we concluded that the main source of nitrate pollution is organic fertilizers. Additionally, large quantities of manure, sewage, and chemical fertilizers are used in Kubi, suggesting a high rate of leaching, resulting in a higher concentration of nitrates in the groundwater system. Table 4 demonstrates the relationship between social factors and groundwater nitrate–nitrogen contamination in Kubi and Ocho villages, with Kubi having a smaller population and agricultural area and a higher concentration of NO3−N levels, primarily due to chemical fertilizers. Compared to Kubi village, Ocho village had lower NO3−N levels despite having a larger population, and the organic fertilizers used in agricultural areas were the primary source of contamination.

5. Conclusions

Japan is currently experiencing rapid depopulation in most cities, and this demographic shift has caused both social and environmental impacts. Agricultural intensification is one of the main causes of elevated nitrate levels in the groundwater sources in Japan, often due to increased fertilizer application. However, other factors, such as land use and cultivation and farming methods, may also contribute to pollution. In comparison to Ocho, which had an average NO3−N concentration of 4.8 mg/L (37.5% of water samples > 5 mg/L), Kubi village had a higher average NO3−N concentration of 6.6 mg/L (66.7% of water samples > 5 mg/L). Analysis of the proportional contributions of each NO3−N source, calculated using SIAR (end-member method), showed that chemical fertilizers (60.1%) were the main contributing NO3−N source in Kubi, while in Ocho, organic fertilizers (50.7%) were the major contributors, with minor contributions observed from sewage. These source distributions were associated with social aspects, such as depopulation and population age, and to a greater extent, land use type. It is suspected that the considerable quantities of chemical fertilizers that are applied by the older farmers in Kubi lead to inefficient nitrogen uptake by the plants, resulting in leaching of N. Thus, the level of contamination of groundwater was higher in Kubi than in Ocho.
There were limited source end-members in this study, and they were not fully representative of both study sites; therefore, some sources may have been underestimated or overestimated. For the sustainable development of local agriculture, improved fertilizer utilization efficiency is crucial considering the characteristics of intensive cultivation areas. To overcome the limitations of this research, further research could include detailed demographic and time-series studies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/w14142277/s1, Figure S1: Outcrop geology of Osakishimojima Island; Table S1: Hydrochemical compositions of the groundwater and surface water sources in Kubi and Ocho village for the 2021 sampling campaign, Table S2: Correlation matrix of physiochemical data of groundwater samples collected from Kubi and Ocho village, Table S3: Isotopic data (‰) of sampled groundwater and surface water sources in Kubi and Ocho village.

Author Contributions

Conceptualization, S.-I.O.; formal analysis, S.B.K., M.T. and Y.T.; funding acquisition, S.-I.O.; resources, S.-I.O.; investigation, S.B.K., S.-I.O., M.S., T.I., M.T. and I.N.; methodology, S.-I.O. and T.I.; project administration, S.-I.O. and M.S.; supervision, S.-I.O., T.I. and M.S.; writing—original draft, S.B.K.; writing—review and editing, S.B.K., S.-I.O., T.I. and M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Asia-Pacific Network for Global Change Research (APN) under grant No. CRRP2019-09MY-Onodera (funder ID: https://doi.org/10.30852/p.4609 and Grant-in-Aid for Scientific Research (A) by the Japan Society for the Promotion of Science (Project No. 18H04151, PI: Shin-ichi Onodera) accessed on 19 July 2022.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study area are available on request from the corresponding author.

Acknowledgments

We express profound gratitude to the Mamena Association, Osakishimojima, for their continued support and assistance throughout this study. We thank Muhammad Shahinur Rahman and Nang Yu War from the Laboratory of Biogeochemistry for their valuable comments and suggestions concerning map creation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hayashi, K.; Shibata, H.; Oita, A.; Nishina, K.; Ito, A.; Katagiri, K.; Shindo, J.; Winiwarter, W. Nitrogen budgets in Japan from 2000 to 2015: Decreasing trend of nitrogen loss to the environment and the challenge to further reduce nitrogen waste. Environ. Pollut. 2021, 286, 117559. [Google Scholar] [CrossRef] [PubMed]
  2. Norio Tase, N. Groundwater contamination in Japan. Environ. Geol. Water Sci. 1992, 20, 15–20. [Google Scholar] [CrossRef]
  3. Onodera, S. Subsurface pollution in Asian megacities. In Groundwater and Subsurface Environments: Human Impacts in Asian Coastal Cities; Springer: Berlin/Heidelberg, Germany, 2011; pp. 159–184. [Google Scholar] [CrossRef]
  4. Haque, S.J.; Onodera, S.; Shimizu, Y. An overview of the effects of urbanization on the quantity and quality of groundwater in South Asian megacities. Limnology 2013, 14, 135–145. [Google Scholar] [CrossRef]
  5. Agrawal, G.D.; Lunkad, S.K.; Malkhed, T. Diffuse agricultural nitrate pollution of groundwaters in India. Water Sci. Technol. 1999, 39, 67–75. [Google Scholar] [CrossRef]
  6. Nolan, B.; Stoner, J. Nutrients in groundwaters of the conterminous United States, 1992–1995. Environ. Sci. Technol. 2000, 34, 1156–1165. [Google Scholar] [CrossRef] [Green Version]
  7. Sacchi, E.; Acutis, M.; Bartoli, M.; Brenna, S.; Delconte, C.A.; Laini, A.; Pennisi, M. Origin and fate of nitrates in groundwater from the central Po plain: Insights from isotopic investigations. Appl. Geochem. 2013, 34, 164–180. [Google Scholar] [CrossRef]
  8. Mcclain, M.; Richey, J.; Pimentel, T. Groundwater nitrogen dynamics at the terrestrial-lotic interface of a small catchment in the Central Amazon Basin. Biogeochemistry 1994, 27, 113–127. [Google Scholar] [CrossRef]
  9. Kumazawa, K. Nitrogen fertilization and nitrate pollution in groundwater in Japan: Present status and measures for sustainable agriculture. Nutr. Cycl. Agroecosystems 2002, 63, 129–137. [Google Scholar] [CrossRef]
  10. Saito, M.; Onodera, S.; Okubo, K.; Takagi, S.; Maruyama, Y.; Jin, G.; Shimizu, Y. Effects of physical and morphometric factors on nutrient removal properties in agricultural ponds. Water Sci. Technol. 2015, 72, 2187–2193. [Google Scholar] [CrossRef] [Green Version]
  11. Kimbi, S.B.; Onodera, S.; Nozaki, S.; Tomozawa, Y.; Wang, K.; Rusydi, A.F.; Saito, M. Impact of citrus agriculture on the quality of water resource in as small steep Island, Seto Inland Sea, Japan. GEOMATE J. 2021, 20, 109–114. [Google Scholar] [CrossRef]
  12. Zhang, Y.; Li, F.; Zhang, Q.; Li, J.; Liu, Q. Tracing nitrate pollution sources and transformation in surface- and ground-waters using environmental isotopes. Sci. Total Environ. 2014, 490, 213–222. [Google Scholar] [CrossRef] [PubMed]
  13. Kendall, C. Tracing nitrogen sources and cycling in catchments. In Isotope Tracers in Catchment Hydrology; Kendall, C., McDonnell, J.J., Eds.; Elsevier Science: Amsterdam, The Netherlands, 1998; pp. 519–576. [Google Scholar]
  14. Bedard-Haughn, A.K.; van Groenigen, J.-W.; van Kessel, C. Tracing 15 N through landscapes: Potential uses and precautions. J. Hydrol. 2003, 272, 175–190. [Google Scholar] [CrossRef]
  15. Knapp, A.N.; Sigman, D.M.; Lipschultz, F.N. Isotopic composition of dissolved organic nitrogen and nitrate at the Bermuda Atlantic time-series study site. Glob. Biogeochem. Cycles 2005, 19, GB1018. [Google Scholar] [CrossRef]
  16. Chen, J.; Taniguchi, M.; Liu, G.; Miyaoka, K.; Onodera, S.; Tokunaga, T.; Fukushima, Y. Nitrate pollution of groundwater in the Yellow River delta, China. Hydrogeol. J. 2007, 15, 1605–1614. [Google Scholar] [CrossRef]
  17. Umezawa, Y.; Hosono, T.; Onodera, S.; Siringan, F.; Buapeng, S.; Delinom, R.; Yoshimizu, C.; Tayasu, I.; Nagata, T.; Taniguchi, M. Sources of nitrate and ammonium contamination in groundwater under developing Asian megacities. Sci. Total Environ. 2008, 404, 361–376. [Google Scholar] [CrossRef]
  18. Peralta, E.M.; Batucan, L.S.; De Jesus, I.B.B.; Triño, E.M.C.; Uehara, Y.; Ishida, T.; Kobayashi, Y.; Ko, C.Y.; Iwata, T.; Borja, A.S.; et al. Nutrient loadings and deforestation decrease benthic macroinvertebrate diversity in an urbanised tropical stream system. Limnologica 2020, 80, 125744. [Google Scholar] [CrossRef]
  19. Rusydi, A.F.; Onodera, S.; Saito, M.; Hyodo, F.; Maeda, M.; Sugiantiuyo, K.; Wibawa, S. Potential sources of ammonium-nitrogen in the coastal groundwater determined from a combined analysis of nitrogen isotope, biological and geological parameters, and land use. Water 2021, 13, 25. [Google Scholar] [CrossRef]
  20. Wang, K.; Onodera, S.; Saito, M. Evaluation of nitrogen loading in the last 80 years in an urbanized Asian coastal catchment through the reconstruction of severe contamination period. Environ. Res. Lett. 2022, 17, 014010. [Google Scholar] [CrossRef]
  21. Xue, D.M.; Botte, J.; Baets, B.D.; Accoe, F.; Nestler, A.; Taylor, P.; Cleemput, O.V.; Berglund, M.; Boeckx, P. Present limitations and future prospects of stable isotope methods for nitrate source identification in surface- and groundwater. Water Res. 2009, 43, 1159–1170. [Google Scholar] [CrossRef]
  22. Kou, X.; Ding, J.; Li, Y.; Li, Q.; Mao, L.; Xu, C.; Zheng, Q.; Zhuang, S. Tracing nitrate sources in the groundwater of an intensive agricultural region. Agric. Water Manag. 2021, 250, 106826. [Google Scholar] [CrossRef]
  23. Kohl, D.H.; Shearer, G.B.; Commoner, B. Fertilizer nitrogen: Contribution to nitrate in surface water in a Corn Belt Watershed. Science 1971, 174, 1331–1334. [Google Scholar] [CrossRef] [PubMed]
  24. Amberger, A.; Schmidt, H.L. Natürliche Isotopengehalte von Nitrat als Indikatoren für dessen Herkunft. Geochim. Cosmochim. Acta 1987, 51, 2699–2705. (In German) [Google Scholar] [CrossRef]
  25. Kendall, C.; Elliott, E.M.; Wankel, S.D. Tracing anthropogenic inputs of nitrogen to ecosystems. In Stable Isotopes in Ecology and Environmental Science; Michener, R.H., Lajtha, K., Eds.; Blackwell: Oxford, UK, 2008; pp. 375–449. [Google Scholar]
  26. Aravena, R.; Robertson, W. Use of multiple isotope tracers to evaluate denitrification in groundwater: Case study of nitrate from a large-flux septic system plume. Ground Water 1998, 36, 975–982. [Google Scholar] [CrossRef]
  27. Wu, D.D.; Jiang, Y.H.; Jia, J.Y.; Cai, H.S. Analyses of pollution source of nitrogen in groundwater with δ15N-NO3 and δ18O-NO3 in Changzhou. Hydrogeol. Eng. Geol. 2006, 3, 11–15. [Google Scholar]
  28. Chen, F.J.; Jia, G.D.; Chen, J.Y. Nitrate sources and watershed denitrification inferred from nitrate dual isotopes in the Beijiang River, south China. Biogeochemistry 2009, 94, 163–174. [Google Scholar] [CrossRef]
  29. Liu, T.; Wang, F.; Michalski, G.; Xia, X.H.; Liu, S.D. Using 15N, 17O, and 18O to determine nitrate sources in the Yellow River, China. Environ. Sci. Technol. 2013, 47, 13412–13421. [Google Scholar] [CrossRef]
  30. Meghdadi, A.; Javar, N. Quantification of spatial and seasonal variations in the proportional contribution of nitrate sources using a multi-isotope approach and Bayesian isotope mixing model. Environ. Pollut. 2018, 235, 207–222. [Google Scholar] [CrossRef]
  31. Ji, X.L.; Xie, R.T.; Hao, Y.; Lu, J. Quantitative identification of nitrate pollution sources and uncertainty analysis based on dual isotope approach in an agricultural water- shed. Environ. Pollut. 2017, 229, 586–594. [Google Scholar] [CrossRef]
  32. Parnell, A.C.; Inger, R.; Bearhop, S.; Jackson, A.L. Source partitioning using stable isotopes: Coping with too much variation. PLoS ONE 2010, 5, e9672. [Google Scholar] [CrossRef]
  33. Fadhullah, W.; Syahirah Yaccob, N.; Syakir, M.I.; Akmal, M.S.; Yue, F.-J.; Li, S.-L. Nitrate sources and processes in the surface water of a tropical reservoir by stable isotopes and mixing model. Sci. Total Environ. 2020, 700, 134517. [Google Scholar] [CrossRef]
  34. Fu, X.M.; Sun, Y.Y.; Su, J.; Zheng, M.X.; Xi, B.D.; Qian, G.R. Source of nitrate in groundwater based on hydrochemical and dual stable isotopes. China Environ. Sci. 2019, 39, 3951–3958. [Google Scholar]
  35. Yu, L.; Zheng, T.; Zheng, X.; Hao, Y.; Yuan, R. Nitrate source apportionment in groundwater using Bayesian isotope mixing model based on nitrogen isotope fractionation. Sci. Total Environ. 2020, 718, 137242. [Google Scholar] [CrossRef]
  36. Torres-Martínez, J.A.; Mora, A.; Knappett, P.S.K.; Ornelas-Soto, N.; Mahlknecht, J. Tracking nitrate and sulfate sources in groundwater of an urbanized valley using a multi-tracer approach combined with a Bayesian isotope mixing model. Water Res. 2020, 182, 115962. [Google Scholar] [CrossRef]
  37. WHO. Guidelines for Drinking-Water Quality: First Addendum to the Fourth Edition; WHO: Geneva, Switzerland, 2017. [Google Scholar]
  38. Kure City. Population Statistical Data (Basic Resident Register). 2021. (In Japanese). Available online: https://www.city.kure.lg.jp/soshiki/36/people.html (accessed on 10 May 2022).
  39. Japan Japan Agricultural Cooperatives (JA), Yutaka, Hiroshima, 2022. Yutaka Town Population Statistics (2005–2021) and Yutaka Town Farmer Age and Population Statistics (1970–2005) Unpublished Data in Japanese. Available online: https://www.city.kure.lg.jp/ (accessed on 10 May 2022).
  40. Ministry of Land, Infrastructure, Transport, and Tourism of Japan. 2021. Available online: https://nlftp.mlit.go.jp/ (accessed on 8 April 2022).
  41. McIlvin, M.R.; Altabet, M.A. Chemical conversion of nitrate and nitrite to nitrous oxide for nitrogen and oxygen isotopic analysis in freshwater and seawater. Anal. Chem. 2005, 77, 5589–5595. [Google Scholar] [CrossRef]
  42. Tu, Y.; Fang, Y.T.; Liu, D.W.; Pan, Y.P. Modifications to the azide method for nitrate isotope analysis. Rapid Commun. Mass Spectrom. 2016, 30, 1213–1222. [Google Scholar] [CrossRef]
  43. Jackson, A.L.; Inger, R.; Bearhop, S.; Parnell, A. Erroneous behaviour of MixSIR, a recently published Bayesian isotope mixing model: A discussion of Moore & Semmens (2008). Ecol. Lett. 2009, 12, E1–E5. [Google Scholar]
  44. Huang, G.; Liu, C.; Sun, J.; Zhang, M.; Jing, J.; Li, L. A regional scale investigation on factors controlling the groundwater chemistry of various aquifers in a rapidly urbanized area: A case study of the Pearl River Delta. Sci. Total Environ. 2018, 625, 510–518. [Google Scholar] [CrossRef]
  45. Zhang, F.; Huang, G.; Hou, Q.; Liu, C.; Zhang, Y.; Zhang, Q. Groundwater quality in the Pearl River Delta after the rapid expansion of industrialization and urbanization: Distributions, main impact indicators, and driving forces. J. Hydrol. 2019, 577, 124004. [Google Scholar] [CrossRef]
  46. Evangelou, V.P. Environmental Soil and Water Chemistry; Wiley: New York, NY, USA, 1998. [Google Scholar]
  47. Earle, S. Physical Geology, 2nd ed.; BC Open Textbook Project; BCcampus: Victoria, BC, USA, 2019; Available online: https://books.google.co.jp/books?id=lSwwywEACAAJ (accessed on 3 May 2022).
  48. Keesari, T.; Ramakumar, K.L.; Chidambaram, S.; Pethperumal, S.; Thilagavathi, R. Understanding the hydrochemical behavior of groundwater and its suitability for drinking and agricultural purposes in Pondicherry area, South India—A step towards sustainable development. Groundw. Sustain. Dev. 2016, 2, 143–153. [Google Scholar] [CrossRef]
  49. Piper, A.M. A graphic procedure in the geochemical interpretation of water analyses. Eos Trans. Am. Geophys. Union 1944, 25, 914–923. [Google Scholar] [CrossRef]
  50. Babiker, I.S.; Mohamed, M.A.A.; Terao, H.; Kato, K.; Ohta, K. Assessment of groundwater contamination by nitrate leaching from intensive vegetable cultivation using geographical information system. Environ. Int. 2004, 29, 1009–1017. [Google Scholar] [CrossRef]
  51. Lasagna, M.; de Luca, D.A. Evaluation of sources and fate of nitrates in the western Po plain groundwater (Italy) using nitrogen and boron isotopes. Environ. Sci. Pollut. Res. 2019, 26, 2089–2104. [Google Scholar] [CrossRef]
  52. Rivett, M.O.; Buss, S.R.; Morgan, P.; Smith, J.W.N.; Bemment, C.D. Nitrate attenuation in groundwater: A review of biogeochemical controlling processes. Water Res. 2008, 42, 4215–4232. [Google Scholar] [CrossRef]
  53. Korom, S.F. Natural denitrification in the saturated zone: A review. Water Resour. Res. 1992, 28, 1657–1668. [Google Scholar] [CrossRef]
  54. Cuoco, E.; Darrah, T.H.; Buono, G.; Verrengia, G.; de Francesco, S.; Eymold, W.K.; Tedesco, D. Inorganic contaminants from diffuse pollution in shallow groundwater of the Campanian plain (southern Italy). Implications for geochemical survey. Environ. Monit. Assess. 2015, 187, 46. [Google Scholar] [CrossRef]
  55. Rufino, F.; Busico, G.; Cuoco, E.; Darrah, T.H.; Tedesco, D. Evaluating the suitability of urban groundwater resources for drinking water and irrigation purposes: An integrated approach in the Agro-Aversano area of Southern Italy. Environ. Monit. Assess. 2019, 191, 768. [Google Scholar] [CrossRef]
  56. Nakagawa, K.; Amano, H.; Asakura, H.; Berndtsson, R. Spatial trends of nitrate pollution and groundwater chemistry in Shimabara City, Nagasaki Prefecture, Japan. Environ. Earth Sci. 2016, 75, 1–17. [Google Scholar] [CrossRef] [Green Version]
  57. Egbi, C.D.; Anornu, G.; Appiah-Adjei, E.K.; Ganyaglo, S.Y.; Dampare, S.B. Evaluation of water quality using hydrochemistry, stable isotopes, and water quality indices in the Lower Volta River Basin of Ghana. Environ. Dev. Sustain. 2019, 21, 3033–3063. [Google Scholar] [CrossRef]
  58. Yadav, R.K.; Goyal, B.; Sharma, R.K.; Dubey, S.K.; Minhas, P.S. Post-irrigation impact of domestic sewage effluent on composition of soils, crops and ground water—A case study. Environ. Int. 2002, 28, 481–486. [Google Scholar] [CrossRef]
  59. Soil Health, 2013. Leaching Characteristics of Organic Fertilizers. LebanonTurf Cumberland ST. Lebanon. Available online: https://www.lebanonturf.com/education-center/soil-health (accessed on 9 June 2022).
  60. Kure City Water Works and Sewerage Bureau, 2021. Business Plan 2020. Pages: 110, Japanese. Available online: https://www.city.kure.lg.jp/site/jougesui/about-jigyougaiyou.html (accessed on 10 May 2022).
  61. Morgan, T.K.; Kadyamakeni, M.D.; Zekri, M.; Schumann, W.A.; Vashisth, T.; Obreza, A.T.; Ferrarezi, S.R. 2021–2022 Florida Citrus Production Guide: Nutrition Management for Citrus Trees; UF/IFAS: Polk, FL, USA, 2020. [Google Scholar]
  62. Moe, K.; Yamakawa, T.; Thu, T.T.P.; Kajihara, Y. Effects of pretransplant basal and split applications of nitrogen on the growth and yield of Manawthukha rice. Commun. Soil Sci. Plant Analysis 2014, 45, 2833–2851. [Google Scholar] [CrossRef]
  63. Zhou, Z.J.; Shen, Y.Z.; Du, C.W.; Zhou, J.M.; Qin, Y.S.; Wu, Y.J. Economic and soil environmental benefits of using controlled-release bulk blending urea in the North China plain. Land Degrad. Dev. 2017, 28, 2370–2379. [Google Scholar] [CrossRef]
Figure 1. (a); Location map of Osakishimojima Island in the Seto Inland Sea showing the sampling points in (b); Kubi village and (c); Ocho village.
Figure 1. (a); Location map of Osakishimojima Island in the Seto Inland Sea showing the sampling points in (b); Kubi village and (c); Ocho village.
Water 14 02277 g001
Figure 2. Land use of Osakishimojima, showing proportions of citrus orchards and residential areas in (a); Kubi and Ocho lowland areas, (b); Kubi sampling area, and (c); Ocho sampling area.
Figure 2. Land use of Osakishimojima, showing proportions of citrus orchards and residential areas in (a); Kubi and Ocho lowland areas, (b); Kubi sampling area, and (c); Ocho sampling area.
Water 14 02277 g002
Figure 3. Piper diagram of surface water and groundwater in (a) Kubi and (b) Ocho (adapted from Piper [49]).
Figure 3. Piper diagram of surface water and groundwater in (a) Kubi and (b) Ocho (adapted from Piper [49]).
Water 14 02277 g003
Figure 4. (a) Osakishimojima Island, Spatial variation of NO3−N and δ15N-NO3 concentrations in groundwater of (b) Kubi and (c) Ocho villages.
Figure 4. (a) Osakishimojima Island, Spatial variation of NO3−N and δ15N-NO3 concentrations in groundwater of (b) Kubi and (c) Ocho villages.
Water 14 02277 g004
Figure 5. NO3−N concentration vs. well depth in shallow wells in the study area.
Figure 5. NO3−N concentration vs. well depth in shallow wells in the study area.
Water 14 02277 g005
Figure 6. Relationship between NO3−N and (a), DOC, (b), Cl, (c), SO42−, (d), Na+, (e), Mg2+, and (f), Ca2+ in groundwater samples of Kubi and Ocho villages.
Figure 6. Relationship between NO3−N and (a), DOC, (b), Cl, (c), SO42−, (d), Na+, (e), Mg2+, and (f), Ca2+ in groundwater samples of Kubi and Ocho villages.
Water 14 02277 g006
Figure 7. Nitrate isotopic signatures for sampling points in Kubi and Ocho with typical nitrate endmembers (adapted from Kendall et al. [25]).
Figure 7. Nitrate isotopic signatures for sampling points in Kubi and Ocho with typical nitrate endmembers (adapted from Kendall et al. [25]).
Water 14 02277 g007
Figure 8. Spatial proportional contributions of NO3 sources estimated using SIAR in (a); Kubi and (b); Ocho village. Boxplots illustrate the 50th, 75th, and 95th percentiles from dark to light.
Figure 8. Spatial proportional contributions of NO3 sources estimated using SIAR in (a); Kubi and (b); Ocho village. Boxplots illustrate the 50th, 75th, and 95th percentiles from dark to light.
Water 14 02277 g008
Table 1. Statistical characteristics of the hydrochemical compositions of the groundwater and surface water sources in Kubi and Ocho village. For values of individual samples, see Table S1 of Supplementary Materials.
Table 1. Statistical characteristics of the hydrochemical compositions of the groundwater and surface water sources in Kubi and Ocho village. For values of individual samples, see Table S1 of Supplementary Materials.
Kubi VillageOcho Village
Spring (n = 1)River (n = 2)Wells
(n = 17)
Borehole (n = 1)Spring (n =2)Wells (n = 30)
ParametersMin/MaxMin.Max.Ave.Med.SDMin.Max.Ave.Med.SDMin/MaxMin.Max.Ave.Med.SDMin.Max.Ave.Med.SD
pH7.37.98.68.38.30.5NM7.46.86.70.47.57.57.57.57.506.28.56.86.70.5
Temp19.123.724.724.224.20.818.126.721.220.72.41917.623.620.620.64.319.228.322.322.31.7
EC185.720627424024048.1107465314.832194.247213321717517559.4165479308.5317.572.7
TDS118.9131.9175.4153.6153.630.868.5297.6201.5205.560.3302.185.2138.911211238.1105.6306.6105.6203.246.5
DO9.28.398.78.70.51.18.754.72.16.17.79.58.68.61.30.89.13.532.3
DOC3.63.85.34.64.61.12.513.35.14.13.14.1NM3.53.53.50NM73.33.21.4
Cl14.79.411.110.310.31.37.63411.99.96.213.44.610.97.87.84.54.137.511.911.26.4
SO42−22.930.550.440.540.514.122.175.344.438.816.680.74.522.913.713.713.1145.137.72728.2
NO37.112.434.223.323.315.514.974.541.535.519.370.32.713.78.28.27.83.566.421.719.713.4
NO3−N1.72.97.85.45.43.53.416.99.58.14.415.90.73.11.91.91.70.815.154.53.1
HCO37.250.661.255.955.97.5121.648.645.324.539.149.36.76.73.84.5150.264.269.745.4
Na+2116.318.317.317.31.59.432.515.713.7731.45.52.85.53.99.457.618.616.79.2
K+2.23.23.63.43.40.30.76.94.14.31.913.28.45.85.83.70.310.54.142.9
Mg2+9.88.210.79.59.51.8717.811.5113.217.81.425.813.613.617.34.922.2107.95
Ca2+24.622.33026.226.25.520.448.431.730.69.144.84.64.64.6050.828.82811.6
Note: n = number of samples; SD = standard deviation; − = below detection limit; NM = not measured; ionic concentrations, DOC, and DO = mg/L; EC = μS/cm; Temp = °/C.
Table 2. Principal component (PC) loadings of 11 chemical variables for groundwater samples in Kubi and Ocho villages.
Table 2. Principal component (PC) loadings of 11 chemical variables for groundwater samples in Kubi and Ocho villages.
PCs Kubi Village (21 Samples)PCs Ocho Village (32 Samples)
VariablesPC1PC2PC3VariablesPC1PC2PC3
pH−0.0980.3960.663pH−0.088−0.2420.768
TDS0.783−0.3790.309TDS0.5420.6860.141
DO−0.2090.7660.105DO0.035−0.7260.382
NO3−N0.903−0.05−0.119NO3−N0.8150.058−0.033
Cl0.4070.398−0.528Cl0.3760.7230.067
SO42−0.914−0.018−0.001SO42−0.9220.041−0.067
HCO3−0.117−0.6030.574HCO3−0.1310.798−0.002
Na+0.8230.3940.117Na+0.6830.517−0.047
K+−0.167−0.747−0.375K+−0.109−0.144−0.787
Mg2+0.917−0.099−0.13Mg2+0.7190.1240.055
Ca2+0.673−0.0510.29Ca2+0.754−0.0230.102
Eigen value4.52.21.5 4.321.3
EV40.919.513 38.21811.7
CV40.960.373.2 38.256.267.8
EV: explained variance (%); CV: cumulative% of variance; bold and italicized numbers: maximum absolute PC loading of one parameter.
Table 3. The δ15N and δ18O values of various nitrate sources.
Table 3. The δ15N and δ18O values of various nitrate sources.
Sourceδ15N (‰)δ18O (‰)
MeanSDMeanSD
b Soil nitrogen (SN)4.40.72.850.2
a Chemical fertilizer (CF)40.32.850.4
a,b Chemical fertilizer_Soil nitrogen (CF_SN)4.20.52.90.3
a Organic fertilizer (OF)100.32.850.4
b Manure and sewage (M&S)190.3−12.510.4
a The theoretical values analyzed and b data sources [25,33,34,35].
Table 4. A summary of social aspects and groundwater NO3−N contamination status in Kubi and Ocho villages.
Table 4. A summary of social aspects and groundwater NO3−N contamination status in Kubi and Ocho villages.
Social InformationUnitsKubi VillageOcho VillageKey References
Population 413830[38]
Average farmer ageyears7673[39]
Agricultural landkm20.61.6[40]
Sewage treatment coverage%700[60]
Summary of NO3−N concentration Figure 4
Average NO3−N concentrationmg/L6.64.8
<5%33.362.5
5–10%47.731.3
>10%196.2
Main NO3−N contribution ratio Figure 8
Chemical fertilizer%60.138.4
Organic fertilizer%30.150.7
Manure and sewage%9.810.9
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kimbi, S.B.; Onodera, S.-I.; Ishida, T.; Saito, M.; Tamura, M.; Tomozawa, Y.; Nagasaka, I. Nitrate Contamination in Groundwater: Evaluating the Effects of Demographic Aging and Depopulation in an Island with Intensive Citrus Cultivation. Water 2022, 14, 2277. https://doi.org/10.3390/w14142277

AMA Style

Kimbi SB, Onodera S-I, Ishida T, Saito M, Tamura M, Tomozawa Y, Nagasaka I. Nitrate Contamination in Groundwater: Evaluating the Effects of Demographic Aging and Depopulation in an Island with Intensive Citrus Cultivation. Water. 2022; 14(14):2277. https://doi.org/10.3390/w14142277

Chicago/Turabian Style

Kimbi, Sharon Bih, Shin-Ichi Onodera, Takuya Ishida, Mitsuyo Saito, Masayuki Tamura, Yusuke Tomozawa, and Itaru Nagasaka. 2022. "Nitrate Contamination in Groundwater: Evaluating the Effects of Demographic Aging and Depopulation in an Island with Intensive Citrus Cultivation" Water 14, no. 14: 2277. https://doi.org/10.3390/w14142277

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