Next Article in Journal
Does BMI Modify the Association between Vitamin D and Pulmonary Function in Children of the Mild Asthma Phenotype?
Next Article in Special Issue
Influence of Manure as a Complex Mixture on Soil Sorption of Pharmaceuticals—Studies with Selected Chemical Components of Manure
Previous Article in Journal
Benefits of Cycling Wheelchair Training for Elderly with Physical Disability: A Prospective Cohort Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Models for Estimating the Sorption of Sulfonamide and Tetracycline Antibiotics in Soils

1
Institute of Mountain Hazards and Environment, Chinese Academy of Sciences, Chengdu 610041, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
3
State Key Laboratory of Subtropical Silviculture, Zhejiang A&F University, Hangzhou 311300, China
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2022, 19(24), 16771; https://doi.org/10.3390/ijerph192416771
Submission received: 12 November 2022 / Revised: 9 December 2022 / Accepted: 11 December 2022 / Published: 14 December 2022

Abstract

:
Sulfonamides (SAs) and tetracyclines (TCs) are two classes of widely used antibiotics. There is a lack of easy models for estimating the parameters of antibiotic sorption in soils. In this work, a dataset of affinity coefficients (Kf and Kd) of seven SA/TC antibiotics (i.e., sulfachlorpyridazine, sulfamethazine, sulfadiazine, sulfamethoxazole, oxytetracycline, tetracycline, and chlortetracycline) and associated soil properties was generated. Correlation analysis of these data showed that the affinity coefficients of the SAs were predominantly affected by soil organic matter and cation exchange capacity, while those of the TCs were largely affected by soil organic matter and pH. Pedotransfer functions for estimating Kf and Kd were built by multiple linear regression analysis and were satisfactorily validated. Their performances would be better for soils having higher organic matter content and lower pH. These pedotransfer functions can be used to aid environmental risk assessment, prioritization of antibiotics and identification of vulnerable soils.

1. Introduction

In recent years, antibiotics have often been detected in soils at elevated levels as a result of their discharge after human and animal use [1,2]. Soil pollution with antibiotics has been recognized as a potential threat, potentially causing the development and spread of antibiotic resistance in soil microbes, impairment of soil ecosystem functions, contamination of agricultural products, and offsite pollution of receiving water bodies via hydrological processes [3,4].
Sorption regulates the distribution of antibiotics between aqueous and solid phase in soils, thus affecting their mobility, bioavailability, and fate [5,6,7]. For instance, sulfonamides (SAs) are the most mobile antibiotics in soils due to their low sorption coefficient, and their residues in soils range from ng kg−1 to μg kg−1 level [8]. Contrastingly, tetracyclines (TCs), as another class of antibiotics commonly used, are less mobile and show higher residues in soils ranging from 12 to 100 μg kg−1 [9]. Sorption behavior is dependent on antibiotic physico-chemical properties (e.g., octanol–water partitioning coefficient (Kow), acid dissociation constant (pKa), and molecular structure) and soil properties (e.g., organic matter content, pH, cation exchange capacity, and texture) [10,11]. Parameters of sorption isotherm models, such as affinity coefficient (Kf) and linearity coefficient (n) of the Freundlich model and affinity/distribution coefficient (Kd) of the linear model, are key components in environmental risk assessment of antibiotics [12]. In an area where a large number of antibiotics are in use, it may be infeasible to experimentally determinate sorption parameters of all antibiotics in all soils of concern, due to the budget and time limitations. There is clearly a need to develop models for estimating sorption parameters.
Antibiotics are mostly polar, ionizable compounds. In soil’s aqueous phase, ionizable antibiotics (e.g., sulfonamides, tetracyclines, and fluoroquinolones) can exist as cation, anionic, neutral, or zwitterion species, depending on their pKa and solution pH [12,13,14]. A number of mechanisms, such as hydrophobic interactions with soil organic matter (SOM), hydrogen and covalent bonding to SOM, exchange of cationic antibiotic species with cations on negatively charged surfaces of SOM and phyllosilicate clay minerals, surface complexation of anionic antibiotic species on surficial Fe/Mn oxides and clay mineral edge sites, cation bridging of anionic antibiotic species to negatively charged sites on clay minerals and organic matter, and electrostatic attraction of anionic antibiotic species with positively charged Fe/Al oxides, can be involved in the sorption of antibiotics [11,12,15,16,17,18,19,20,21,22]. For a given antibiotic, it is reasonable to consider parameters concerning antibiotic speciation, soil components, and environmental conditions as potential inputs of models for estimating sorption parameters. Differences in model formulation among antibiotics can reflect differences in main sorption mechanisms and their relative contributions.
Models for estimating sorption parameters can be developed using different approaches. Statistical regression analyses are traditional approaches to establishing linear or nonlinear quantitative models relating to the sorption parameters of antibiotics with soil properties [23,24]. Regression-based models using antibiotic physico-chemical properties alone as inputs were also developed, and the performance of such models can be improved by also including soil properties as inputs [25,26,27]. Moreover, satisfactory estimation of sorption parameters was obtained using machine learning approaches (e.g., artificial neural network, and random forest), which can involve many more inputs than regression-based approaches can do [28,29]. It should be noted that some soil properties (e.g., exchangeable K, Na, and Mg) used as inputs of machine learning models are not commonly reported in the literature, potentially limiting their broader applications. Overall, regression-based models are recognized as the most practical tools for estimating antibiotic sorption parameters as they are explicitly programmed and require only readily available inputs. Nevertheless, such models developed in each of most previous studies were based only on sorption data for soils of a single country or geological region [2,6,10,13,24,26,30,31,32,33], and their application in soils of other countries/regions may be problematic due to their inherent site-specific nature.
The aim of this study was therefore to develop new regression-based models for estimating sorption parameters of seven widely used SA/TC antibiotics, which were expected to be applicable in soils of different countries/regions. The specific objectives of this study were to (a) establish a dataset of sorption parameters for the target antibiotics in a wide range of soil properties based on data from the literature; (b) identify key factors affecting antibiotic sorption in soils and underlying mechanisms; and (c) develop and validate new models for estimating the sorption parameters of the target antibiotics which can be used in combination with spatial information on soil properties to evaluate the environmental risk of antibiotics on a global scale.

2. Materials and Methods

2.1. Physical and Chemical Properties of Antibiotics

Seven antibiotics, including four SAs (sulfachlorpyridazine (SCP), sulfamethazine (SMT), sulfadiazine (SDZ), and sulfamethoxazole (SMX)) and three TCs (oxytetracycline (OTC), tetracycline (TC), and chlortetracycline (CTC)), were selected for this study. Their physiochemical properties are provided in Table 1. The proportion of their species in soil water is dependent on their pKa and soil pH [34]. All the SAs are hydrophilic (logKow ≤ 0.89), and so are the TCs, except CTC which has the greatest hydrophobicity (logKow = 2.07).

2.2. Data Collection

A total of 104 publications, of which 10, 21, 20, 18, 38, 22, and 17 were related to SCP, SMT, SDZ, SMX, OTC, TC, and CTC, respectively, were reviewed to create a dataset consisting of sorption parameters of sulfonamide and tetracycline antibiotics and properties of tested soils. This dataset covers 5 continents (Asia, Europe, South America, North America, and Oceania), 14 countries, 3 temperature zones (southern temperate, northern temperate, and tropical), and 10 climate types. Only sorption studies in natural soils were included. Three sorption parameters (Kf and n of the Freundlich model, and Kd of the linear model) and six soil properties (pH, organic matter (OM) content, cation exchange capacity (CEC) and soil texture (sand, silt, and clay content)) were selected, and their mean values are shown Tables S2–S8. Soil OM content was converted to soil organic carbon (OC) content using the relationship %OM = 1.724 × (%OC) when needed [39]. A high diversity of texture and properties was represented by the soils in this dataset. Experimental parameters (i.e., initial antibiotic concentration in aqueous phase, solid/liquid ratio), which could affect antibiotic sorption for batch experiments [31,40], were also included. The initial antibiotic concentration and solid/liquid ratio ranged from 0.04 to 14,236.00 mg L−1 and from 1:1 to 1:625, respectively. This dataset was divided into four independent sub-datasets (“A” and “a” for Kf, or “B” and “b” for Kd). Sub-datasets “A” and “B” (number of Kf data: 68, 107, 53, 49, 104, 84, and 73 for SCP, SMT, SDZ, SMX, OTC, TC, and CTC, respectively; number of Kd data: 80, 114, 83, 57, 94, 67, and 72 for SCP, SMT, SDZ, SMX, OTC, TC, and CTC, respectively) was used to build pedotransfer functions for estimating the affinity coefficients of the seven target antibiotics, and sub-datasets “a” and “b” (number of Kf data: 15, 35, 24, 10, 29, 23, and 20 for SCP, SMT, SDZ, SMX, OTC, TC, and CTC, respectively; number of Kd data: 18, 39, 21, 15, 27, 23, and 16 for SCP, SMT, SDZ, SMX, OTC, TC, and CTC, respectively) were used to validate the established models.

2.3. Sorption Isotherms

The sorption of SAs and TCs on soil is usually described by the linear or Freundlich models, which can be written as Equations (1) and (2), respectively.
Q e = K d C e
Q e = K f C e 1 n
where Qe (mg kg−1) is the amount of antibiotic sorbed on the soil at equilibrium; Ce (mg L−1) is the equilibrium concentration of antibiotic in aqueous phase; Kd (L kg−1) is the linear affinity/distribution coefficient; Kf (mg1−1/n L1/n kg−1) is the Freundlich affinity coefficient; and n is the Freundlich linearity index. When the value of n is close to 1, Freundlich models are approximately equal to linear models. n > 1 indicates the saturation of sorption sites at high concentrations, which hinders the sorption process. n < 1 indicates that the previously sorbed antibiotic increases the sorption power of the soil [41].
The Qe can be calculated as follows [6]:
Q e = ( C i C e ) V w m s
where Ci is the initial aqueous antibiotic concentration (mg L−1), Vw is the aqueous volume (mL), and ms is the soil mass (g).
For studies in which only the Freundlich model was used, Equations (1)–(3) were used to estimate Kd by re-fitting to the data of initial concentrations and equilibrium concentrations estimated from the reported Kf and n of the Freundlich model. Only the Kd values estimated with good fittings (p < 0.05) were included in the dataset.

2.4. Statistics and Modeling

Using the data shown in Tables S2–S8, Pearson correlations of sorption parameters with soil properties were analyzed to reveal the governing factors and mechanisms of antibiotic sorption in soils. Subsequently, multiple linear regressions were performed using SPSS 22.0 software (IBM Corp., Armonk, NY, USA) to develop pedotransfer functions for estimating the parameters of the linear and Freundlich models. Regression-based modeling using both edaphic and non-edaphic variable(s) as inputs was also conducted.

2.5. Model Evaluation

The applicability and accuracy of the pedotransfer functions were assessed using the adjusted determination coefficient (r2), Nash-Sutcliffe efficiency (NSE), root-mean-square error (RMSE), and absolute error (AE). NSE was calculated using Equation (4) to assess the model predictive capability [23]. RMSE and AE were obtained to measure the average magnitude of error in estimation using Equations (5) and (6), respectively. In addition, the percentage of RMSE over the standard deviation (SD) of the reported/re-fitted model parameters was also calculated.
NSE = 1 [ i = 1 N ( M i E i ) 2 i = 1 N ( M i M mean ) 2 ]
RMSE = i = 1 N ( M i E i ) 2 N
AE = | M i E i M i |
where Mi and Ei are the ith measured and estimated values, respectively. Mmean is the average of measured data, and N is the number of measurements. NSE, which can range from −∞ to 1, was used to evaluate how well the estimation was. The closer NSE is to 1, the better the model can perform. An RMSE value of 0 indicates a perfect fit.

3. Results

3.1. Distribution of Soil Properties

With respect to individual antibiotics, basic descriptive statistics of the physicochemical properties of the selected soils in the dataset generated in our study are presented in Table S1. For a specific target antibiotic, the number of soils used was in the range of 79–159. The dataset covers a broad range of soil characteristics, reflecting a large variability in source and nature of soils. Soil pH varied from 2.75 to 9.40, with 75.0% of the soils being in the acidic range. pH distribution of the soils used for each target antibiotic and main antibiotic species in soil water are shown in Table S9. In the soils included in the dataset, both the SAs and the TCs were mainly present in neutral/zwitterionic and/or anionic species, with SMX showing the highest dominance of anion forms. Soil OC content ranged from 0.1% to 21.3%, with the median being below 3.0%. Soil CEC was highly variable between 3.40 and 740.00 mmol kg−1, with the median being in the range of 61.00–155.00 mmol kg−1. Regarding soil texture, more than 50.0% of the total number of soils belong to the clay loam group according to the international soil classification system (Tables S2–S8).

3.2. Distribution of Antibiotic Sorption

Sorption parameters for each antibiotic varied greatly among the soils, especially those for the TCs that exhibited high sorption in the soils (Tables S2–S8). Table 2 shows basic descriptive statistics of the reported sorption parameters. Across the SAs, Kf and Kd varied from 0.13 to 16.00 mg1−1/n L1/n kg−1 and from 0.02 to 28.50 L kg−1, respectively, and were mainly in the low value ranges (1.65–5.25 mg1−1/n L1/n kg−1 and 2.05–4.42 L kg−1, respectively). For the SA antibiotics, the median of Kf and Kd followed the order of SCP > SMT > SMX > SDZ and SCP > SMT > SMX ≈ SDZ, respectively. Both mean and median of n for the SAs were close to 1 (i.e., linear isotherms), indicating that the distribution between the aqueous and solid phase was independent of the amount of antibiotic addition [30]. Both Kf and Kd of the TCs (0.28–8176.99 mg1−1/n L1/n kg−1 and 10.06–4473.20 L kg−1, respectively) were, at a maximum, three orders of magnitude greater than those of the SAs, which can be mainly attributed to the greater aromaticity of the TCs. It has been known that the antibiotics of greater aromaticity can be more strongly sorbed by soil organic matter, which is known to be highly aromatic [42,43]. For the TC antibiotics, the median of both Kf and Kd followed the same order of CTC > OTC > TC. Both mean and median of n for the TCs were about 2 (i.e., nonlinear isotherms), indicating that a decreasing tendency for sorption on heterogeneous soil surfaces with increasing initial TC concentration [30]. Moreover, the Freundlich model would provide a better fit for the sorption isotherms of the TCs than the linear model, which was also found by previous studies [30,44].

3.3. Correlations between Antibiotic Sorption and Soil Properties

The large variability of sorption parameters across the very dissimilar soils allows the analysis of their correlations with soil properties. Results of the Pearson correlation analysis are presented in Table 3.
Kf and Kd of the SAs were positively correlated with OC and CEC (p < 0.05), with an exception of the Kf of SDZ showing no correlation with CEC, indicating hydrophobic interactions and cation exchange were two main sorption mechanisms. The markedly higher correlation coefficient (r) values of OC with Kf/Kd, compared with any other soil properties, imply the predominant role of hydrophobic interactions with soil organic matter in the sorption of the SAs. Both Kf and Kd of SMX were negatively correlated with soil pH (p < 0.01), indicating hydrogen bonding might play a more important role in its sorption to the soils (particularly in the acidic soils) compared with the other three SAs. A previous study with 13 soils with pH ranging from 5.3 to 8.7 also reported a negative correlation of Kf with soil pH for SMX [10]. Antibiotic sorption may also be affected by soil texture [45]. The most significant negative correlation of sand content with Kf and Kd was observed for SDZ. Positive correlations of clay content with Kf and Kd were found for both the most strongly sorbing SCP and the most weakly sorbing SDZ (p < 0.01). Similarly, a number of previous studies in acidic soils have reported positive correlations of Kd and/or Kf for SAs with OC, CEC, and clay content but negative correlations with sand content [14,31,32,46].
Kf and/or Kd of the TCs showed a positive correlation with OC (p < 0.01), indicating that their interactions with SOM through hydrophobic interactions (e.g., π-π electron donor–acceptor interaction and van der Waals attractions) were important sorption mechanisms, as also reported previously [47,48,49]. Notably, given the similar correlation coefficients of Kf or Kd with OC for the three TCs, CTC’s much higher Kf and Kd than those of OTC and TC (Table 2) can be attributed to the highest logKow (a key parameter of a hydrophobic antibiotic) of CTC (Table 1). On the other hand, Kf of all the three TCs were negatively correlated with soil pH, implying that cation exchange also played a key role in the sorption of the TCs (p < 0.01). Despite the inconsistent relationships between Kf/Kd and CEC observed among different TCs (Table 3), it can be inferred that cation exchange between soil surfaces and the protonated amine groups of TCs was the main sorption mechanism at pH lower than their pKa1 [50,51]. Moreover, Kd of the least strongly sorbing TC showed a strong negative correlation with sand content, while showing positive correlations with clay and silt content as well as CEC (p < 0.01). The observed texture effect agrees with a previous finding that both Kf and Kd of OTC in a clay loam soil were higher than in a loamy sand soil [24].
Overall, the main soil properties influencing the sorption of the SAs were OC and CEC, while key influential soil properties for TC sorption were OC and pH. Apparently, the effect of soil texture on antibiotic sorption was inconsistent and antibiotic specific.

3.4. Model Development and Validation

Pedotransfer functions developed from sub-datasets “A” and “B” are presented in Table 4, and the results of model validation with sub-datasets “a” and “b” are shown in Figure 1. The models for all target antibiotics yielded good estimations of both Kf and Kd. For sub-datasets “A” and “B”, RMSE of the pedotransfer functions for Kf ranged from 0.39 to 2.01 and from 612.94 to 1340.46 for the SAs and the TCs, respectively; whereas, their RMSE/SD ratios for all target antibiotics fell within a narrow range (56.2–77.2%). Irrespective of antibiotic type, NSE of the pedotransfer functions for Kf ranged from 0.40 to 0.69 for sub-dataset “A”, reflecting good model performances. The values of soil properties in sub-datasets “a” and “b” were mostly within the ranges of sub-datasets “A” and “B”. For sub-dataset “a”, RMSE of the pedotransfer functions for Kf ranged from 1.08 to 4.29 and from 1210.67 to 1281.56 for the SAs and the TCs, respectively; their RMSE/SD ratios ranged from 60.4% to 99.4% and from 74.0% to 138.1% for the SAs and the TCs, respectively. Regardless of antibiotic type, NSE of the pedotransfer functions for Kf ranged from –0.91 to 0.64 for sub-dataset “a”, which indicated good estimations and were only slightly lower than those for sub-dataset “A”. The pedotransfer functions for Kf were thus validated by their satisfactory performances observed for sub-dataset “a”. Similarly, the pedotransfer functions developed for Kd with sub-dataset “B” were also satisfactorily validated with sub-dataset “b”.
In addition to basic soil properties, differences in experimental methods for soil characterization (especially for soil texture and CEC), maximal initial concentration (Cimax), and solid/liquid ratio (SLR) for batch sorption test, may be partly responsible for the variations in measured affinity coefficients (Kf, Kd), but these experimental parameters have not been used as input variable(s) for multiple linear regression analysis in previous studies [52]. Moreover, the inclusion of parameters of antibiotic species in multiple linear regression analysis may also help develop better models for estimating affinity coefficients [31]. In this study, some of these parameters, including Cimax, SLR, and percentage of antibiotic form(s) at a given pH (α+, α0, or α, representing cationic, neutral/zwitterion, or anionic species, respectively), were considered as additional independent variables, and eight models with better performances in estimating Kf and/or Kd were thus developed for all the target antibiotics except SMT (Table 5). Performance of the improved models for SCP was slightly better (as indicated by a 0.3% and 0.6% increase in r2 for Kf and Kd, respectively), and moderately better model performances (as indicated by 2.5–7.2% increase in r2) were achieved for SDZ, SMX, and the TCs. This improvement in model performance for the three SAs can be explained by the dependence of relative importance of various sorption mechanisms on antibiotic species distribution at a given pH. As for the TCs, these improvements can be attributed to the inclusion of not only species distribution parameters, but also SLR and/or Cimax, which can reflect the non-linear sorption behavior of TCs. This is in line with the finding that the smaller the SLR, the fewer the sorption sites, as a result of more rapid saturation of the sorption sites with increasing pesticide concentration [52]. Similarly, the lowering of SLR from 1:10 to 1:50 was found to cause decreases in sorption of TCs by 75% and 43% for alfisol and ultisol, respectively [40]. It should be noted that this study was not able to obtain improved models for SMT, and improved models were successfully built for estimating either Kf or Kd (not both) for four antibiotics (SDZ, SMX, OTC, and CTC) and both Kf and Kd for the other two antibiotics (SCP and TC). Apparently, the improvements in model performance achieved by incorporating additional non-edaphic variable(s) were limited, and therefore the validation of these models was not conducted further.
Soil OC and pH are the two most useful edaphic variables that can be used to estimate Kf and Kd of a specific SA or TC antibiotic in dissimilar soils. The effects of OC and pH variation on the performance of the pedotransfer functions (Table 4) were evaluated using sub-datasets “A” and “B” in terms of AE, and the AE values obtained for different OC or pH ranges are shown in Figure 2. It was found that the soils with higher OC content showed lower AE in estimating Kf and Kd, with the lowest AE (19.8% and 22.0% for Kf and Kd, respectively) being observed in the soils with OC content greater than 5%. Since SOM was the most influential soil property for the sorption of SAs and TCs (Table 3), a higher OC content can lead to a greater ability of SOM to estimate affinity coefficients. Contrastingly, a higher soil pH was associated with a poorer model performance (i.e., a higher AE), which can be attributed to a decreased importance of strong sorption mechanisms (e.g., cation exchange and electrostatic attraction) but an increased importance of weak sorption mechanisms (e.g., π-π interaction, and van der Waals forces) at an increased pH. For sub-datasets “a” and “b”, similar AE distributions across different OC or pH ranges were found for SAs, TC, and CTC (Figure S1). Nevertheless, AE of OTC estimation using sub-datasets “a” and “b” showed opposite trends with increasing OC and pH, which might be caused by the limitation of the small observation number in sub-datasets “a” and “b” and the generally lower NSE for OTC than the other antibiotics. Apparently, the pedotransfer functions of OTC did not perform as well as those for the other antibiotics (Figure 1). It is expected that the pedotransfer functions developed in this study can give better estimation of Kf and Kd for soils with higher OC and lower pH.

4. Discussion

4.1. Governing Factors and Mechanisms of Antibiotic Sorption

Soil pH, OC, CEC, and texture are readily available parameters that may be correlated with sorption capacity (i.e., Kf and/or Kd) of antibiotics [10,11,12,31,32,41,53,54,55]. Among these edaphic parameters, OC showed the highest positive correlation with either Kf or Kd for most antibiotics investigated in this study, implying the dominant role of hydrophobic interactions (e.g., π-π electron donor–acceptor interaction, and van der Waals interactions) in sorption to soil organic matter [42]. In addition, other sorption mechanisms include hydrogen bonding of antibiotics with hydroxyl groups on soil organic matter, particularly in acidic soils [17,22], and electrostatic interactions (e.g., cation exchange, surface complexation with potential contribution of cation bridging, charge transfer, and ligand exchange) of antibiotics with negatively charged surfaces [10,23]. It should be noted that electrostatic forces are stronger than hydrophobic interactions [37,56].
The SAs are acidic and largely uncharged, or negatively charged at natural soil pH (e.g., ≥3.7 in this study), as indicated by pKa1 ≤ 2.1 (Table 1). It has been well recognized that hydrophobic partitioning of soil organic matter is a main mechanism for the sorption of SAs [32,57,58,59]. Hydrophobic interactions were more important sorption mechanisms for the SAs than for the TCs, as reflected by the observed higher r values between Kf or Kd and OC for the SAs (Table 3). In addition to OC, CEC was another predominant edaphic parameter affecting the sorption of the SAs, probably as a result of the exchange of cationic SA species on negatively charged sites of clay mineral surfaces or organic matter, and surface complexation of anionic SA species on the edges of layered clay minerals and radical fragments of humus via cation bridges [16]. The important role of hydrogen bonding in sorption was demonstrated by FTIR spectrum analysis for SDZ and SMX [60].
The TCs, which are basic (represented by high pKa3) and thus have higher contents of cation form than the SAs at natural soil pH, could show more significant non-hydrophobic interactions (e.g., cation exchange, ligand exchange, surface complexation, and H bonding) in soils [61]. The negative correlations of soil pH with Kf for the TCs observed in this study agree with the results of many previous studies [13,17]. For instance, in a study with 63 soils, a negative correlation of Kf with soil pH was observed for OTC (p < 0.01) but not for CTC [13]. Similarly, negative correlations with KTe of the Temkin model with soil pH were reported for TC [17]. It should be noted that the predominant sorption mechanism of TCs may vary with soil pH. For instance, cation exchange might be the predominant sorption mechanism of TC in alkaline soils while hydrophobic interactions might be its primary sorption mechanism in acidic soils [11,17].
Overall, SOM dominates the sorption of SAs and TCs through hydrophobic interactions with neutral/zwitterion species of the antibiotics [17,31,32,58,62], hydrogen bonding of its protonated sites with polar groups of the antibiotics [19], and forming complexes with the antibiotics [7]. Moreover, both SA and TC antibiotics can covalently attach SOM [15,17,63]. Clay minerals may have a positive effect on antibiotic sorption due to their greater amounts of negative charge, larger surface area and higher CEC, as well as more preferential association with organic matter, while sand often shows a negative effect [13,14,17]. In the low hydrophobicity range, clay minerals in soils may play a significant role in sorption through cation exchange and cation bridging [64,65,66]. Notably, cation bridges, which can form on the surfaces of both soil minerals (aluminosilicates, metal oxides) and organic matter [67,68], may play a more important role in complexation with anionic species of SAs, compared with TCs which are less negatively charged, particularly in acidic soils. The relative importance of different mechanisms is dependent on the physicochemical properties of soils and antibiotics, and environmental factors (e.g., pH, ionic strength, organic matter, and temperature) [69]. A decrease in soil solution pH can lead to an increased proportion of the cationic species of SAs and TCs and thus an enhanced sorption of antibiotics via cation exchange [66,70,71,72].

4.2. Model Performance

The pedotransfer functions established in this study are very useful, as most of them (11 of 14) could explain more than 50% of the variance of Kf or Kd (Table 4). The multiple linear regression analysis indicated that OC was the only edaphic variable commonly included in all the established functions except that for Kd of TC, with explanation of the variance being 58.9% (Kf) and 67.3% (Kd) for SCP, 56.5% (Kf) and 51.0% (Kd) for SMT, 20.4% (Kf) and 46.4% (Kd) for SDZ, 53.5% (Kf) and 38.1% (Kd) for SMX, 4.9% (Kf) and 44.4% (Kd) for OTC, 33.5% (Kf) for TC, and 49.8% (Kf) and 32.9% (Kd) for CTC. pH was the secondary edaphic variable, which was included in 8 of the 14 models and could achieve a maximum explanation (55.7%) of the variance of Kf for OTC. CEC could explain only the variance of Kf for SMX and the variance of Kf and Kd for TC by smaller percentages, reflecting the lesser ability of CEC to estimate antibiotic sorption than OC and pH. The variance of Kf for SDZ and SMX and the variance of Kd for TC and CTC could be partly explained by selected soil textural parameters (e.g., sand, silt, and clay content).
In most previous studies, linear models were established based on batch sorption experiments in a suit of different soils for a single or a few antibiotics, such as sulfachloropyridazine [24,30,32], sulfadiazine [31], sulfamethazine [30,32], oxytetracycline [2,13,24], chlortetracycline [13], and tylosin [24]. The performance of selected published models was evaluated using sub-datasets “a” and “b” and the results are shown in Table 6. For SCP and SMT, RMSE, RMES/SD and NSE of published models were very close to those of our models (Figure 1); however, for SDZ and OTC, the performances of published models were poorer than those of our models (Figure 1). This result could be explained by the following: on the one hand, SOM played a more dominant role in the sorption of SCP and SMT compared with SDZ and OTC; on the other hand, in addition to SOM, the sorption of SDZ and OTC was also affected by soil pH and texture, which were more effectively represented in our models. It should be noted that the data used for the development of these published models were limited and regionally constrained. For instance, the models for SCP and SDZ were established with soils showing OC and pH in the range of 1.1–10.9 and 3.7–6.2, respectively [31,32]; the model for SMT was established with soils showing OC in the range of 0.1–3.8 [56]; and the model for OTC was established with soils showing OC and CEC in the range of 1.1–10.9 and 3.8–30.31, respectively [13]. Apparently, these published models were applicable only to a narrower range of soils compared with our models (Table S1).
The pedotransfer functions developed for two major groups of antibiotics in this study are simple and can be applied in environmental risk assessment of antibiotics in soils. The different effects of inclusion of non-edaphic variable(s) on model performance among antibiotics and affinity coefficients indicated that more than one sorption mechanism might dominate and the relative importance of one mechanism over another depended on, in addition to soil properties, antibiotic species and environmental conditions (e.g., pollution level and soil to water ratio). Given the complex relationships of affinity coefficients with varying properties/parameters, some previous studies employed machine learning approaches (artificial neural network, random forest, and support vector machine) to develop nonlinear models for antibiotics (together with non-antibiotic pharmaceuticals), and the best performance was achieved by a random forest-based model using antibiotic and soil properties as the independent variable(s) [25,73,74]. Notably, the random forest model can be utilized to reveal the relative importance order of variables and thus may help select the top contributing variables for the development of new models [74]. Compared with traditional regression models, machine learning models are more complex in nature and less transparent for users. From a regulatory perspective, simple and transparent methods would be preferred.
In addition, the applicability of our model for SMT was tested for sulfadimethoxine (SDM), which is an SA antibiotic that has similar physico-chemical properties to SMT [75]. Results showed that it performed well in predicting Kf of SDM, with RMSE, RMSE/SD, and NSE being 2.23, 34.1%, and 0.88, respectively (Table S10). Nevertheless, this potential for a wider application needs to be verified with more data in future studies.

4.3. Future Perspectives

It would be costly and time consuming to experimentally measure the sorption parameters of all antibiotics in all soils in an area of interest or globally. Given the complexities of antibiotic–soil interactions, future efforts to improve the predictive performance of new models should be directed to the following: generation of a bigger high-quality dataset of antibiotic sorption and associated soil properties with standard experimental protocols and development of conversion methods for results obtained under varying non-standard experimental conditions (e.g., initial aqueous antibiotic concentrations, soil to solution ratio, and solution electrolyte composition, temperature, and pH) and analytical methods; comparison of traditional regression models and machine learning models with independent datasets to identify their suitability for different antibiotics and soils; and the building of different model options that can meet varying requirements of accuracy.

5. Conclusions

A dataset of sorption parameters for 4 SAs and 3 TCs in soils collected from the literature was built, and key soil factors (OC, pH and CEC) affecting antibiotic sorption were identified using correlation analysis. Linear pedotransfer functions for estimating Kf and Kd were successfully established by multiple linear regression analysis and were satisfactorily validated. The new pedotransfer functions developed in this study can be used as an easy tool for environmental risk assessment, prioritization of antibiotics and identification of vulnerable soils in an area of interest, which could help develop mitigation measures to minimize the adverse impacts of antibiotic pollutants on human and environmental health.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijerph192416771/s1, Figure S1: Absolute errors (AE) distributions of Kf and Kd estimation across different OC and pH ranges in sub-datasets “a” and “b” using the pedotransfer equations given in Table 4; Table S1: Statistical characteristics of basic properties of the soils used for individual target antibiotics; Table S2: SCP sorption parameters and associated soil properties; Table S3: SMT sorption parameters and associated soil properties; Table S4: SDZ sorption parameters and associated soil properties; Table S5: SMX sorption parameters and associated soil properties; Table S6: OTC sorption parameters and associated soil properties; Table S7: TC sorption parameters and associated soil properties; Table S8: CTC sorption parameters and associated soil properties; Table S9: pH distribution of the soils in sub-datasets “A” (for Kf) and “B” (for Kd) and main antibiotic species in soil water; Table S10: SDM sorption parameters, associated soil properties, and the performance of the model established for SMT in this study. References [76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146] are cited in the supplementary materials.

Author Contributions

Conceptualization, J.H. and X.T.; methodology, J.H. and X.T.; software, J.H.; validation, J.H. and X.T.; formal analysis, J.H. and M.Q.; investigation, J.H. and M.Q.; resources, X.T. and J.C.; data curation, J.H. and M.Q.; writing—original draft preparation, J.H.; writing—review and editing, J.H. and X.T.; visualization, J.H.; supervision, X.T.; project administration, X.T. and J.C.; funding acquisition, X.T. and J.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Zhejiang Provincial Natural Science Foundation of China (Grant No. LD21D010001), the National Natural Science Foundation of China (Grant Nos. 42177379 and 42007361), and the ZAFU Scientific Research Development Foundation (Grant Nos. 2020FR040 and 2021LFR045).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Xu, X.Y.; Ma, W.J.; An, B.Y.; Zhou, K.X.; Mi, K.; Huo, M.X.; Liu, H.Y.; Wang, H.Y.; Liu, Z.L.; Cheng, G.Y.; et al. Adsorption/desorption and degradation of doxycycline in three agricultural soils. Ecotox. Environ. Saf. 2021, 224, 112675. [Google Scholar] [CrossRef] [PubMed]
  2. Jones, A.D.; Bruland, G.L.; Agrawal, S.G.; Vasudevan, D. Factors influencing the sorption of oxytetracycline to soils. Environ. Toxicol. Chem. 2010, 24, 761–770. [Google Scholar] [CrossRef] [PubMed]
  3. Bartikova, H.; Podlipna, R.; Skalova, L. Veterinary drugs in the environment and their toxicity to plants. Chemosphere 2016, 144, 2290–2301. [Google Scholar] [CrossRef] [PubMed]
  4. Luo, Y.; Xu, L.; Rysz, M.; Wang, Y.; Zhang, H.; Alvarez, P.J. Occurrence and transport of tetracycline, sulfonamide, quinolone, and macrolide antibiotics in the Haihe River Basin, China. Environ. Sci. Technol. 2011, 45, 1827–1833. [Google Scholar] [CrossRef] [PubMed]
  5. Conde-Cid, M.; Ferreira-Coelho, G.; Fernández-Calviño, D.; Núñez-Delgado, A.; Fernández-Sanjurjo, M.J.; Arias-Estévez, M.; Álvarez-Rodríguez, E. Single and simultaneous adsorption of three sulfonamides in agricultural soils: Effects of pH and organic matter content. Sci. Total Environ. 2020, 744, 140872. [Google Scholar] [CrossRef]
  6. Teixidó, M.; Granados, M.; Prat, M.D.; Beltran, J.L. Sorption of tetracyclines onto natural soils: Data analysis and prediction. Environ. Sci. Pollut. Res. 2012, 19, 3087–3095. [Google Scholar] [CrossRef]
  7. Thiele-Bruhn, S.; Aust, M.O. Effects of pig slurry on the sorption of sulfonamide antibiotics in soil. Arch. Environ. Contam. Toxicol. 2004, 47, 31–39. [Google Scholar] [CrossRef]
  8. Tang, W.; Jing, F.Q.; Laurent, Z.; Liu, Y.Y.; Chen, J.W. High-temperature and freeze-thaw aged biochar impacts on sulfonamide sorption and mobility in soil. Chemosphere 2021, 276, 130106. [Google Scholar] [CrossRef]
  9. Hamscher, G.; Pawelzick, H.T.; Hoper, H.; Nau, H. Different behavior of tetracyclines and sulfonamides in sandy soils after repeated fertilization with liquid manure. Environ. Toxicol. Chem. 2005, 24, 861–868. [Google Scholar] [CrossRef]
  10. Kodešová, R.; Grabic, R.; Kočárek, M.; Klement, A.; Golovko, O.; Fér, M.; Jakšík, O. Pharmaceuticals’ sorptions relative to properties of thirteen different soils. Sci. Total Environ. 2015, 511, 435–443. [Google Scholar] [CrossRef]
  11. Sassman, S.; Lee, L. Sorption of three tetracyclines by several soils: Assessing the role of pH and cation exchange. Environ. Sci. Technol. 2005, 39, 7452–7459. [Google Scholar] [CrossRef] [PubMed]
  12. Srinivasan, P.; Sarmah, A.K.; Manley-Harris, M. Sorption of selected veterinary antibiotics onto dairy farming soils of contrasting nature. Sci. Total Environ. 2014, 472, 695–703. [Google Scholar] [CrossRef] [PubMed]
  13. Conde-Cid, M.; Fernández-Calviño, D.; Núñez-Delgado, A.; Fernández-Sanjurjo, M.J.; Arias-Estévez, M.; Álvarez-Rodríguez, E. Estimation of adsorption/desorption Freundlich’s affinity coefficients for oxytetracycline and chlortetracycline from soil properties: Experimental data and pedotransfer functions. Ecotox. Environ. Saf. 2020, 196, 110584. [Google Scholar] [CrossRef]
  14. Leal, R.M.P.; Alleoni, L.R.F.; Tornisielo, V.L.; Regitano, J.B. Sorption of fluoroquinolones and sulfonamides in 13 Brazilian soils. Chemosphere 2013, 92, 979–985. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Franklin, A.M.; Williams, C.; Andrews, D.M.; Watson, J.E. Sorption and desorption behavior of four antibiotics at concentrations simulating wastewater reuse in agricultural and forested soils. Chemosphere 2022, 289, 133038. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, H.; Gao, B.; Li, H.; Ma, L.Q. Effects of pH and ionic strength on sulfamethoxazole and ciprofloxacin transport in saturated porous media. J. Contam. Hydrol. 2011, 126, 29–36. [Google Scholar] [CrossRef]
  17. Chen, Y.X.; Hu, C.Y.; Deng, D.H.; Li, Y.G.; Luo, L. Factors affecting sorption behaviors of tetracycline to soils: Importance of soil organic carbon, pH and Cd contamination. Ecotox. Environ. Saf. 2020, 197, 110572. [Google Scholar] [CrossRef]
  18. Álvarez-Esmorís, C.; Rodríguez-López, L.; Núñez-Delgado, A.; Álvarez-Rodríguez, E.; Fernández-Calviño, D.; Arias-Estévez, M. Influence of pH on the adsorption-desorption of doxycycline, enrofloxacin, and sulfamethoxypyridazine in soils with variable surface charge. Environ. Res. 2022, 214, 114071. [Google Scholar] [CrossRef]
  19. Pils, J.R.; Laird, D.A. Sorption of tetracycline and chlortetracycline on K– and Ca–saturated soil clays, humic substances, and clay-humic complexes. Environ. Sci. Technol. 2007, 41, 1928–1933. [Google Scholar] [CrossRef]
  20. Park, J.Y.; Huwe, B. Effect of pH and soil structure on transport of sulfonamide antibiotics in agricultural soils. Environ. Pollut. 2016, 213, 561–570. [Google Scholar] [CrossRef]
  21. Vasudevan, D.; Bruland, G.L.; Torrance, B.S.; Upchurch, V.G.; MacKay, A.A. pH-dependent ciprofloxacin sorption to soils: Interaction mechanisms and soil factors influencing sorption. Geoderma 2009, 151, 68–76. [Google Scholar] [CrossRef]
  22. Zhao, Z.; Nie, T.; Yang, Z.; Zhou, W. The role of soil components in the sorption of tetracycline and heavy metals in soils. RSC Adv. 2018, 8, 32178–32187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Li, L.L.; Huang, L.D.; Chung, R.S.; Fok, K.H.; Zhang, Y.S. Sorption and dissipation of tetracyclines in soils and compost. Pedosphere 2010, 20, 807–816. [Google Scholar] [CrossRef]
  24. Ter Laak, T.L.; Gebbink, W.A.; Tolls, J. Estimation of soil sorption coefficients of veterinary pharmaceuticals from soil properties. Environ. Toxicol. Chem. 2006, 25, 933–941. [Google Scholar] [CrossRef] [PubMed]
  25. Carter, L.J.; Wilkinson, J.L.; Boxall, A.B.A. Evaluation of existing models to estimate sorption coefficients for ionisable pharmaceuticals in soils and sludge. Toxics 2020, 8, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Droge, S.T.J.; Goss, K.U. Development and evaluation of a new sorption model for organic cations in soil: Contributions from organic matter and clay minerals. Environ. Sci. Technol. 2013, 47, 14233–14241. [Google Scholar] [CrossRef]
  27. Franco, A.; Trapp, S. Estimation of the soil–water partition coefficient normalized to organic carbon for ionizable organic chemicals. Environ. Toxicol. Chem. 2008, 27, 1995–2004. [Google Scholar] [CrossRef]
  28. Barron, L.; Havel, J.; Purcell, M.; Szpak, M.; Kelleher, B.; Paull, B. Predicting sorption of pharmaceuticals and personal care products onto soil and digested sludge using artificial neural networks. Analyst 2009, 134, 663–670. [Google Scholar] [CrossRef]
  29. Li, X.; Song, Y.; Jia, M.; Wang, F.; Bian, Y.; Jiang, X. Sorption and desorption characteristics of sulfamethazine in three different soils before and after removal of organic matter. Pedosphere 2021, 31, 796–806. [Google Scholar] [CrossRef]
  30. Conde-Cid, M.; Fernández-Calviño, D.; Nóvoa-Muñoz, J.C.; Núñez-Delgado, A.; Fernández-Sanjurjo, M.J.; Arias-Estévez, M.; Álvarez-Rodríguez, E. Experimental data and model prediction of tetracycline adsorption and desorption in agricultural soils. Environ. Res. 2019, 177, 108607. [Google Scholar] [CrossRef]
  31. Conde-Cid, M.; Nóvoa-Muñoz, J.C.; Fernández-Calviño, M.J.; Núñez-Delgado, A.; Álvarez-Rodríguez, E.; Arias-Estévez, M. Pedotransfer functions to estimate the adsorption and desorption of sulfadiazine in agricultural soils. Sci. Total Environ. 2019, 691, 933–942. [Google Scholar] [CrossRef] [PubMed]
  32. Conde-Cid, M.; Nóvoa-Muñoz, J.C.; Núñez-Delgado, A.; Fernández-Sanjurjo, M.J.; Arias-Estévez, M.; Álvarez-Rodríguez, E. Experimental data and modeling for sulfachloropyridazine and sulfamethazine adsorption/desorption on agricultural acid soils. Microporous Mesoporous Mat. 2019, 288, 109601. [Google Scholar] [CrossRef]
  33. Chu, B.; Goyne, K.W.; Anderson, S.H.; Lin, C.H.; Lerch, R.N. Sulfamethazine sorption to soil: Vegetative management, pH, and dissolved organic matter effects. J. Environ. Qual. 2013, 42, 794–805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Schaffer, M.; Licha, T. A framework for assessing the retardation of organic molecules in groundwater: Implications of the species distribution for the sorption-influenced transport. Sci. Total Environ. 2015, 524–525, 187–194. [Google Scholar] [CrossRef] [PubMed]
  35. Liu, Z.; Han, Y.; Jing, M.; Chen, J. Sorption and transport of sulfonamides in soils amended with wheat straw-derived biochar: Effects of water pH, coexistence copper ion, and dissolved organic matter. J. Soils Sediments 2015, 17, 771–779. [Google Scholar] [CrossRef]
  36. ElSayed, E.M.; Prasher, S.O. Sorption/desorption behavior of oxytetracycline and sulfachloropyridazine in the soil water surfactant system. Environ. Sci. Pollut. Res. 2014, 21, 3339–3350. [Google Scholar] [CrossRef]
  37. Jiang, Y.F.; Zhang, Q.; Deng, X.R.; Nan, Z.J.; Liang, X.R.; Wen, H.; Huang, K.; Wu, Y.Q. Single and competitive sorption of sulfadiazine and chlortetracycline on loess soil from Northwest China. Environ. Pollut. 2020, 263, 114650. [Google Scholar] [CrossRef]
  38. Schwarzenbach, R.P.; Gschwend, P.M.; Imboden, D.M. Environmental Organic Chemistry; J Wiley: New York, NY, USA, 1993; Volume 8, pp. 245–274. [Google Scholar]
  39. Weber, J.B.; Wilkerson, G.G.; Reinhardt, C.F. Calculating pesticide sorption coefficients (Kd) using selected soil properties. Chemosphere 2004, 55, 157–166. [Google Scholar] [CrossRef]
  40. Bao, Y.; Zhou, Q.; Wan, Y.; Yu, Q.; Xie, X. Effects of soil/solution ratios and cation types on adsorption and desorption of tetracycline in soils. Soil Sci. Soc. Am. J. 2010, 74, 1553–1561. [Google Scholar] [CrossRef] [Green Version]
  41. Conde-Cid, M.; Fernández-Calviño, D.; Fernández-Sanjurjo, M.J.; Núñez-Delgado, A.; Álvarez-Rodríguez, E.; Arias-Estévez, M. Adsorption/desorption and transport of sulfadiazine, sulfachloropyridazine, and sulfamethazine, in acid agricultural soils. Chemosphere 2019, 234, 978–986. [Google Scholar] [CrossRef]
  42. Guo, X.Y.; Shen, X.F.; Zhang, M.; Zhang, H.Y.; Chen, W.X.; Wang, H.; Koelmans, A.A.; Cornelissen, G.; Tao, S.; Wang, X.L. Sorption mechanisms of sulfamethazine to soil humin and its subfractions after sequential treatments. Environ. Pollut. 2017, 221, 266–275. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, X.L.; Cook, R.; Tao, S.; Xing, B.S. Sorption of organic contaminants by biopolymers: Role of polarity, structure and domain spatial arrangement. Chemosphere 2007, 66, 1476–1484. [Google Scholar] [CrossRef] [PubMed]
  44. Chu, B.; Goyne, K.W.; Anderson, S.H.; Lin, C.H.; Udawatta, R.P. Veterinary antibiotic sorption to agroforestry buffer, grass buffer and cropland soils. Agroforest. Syst. 2010, 79, 67–80. [Google Scholar] [CrossRef]
  45. Xu, Y.; Yu, W.; Ma, Q.; Zhou, H. Interactive effects of sulfadiazine and Cu(II) on their sorption and desorption on two soils with different characteristics. Chemosphere 2015, 138, 701–707. [Google Scholar] [CrossRef] [PubMed]
  46. Vieira, A.P.; Rath, S.; Fostier, A.H. Sorption of sulfachloropyridazine in Brazilian soils. J. Braz. Chem. Soc. 2017, 28, 158–167. [Google Scholar] [CrossRef]
  47. Feng, Y.; Li, Z.; Hao, X. Impacts of soil organic matter, iron-aluminium oxides and pH on adsorption-desorption behaviors of oxytetracycline. Res. J. Biotechnol. 2016, 11, 121–131. [Google Scholar]
  48. He, Y.; Liu, C.; Tang, X.Y.; Xian, Q.S.; Zhang, J.Q.; Guan, Z. Biochar impacts on sorption-desorption of oxytetracycline and florfenicol in an alkaline farmland soil as affected by field ageing. Sci. Total Environ. 2019, 671, 928–936. [Google Scholar] [CrossRef]
  49. Zhang, D.; Yang, S.K.; Wang, Y.N.; Yang, C.Y.; Chen, Y.Y.; Wang, R.Z.; Wang, Z.Z.; Yuan, X.Y.; Wang, W.K. Adsorption characteristics of oxytetracycline by different fractions of organic matter in sedimentary soil. Environ. Sci. Pollut. Res. 2019, 26, 5668–5679. [Google Scholar] [CrossRef]
  50. Jiang, W.T.; Chang, P.H.; Wang, Y.S.; Tsai, Y.; Jean, J.S.; Li, Z. Sorption and desorption of tetracycline on layered manganese dioxide birnessite. Int. J. Environ. Sci. Technol. 2015, 12, 1695–1704. [Google Scholar] [CrossRef] [Green Version]
  51. Kong, W.; Li, C.; Dolhi, J.M.; Li, S.; He, J.; Qiao, M. Characteristics of oxytetracycline sorption and potential bioavailability in soils with various physical-chemical properties. Chemosphere 2012, 87, 542–548. [Google Scholar] [CrossRef]
  52. Dollinger, J.; Dagès, C.; Voltz, M. Glyphosate sorption to soils and sediments predicted by pedotransfer functions. Environ. Chem. Lett. 2015, 13, 293–307. [Google Scholar] [CrossRef]
  53. Doretto, K.M.; Peruchi, L.M.; Rath, S. Sorption and desorption of sulfadimethoxine, sulfaquinoxaline and sulfamethazine antimicrobials in Brazilian soils. Sci. Total. Environ. 2014, 476–477, 406–414. [Google Scholar] [CrossRef] [PubMed]
  54. Ter Laak, T.L.; Gebbink, W.A.; Tolls, J. The effect of pH and ionic strength on the sorption of sulfachloropyridazine, tylosin, and oxytetracycline to soil. Environ. Toxicol. Chem. 2006, 25, 904–911. [Google Scholar] [CrossRef]
  55. Zhang, W.; Tang, X.; Thiele-Bruhn, S. Interaction of pig manure-derived dissolved organic matter with soil affects sorption of sulfadiazine, caffeine and atenolol pharmaceuticals. Environ. Geochem. Health. 2021, 43, 4299–4313. [Google Scholar] [CrossRef]
  56. Lertpaitoonpan, W.; Ong, S.K.; Moorman, T.B. Effect of organic carbon and pH on soil sorption of sulfamethazine. Chemosphere 2009, 76, 558–564. [Google Scholar] [CrossRef] [PubMed]
  57. Rath, S.; Fostier, A.H.; Pereira, L.A.; Dionisio, A.C.; Ferreira, F.O.; Doretto, K.M.; Peruchi, L.M.; Viera, A.; Neto, O.F.O.; Bosco, S.M.D.; et al. Sorption behaviors of antimicrobial and antiparasitic veterinary drugs on subtropical soils. Chemosphere 2019, 214, 111–122. [Google Scholar] [CrossRef]
  58. Srinivasan, P.; Sarmah, A.K. Assessing the sorption and leaching behaviour of three sulfonamides in pasture soils through batch and column studies. Sci. Total Environ. 2014, 493, 535–543. [Google Scholar] [CrossRef]
  59. Wegst-Uhrich, S.R.; Navarro, D.A.G.; Zimmerman, L.; Aga, D.S. Assessing antibiotic sorption in soil: A literature review and new case studies on sulfonamides and macrolides. Chem. Cent. J. 2014, 8, 5–16. [Google Scholar] [CrossRef] [Green Version]
  60. Hu, S.Q.; Zhang, Y.; Shen, G.X.; Zhang, H.C.; Yuan, Z.J.; Zhang, W. Adsorption/desorption behavior and mechanisms of sulfadiazine and sulfamethoxazole in agricultural soil systems. Soil Tillage Res. 2019, 186, 233–241. [Google Scholar] [CrossRef]
  61. Tolls, J. Sorption of veterinary pharmaceuticals in soils: A review. Environ. Sci. Technol. 2001, 35, 3397–3406. [Google Scholar] [CrossRef]
  62. Kah, M.; Sigmund, G.; Xiao, F.; Hofmann, T. Sorption of ionizable and ionic organic compounds to biochar, activated carbon and other carbonaceous materials. Water Res. 2017, 124, 673–692. [Google Scholar] [CrossRef] [PubMed]
  63. Lou, Y.Y.; Ye, Z.L.; Chen, S.H.; Ye, X.; Deng, Y.J.; Zhang, J.Q. Sorption behavior of tetracyclines on suspended organic matters originating from swine wastewater. J. Environ. Sci. 2018, 65, 144–152. [Google Scholar] [CrossRef] [PubMed]
  64. Gao, J.A.; Pedersen, J.A. Adsorption of sulfonamide antimicrobial agents to clay minerals. Environ. Sci. Technol. 2005, 39, 9509–9516. [Google Scholar] [CrossRef]
  65. Kahle, M.; Stamm, C. Time and pH-dependent sorption of the veterinary antimicrobial sulfathiazole to clay minerals and ferrihydrite. Chemosphere 2007, 68, 1224–1231. [Google Scholar] [CrossRef] [PubMed]
  66. Sithole, B.B.; Guy, R.D. Models for tetracycline in aquatic environments. I. interaction with bentonite clay systems. Water Air Soil Pollut. 1987, 32, 303–314. [Google Scholar] [CrossRef]
  67. Wang, S.; Wang, H. Adsorption behavior of antibiotic in soil environment: A critical review. Front. Environ. Sci. Eng. 2015, 9, 565–574. [Google Scholar] [CrossRef]
  68. Zhao, Y.; Geng, J.; Wang, X.; Gu, X.; Gao, S. Tetracycline adsorption on kaolinite: pH, metal cations and humic acid effects. Ecotoxicology 2011, 20, 1141–1147. [Google Scholar] [CrossRef]
  69. Xu, Y.B.; Yu, X.Q.; Xu, B.L.; Peng, D.; Guo, X.T. Sorption of pharmaceuticals and personal care products on soil and soil components: Influencing factors and mechanisms. Sci. Total Environ. 2021, 753, 141891. [Google Scholar] [CrossRef]
  70. Figueroa, R.A.; Leonard, A.; MacKay, A.A. Modeling tetracycline antibiotic sorption to clays. Environ. Sci. Technol. 2004, 38, 476–483. [Google Scholar] [CrossRef]
  71. Kim, Y.; Lim, S.; Han, M.; Cho, J. Sorption characteristics of oxytetracycline, amoxicillin, and sulfathiazole in two different soil types. Geoderma 2012, 185–186, 97–101. [Google Scholar] [CrossRef]
  72. Kurwadkar, S.T.; Adams, K.C.; Meyer, M.T.; Kolpin, D.W. Effects of sorbate speciation on sorption of selected sulfonamides in three loamy soils. J. Agric. Food Chem. 2007, 55, 1370–1376. [Google Scholar] [CrossRef] [PubMed]
  73. Li, J.; Carter, L.J.; Boxall, A.B.A. Evaluation and development of models for estimating the sorption behaviour of pharmaceuticals in soils. J. Hazard. Mater. 2020, 392, 122469. [Google Scholar] [CrossRef] [PubMed]
  74. Li, J.; Wilkinson, J.L.; Boxall, A.B.A. Use of a large dataset to develop new models for estimating the sorption of active pharmaceutical ingredients in soils and sediments. J. Hazard. Mater. 2021, 415, 125688. [Google Scholar] [CrossRef] [PubMed]
  75. Anna, B.B.; Joanna, M.; Wojciech, M.; Agata, B.; Marta, K.; Richard, P.; Piotr, S.; Jolanta, K. Sulfadimethoxine and sulfaguanidine: Their sorption potential on natural soils. Chemosphere 2012, 86, 1059–1065. [Google Scholar]
  76. Chen, H.; Zhang, J.Q.; Zhong, M.; Li, S.S.; Dong, Y.H. Adsorption of sulfonamides on paddy soil of Taihu Lake regeion. China Environ. Sci. 2008, 28, 309–312. [Google Scholar]
  77. Zeng, Q.H.; Wang, Y.W.; Li, L. Effect of acidification on adsorption behavior of sulfachloropyridazine (SCP) by black soil. Soils 2019, 51, 359–365. [Google Scholar]
  78. Fan, Z.; Casey, F.X.M.; Hakk, H.; Larsen, G.L.; Khan, E. Sorption, fate, and mobility of sulfonamides in soils. Water Air Soil Pollut. 2010, 218, 49–61. [Google Scholar] [CrossRef]
  79. Liu, Z.F. Study on Sorption-Desorption and Transportation of Typical Antibiotics in Soils. Master’s Thesis, China University of Geosciences, Beijing, China, 2016. [Google Scholar]
  80. Mutavdzic Pavlovic, D.; Curkovic, L.; Blazek, D.; Zupan, J. The sorption of sulfamethazine on soil samples: Isotherms and error analysis. Sci. Total Environ. 2014, 497–498, 543–552. [Google Scholar] [CrossRef] [Green Version]
  81. Vithanage, M.; Rajapaksha, A.U.; Tang, X.Y.; Thiele-Bruhn, S.; Kim, K.H.; Lee, S.E.; Ok, Y.S. Sorption and transport of sulfamethazine in agricultural soils amended with invasive-plant-derived biochar. J. Environ. Manag. 2014, 141, 95–103. [Google Scholar] [CrossRef]
  82. Pinna, M.V.; Castaldi, P.; Deiana, P.; Pusino, A.; Garau, G. Sorption behavior of sulfamethazine on unamended and manure-amended soils and short-term impact on soil microbial community. Ecotox. Environ. Saf. 2012, 84, 234–242. [Google Scholar] [CrossRef]
  83. Ren, M.; Tang, X.Y.; Geng, C.N.; Guan, Z.; Liu, C.; Xian, Q.S. Effects of biochar on adsorption-desorption and migration of antibiotics in slope farmland of purple soil. Soils 2020, 52, 978–986. [Google Scholar]
  84. Wang, N.; Guo, X.Y.; Xu, J.; Hao, L.J.; Kong, D.Y.; Gao, S.X. Sorption and transport of five sulfonamide antibiotics in agricultural soil and soil-manure systems. J. Environ. Sci. Health B 2015, 50, 23–33. [Google Scholar] [CrossRef] [PubMed]
  85. Wang, C. Research on the Features of Sorption and Desorption for Fluoroquinolones and Sulfonamides on Purple Soil. Master’s Thesis, Chongqing University, Chongqing, China, 2018. [Google Scholar]
  86. Zhou, Z.Q.; Liu, C.; Yang, H.W.; Xian, Q.S.; Tang, X.Y. Effects of biochar application on sorption-desorption process and leaching behaviour of sulfonamide antibiotics. Soils 2018, 50, 353–360. [Google Scholar]
  87. Doretto, K.M.; Rath, S. Sorption of sulfadiazine on Brazilian soils. Chemosphere 2013, 90, 2027–2034. [Google Scholar] [CrossRef] [Green Version]
  88. Kasteel, R.; Mboh, C.M.; Unold, M.; Groeneweg, j.; Vanderborght, J.; Vereecken, H. Transformation and sorption of the veterinary antibiotic sulfadiazine in two soils a short-term batch study. Environ. Sci. Technol. 2010, 44, 4651–4657. [Google Scholar] [CrossRef]
  89. Kong, J.J.; Pei, Z.G.; Wen, P.; Shan, X.Q.; Chen, Z.L. Adsorption of sulfadiazine and sulfathiazole by soils. Environ. Chem. 2008, 27, 736–741. [Google Scholar]
  90. Li, S.H.; Liu, C.; Tang, X.Y.; Yang, H.W. Leaching characteristics of dissolved organic matter in chicken manure and its effect on antibiotic migration in orchard. Trans. Chin. Soc. Agric. Eng. 2020, 36, 37–46. [Google Scholar]
  91. Lou, F.L. Effects of Dissolved Organic Matter form Pig Manure on Sorption and Migration of Antibiotics in Purple Soil. Master’s Thesis, Southwest Jiaotong University, Chengdu, China, 2019. [Google Scholar]
  92. Shao, Z.Z.; Li, Q.; Xu, S.H. Effect of silica colloids on adsorption and migration of sulfadiazine in soil relative to ionic intensity. Acta Petrol. Sin. 2018, 55, 411–421. [Google Scholar]
  93. Shao, Z.Z. Adsorptin, Migration and Numercical Simulation of Sulfadiazine in Soil under Silica Colloids. Master’s Thesis, Qingdao University, Qingdao, China, 2018. [Google Scholar]
  94. Sukul, P.; Lamshoft, M.; Zuhlke, S.; Spiteller, M. Sorption and desorption of sulfadiazine in soil and soil-manure systems. Chemosphere 2008, 73, 1344–1350. [Google Scholar] [CrossRef]
  95. Xu, Z.; Lv, S.; Hu, S.; Chao, L.; Rong, F.; Wang, X.; Dong, M.; Liu, K.; Li, M.; Liu, A. Effect of soil solution properties and Cu2+ co-existence on the adsorption of sulfadiazine onto paddy soil. Int. J. Environ. Res. Public Health 2021, 18, 13383. [Google Scholar] [CrossRef]
  96. Zhang, C.L.; Wang, Y.; Wen, C.B.; Wang, F.A. Study on the adsorption for sulfadiazine in the different type soils. J. Agric. Mech. Res. 2007, 9, 143–146. [Google Scholar]
  97. Kocarek, M.; Kodesova, R.; Vondrackova, L.; Golovko, O.; Fer, M.; Klement, A.; Nikodem, A.; Jaksik, O.; Grabic, R. Simultaneous sorption of four ionizable pharmaceuticals in different horizons of three soil types. Environ. Pollut. 2016, 218, 563–573. [Google Scholar] [CrossRef] [PubMed]
  98. Schmidtova, Z.; Kodesova, R.; Grabicova, K.; Kocarek, M.; Fer, M.; Svecova, H.; Klement, A.; Nikodem, A.; Grabic, R. Competitive and synergic sorption of carbamazepine, citalopram, clindamycin, fexofenadine, irbesartan and sulfamethoxazole in seven soils. J. Contam. Hydrol. 2020, 234, 103680. [Google Scholar] [CrossRef] [PubMed]
  99. Srinivasan, P.; Sarmah, A.K.; Manley-Harris, M. Co-contaminants and factors affecting the sorption behaviour of two sulfonamides in pasture soils. Environ. Pollut. 2013, 180, 165–172. [Google Scholar] [CrossRef]
  100. Srinivasan, P.; Sarmah, A.K. Characterisation of agricultural waste-derived biochars and their sorption potential for sulfamethoxazole in pasture soil: A spectroscopic investigation. Sci. Total Environ. 2015, 502, 471–480. [Google Scholar] [CrossRef]
  101. Wang, X.S. The Characteristic Study of Different Type of Soils Absorption/Desorption for Sulfamethoxazole. Master’s Thesis, Jilin Agricultural University, Changchun, China, 2017. [Google Scholar]
  102. Wang, J.Q. Adsorption-Desorption and Vertical Migration of Sulfamethoxazole in a Soil under the Impact of Microplastics. Master’s Thesis, Zhejiang A&F University, Hangzhou, China, 2021. [Google Scholar]
  103. Wu, Y.; Chen, D.H.; Huang, M.H. Sorption and pH effect on selected antibiotics in soils. Adv. Mater. Res. 2011, 356–360, 35–38. [Google Scholar] [CrossRef]
  104. Bao, Y.Y.; Zhou, Q.X.; Wang, Y.; Xie, X.J. Effect of soil organic matter on adsorption and desorption of oxytetracycline in soils. China Environ. Sci. 2009, 29, 651–655. [Google Scholar]
  105. Bao, Y.Y.; Zhou, Q.X.; Zhang, H. Influences of cation species on adsorption and desorption of oxytetracycline in two typical soils of China. Environ. Sci. 2009, 30, 551–556. [Google Scholar]
  106. Bao, Y.Y.; Zhou, Q.X.; Wang, Y.; Yu, Q.; Xie, X.J. Adsorption and desorption of three tetracycline antibiotics in cinnamon soils of China. China Environ. Sci. 2010, 30, 1383–1388. [Google Scholar]
  107. Cao, Z.L.; Yu, H.M.; Ge, C.J.; Luo, J.W.; Wang, P.; Zhao, Y.Y.; Li, J.T. Effects of dissolved organic matter on adsorption-desorption behavior of oxytetracycline in soil system. Chin. J. Trop. Crop. 2018, 39, 825–831. [Google Scholar]
  108. Chen, W.W.; Chen, T.; Yang, P.; Guo, P.; Zhang, W.Q.; Zhang, X.Y. Adsorption behavior of oxytetracycline on black soil under the influence of coexisting Cu2+. Jiangsu Environ. Sci. 2017, 45, 240–244. [Google Scholar]
  109. Conde-Cid, M.; Ferreira-Coelho, G.; Núñez-Delgado, A.; Fernández-Calviño, D.; Arias-Estévez, M.; Álvarez-Rodríguez, E.; Fernández-Sanjurjo, M.J. Competitive adsorption of tetracycline, oxytetracycline and chlortetracycline on soils with different pH value and organic matter content. Environ. Res. 2019, 178, 108669. [Google Scholar] [CrossRef] [PubMed]
  110. He, H. Effects of Biochar on Sorption and Transportation of Florfenicol in a Purple Soil. Master’s Thesis, Southwest Jiaotong University, Chengdu, China, 2019. [Google Scholar]
  111. Li, G.A.; Chen, Z.H.; Liu, Z.F.; Li, Y. Adsorption analysis of oxytetracycline on fluvo-aquic soils in Beijing. Geoscience 2015, 29, 377–382. [Google Scholar]
  112. Li, X.; Wang, D.S.; Zhang, T. Adsorption-desorption behavior of oxytetracycline (OTC) and chlortetracycline (CTC) adsorption-desorption. J. Earth Environ. 2015, 6, 317–322. [Google Scholar]
  113. Li, X. The Study on Environmental Behavior of Tetracycline Antibiotics in Different Texture Soils. Master’s Thesis, Liaoning Technical University, Fuxin, China, 2015. [Google Scholar]
  114. Li, Y.; Pan, T.; Miao, D.; Chen, Z.; Tao, Y. Sorption–desorption of typical tetracyclines on different soils: Environment hazards analysis with partition coefficients and hysteresis index. Environ. Eng. Sci. 2015, 32, 865–871. [Google Scholar] [CrossRef]
  115. Li, J.; Yu, S.G.; Shen, L.E.; Cui, M.; Wang, Y.L. Influence of microplastics on sorption behaviors of oxytetracycline onto soils: A preliminary study. Environ. Chem. 2021, 40, 3133–3143. [Google Scholar]
  116. Liu, B.; Bao, Y.Y.; Zhou, Q.X.; Zhang, C.D. Effect of N, P fertilizers on adsorption of oxytetracycline to cinnamon soil. China Environ. Sci. 2014, 34, 2057–2062. [Google Scholar]
  117. Mette, R.; Spliid, N.H. Sorption and mobility of metronidazole, olaquindox, oxytetracycline and tylosin in soil. Chemosphere 2000, 40, 715–722. [Google Scholar]
  118. Ming, L. Studies of the Effect of DOM on the Adsorption of Cu and OTC in Combined System. Master’s Thesis, Jilin University, Changchun, China, 2014. [Google Scholar]
  119. Peng, F.J.; Ying, G.G.; Zhou, L.J.; Liu, Y.S.; Pan, C.G.; Liang, Q.Y. Adsorption and desorption of oxytetracycline on typical soils and soil-adsorbed oxytetracycline’s bioavailability. Geochimica 2015, 44, 71–78. [Google Scholar]
  120. Qi, H.M. Sorption of Oxytetracycline to Wushan Soil, Red Soil and Their Main Components. Master’s Thesis, Dalian University of Technology, Dalian, China, 2008. [Google Scholar]
  121. Shi, L.P.; Jiang, Y.F.; Guang, A.L.; Yuan, L.M.; Liu, L.L.; Zhan, H.Y. Effect of natural organic matter on adsorption behavior of oxytetracycline onto sierozem soils in northwest China. Res. Environ. Sci. 2019, 32, 1584–1593. [Google Scholar]
  122. Wang, D.S.; Zhang, T.; Chao, Y. Influence of different strength and species of cation on adsorption of oxytetracycline in meadow soils. Ecol. Environ. Sci. 2014, 23, 870–875. [Google Scholar]
  123. Wang, D.S.; Li, X. Adsorption and desorption features of tetracycline antibiotics in different texture soils. J. Saf. Environ. 2017, 17, 227–231. [Google Scholar] [CrossRef] [PubMed]
  124. Bao, Y.Y.; Wan, Y.; Zhou, Q.X.; Li, W.M.; Liu, Y.X. Competitive adsorption and desorption of oxytetracycline and cadmium with different input loadings on cinnamon soil. J. Soils Sediments 2012, 13, 364–374. [Google Scholar] [CrossRef]
  125. Yan, L.; Pan, D.Q.; Jiang, X.X.; Ji, X.L.; Yang, H.H.; Li, S.Q.; Yu, M. Adsorption behaviour of tetracycline antibiotics in black soil and albic soil. J. Northeast Agric. Univ. 2017, 48, 54–62. [Google Scholar]
  126. Yao, Y. Adsorption of OTC on Soils. Master’s Thesis, Liaoning Technical University, Fuxin, China, 2013. [Google Scholar]
  127. Yi, L.L.; Jiao, W.T.; Chen, W.P. Adsorption characteristics of three types of antibiotics in the soil profiles. Environ. Chem. 2013, 32, 2357–2363. [Google Scholar]
  128. Zhang, M.K.; Wang, L.P.; Zheng, S.A. Adsorption and transport characteristics of two exterior-source antibiotics in some agricultural soils. Acta Ecol. Sin. 2008, 28, 761–766. [Google Scholar]
  129. Zhu, W.G.; Duan, Y.Y.; Meng, G.J.; Guo, R.C.; Li, X.H. Adsorption-desorption of tetracycline and oxytetracycline in Cu contaminated soil. J. Henan Univ. 2020, 50, 11–18. [Google Scholar]
  130. Alvarez-Esmoris, C.; Conde-Cid, M.; Fernandez-Sanjurjo, M.J.; Nunez-Delgado, A.; Alvarez-Rodriguez, E.; Arias-Estevez, M. Environmental relevance of adsorption of doxycycline, enrofloxacin, and sulfamethoxypyridazine before and after the removal of organic matter from soils. J. Environ. Manag. 2021, 287, 112354. [Google Scholar] [CrossRef]
  131. Bao, Y.; Zhou, Q.; Wang, Y. Adsorption characteristics of tetracycline by two soils: Assessing role of soil organic matter. Aust. J. Soil Res. 2009, 47, 286–295. [Google Scholar] [CrossRef]
  132. Bao, Y.J.; Ding, H.S.; Bao, Y.Y. Effects of temperature on the adsorption and desorption of tetracycline in soils. Adv. Mater. Res. 2013, 726–731, 344–347. [Google Scholar] [CrossRef]
  133. Jiao, S.J.; Sum, Z.H.; Zheng, S.R.; Yi, D.Q.; Pu, H.Q.; Chen, L.Y. Sorption and desorption of tetracycline on Wushantu soil. J. Agro-Environ. Sci. 2008, 27, 1732–1736. [Google Scholar]
  134. Wan, Y.; Bao, Y.Y.; Zhou, Q.X. Simultaneous adsorption and desorption of cadmium and tetracycline on cinnamon soil. Chemosphere 2010, 80, 807–812. [Google Scholar] [CrossRef] [PubMed]
  135. Wan, Y.; Bao, Y.Y.; Zhou, Q.X. Adsorption and desorption of tetracycline and effect of cadmium on these in two typical soils of China. J. Agro-Environ. Sci. 2010, 29, 85–90. [Google Scholar]
  136. Wu, T.X.; Zhou, M.; Guo, H.D.; Duan, H.; Chen, H. Adsorption of tetracycline on loess soils. Acta Sci. Circum. 2008, 28, 2311–2314. [Google Scholar]
  137. Zhang, G.X.; Liu, X.T.; Sun, K.; Zhao, Y.; Lin, C.Y. Sorption of tetracycline to sediments and soils: Assessing the roles of pH, the presence of cadmium and properties of sediments and soils. Front. Environ. Sci. Eng. China 2010, 4, 421–429. [Google Scholar] [CrossRef]
  138. Ding, Y.H.; Hu, Y.Z.; Chen, J.Q. Research on the adsorption and desorption behavior of chlorotetracycline in silty clay loam. J. Anhui Agric. Sci. 2011, 39, 8979–8982. [Google Scholar]
  139. Jiang, Y.F.; Liang, X.R.; Yuan, L.M.; Nan, Z.J.; Deng, X.R.; Wu, Y.Q.; Ma, F.F.; Diao, J.R. Effect of livestock manure on chlortetracycline sorption behaviour and mechanism in agricultural soil in Northwest China. Chem. Eng. J. 2021, 415, 129020. [Google Scholar] [CrossRef]
  140. Liu, X.C.; Dong, Y.H. Adsorption-desorption of chlortetracycline in various cultivated soils in China. Acta Pedologica Sinica 2009, 46, 861–868. [Google Scholar]
  141. Liu, X.C.; Dong, Y.H.; Liu, H.J. Competitive absorption of chlortetracycline by cation in typical soils of China. Acta Pedol. Sin. 2010, 47, 781–785. [Google Scholar]
  142. Liu, F.; Liu, H.J.; Yang, S.J.; Dong, Y.H.; Zhang, Z.L.; Sun, W.T.; Liu, X.C. Adsorption and desorption behavior of chlortetracycline in brown soils of different farming conditions. J. Anhui Agric. Sci. 2011, 39, 16053–16055, 16065. [Google Scholar]
  143. Nan, Z.J.; Jiang, Y.F.; Mao, H.H.; Liang, X.R.; Deng, X.R. Effect of corn stalk biochar on the adsorption of aureomycin from sierozem. Environ. Sci. 2021, 42, 5896–5904. [Google Scholar]
  144. Wan, Y.; Bao, Y.Y.; Zhou, Q.X. Effect of soil organic matter and cadmium (II) on adsorption and desorption of chlortetracycline in soil. Environ. Sci. 2010, 31, 3050–3055. [Google Scholar]
  145. Yang, Y.B. Study on Adsorption Characteristics of Chlorotetracycline in Different Soil in Jilin Province. Master’s Thesis, Jilin Agricultural University, Changchun, China, 2016. [Google Scholar]
  146. Zhang, Y.L.; Lin, S.S.; Dai, C.M.; Shi, L.; Zhou, X.F. Sorption-desorption and transport of trimethoprim and sulfonamide antibiotics in agricultural soil: Effect of soil type, dissolved organic matter, and pH. Environ. Sci. Pollut. Res. 2014, 21, 5827–5835. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Measured Kf and Kd in sub-datasets “a” and “b” against estimated Kf and Kd from the pedotransfer functions. The solid line represents a perfect model fit (1:1 line), and the dashed lines represent a difference of one order of magnitude, which indicate satisfactorily estimated values.
Figure 1. Measured Kf and Kd in sub-datasets “a” and “b” against estimated Kf and Kd from the pedotransfer functions. The solid line represents a perfect model fit (1:1 line), and the dashed lines represent a difference of one order of magnitude, which indicate satisfactorily estimated values.
Ijerph 19 16771 g001
Figure 2. Absolute error (AE) distributions of Kf and Kd estimation across different OC and pH ranges in sub-datasets “A” and “B” using the pedotransfer equations given in Table 4.
Figure 2. Absolute error (AE) distributions of Kf and Kd estimation across different OC and pH ranges in sub-datasets “A” and “B” using the pedotransfer equations given in Table 4.
Ijerph 19 16771 g002
Table 1. Physicochemical properties of the target antibiotics.
Table 1. Physicochemical properties of the target antibiotics.
AntibioticPhysical-Chemical PropertiesChemical StructureSpecies Distribution f
SCPMolecular formulaC10H9ClN4O2SIjerph 19 16771 i001Ijerph 19 16771 i002
Solubility (mg L−1)7000.00 c
Mw (g mol−1)284.72
logKow–0.80 a
pKa11.87 a
pKa25.45 a
SMTMolecular formulaC12H14N4O2SIjerph 19 16771 i003Ijerph 19 16771 i004
Solubility (mg L−1)1500.00 a
Mw (g mol−1)278.34
logKow0.14 a
pKa12.07 a
pKa27.49 a
SDZMolecular formulaC10H10N4O2SIjerph 19 16771 i005Ijerph 19 16771 i006
Solubility (mg L−1)77.00 a
Mw (g mol−1)250.30
logKow−1.05 a
pKa12.10 a
pKa26.28 a
SMXMolecular formulaC10H11N3O3SIjerph 19 16771 i007Ijerph 19 16771 i008
Solubility (mg L−1)370.00 b
Mw (g mol−1)253.28
logKow0.89 b
pKa11.83 b
pKa25.62 b
OTCMolecular formulaC22H24N2O9Ijerph 19 16771 i009Ijerph 19 16771 i010
Solubility (mg L−1)1000.00 c
Mw (g mol−1)460.40
logKow–0.12 c
pKa13.30 c
pKa27.30 c
pKa39.10 c
TCMolecular formulaC22H24N2O8Ijerph 19 16771 i011Ijerph 19 16771 i012
Solubility (mg L−1)231.00 d
Mw (g mol−1)444.43
logKow−1.37 d
pKa13.20 d
pKa27.80 d
pKa39.60 d
CTCMolecular formulaC22H23ClN2O8Ijerph 19 16771 i013Ijerph 19 16771 i014
Solubility (mg L−1)4120.00 e
Mw (g mol−1)479.00
logKow2.07 e
pKa13.30 e
pKa27.44 e
pKa39.27 e
Note: a [5]; b [35]; c [36]; d [30]; e [37]; f [38].
Table 2. Statistical characteristics of sorption parameters of individual antibiotics in the soils.
Table 2. Statistical characteristics of sorption parameters of individual antibiotics in the soils.
AntibioticParameterStatistics
MaxMinMeanMedianNobs *
SCPKf (mg1−1/n L1/n kg−1)15.330.605.254.6083
n2.980.901.181.2083
Kd (L kg−1)23.100.304.423.2598
SMTKf (mg1−1/n L1/n kg−1)16.000.133.352.67142
n4.000.411.201.20142
Kd (L kg−1)14.200.112.451.62153
SDZKf (mg1−1/n L1/n kg−1)4.860.451.651.4577
n2.170.281.001.0077
Kd (L kg−1)12.700.092.051.40104
SMXKf (mg1−1/n L1/n kg−1)12.600.1332.762.0059
n2.380.481.281.2059
Kd (L kg−1)28.500.022.491.4072
OTCKf (mg1−1/n L1/n kg−1)5110.0074.001901.521814.00133
n6.250.302.091.85126
Kd (L kg−1)2191.0016.76692.59601.17121
TCKf (mg1−1/n L1/n kg−1)6928.090.281707.471640.61107
n6.850.392.102.13107
Kd (L kg−1)1940.8910.06481.53362.3090
CTCKf (mg1−1/n L1/n kg−1)8176.99302.003220.163131.5693
n4.000.432.202.3393
Kd (L kg−1)4473.20147.081596.771219.2188
* Number of reported observations.
Table 3. Pearson correlation coefficients (r) of sorption parameters with soil properties.
Table 3. Pearson correlation coefficients (r) of sorption parameters with soil properties.
AntibioticSoil PropertyKfnKd
SCPpH0.0710.248 *0.080
OC0.772 **–0.1200.631 **
CEC0.338 **0.215 *0.200 *
Sand–0.207 *–0.122–0.170
Silt0.0750.081–0.080
Clay0.310 **0.1070.372 **
SMTpH–0.0090.0560.011
OC0.754 **–0.172 *0.717 **
CEC0.409 **0.0350.269 **
Sand–0.137–0.211 *0.040
Silt0.1240.179 *–0.062
Clay0.1010.163 *0.014
SDZpH0.0680.1600.267 **
OC0.364 **–0.1510.686 **
CEC–0.1190.0880.353 **
Sand–0.439 **–0.263 *–0.415 **
Silt0.287 *0.1950.203 *
Clay0.380 **0.1950.442 **
SMXpH–0.381 **0.082–0.389 **
OC0.738 **–0.1800.626 **
CEC0.435 **0.0260.270 *
Sand0.0520.109–0.008
Silt–0.214–0.028–0.063
Clay0.235–0.1500.119
OTCpH–0.749 **–0.612 **–0.268 **
OC0.587 **0.216 *0.670 **
CEC–0.488 **–0.490 **0.146
Sand0.350 **0.374 **–0.030
Silt–0.313 **–0.293 **0.011
Clay–0.174 *–0.252 **0.048
TCpH–0.570 **–0.727 **0.283 *
OC0.585 **0.1350.151
CEC–0.093–0.603 **0.480 **
Sand0.186 *0.474 **–0.704 **
Silt–0.175–0.463 **0.457 **
Clay–0.104–0.265 **0.669 **
CTCpH–0.457 **–0.737 **–0.149
OC0.684 **0.302 **0.419 **
CEC–0.240 *–0.672 **0.073
Sand0.1270.576 **–0.061
Silt–0.081–0.417 **–0.049
Clay–0.104–0.426 **0.226 *
Note: * and ** represent significance at the 0.05 and 0.01 probability level, respectively.
Table 4. Statistical characteristics of sorption parameters of individual antibiotics in the soils.
Table 4. Statistical characteristics of sorption parameters of individual antibiotics in the soils.
AntibioticPedotransfer Functionr2RMSE (RMSE/SD)NSENobs 1
SCP K f = 4.198 + 1.666 OC 0.735 pH 0.616 ** 22.01(61.1%)0.6368
K d = 2.769 + 1.668 OC 0.689 pH 0.689 **2.34 (55.1%)0.7080
SMT K f = 0.820 + 0.818 OC 0.565 **1.24 (65.6%)0.57107
K d = 0.099 + 0.789 OC 0.510 **1.58 (69.6%)0.52114
SDZ K f = 1.951 + 0.239 OC 0.018 Sand 0.380 **0.39 (77.2%)0.4053
K d = 0.480 + 0.484 OC + 0.201 pH 0.507 **0.73 (69.4%)0.5283
SMX K f = 4.717 + 0.464 OC 0.565 pH + 0.006 CEC 0.027 Silt 0.666 **1.07 (56.2%)0.6949
K d = 3.208 + 0.519 OC 0.457 pH 0.525 **0.88 (67.2%)0.5457
OTC K f = 4428.177 + 330.323 OC 543.318 pH 0.606 **851.94 (62.2%)0.61104
K d = 182.875 + 239.030 OC 0.444 **328.77 (74.2%)0.4594
TC K f = 2740.451 + 215.512 OC 363.881 pH + 3.331 CEC 0.509 **612.94 (68.8%)0.5384
K d = 274.636 + 1.151 CEC 5.607 Sand + 14.741 Clay 0.607 **179.22 (61.3%)0.6267
CTC K f = 2345.591 + 1205.573 OC 281.455 pH 0.524 **1340.46 (68.0%)0.5473
K d = 588.94 + 561.887 OC + 38.582 Clay 0.371 **811.11 (78.2%)0.3972
1 Number of reported observations. 2 ** represents significance at the 0.01 probability level.
Table 5. Improved models by inclusion of additional independent variable(s) for estimating Kf and Kd of antibiotics in soils (based on sub-datasets “A” and “B”).
Table 5. Improved models by inclusion of additional independent variable(s) for estimating Kf and Kd of antibiotics in soils (based on sub-datasets “A” and “B”).
AntibioticPedotransfer FunctionR2
SCP K f = 0.807 + 1.657 OC + 2.023 α 0 0.619 **
K d = 1.149 + 1.657 OC + 183.089 α + 0.695 **
SDZ K d = 0.980 + 0.398 OC + 0.200 Clay 0.949 α 0 0.579 **
SMX K d = 1.652 + 0.485 OC 1.91 α 0.552 **
OTC K f = 4595.375 701.271 pH + 397.550 OC + 16589.89 SLR + 5204.082 α 0.647 **
TC K f = 2863.937 + 288.648 OC 265.391 pH 20202.101 SLR 0.545 **
K d = 325.965 + 51.147 OC 6.099 Sand + 15.714 Clay 1058.792 α + 0.633 **
CTC K f = 256.377 + 1219.651 OC + 6.639 C imax 0.549 **
Note: ** represents significance at the 0.01 probability level.
Table 6. Performance of previously published models for sub-datasets “a” and “b”.
Table 6. Performance of previously published models for sub-datasets “a” and “b”.
AntibioticPedotransfer Functionr2OriginNobs 1RMSE
(RMSE/SD)
NSEReference
SCP K f = 8.810 + 1.967 OC 2.028 pH 0.829 ** 2Galicia (Spain)501.35
(62.4%)
0.61[32]
SMT K d = 0.38 + 0.81 OC 0.92 **Iowa (USA)51.62
(63.0%)
0.60[56]
SDZ K f = 3.493 + 0.780 OC 0.819 pH 0.675 **Galicia (Spain)501.81
(149.6%)
−1.20[31]
OTC K f = 96.924 + 701.607 OC + 8118.902 CEC 1 0.349 **Galicia (Spain)632052.63
(221.87%)
−3.90[13]
1 Number of reported observations used to evaluate the performance of previously published models. 2 ** represents significance at the 0.01 probability level.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hu, J.; Tang, X.; Qi, M.; Cheng, J. New Models for Estimating the Sorption of Sulfonamide and Tetracycline Antibiotics in Soils. Int. J. Environ. Res. Public Health 2022, 19, 16771. https://doi.org/10.3390/ijerph192416771

AMA Style

Hu J, Tang X, Qi M, Cheng J. New Models for Estimating the Sorption of Sulfonamide and Tetracycline Antibiotics in Soils. International Journal of Environmental Research and Public Health. 2022; 19(24):16771. https://doi.org/10.3390/ijerph192416771

Chicago/Turabian Style

Hu, Jinsheng, Xiangyu Tang, Minghui Qi, and Jianhua Cheng. 2022. "New Models for Estimating the Sorption of Sulfonamide and Tetracycline Antibiotics in Soils" International Journal of Environmental Research and Public Health 19, no. 24: 16771. https://doi.org/10.3390/ijerph192416771

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