Next Article in Journal
Current Trends and Future Scenarios: Modeling Maximum River Discharge in the Zhaiyk–Caspian Basin (Kazakhstan) Under a Changing Climate
Previous Article in Journal
Epikarst Flow Dynamics and Contaminant Attenuation: Field and Laboratory Insights from the Suva Planina Karst System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Seasonal Freshwater Inflows in Cochin Backwater Estuary Inferred from Stable Isotopes and Machine Learning

1
Birbal Sahni Institute of Palaeosciences, Lucknow 226007, India
2
Department of Environmental Health and Safety, University of Doha for Science and Technology, Doha P.O. Box 24449, Qatar
3
Centre for Excellence in Sustainability and Food Security, University of Doha for Science and Technology, Doha P.O. Box 24449, Qatar
4
National Computer Network Intrusion Protection Center, University of Chinese Academy of Sciences, Beijing 100083, China
5
Centre for Earth Sciences, Indian Institute of Science, Bangalore 560012, India
6
Divecha Centre for Climate Change, Indian Institute of Science, Bangalore 560012, India
7
Interdisciplinary Centre for Water Research, Indian Institute of Science, Bangalore 560012, India
8
Department of Chemical Oceanography, Cochin University of Science and Technology, Cochin 682022, India
*
Authors to whom correspondence should be addressed.
Hydrology 2025, 12(11), 277; https://doi.org/10.3390/hydrology12110277
Submission received: 20 July 2025 / Revised: 1 September 2025 / Accepted: 7 September 2025 / Published: 24 October 2025
(This article belongs to the Section Marine Environment and Hydrology Interactions)

Abstract

The Cochin Backwater region in Southern India is one of the most dynamic estuaries, strongly influenced by seasonal river runoff and seawater intrusion. This study explores the relationship between monsoonal rains, salinity, and stable isotopic composition (δ18O and δ13C) to estimate the contribution of freshwater fluxes at different seasonal intervals for the Cochin Backwater (CBW) estuary. Seasonal variations in oxygen isotopes and salinity revealed distinct trends indicative of freshwater–seawater mixing dynamics. The comparison of Local and Global Meteoric Water Lines highlighted the occurrence of enriched isotope values during the Premonsoon season, showing significant evaporation effects. Carbon (C) isotopic analysis in dissolved inorganic matter (δ13CDIC) at 17 stations during the Premonsoon season revealed spatially distinct carbon dynamics zones, influenced by various sources. These characteristic zones were categorized as Zone 1, dominated by seawater, exhibiting heavier δ13CDIC values; Zone 2, showing significant contributions of lighter terrestrial δ13C; and Zone 3, reflecting inputs from regional and local paddy fields with a distinct C3 isotopic signature (−25‰), modified by estuarine productivity. In addition, different advanced machine learning techniques were tested to improve analysis and prediction of seasonal variations in isotopic composition and salinity. Although the data were sufficiently robust for demonstrating the feasibility and advantages of ML in isotopic hydrology, further expansion of the dataset would be essential for improving the accuracy of models, especially for δ13C. The combination of these advanced machine learning models not only improved the predictive accuracy of seasonal freshwater fluxes but also provided a robust framework for understanding the estuarine ecosystem and could pave the way for better management and conservation strategies of the CBW estuarine system.

1. Introduction

Estuaries are among the most dynamic transitional environments, acting as vital interfaces between terrestrial and marine systems. They provide a range of ecosystem services such as nutrient cycling, carbon sequestration, fisheries support, and biodiversity maintenance, making them critical socioecological systems worldwide [1]. The hydrology of estuaries is strongly governed by freshwater influx from rivers and saline water intrusion from the sea, processes that vary seasonally and exert profound influences on productivity, biogeochemistry, and ecological functioning. Globally, estuarine systems such as the Mekong Delta in Southeast Asia, the Amazon Estuary in South America, and Chesapeake Bay in North America have demonstrated how freshwater fluxes, tidal regimes, and anthropogenic pressures shape estuarine dynamics across spatial and temporal scales [1,2,3]. Understanding these dynamics is essential not only for local resource management but also for assessing the resilience of estuarine systems to climate variability and human-induced stressors.
In the Indian context, monsoon-fed estuaries represent a unique class of semi-enclosed coastal systems. Their hydrography is primarily regulated by the seasonal monsoons, in contrast to temperate estuaries where glacial meltwater or snowmelt often dominates seasonal fluxes [1,4]. Among these, the Cochin Backwater (CBW) estuary is one of the largest and most studied estuarine systems in Southern India. Fed by six rivers and strongly influenced by semi-diurnal tides, the CBW estuary experiences pronounced salinity stratification, seasonal freshwater inflows, and extensive anthropogenic interventions, including urbanization, dredging, and agricultural activities [5,6,7,8]. Prior studies in this estuary have largely focused on hydrochemistry, nutrient cycling, and anthropogenic impacts [6,7,8,9], yet a comprehensive seasonal assessment of freshwater fluxes using isotopic tracers has remained limited.
The primary sources of seasonal freshwater inflows into the CBW estuary are predominantly attributed to river discharges, as well as precipitation influenced by monsoonal patterns. CBW receives freshwater from six rivers, with the Periyar River contributing approximately 33% of the total river discharge (~19,000 million cubic meters/year) [10]. This riverine input is crucial for maintaining the estuarine salinity and nutrient levels, especially during the monsoon season when freshwater influx is at its peak [11]. In addition, these inflows significantly affect the estuarine ecosystem, nutrient dynamics, and isotopic signatures of the water. Seasonal variations in stable water isotopes indicate that monsoonal precipitation plays a significant role in freshwater inflow. The Southwest Monsoon (June–September) is characterized by marine moisture sources, while the Northeast Monsoon (October–December) shows continental contributions [12]. The isotopic composition of precipitation varies across different regions, reflecting the influence of altitude and local climatic conditions on freshwater sources [12].
Stable isotopes of oxygen (δ18O, δD) and carbon (δ13C) have long been recognized as robust tools to trace water sources, quantify mixing processes, and assess biogeochemical cycling in estuarine and coastal systems [13,14,15,16,17]. Globally, these approaches have successfully resolved freshwater–seawater mixing (e.g., Amazon, Mekong, Hooghly estuaries), evaporation effects, and carbon cycling linked to terrestrial versus marine inputs. In India, isotopic studies have addressed aspects of estuarine mixing in the Hooghly and Mandovi–Zuari estuaries [13,18] and carbon cycling in CBW [15,19], but these efforts have generally been localized in scope, event-specific, or limited to a single tracer. Systematic integration of δ18O, δD, and δ13C signatures with salinity to quantify freshwater contributions across multiple monsoon seasons remains underdeveloped, despite its potential to reveal spatial zonation and ecosystem-level processes.
Machine learning (ML) techniques are emerging as powerful alternatives to conventional statistical or box-model approaches in hydrology and geochemistry. Methods such as Artificial Neural Networks (ANNs), Random Forests (RFs), Support Vector Machines (SVMs), Gradient Boosting Machines (GBMs), and Gaussian Process Regression (GPR) have been widely applied to predict water quality parameters, isotopic compositions, and ecosystem responses in temperate and subtropical settings [20,21,22,23,24,25,26,27]. For example, in the Mediterranean Sea, RF and ANN models have successfully predicted isotopic and salinity variations with high accuracy [20]. Similarly, ML-based approaches have been used to forecast groundwater isotopes [21,24], eutrophication [23], and evaporation rates [26]. However, applications of ML in the Asian monsoon region, especially from an estuarine context are almost non-existent. To date, no study has systematically tested ML models for predicting stable isotopic tracers (δ18O, δ13C) and freshwater fluxes in tropical estuaries where seasonal variability is extreme.
Stable isotopic signatures, such as δ2H and δ18O, provide valuable insights into hydrological processes, including evaporation, precipitation, and mixing in complex tidal environments. The integration of ML models with isotopic data can enhance the accuracy of predictions by capturing the dynamic interactions within estuarine systems. Various ML models, including RF and Multivariate Adaptive Regression Spline (MARS), have been effectively used to reconstruct hydrological data, outperforming traditional methods like stage-discharge rating curves. These models have demonstrated reduced errors in predicting discharge, indicating their potential applicability in estuarine environments [28]. The use of stable isotopes in conjunction with machine learning can further refine predictions by providing continuous, high-resolution data on water residence times and mixing processes, which are critical for accurate inflow predictions [29]. While machine learning models and isotopic data offer promising tools for predicting freshwater inflows, challenges remain in accurately capturing the complex interactions in estuarine systems. Factors such as tidal influences, anthropogenic changes, and climatic variability can introduce uncertainties that require further refinement of models and methodologies. Nonetheless, the integration of these advanced techniques holds significant potential for enhancing our understanding and management of estuarine water resources.
In the above context, the present study provides the first combined application of stable isotopes and multiple machine learning models to estimate seasonal freshwater fluxes in the CBW estuary. The novelty of our work lies in comprehensive seasonal assessment to quantify freshwater contributions to CBW across Northeast Monsoon, Southwest Monsoon, and Premonsoon periods, using δ18O–salinity mass balance complemented by δ13C-based carbon cycling analysis. We used existing data from the Arabian Sea to establish seawater δ18O and salinity baselines and adopted seasonal rainwater composition from previous studies [30,31] for understanding the stable isotope mixing patterns. Furthermore, isotopic zonation of estuarine dynamics was studied using spatial field sampling to understand the effects of seawater dominance, terrestrial inputs, and agricultural runoff, with the aim of providing new insights into the biogeochemical stratification of the CBW estuary. In addition, with the aim of integrating ML techniques with isotope hydrology, we rigorously evaluated eight ML algorithms (RF, KNN, SVM, GBM, CART, GPR, ELM, RBNN) for predicting δ18O, δ13C, and salinity in the estuary. This is the first study in the Indian monsoon estuarine setting to systematically benchmark ML performance against isotope–salinity datasets. By doing so, this study not only addresses a critical knowledge gap in the estuarine isotope hydrology of the Indian subcontinent but also contributes a methodological advancement by coupling high-resolution geochemical tracers with state-of-the-art data-driven ML modeling approaches.

2. Materials and Methods

2.1. Study Area

Based on its dimensions, CBW is the largest estuarine system along the west coast of India and belongs to the Vembanad-Kol wetland system, one of the three Ramsar sites in Kerala that extends from Munambam (10°10′00″ N, 76°10′15″ E) in the north to Alappuzha (9°30′00″ N, 76°28′10″ E) in the south totalling 96.5 km in length (Figure 1). The width varies from 450 m to 4 km, and the depth ranges from 15 m at the Cochin Inlet to 3 m near the head, with an average depth of 1.5 m. The barrier spits are interrupted by tidal inlets at two locations: (i) Munambam Inlet in the north and (ii) Cochin Inlet in the middle.
The present study was conducted on CBW and Vembanad Lake, situated at the southern tip of the Indian subcontinent. This region experiences three distinct seasons: Premonsoon (PM, April–May), Southwest Monsoon (SWM, June–September), and Northeast Monsoon (NEM, October–December). During the monsoon season, the influx of freshwater significantly increases due to heavy precipitation, while the non-monsoon season (January–March) is characterized by reduced riverine input and dominant tidal forcing, leading to higher salinity levels in the estuary [14,32]. Water samples for δ18O and salinity analysis were collected fortnightly during high and low tides for a 1-year period between October and September of the subsequent year (Table 1). Fortnightly sampling during both high and low tides over a full annual cycle was essential to capture the temporal variability driven by monsoonal rainfall, tidal influences, and seasonal freshwater influx. This approach ensured high-resolution data reflecting both short-term tidal fluctuations and long-term seasonal shifts, which are critical for accurately characterizing estuarine mixing dynamics and validating isotope–salinity relationships. Additionally, a spatial survey covering 17 sites (Figure 1, Table 2) was conducted in the Premonsoon period (March–May) to document variability in isotopic composition and salinity. This spatial survey was conducted only in the Premonsoon period to capture the peak influence of seawater intrusion and evaporation-driven enrichment, which are most pronounced during the Premonsoon period when freshwater inflow is minimal. Documenting isotopic and salinity variability during this time provides critical insights into baseline estuarine dynamics before the onset of monsoonal dilution.

2.2. Sample Collection and Analysis

Surface water samples were collected in 50 mL HDPE bottles at a depth of 15 cm from the surface to avoid immediate surface film contaminants and stored until analysis. Sampling was conducted from the middle of the channel at each location to avoid bank effects. For the temporal study (Table 1), samples were collected fortnightly between the years 2009 and 2010 for one full year, at a single key station (Cochin Inlet, Station 1) during both high and low tides to capture the full range of tidal and seasonal variability at a strategic point of maximum seawater exchange. In total, 44 water samples were collected as shown in Table 1. For the spatial survey (Table 2), samples were collected from 17 stations/sampling points along the main axis of the estuary. It is important to note that this comprehensive spatial survey was conducted only during the Premonsoon period (March–May of 2015 for one year only). This period was specifically chosen to capture the baseline conditions of maximum seawater intrusion and salinity stratification before the onset of heavy monsoonal runoff. While this design provides a detailed spatial ‘snapshot’ under specific conditions, we acknowledge that it limits our ability to make direct spatial comparisons across all seasons, and that will be the focus of the future work.
Measurements of δ18O were conducted using the CO2-H2O equilibration method [33,34] on a Thermo Fisher MAT-253 isotope ratio mass spectrometer coupled with a GasBench II peripheral in a continuous flow mode, with a measurement reproducibility of 0.08‰. δD measurements were also carried out on the water samples with an aim to understand the dual water isotope behavior at the CBW estuary and compare with the well-established Meteoric Water Lines of the region. For δD analyses, samples were flushed with a He + H2 (2.5%) mixture and equilibrated for 70 min at 25 °C in the presence of catalytic platinum sticks. The obtained reproducibility using this method was 1.2‰. Water samples were analyzed immediately after collection using a conductivity probe (Orion, range 0.1 to 42, accuracy ±0.1) connected to a Thermo Scientific (Bremen, Germany) Orion 5-star multimeter. The conductivity probe was standardized with Orion conductivity standards (147 μS/cm, 1413 μS/cm, and 12.9 mS/cm). For δ13CDIC, water samples were collected in glass amber bottles with butyl rubber septa, treated with 1 mL of saturated HgCl2 solution to inhibit biological activity, and stored. δ13CDIC was measured by acidifying 2 mL of water with 0.5 mL of 100% orthophosphoric acid [35]. Standards, including NBS19 and MARJ1, were analyzed for calibration with a standard deviation of 0.09‰ for δ13C [36]. Rainfall data were obtained from the nearest weather station on Willingdon Island (latitude 9°57′14″ N, longitude 76°16′06″ E) (http://www.tutiempo.net, accessed on 12 June 2025) for comparison.

2.3. ML Methodology

To model and predict isotopic compositions and salinity, a suite of eight machine learning models was strategically selected to cover a range of learning approaches, from conventional algorithms to more advanced techniques. The rationale for this multimodel approach was to conduct a robust comparative analysis across diverse algorithmic families to assess their feasibility for this specific application. Given the relatively limited size of our dataset, we deliberately prioritized models that are less prone to overfitting and do not require extensive training data. Therefore, highly parameterized models such as Artificial Neural Networks (ANNs), which carry a greater risk of overfitting with smaller datasets, were excluded from this analysis in favor of a broader suite of benchmark and ensemble techniques. The selected models were chosen to represent a wide spectrum of learning strategies and are listed below:
  • Conventional Models: K-Nearest Neighbor (KNN), Random Forest (RF), and Sup-port Vector Machine (SVM) were chosen as they have been previously applied in stable isotope modeling studies and serve as established benchmarks [20].
  • Advanced Tree-Based Models: Gradient Boosting Machine (GBM) and Classification and Regression Tree (CART) were selected for their ability to capture complex non-linear interactions and provide insights into feature importance, which is crucial for understanding the environmental drivers.
  • Other Advanced Models: Gaussian Process Regression (GPR) was included for its strength in quantifying prediction uncertainty. Extreme Learning Machines (ELMs) and Radial Basis Function Networks (RBNNs) were tested for their efficiency and effectiveness in modeling spatial gradients and patterns with potentially limited data.
The foundation of our predictive modeling workflow was the careful curation and preprocessing of the dataset. We selected four key predictor variables—latitude, longitude, month, and tide phase—as proxies for the spatial and temporal drivers of estuarine dynamics. These features were used to predict three target variables: salinity, δ18O, and δ13C. Prior to model training, a rigorous preprocessing pipeline was implemented to handle the diverse nature of this data. Continuous variables were normalized using Min–Max scaling to prevent features with larger numeric ranges from disproportionately influencing model outcomes. Categorical data, such as the tide phase, were numerically encoded to be machine-readable. Furthermore, the single missing data point for δ18O was addressed using Gaussian Process Regression, a method chosen for its ability to leverage the strong underlying correlation with salinity. For complete methodological transparency and to facilitate the replication of our results, all variables, their roles, and these specific preprocessing steps are consolidated and detailed in Table 3.
The overall workflow is depicted in Figure 2 and Table 3. The input features selected for the models included temporal variables (e.g., month, tidal phase) and spatial variables (latitude, longitude), which serve as proxies for the underlying environmental gradients. Before training, missing values for δ18O in the spatial dataset (one station) were imputed using Gaussian Process Regression, leveraging the strong covariance observed between salinity and δ18O. All continuous variables were normalized to a common scale to ensure compatibility across models. The dataset was partitioned into training (60%), validation (20%), and test (20%) subsets. Model hyperparameters were tuned using a Bayesian optimization approach over 100 iterations with the objective of minimizing the Root Mean Square Error (RMSE). Robustness was ensured through k-fold cross-validation during the training process.
The data preprocessing comprised all continuous variables, including salinity and isotopic ratios. All the data were normalized to ensure compatibility across models, and the missing values were addressed using statistical imputation methods. Hyperparameters were tuned via Bayesian optimization (GPyOpt library) with 100 iterations, prioritizing minimization of hydrological prediction uncertainty. A tiered data imputation of δ18O values (Table 2) was reconstructed via Gaussian Process Regression using salinity-δ18O covariance kernels. The Model Training and Evaluation step included dividing the data into training (60%), validation (20%), and test (20%) subsets. K-fold cross-validation was employed to ensure robust model evaluation. The determination of performance metrics included Root Mean Square Error (RMSE) (Equation (1)), Mean Absolute Percentage Error (MAPE) (Equation (2)), and Coefficient of Determination (R2) (Equation (3)) measures. The equations used for the individual performance metric estimations are as below:
R M S E = 1 n i = 1 n y i ^ y i 2  
M A P E = 1 n i = 1 n | y i ^ y i |
R 2 = 1 y i y i ^ 2 y i y ¯ 2
where n is the total number of observation data points, yᵢ is the observed (measured) value for the i-th data point, ŷᵢ is the predicted value for the i-th data point generated by the model, ȳ is the mean of the observed values, (ŷᵢ − yᵢ)2 is the squared error between the predicted and observed values, and |ŷᵢ − yᵢ| is the absolute error between the predicted and observed values.
Feature analysis and interpretation was performed using GBM that provided feature importance rankings to identify key drivers of isotopic and salinity variations. GPR and CART enabled spatial and temporal classification of estuarine zones, revealing insights into monsoonal and non-monsoonal dynamics.

3. Results

3.1. Seasonal Variations in δ18O and Salinity

The seasonal δ18O and salinity data revealed significant oscillations in δ18O and salinity values, irrespective of tidal phases, as shown in Figure 3. During the Northeast Monsoon (NEM) period (October–December), δ18O values ranged from −4.69‰ to −1.15‰, with salinity values between 20 and 0.2. The Premonsoon (April–May) period exhibited δ18O values from −1.1‰ to −0.47‰, with corresponding salinity values from 21.6 to 4.9. In the Southwest Monsoon (SWM; gray region—Figure 3), δ18O reached its lowest value of −5‰ during high tide. The positive correlation between δ18O and salinity was evident across all seasons (Figure 3).

3.2. Spatial Variations in δ18O, δ13C, and Salinity

Spatial variations were pronounced, with δ18O values ranging from −1.75‰ to 3.31‰, and salinity values spanning from 0.1 to 28.0 across different stations (Table 2). Salinity and oxygen isotopic values varied significantly across the studied area, and three distinct spatial zones were evident. These were classified into Zone 1 (influenced by the Periyar River), with salinity ranging from 0.1 to 28.0, δ18O ranging between 0.7‰ and −0.6‰, and δ13C ranging from −2.9‰ to −10.5‰ (Figure 4); Zone 2 (influenced by other freshwater sources), with salinity ranging from 10.3 to 15.7, δ18O ranging from 2.2‰ to −0.5‰, and δ13C ranging from −7.7‰ to −11.9‰; and Zone 3 (Vembanad Lake), with salinity ranging from 0.1 to 3.4, δ18O ranging from 2.0‰ to −0.7‰, and δ13C ranging from −14.1‰ to −21.3‰ (Figure 4 and Table 2). The zones were classified following previous studies from the same estuary [15,19]. The results of the observed zonation are further discussed in Section 4.2 and Section 4.5.

3.3. Freshwater Flux in Comparison with Seasonal Rainfall

In an estuarine setting, the source of freshwater can vary, and each of these freshwater sources have unique δ18O values, although the salinity of these sources is minimal. Previously reported δ18O values of this region are −10‰ for the NEM season, −5‰ for the SWM season, and −2‰ for the Premonsoon season [37,38]. These values are close (−1.8 to −5‰) to the river water composition measured at a seasonal interval in the region [39]. Using these δ18O values as freshwater end members and the measured average δ18O of estuarine waters, the relative contribution of freshwater can be ascertained for both the high tide (HT) and low tide (LT) conditions using the below mass balance equations:
Fsw × δ18Osw + Frw × δ18Orw = δ18Oew
Fsw + Frw = 100
By rearranging Equations (4) and (5),
Percentage (%) = [δ18Oew − δ18Osw]/[δ18Orw − δ18Osw]
where sw, rw, and ew refer to seawater, rainwater, and estuarine water, respectively, and F indicates the flux parameter.
The region receives rainfall distributed across three distinct rainfall seasons and is characterized by unique isotopic ratios [37,38,40]. The monthly mean and daily rainfall data of the study location (Willingdon weather station, located within ~3 km of the site of monthly tidal surface water collection) indicated that in almost all years, rainfall is maximum from June to August, coinciding with the SW Monsoon (Figure 5—bottom panel). Based on the amount of rainfall and δ18O offset, the percentage contribution of freshwater was estimated for HT and LT sampling and is shown in Figure 5 (top panel). It was evident that freshwater flux into the CBW estuary is primarily driven by runoff, which is a major contributor during the SWM season but drops significantly during the Premonsoon and Postmonsoon periods. The water δ18O values of seasonal surface water were adapted from the tropical Chaliyar River basin to constrain the end member composition. The catchment area of this river is the Western Ghats region in India, which receives rainfall from the source that feeds the rivers flowing into the CBW [34]. The surface water δ18O values for the Chaliyar river were −1.8‰, −1.2‰, and −4.5‰ for the NEM, Premonsoon, and SWM seasons. respectively [39], and the corresponding freshwater percentages estimated using Equation (6) were 20–52% during the NEM season, 53–67% during the Premonsoon season, and 42–80% during the SWM season, similar to earlier reported studies [41].

3.4. Performance Metrics for ML Models

The performance metrics of all the used ML models are summarized in Table 4. The results revealed a stark difference in model predictability across the three target parameters (salinity, δ18O, and δ13C). For salinity prediction, the GBM model showed the highest accuracy, with the lowest RMSE (0.0993) and the highest R2 (0.9563, p-value: <0.001). For δ18O prediction, the KNN model performed best, achieving an RMSE of 0.1703, an R2 of 0.5039, and a MAPE of 29.87%. The RF model also showed acceptable performance (RMSE: 0.2101, R2: 0.2451, p-value: <0.001), while several other models, including SVM, GPR, and RBNN, failed entirely, producing negative R2 values. For δ13C prediction, models such as CART, ELM, and RBNN performed poorly, with negative R2 values suggesting overfitting or insufficient training data. A key advantage of tree-based models like GBM is their ability to provide insights into the drivers of their predictions through feature importance rankings. For our best performing model (GBM for salinity prediction), we analyzed these scores to identify the most influential hydrographic parameters. As illustrated in the bar chart in Figure 6, temporal factors were the dominant predictors. The month of the year, which serves as a proxy for the monsoon cycle, was the single most important feature, followed by the tidal phase. This finding aligns with our understanding of the estuary, where seasonal freshwater influx and tidal exchange are the primary drivers of salinity variation. Geographic location (latitude and longitude) also contributed to the prediction but was of secondary importance.
It is important to note that in models like GPR, CART, and RBNN negative R2 values were obtained, and this poor performance may be attributed to their sensitivity to small datasets and high feature noise, which often leads to overfitting or underfitting, as observed in our limited dataset. Additionally, certain input features may have limited predictive powers affecting the models more severely. A clearer analysis of feature importance and model complexity would provide valuable insights into these performance discrepancies. Furthermore, there exists a necessity for further dataset expansion and model refinement using additional ecological and geochemical variables.

4. Discussion

4.1. Seasonal and Spatial Variations in δ18O and Salinity

The salinity–δ18O relationship is a valuable tool for understanding the temporal mixing pattern of freshwater and seawater in estuaries around the world [42,43,44]. The surface-water salinity–oxygen isotopes of CBW showed two distinct trends defining the summer and winter composition. This large difference between salinity and water δ18O is due to the mixing of a variable amount of fresh water by the rivers draining the catchment and the saline Arabian Sea water. The rainwater δ18O during the monsoon season dominates the river water value, which varies between −1.0‰ and −3‰, while groundwater remains nearly invariant seasonally with a composition maximally ranging from −3.7 to −5.2‰ in the region [39]. The δ18O value of the seawater input into the estuary is isotopically heavier, with values between 1 and 0‰ in the Arabian Sea region. The difference between the salinity values measured during high and low tides was maximum during the summer season and minimum during the monsoon season. This implies that the extent of mixing of surface water is maximum in the monsoon season and minimum in the summer. The overall average empirical relationship between water δ18O and salinity for the CBW was δ18O = 0.1259 (Salinity) −3.31, with a strong regression coefficient value of 0.7 (Figure 7). One-sided ANOVA was carried out to understand the significance of the δ18O–salinity relationship, and the obtained statistical summary, indicating robustness of the regression lines, is given in Supplementary Table S1. Our observed δ18O–salinity relationship was similar to other monsoon estuaries in the Indian monsoon region, with a slightly lower slope for the Mandovi-Zuari [18] and Hooghly estuaries [13].

4.2. δ18O and δ13C Relationship with Salinity

Surface water salinity depends mainly on freshwater drainage, which carries a signature of rainwater during the wet period and groundwater during the dry period. The annual observed salinity varies from 21.7 to 0.1 (Table 1), with the lowest surface salinity coinciding with a period of the SWM, where precipitation and catchment runoff reach their maximum. Maximum observed salinity was in the Premonsoon period, wherein seawater intrusion was prominent. It was evident that the CBW region experienced diurnal and monthly tidal influences, which varied seasonally with the extent of seawater intrusion into the coastal area.
The tidal influence and sea water intrusion strongly stratify the CBW estuary, leading to the observed spatial zonation (Zone 1 to 3; Figure 4 left panel) observed in the measured hydrochemical parameters (Table 2) and presented in Section 3.2. In Zone 1, the region from the Cochin Inlet to Perumpalam Island, the influence of seawater from the Arabian Sea is pronounced with a δ18O–salinity relationship slope of 0.03 and a positive correlation (R2 = 0.37). Zone 2, from Perumpalam Island to Thanneermukkam Bund, exhibits the mixing of freshwater from riverine input and seawater from the Arabian Sea with a steep slope of 0.5 and a strong positive correlation (R2 = 0.6). Zone 3 is dominated by freshwater influx and evaporative effects, as seen in Figure 4 (left panel), and had a negative slope of −0.9 with insignificant correlation (R2 = 0.05). A summary table of various parameters in each of these zones is given in Supplementary Table S2. It is important to note that this classification of zones is spatial and not temporal. In addition, further studies with expanded sampling and statistical validation are needed to accurately delineate freshwater input zones and establish robustness.

4.3. δ18O–δD Relationship of CBW Estuary

A total of 17 surface water samples collected during the Premonsoon season were analyzed for δD and δ18O to understand the δ18O–δD relationship. The results are plotted in Figure 8 along with measurements from previously published studies in the region [14], in addition to both the Global Meteoric Water Line (GMWL) and Local Meteoric Water Line (LMWL) from the published literature [40]. In the previous study documenting the stable isotopic values of Vembanad Lake water samples, the reported values were −20.2 to +17.0‰ and −5.6 to +3.3‰ for δD and δ18O, respectively [14]. The most isotopically depleted values were observed during the monsoon season, which could be attributed to the amount effect. In contrast, an enrichment in isotopic values was observed during the Premonsoon period due to salinity mixing or evaporation. In our study, we carried out spatial sampling only during the Premonsoon season, and the δD vs. δ18O plot was characterized by enriched δ values similar to earlier observations, as shown in Figure 8 [14]. The ingression of saline seawater with enriched isotope values also contributed to overall isotope observations in the CBW estuary. The gradual decrease in δ18O and δD values observed within even a single Premonsoon season indicates either excess evaporation or an enhanced contribution of seawater, causing a compositional shift during the Premonsoon season.

4.4. Carbon Dynamics Using Salinity and δ13CDIC

Freshwater input to the estuary affects the net carbon isotopic composition of dissolved inorganic carbon, which is either derived from the degradation of organic matter present in the regional soil ecosystem or further modified due to water column productivity. Thus, δ13CDIC can act as a proxy for productivity in estuarine systems. However, the inorganic carbon reservoir in estuary water can vary across seasons, recording the flux of carbon due to silicate weathering [45]. The δ13CDIC observed during the present study ranged from −2.9‰ to −21.3‰, with an average δ13CDIC of −10.4‰. These obtained values are comparable with a previous study of δ13CDIC in the region [19]. Figure 4 (right panel) shows the conservative mixing curve of δ13CDIC, along with the respective actual values and salinity in the three different spatial zones that were identified and defined in this study for the CBW estuary (Figure 4, right panel). Among the five stations sampled in Zone 1, four stations were closest to the estimated conservative mixing line, indicating the role of the mixing of freshwater and seawater end members. The samples from Zone 2 were characterized by lower δ13CDIC values relative to the respective conservative mixing curve values (Figure 4, right panel). As observed in previous studies [19,45], the possible reason for the low δ13CDIC at these stations might be the dominant contribution from terrestrial inorganic carbon from the degradation of organics. The samples from Zone 3 were characterized by the lowest δ13CDIC relative to the respective conservative mixing curves. This indicates that the carbon dynamics in Zone 3 are dictated by processes like runoff from the nearby paddy fields with a lighter C3 signature (−25‰), which is further modified owing to the productivity of the estuary [19,46], resulting in δ13CDIC values ranging from −13 to −21‰. This is further evidenced by higher nutrients (nitrate, phosphate, silicate) in this zone during the monsoon and Postmonsoon season [47].

4.5. Evaluating ML Models for Salinity and Stable Isotopic Prediction

This study compared multiple machines learning models, including GBM, KNN, RF, SVM, CART, ELM, and RBNN, to predict salinity and isotopic ratios. The comparison of various machine learning models highlights the challenges and opportunities in predicting salinity and isotopic compositions [20,21]. The varying success of the machine learning models in predicting salinity, δ18O, and δ13C offers significant insights into not only the models themselves but also the fundamental nature of the processes governing each parameter. The GBM model emerged as the most reliable for salinity prediction, demonstrating the highest accuracy and predictive strength (Table 4). Salinity in the estuary, while dynamic, behaves conservatively and is a direct function of the mixing ratio between freshwater and seawater. This ratio is strongly governed by predictable spatial (latitude, longitude) and temporal (month, tidal phase) factors. The GBM model, with its tree-based architecture, excels at capturing this complex, non-linear, but deterministic relationship, making it an ideal tool for this parameter.
For δ18O prediction, the KNN model outperformed others, although overall performance metrics indicated room for improvement. δ18O variability is largely controlled by conservative mixing, but it is subject to additional, localized processes like evaporation. The KNN algorithm, which makes predictions based on the values of its ‘neighbors’, is well-suited to capturing such strong spatial autocorrelation. However, the fact that even the best model could only explain 50% of the variance highlights a key limitation: our input features do not account for stochastic factors like the variable isotopic composition of individual rainfall events, which introduces noise that the model cannot possibly resolve. The low performance of models like SVM for δ18O prediction suggests that certain ML approaches may not be well-suited for isotopic modeling without extensive parameter tuning and preprocessing (Table 4). The success of GBM/KNN models in salinity and δ18O prediction with acceptable correlation values and significant p-values (p-value: <0.001) could be attributed to the capturing of tidal hysteresis effects via lagged salinity features and evaporation enrichment through latitude–temperature interactions (feature importance > 0.85). This aligns with studies where tree-based models outperformed SVMs in monsoon-driven estuaries [21,27].
δ13C prediction faced significant challenges, with all tested models failing to produce a meaningful predictive relationship, as indicated by the negative R2 values (Table 4). A negative R2 indicates that the model’s predictions are worse than a simple baseline model that always predicts the mean of the target variable, signifying a complete failure of the model to generalize. This outcome is a critical finding of our study and is not merely an issue of dataset size, but rather a result of cascading issues stemming from both the data and the modeling process itself. The primary cause is the absence of key biogeochemical predictor variables in our dataset, leading to a very low signal-to-noise ratio. From a machine learning perspective, this leads to several technical failures:
  • Severe Overfitting: Flexible models, in an attempt to find a pattern, likely learned the random noise within the training data. This results in a model that is perfectly tuned to the training set but has no predictive power on unseen data.
  • Unsuitable Model Architecture: The complex and potentially discontinuous nature of δ13C dynamics may require more specialized model architectures than those tested. The ‘step-function’ discontinuities introduced by biological fractionation are particularly challenging for standard regression algorithms.
  • Hyperparameter Tuning Failure: In a low-signal environment, the Bayesian optimization process is susceptible to finding a ‘fluke’ set of hyperparameters that perform well on the validation set by chance but fail to capture any generalizable relationship, leading to the observed poor performance on the final test set.
Strong correlation between δ18O and salinity reinforces the utility of isotopes in tracing hydrochemical processes. Geographic parameters could also play a crucial role in model performance, indicating spatial influences on isotopic variations. The findings emphasize that while ML can be a powerful tool for hydrochemical predictions, robust data collection and model refinement are essential to improve generalizability and reliability across different environmental conditions. In addition, robust data collection, by means of future extended sampling campaigns spatially and temporally, and understanding tidal influences are essential to further constrain and understand the biogeochemical zone stratification discussed here.

5. Conclusions

This study provides the first combined application of stable isotopes and machine learning (ML) models to quantify seasonal freshwater inflows into the Cochin Backwater (CBW) estuary. The summary of the key findings of the study are as follows: (1) Seasonal Hydrography: The δ18O–salinity relationship (δ18O = 0.1259 × Salinity − 3.31; R2 = 0.7) clearly captured the seasonal freshwater–seawater mixing dynamics. Freshwater contributions were highest during the Southwest Monsoon (42–80%), moderate during the Pre-monsoon (53–67%), and lowest during the Northeast Monsoon (20–52%). (2) Spatial Estuarine Zonation: Spatial surveys during the Premonsoon identified three distinct hydro-biogeochemical zones: Zone 1: Dominated by seawater with heavier δ13C signatures; Zone 2: Mixed inputs from rivers and terrestrial sources; and Zone 3: Strong influence of paddy field runoff and estuarine productivity, marked by lighter δ13C values. (3) Carbon Dynamics: δ13CDIC analysis revealed clear spatial heterogeneity, reflecting the combined influence of seawater mixing, terrestrial carbon input, and agricultural runoff. (4) Among the eight ML models tested, Gradient Boosting Machine (GBM) showed the highest accuracy for salinity prediction (R2 = 0.96), while K-Nearest Neighbor (KNN) performed best for δ18O (R2 = 0.50). In contrast, δ13C predictions with ML models were unsuccessful due to the absence of key biogeochemical drivers in the dataset. Overall, this study demonstrates that coupling stable isotope geochemistry with ML models provides a powerful framework for quantifying seasonal freshwater fluxes and estuarine zonation in monsoon-regulated systems. The approach offers a scalable pathway for predictive modeling under changing climatic and anthropogenic pressures and can inform better management and conservation of tropical estuarine ecosystems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/hydrology12110277/s1, Table S1: Statistical Summary for δ18O and Salinity linear relationships; Table S2: Summary table showing Estuary Zonation.

Author Contributions

Conceptualization, P.K., P.G. and R.R.; methodology, H.R., P.K. and F.T.; writing—original draft preparation, P.K., R.R., P.G. and H.R.; writing—review and editing, R.R. and P.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

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

Acknowledgments

The authors are thankful for all the help obtained during the sampling and analysis throughout the project from the project members. RR is thankful to the research support provided through the grant CCEC01-1108-230167 from the Qatar Research Development and Innovation Council.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
CBWCochin Backwater
DICDissolved Inorganic Carbon
PMPremonsoon
SWMSouthwest Monsoon
NEMNortheast Monsoon
HTHigh Tide
LTLow Tide
NBS19National Bureau of Standards-19
HDPEHigh-Density Polyethylene
MLMachine Learning
ANNArtificial Neural Network
ANFISAdaptive Neuro-Fuzzy Inference System
SVMSupport Vector Machine
RBNNRadial Function Based Neural Network
RFRandom Forest
KNNK-Nearest Neighbor
GBMGradient Boosting Machine
GPRGaussian Process Regression
CARTClassification and Regression Tree
ELMExtreme Learning Machine
RMSERoot Mean Square Error
MAPEMean Absolute Percentage Error

References

  1. Day, J.W., Jr.; Kemp, W.M.; Yáñez-Arancibia, A.; Crump, B.C. (Eds.) Estuarine Ecology; John Wiley & Sons: Hoboken, NJ, USA, 2012. [Google Scholar]
  2. Dittmar, T.; Lara, R.J. Driving forces behind nutrient and organic matter dynamics in a mangrove tidal creek in North Brazil. Estuar. Coast. Shelf Sci. 2001, 52, 249–259. [Google Scholar] [CrossRef]
  3. Cloern, J.E.; Jassby, A.D. Patterns and scales of phytoplankton variability in estuarine–coastal ecosystems. Estuaries Coasts 2010, 33, 230–241. [Google Scholar]
  4. Vijith, V.; Sundar, D.; Shetye, S.R. Time-dependence of salinity in monsoonal estuaries. Estuar. Coast. Shelf Sci. 2009, 85, 601–608. [Google Scholar] [CrossRef]
  5. Qasim, S.; Gopinathan, C. Tidal cycle and the environmental features of Cochin Backwater (a tropical estuary). Proc. Indian Acad. Sci. Sect. A Part 3 Math. Sci. 1969, 69, 336–348. [Google Scholar]
  6. Haldar, R.; Khosa, R.; Gosain, A.K. Impact of Anthropogenic Interventions on the Vembanad Lake System. In Water Resources and Environmental Engineering I; Rathinasamy, M., Chandramouli, S., Phanindra, K., Mahesh, U., Eds.; Springer: Singapore, 2019. [Google Scholar]
  7. Kulk, G.; George, G.; Abdulaziz, A.; Menon, N.; Theenathayalan, V.; Jayaram, C.; Brewin, R.J.; Sathyendranath, S. Effect of Reduced Anthropogenic Activities on Water Quality in Lake Vembanad, India. Remote Sens. 2020, 13, 1631. [Google Scholar] [CrossRef]
  8. Pranav, P.; Roy, R.; Jayaram, C.; D’Costa, P.M.; Choudhury, S.B.; Menon, N.N.; Nagamani, P.; Sathyendranath, S.; Abdulaziz, A.; Sai, M.S.; et al. Seasonality in carbon chemistry of Cochin backwaters. Reg. Stud. Mar. Sci. 2021, 46, 101893. [Google Scholar] [CrossRef]
  9. Shivaprasad, A.; Vinita, J.; Revichandran, C.; Reny, P.D.; Deepak, M.P.; Muraleedharan, K.R.; Naveen Kumar, K.R. Seasonal stratification and property distributions in a tropical estuary (Cochin estuary, west coast, India). Hydrol. Earth Syst. Sci. 2013, 17, 187–199. [Google Scholar] [CrossRef]
  10. Balachandran, K.K.; Joseph, T.; Nair, K.K.C.; Nair, M.; Joseph, P.S. The Complex Estuarine Formation of Six Rivers (Cochin Backwaters System on West Coast of India)—Sources and Distribution of Trace Metals and Nutrients. 2002. Available online: https://drs.nio.org/drs/handle/2264/1340 (accessed on 30 August 2025).
  11. Sankaranarayanan, V.N.; Qasim, S.Z. Nutrients of the Cochin Backwater in relation to environmental characteristics. Mar. Biol. 1969, 2, 236–247. [Google Scholar] [CrossRef]
  12. Aneesh, T.D.; Srinivas, R.; Singh, A.T.; Resmi, T.R.; Nair, A.M.; Redkar, B.L. Stable water isotope signatures of dual monsoon precipitation: A case study of Greater Cochin region, south-west coast of India. J. Earth Syst. Sci. 2019, 128, 210. [Google Scholar] [CrossRef]
  13. Ghosh, P.; Chakrabarti, R.; Bhattacharya, S.K. Short- and long-term temporal variations in salinity and the oxygen, carbon and hydrogen isotopic compositions of the Hooghly Estuary water, India. Chem. Geol. 2013, 335, 118–127. [Google Scholar] [CrossRef]
  14. Nasir, U.P.; Harikumar, P.S. Hydrochemical and isotopic investigation of a tropical wetland system in the Indian sub-continent. Environ. Earth Sci. 2012, 66, 111–119. [Google Scholar]
  15. Bhavya, P.S.; Kumar, S.; Gupta, G.V.M.; Sudharma, K.V.; Sudheesh, V.; Dhanya, K.R. Carbon isotopic composition of suspended particulate matter and dissolved inorganic carbon in the Cochin estuary during post-monsoon. Curr. Sci. 2016, 110, 1539–1543. [Google Scholar]
  16. Raymond, P.A.; Bauer, J.E.; Caraco, N.F.; Cole, J.J.; Longworth, B.; Petsch, S.T. Controls on the variability of organic matter and dissolved inorganic carbon ages in northeast US rivers. Mar. Chem. 2004, 92, 353–366. [Google Scholar] [CrossRef]
  17. Bouillon, S.; Borges, A.V.; Castañeda-Moya, E.; Diele, K.; Dittmar, T.; Duke, N.C.; Kristensen, E.; Lee, S.Y.; Marchand, C.; Middelburg, J.J.; et al. Mangrove production and carbon sinks: A revision of global budget estimates. Glob. Biogeochem. Cycles 2008, 22, GB2013. [Google Scholar] [CrossRef]
  18. Ghosh, P.; Ramdas, L.; Banerjee, Y.; Thamizharasan, S.; Banerjee, S. Seasonal freshwater flux estimation using mollusc from the tropical Mandovi Zuari estuary, Goa, India. J. Earth Syst. Sci. 2021, 130, 107. [Google Scholar] [CrossRef]
  19. Bhavya, P.S.; Kumar, S.; Gupta, G.V.M.; Sudharma, K.V.; Sudheesh, V. Spatio-temporal variation in δ13C-DIC of a tropical eutrophic estuary (Cochin estuary, India) and adjacent Arabian Sea. Cont. Shelf Res. 2018, 153, 75–85. [Google Scholar] [CrossRef]
  20. Astray, G.; Soto, B.; Barreiro, E.; Gálvez, J.F.; Mejuto, J.C. Machine learning applied to the oxygen-18 isotopic composition, salinity and temperature/potential temperature in the Mediterranean Sea. Mathematics 2021, 9, 2523. [Google Scholar] [CrossRef]
  21. Cemek, B.; Arslan, H.; Küçüktopcu, E.; Simsek, H. Comparative analysis of machine learning techniques for estimating groundwater deuterium and oxygen-18 isotopes. Stoch. Environ. Res. Risk Assess. 2022, 36, 4271–4285. [Google Scholar] [CrossRef]
  22. Noori, N.; Kalin, L.; Isik, S. Water quality prediction using SWAT-ANN coupled approach. J. Hydrol. 2020, 590, 125220. [Google Scholar] [CrossRef]
  23. Ly, Q.V.; Nguyen, X.C.; Lê, N.C.; Truong, T.D.; Hoang, T.H.T.; Park, T.J.; Maqbool, T.; Pyo, J.-C.; Cho, K.H.; Lee, K.-S.; et al. Application of machine learning for eutrophication analysis and algal bloom prediction in an urban river: A 10-year study of the Han River, South Korea. Sci. Total Environ. 2021, 797, 149040. [Google Scholar]
  24. Samalavičius, V.; Gadeikienė, S.; Žaržojus, G.; Gadeikis, S.; Lekstutytė, I. Oxygen-18 prediction using machine learning in the Baltic Artesian Basin groundwater. Stoch. Environ. Res. Risk Assess. 2024, 39, 765–787. [Google Scholar] [CrossRef]
  25. Albuquerque, L.G.; de Oliveira Roque, F.; Valente-Neto, F.; Koroiva, R.; Buss, D.F.; Baptista, D.F.; Hepp, L.U.; Kuhlmann, M.L.; Sundar, S.; Covich, A.P.; et al. Large-scale prediction of tropical stream water quality using Rough Sets Theory. Ecol. Inform. 2021, 61, 101226. [Google Scholar] [CrossRef]
  26. Al Sudani, Z.A.; Salem, G.S.A. Evaporation rate prediction using advanced machine learning models: A comparative study. Adv. Meteorol. 2022, 2022, 1433835. [Google Scholar] [CrossRef]
  27. Dagher, D.H. Assessment of using machine and deep learning applications in surface water quantity and quality predictions: A review. J. Water Resour. Geosci. 2024, 3, 18–48. [Google Scholar]
  28. Van Binh, D. Reconstructing Daily Discharge in a Megadelta Using Machine Learning Techniques. Water Resour. Res. 2022, 58, e2021WR031048. [Google Scholar] [CrossRef]
  29. Downing, B.D.; Bergamaschi, B.A.; Kendall, C.; Kraus, T.E.C.; Dennis, K.J.; Carter, J.A.; von Dessonneck, T.S. Using Continuous Underway Isotope Measurements to Map Water Residence Time in Hydrodynamically Complex Tidal Environments. Environ. Sci. Technol. 2016, 50, 13387–13396. [Google Scholar] [CrossRef] [PubMed]
  30. Singh, A.; Jani, R.A.; Ramesh, R. Spatiotemporal variations of the δ18O-salinity relation in the northern Indian Ocean. Deep Sea Res. Part I Oceanogr. Res. Pap. 2010, 57, 1422–1431. [Google Scholar] [CrossRef]
  31. Deshpande, R.D.; Muraleedharan, P.M.; Singh, R.L.; Kumar, B.; Rao, M.S.; Dave, M.; Sivakumar, K.U.; Gupta, S.K. Spatio-temporal distributions of δ18O, δD and salinity in the Arabian Sea: Identifying processes and controls. Mar. Chem. 2013, 157, 144–161. [Google Scholar] [CrossRef]
  32. Rasheed, K.; Joseph, K.A.; Balchand, A.N. Impacts of harbour dredging on the coastal shoreline features around Cochin. In Proceedings of the International Conference on Coastal Change, Bordeaux, France, 10–16 February 1995; pp. 943–948. [Google Scholar]
  33. Epstein, S.; Mayeda, T. Variation of O18 content of waters from natural sources. Geochim. Cosmochim. Acta 1953, 4, 213–227. [Google Scholar] [CrossRef]
  34. Rangarajan, R.; Ghosh, P. Role of water contamination within the GC column of a GasBench II peripheral on the reproducibility of 18O/16O ratios in water samples. Isot. Environ. Health Stud. 2011, 47, 498–511. [Google Scholar] [CrossRef]
  35. Assayag, N.; Rivé, K.; Ader, M.; Jézéquel, D.; Agrinier, P. Improved method for isotopic and quantitative analysis of dissolved inorganic carbon in natural water samples. Rapid Commun. Mass Spectrom. 2006, 20, 2243–2251. [Google Scholar] [CrossRef]
  36. Rangarajan, R.; Pathak, P.; Banerjee, S.; Ghosh, P. Floating boat method for carbonate stable isotopic ratio determination in a GasBench II peripheral. Rapid Commun. Mass Spectrom. 2021, 35, e9115. [Google Scholar]
  37. Lekshmy, P.R.; Midhun, M.; Ramesh, R.; Jani, R.A. 18O depletion in monsoon rain relates to large-scale organized convection rather than the amount of rainfall. Sci. Rep. 2014, 4, 5661. [Google Scholar] [CrossRef]
  38. Rahul, P.; Ghosh, P. Long term observations on stable isotope ratios in rainwater samples from twin stations over Southern India; identifying the role of amount effect, moisture source and rainout during the dual monsoons. Clim. Dyn. 2019, 52, 6893–6907. [Google Scholar] [CrossRef]
  39. Hameed, A.S.; Resmi, T.R.; Suraj, S.; Warrier, C.U.; Sudheesh, M.; Deshpande, R.D. Isotopic characterization and mass balance reveals groundwater recharge pattern in Chaliyar river basin, Kerala, India. J. Hydrol. Reg. Stud. 2015, 4, 48–58. [Google Scholar] [CrossRef]
  40. Warrier, C.U.; Babu, M.P.; Manjula, P.; Velayudhan, K.T.; Hameed, A.S.; Vasu, K. Isotopic characterization of dual monsoon precipitation—Evidence from Kerala, India. Curr. Sci. 2010, 99, 1487–1495. [Google Scholar]
  41. Seena, G.; Muraleedharan, K.R.; Revichandran, C.; Abdul Azeez, S.; John, S. Seasonal spreading and transport of buoyant plumes in the shelf off Kochi, southwest coast of India—A modeling approach. Sci. Rep. 2019, 9, 19956. [Google Scholar] [CrossRef] [PubMed]
  42. Torgersen, T. Isotopic composition of river runoff on the US East Coast: Evaluation of stable isotope versus salinity plots for coastal water mass identification. J. Geophys. Res. Ocean 1979, 84, 3773–3775. [Google Scholar] [CrossRef]
  43. Fairbanks, R.G.; Sverdlove, M.; Free, R.; Wiebe, P.H.; Bé, A.W. Vertical distribution and isotopic fractionation of living planktonic foraminifera from the Panama B—Asin. Nature 1982, 298, 841. [Google Scholar] [CrossRef]
  44. Khim, B.-K.; Park, B.-K.; Yoon, H.-I. Oxygen isotopic compositions of seawater in the Maxwell Bay of King George Island, west Antarctica. Geosci. J. 1997, 1, 115. [Google Scholar] [CrossRef]
  45. Samanta, S.; Dalai, T.K.; Tiwari, S.K.; Rai, S.K. Quantification of source contributions to the water budgets of the Ganga (Hooghly) River estuary, India. Mar. Chem. 2018, 207, 42–54. [Google Scholar] [CrossRef]
  46. Kaushal, R.; Ghosh, P.; Geilmann, H. Fingerprinting environmental conditions and related stress using stable isotopic composition of rice (Oryza sativa L.) grain organic matter. Ecol. Indic. 2016, 61, 941–951. [Google Scholar] [CrossRef]
  47. Nagarathinam, A.; Retnamma, J.; Loganathan, J.; Singaram, P.; Madam, S.M.K.; Jose, A.K.; Subramanian, P.R. Implications of an extensive salt water barrage on the distribution of black clam in a tropical estuarine system, Southwest coast of India. Oceanologia 2021, 63, 343–355. [Google Scholar] [CrossRef]
Figure 1. Map of Cochin Backwater with sampling locations. The temporal sampling location is denoted by a filled circle (samples in Table 1), whereas hollow circles represent the locations of spatial sampling (samples in Table 2).
Figure 1. Map of Cochin Backwater with sampling locations. The temporal sampling location is denoted by a filled circle (samples in Table 1), whereas hollow circles represent the locations of spatial sampling (samples in Table 2).
Hydrology 12 00277 g001
Figure 2. Workflow of advanced ML techniques that were employed to estimate and predict δ18O and δ13C isotopic compositions and salinity in the estuary.
Figure 2. Workflow of advanced ML techniques that were employed to estimate and predict δ18O and δ13C isotopic compositions and salinity in the estuary.
Hydrology 12 00277 g002
Figure 3. The top panel shows high and low tide water δ18O, while the bottom panel shows salinity data of the collected water samples. Filled and hollow squares represent δ18O values of high and low tide, respectively. Filled and hollow diamonds represent salinity values of high and low tide, respectively (Green Region—NEM; Light Pink—PM; Gray—SWM).
Figure 3. The top panel shows high and low tide water δ18O, while the bottom panel shows salinity data of the collected water samples. Filled and hollow squares represent δ18O values of high and low tide, respectively. Filled and hollow diamonds represent salinity values of high and low tide, respectively (Green Region—NEM; Light Pink—PM; Gray—SWM).
Hydrology 12 00277 g003
Figure 4. (a) δ18O–salinity relationship indicating three different spatial zones for collected Premonsoon samples depicted by hollow red squares, green triangles, and purple diamonds representing δ18O and salinity values from Zones 1, 2, and 3, respectively. (b) δ13C–salinity relationship of the spatial zones along with the conservative mixing curve. Hollow red squares, green triangles, and purple diamonds represent δ13C and salinity values from Zones 1, 2, and 3, respectively, while hollow black circles and the corresponding dotted line represent the conservative mixing curve. The grey circles in both (a,b) refers to the identified zonation’s in the estuary.
Figure 4. (a) δ18O–salinity relationship indicating three different spatial zones for collected Premonsoon samples depicted by hollow red squares, green triangles, and purple diamonds representing δ18O and salinity values from Zones 1, 2, and 3, respectively. (b) δ13C–salinity relationship of the spatial zones along with the conservative mixing curve. Hollow red squares, green triangles, and purple diamonds represent δ13C and salinity values from Zones 1, 2, and 3, respectively, while hollow black circles and the corresponding dotted line represent the conservative mixing curve. The grey circles in both (a,b) refers to the identified zonation’s in the estuary.
Hydrology 12 00277 g004
Figure 5. (a) Relative contribution of freshwater to the Cochin Backwater at different times of the year calculated using the δ18O of the freshwater end member and seawater. Red and blue squares represent the percentage of rainwater at low and high tide, respectively. (b) Monthly (continuous and broken black line) and daily (gray) mean rainfall in mm.
Figure 5. (a) Relative contribution of freshwater to the Cochin Backwater at different times of the year calculated using the δ18O of the freshwater end member and seawater. Red and blue squares represent the percentage of rainwater at low and high tide, respectively. (b) Monthly (continuous and broken black line) and daily (gray) mean rainfall in mm.
Hydrology 12 00277 g005
Figure 6. Feature importance for salinity prediction. The relative importance of predictor variables as determined by the Gradient Boosting Machine (GBM) model. Importance is calculated as the mean decrease in impurity; higher values indicate greater importance.
Figure 6. Feature importance for salinity prediction. The relative importance of predictor variables as determined by the Gradient Boosting Machine (GBM) model. Importance is calculated as the mean decrease in impurity; higher values indicate greater importance.
Hydrology 12 00277 g006
Figure 7. Correlation plot between δ18O and salinity in the collected CBW samples indicating the mixing between freshwater and saline water (VSMOW—Vienna Standard Mean Ocean Water). Hollow green and red squares represent δ18O and salinity values of high and low tide, respectively, whereas filled gray circles represent the average value.
Figure 7. Correlation plot between δ18O and salinity in the collected CBW samples indicating the mixing between freshwater and saline water (VSMOW—Vienna Standard Mean Ocean Water). Hollow green and red squares represent δ18O and salinity values of high and low tide, respectively, whereas filled gray circles represent the average value.
Hydrology 12 00277 g007
Figure 8. Comparison of the δ18O–δD regression line based on 18 Premonsoon surface water samples (Table 1) collected and compared with the Global Meteoric Water Line (GMWL) and Local Meteoric Water Line (LMWL) from published literature (Nasir and Harikumar, 2012 [14] and Warrier et al., 2010 [41]).
Figure 8. Comparison of the δ18O–δD regression line based on 18 Premonsoon surface water samples (Table 1) collected and compared with the Global Meteoric Water Line (GMWL) and Local Meteoric Water Line (LMWL) from published literature (Nasir and Harikumar, 2012 [14] and Warrier et al., 2010 [41]).
Hydrology 12 00277 g008
Table 1. Measured salinity and oxygen isotopic composition of Cochin Backwater collected biweekly during both low and high tides at sample site 1 as indicated in Figure 1.
Table 1. Measured salinity and oxygen isotopic composition of Cochin Backwater collected biweekly during both low and high tides at sample site 1 as indicated in Figure 1.
Date of Sample CollectionSeasonsδ18OVSMOW (‰)Salinity
High TideLow TideHigh TideLow Tide
4-OctoberNortheast Monsoon−2.84 0.2
18-October−1.87−1.98.55.1
2-November−1.19−3.4912.611.1
16-November−4.69−4.211.91.3
2-December−2.44−311.57.1
16-December−1.15−2.2219.412.3
31-December−1.15−1.542016.2
15-January−0.96−1.8619.617.8
30-January−1.23−0.721.118
13-FebruaryPremonsoon−0.41−0.7421.717.1
28-February−0.3−1.2420.518.6
15-March−0.58−0.7818.518.1
15-April−0.75−1.120.516.6
14-May−0.72−0.4721.620.3
27-May−1.01−1.097.84.9
14-JuneSouthwest Monsoon−3.72−3.690.20.1
27-June−3.2−3.290.20.2
11-July−2.83−2.6812.6
26-July−2.66−2.582.50.5
11-August−2.26−2.22.84
26-August−2.74−2.580.70.7
8-September−2.01−2.4210.22.9
25-September−5.01 0.1
Table 2. Hydrological parameters of Cochin Backwater collected during the Premonsoon season.
Table 2. Hydrological parameters of Cochin Backwater collected during the Premonsoon season.
Sl. NoLocation Lat (oN)Long (oE)Salinity (PSU)δ18O
(‰ VSMOW)
δ13CDIC
(‰ VPDB)
1Thevara ferry9.92676.3040.100.42−5.60
2Panangad9.88376.33118.800.54−5.64
3Arror9.88276.30717.600.89−10.52
4Kudapuram (Eramallor)9.82976.32015.70NA−9.01
5Kodamthuruthu (Kuthiathodu)9.80376.32613.102.49−9.60
6Thykkatusherry9.77376.33111.703.31−11.10
7AVyalar9.71876.34510.401.52−9.62
7BVyalar9.71876.35010.400.43−8.58
8Punnamada9.50876.3532.100.50−14.09
9Aaryad9.54476.3530.101.76−17.23
10Pallathuserry9.56376.3560.100.53−21.34
11Muhamma9.60476.3623.401.23−17.03
12Thalayazham (Puthanpalam)9.69276.41310.402.00−11.97
13Vaikom9.74976.38911.603.06−9.33
14Kulasekaramagalam (Mekara)9.79976.37911.900.78−8.04
15Punnakkaveli (South Paravoor)9.85576.37912.452.45−7.74
16Chavakakadavuamera (Udayamperoor)9.89476.36316.401.59−7.20
17Fort Kochi9.96876.24428.00−1.75−2.90
Table 3. Description and preprocessing of features for machine learning models.
Table 3. Description and preprocessing of features for machine learning models.
Feature NameRoleTypeUnit/DescriptionPreprocessing Steps
LatitudePredictorContinuousDegrees North (°N)Normalized to a range using Min–Max scaling
LongitudePredictorContinuousDegrees East (°E)Normalized to a range using Min–Max scaling
MonthPredictorContinuousNumerical representation of the month (1 = January, …, 12 = December).Normalized to a range using Min–Max scaling
Tide PhasePredictorCategoricalThe tidal state during sampling (‘High Tide’ or ‘Low Tide’).Converted to a binary numerical format: High Tide = 1, Low Tide = 0
SalinityTargetContinuousPractical Salinity Units (PSU)Normalized to a range using Min–Max scaling
δ18OTargetContinuousIsotopic ratio relative to Vienna Standard Mean Ocean Water (‰ VSMOW).Normalized to a range using Min–Max scaling. The single missing value from the spatial dataset was imputed via Gaussian Process Regression, using salinity as a covariance kernel
δ13CTargetContinuousIsotopic ratio relative to Vienna Pee Dee Belemnite (‰ VPDB).Normalized to a range using Min–Max scaling
Table 4. Performance metrics of all ML models used in this study and the target parameters modeled.
Table 4. Performance metrics of all ML models used in this study and the target parameters modeled.
ModelTargetRMSER2MAPE (%)T-Test (p-Value)
GBMSalinity0.09930.9563N/A<0.001
GPRδ18O0.6298−5.7860N/A0.045
CARTδ13C0.3449−2.0460N/A0.089
ELMδ18O0.9187−13.440N/A0.103
ELMδ13C0.7626−13.890N/A0.097
RBNNδ18O0.2869−0.4080N/A<0.001
RBNNδ13C0.2626−0.7660N/A<0.001
RFδ18O0.21010.245136.19<0.001
RFδ13C0.2489−0.586934.900.032
SVMδ18O0.2500−0.069539.160.071
SVMδ13C0.2556−0.672225.880.089
KNNδ18O0.17030.503929.87<0.001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

K., P.; Rangarajan, R.; Thabit, F.; Ghosh, P.; Rahman, H. Seasonal Freshwater Inflows in Cochin Backwater Estuary Inferred from Stable Isotopes and Machine Learning. Hydrology 2025, 12, 277. https://doi.org/10.3390/hydrology12110277

AMA Style

K. P, Rangarajan R, Thabit F, Ghosh P, Rahman H. Seasonal Freshwater Inflows in Cochin Backwater Estuary Inferred from Stable Isotopes and Machine Learning. Hydrology. 2025; 12(11):277. https://doi.org/10.3390/hydrology12110277

Chicago/Turabian Style

K., Prasanna, Ravi Rangarajan, Fursan Thabit, Prosenjit Ghosh, and Habeeb Rahman. 2025. "Seasonal Freshwater Inflows in Cochin Backwater Estuary Inferred from Stable Isotopes and Machine Learning" Hydrology 12, no. 11: 277. https://doi.org/10.3390/hydrology12110277

APA Style

K., P., Rangarajan, R., Thabit, F., Ghosh, P., & Rahman, H. (2025). Seasonal Freshwater Inflows in Cochin Backwater Estuary Inferred from Stable Isotopes and Machine Learning. Hydrology, 12(11), 277. https://doi.org/10.3390/hydrology12110277

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