Next Article in Journal
Mapping and Assessing Groundwater Quality in Bourgogne-Franche-Comté (France): Toward Optimized Monitoring and Management of Groundwater Resource
Next Article in Special Issue
Integrating Spatial Concerns into Water Reuse Regulations: Insights from the European Union and the Iberian Peninsula
Previous Article in Journal
Bedload Dynamics in a Partially Glaciated Catchment: Insights from over One Decade of Measuring Bedload Transport Processes and Future Perspectives Under Climate Change
Previous Article in Special Issue
Improved Oil/Water Separation by Employing Packed-Bed Filtration of Modified Quartz Particles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Biomimicry-Inspired Automated Machine Learning Fit-for-Purpose Wastewater Treatment for Sustainable Water Reuse

1
Department of Learning, Informatics, Management and Ethics, Karolinska Institutet, 171 77 Stockholm, Sweden
2
MLV Research Group, Department of Informatics, Democritus University of Thrace, 65404 Kavala, Greece
3
College of Pharmacy, Auburn University Harrison, Auburn, AL 36849, USA
4
College of Agricultural, Consumer and Environmental Sciences, University of Illinois Urbana-Champaign, Urbana, IL 61801, USA
5
Mayo Clinic Artificial Intelligence & Discovery, Rochester, MN 55905, USA
6
Department of Product and Systems Design Engineering, University of Aegean, 82132 Chios, Greece
*
Author to whom correspondence should be addressed.
Water 2025, 17(9), 1395; https://doi.org/10.3390/w17091395
Submission received: 8 April 2025 / Revised: 28 April 2025 / Accepted: 5 May 2025 / Published: 6 May 2025

Abstract

:
The growing global freshwater scarcity urgently requires innovative wastewater treatment technologies. This study hypothesized that biomimicry-inspired automated machine learning (AML) could effectively manage wastewater variability through adaptive processing techniques. Utilizing decentralized swarm intelligence, specifically the Respected Parametric Insecta Swarm (RPIS), the system demonstrated robust adaptability to fluctuating influent conditions, maintaining stable effluent quality without centralized control. Bio-inspired oscillatory control algorithms maintained stability under dynamic influent scenarios, while adaptive sensor feedback enhanced real-time responsiveness. Machine learning (ML) methods inspired by biological morphological evolution accurately classified influent characteristics (F1 score of 0.91), optimizing resource allocation dynamically. Significant reductions were observed, with chemical consumption decreasing by approximately 11% and additional energy usage declining by 14%. Furthermore, bio-inspired membranes with selective permeability substantially reduced fouling, maintaining minimal fouling for up to 30 days. Polynomial chaos expansions efficiently approximated complex nonlinear interactions, reducing computational overhead by approximately 35% through parallel processing. Decentralized swarm algorithms allowed the rapid recalibration of system parameters, achieving stable pathogen removal and maintaining effluent turbidity near 3.2 NTU (Nephelometric Turbidity Units), with total suspended solids consistently below 8 mg/L. Integrating biomimicry with AML thus significantly advances sustainable wastewater reclamation practices, offering quantifiable improvements critical for resource-efficient water management.

1. Introduction

Biomimicry-inspired automated machine learning (AML) emerges as a promising solution for sustainable fit-for-purpose wastewater treatment, responding effectively to the rising demands of freshwater scarcity [1]. This innovative approach addresses variable wastewater qualities through adaptive, bio-inspired methodologies. Leveraging decentralized biomimetic swarm intelligence, systems such as the Respected Parametric Insecta Swarm (RPIS) successfully mimic natural biological processes [2,3]. A methodical approach is utilized by RPIS through local agent interactions subject to consistent rules. Behavioral formulation is achieved through iterative processes, which optimize resource allocation while preserving adaptation. After initial observation, a shift in collective decision-making is triggered, generating a fresh transformation inside agent patterns. Implementation remains decentralized, with no supervisory elements controlling progression. A stable configuration is maintained through constant adjustments, considering fluctuations in agent positions. Communication among participants secures multi-path exploration that refines search efficiency for data-intensive tasks. Swarm intelligence algorithms utilize local feedback loops, facilitating rapid adjustments to influent variations without central oversight. Natural oscillatory feedback mechanisms inspire control algorithms, optimizing process stability and efficiency. Furthermore, accretion phenomena observed in biological systems inform the development of biodynamic wastewater infrastructure, leading to durable and resilient architectures. Layered system designs derived from biological analogs substantially mitigate ecological impacts and operational burdens by reducing resource consumption. Adaptive feedback from sensor data streams ensures stable effluent quality, significantly decreasing operational ambiguity [4,5]. Swarm Intelligence utilizes adaptive feedback mechanisms conveying Incremental adjustments based on sensor data streams. Informational completeness achieved through decentralized agent interactions restrains system instability. Imperative adaptive methodologies like Particle Swarm Optimization (PSO) enable dynamic responsiveness by adjusting trajectories incrementally to optimize real-time sensor feedback integration, notably within IoT and Big Data preprocessing contexts [6]. This adaptive capacity mitigates suboptimal convergence, ensuring sustained efficiency amidst fluctuating environmental parameters and multi-objective constraints inherent to sensor-driven decentralized systems.
Furthermore, the widespread application of biomimicry principles has encouraged new methods in water processing. Machine learning (ML) routines inspired by morphological evolution display efficient classification within real-time monitoring systems [7]. Biomimetic scheduling modules allocate energy inputs based on diurnal fluctuations, ensuring resource conservation [8]. This technique has been termed a prime example of Bio-inspired engineering, given its resemblance to self-regulation in biological entities. Specialized modules integrate ephemeral data into predictive frameworks, thereby adjusting retention times. That approach yields minimized chemical usage across diverse feed compositions while meeting local regulatory mandates.
Fit-for-purpose wastewater treatment has been facilitated through targeted adjustments at each processing stage [9,10]. Automated ML algorithms evaluate sensor datasets, producing iterative modifications that align effluent characteristics with intended reuse goals. A direct correlation between morphological patterns coupled with removal efficiency has been observed, proving the influence of biomimicry in system design. Real-time analytics optimize parameter thresholds, circumventing excessive resource usage. Reuse strategies benefit from load-specific modules, tailored to irrigation demand followed by cooling requirements, among additional end uses. Biomimetic prototypes, equipped with self-healing features, present extended durability under fluctuating conditions. Furthermore, pilot investigations underscore consistent pathogen inactivation rates, confirming resilience.
On the other side, the resilience of RPIS attributes is apparent, though reliable comparative data remain absent. Previous evaluations [11] documented partial metrics, yet an accumulated perspective was not supplied. After initial phases of accumulation, distinct properties were recognized, with expected accretion-shaping performance. Then, a systematic approach to swarm-based processes [12] revealed that comparable optimization frameworks, including established Particle Swarm variations, sometimes surpass unrefined RPIS. Meanwhile, distinct cohesion elements have been explored by preliminary research, although findings remain scattered. Observers hypothesize that the continuous expansions of parameter sets, arranged through prisma-based data structuring, might elevate RPIS reliability [13]. Subsequently, enlarging agent-based interactions has been investigated to advance differentiation across solution spaces, fostering a new compilation of adaptive properties [14]. Experimental outcomes remain preliminary, so a direct contrast with methods from ABC, or BFO, families is not yet confirmed.
The confluence of scientific disciplines, including process engineering, ecology, and data analytics, has shaped these innovations. Discrete optimization protocols enhance microbial stability, culminating in multifarious breakdown pathways. By harnessing biomimicry themes, next-generation modules feature oscillatory nutrient dosing, promoting natural attenuation [15]. That configuration underscores biodynamic cooperation between microbial consortia alongside advanced sensor loops. Bio-inspired membranes enable selective permeability, yielding controlled flow rates with minimal fouling. Adaption strategies enable ongoing evolution, counteracting operational drift. Accretion-based layering, utilized in reactor scaffolds, preserves robust architecture. Ambiguity within data streams is mitigated by adaptive algorithms, securing reliable outputs.
Oscillatory nutrient dosing differs from conventional nutrient strategies in fundamental physiological responsiveness and integration pathways [16,17]. Oscillatory dosing, also termed pulsatile dosing, entails sequential nutrient pulses, creating periodic fluctuations that mimic biological rhythms and thus influencing selective gene expression and metabolic transformation differently than continuous dosing [18]. Conventional methods maintain stable nutrient symmetry through continuous administration, providing uniform availability without temporal separation or bifurcation [19]. This temporal scattering inherent in oscillatory dosing induces complex regulatory feedback mechanisms in biological systems, where gene expressions initially decrease, followed by significant overshoots before attaining steady-state equilibrium [17]. Conversely, conventional nutrient strategies typically lack this dynamic vertex of responsiveness, offering simpler but less adaptive nutrient integration [20,21]. Additionally, oscillatory dosing allows certain regulatory pathways to selectively activate different messenger RNA subsets, whereas conventional dosing exhibits less differentiation in downstream biological responsiveness [18]. Thus, the formulated proposition emerges clearly: Fundamental physiological differences between these interconnected dosing methodologies manifest prominently at the division of nutrient delivery timing and biological feedback responsiveness [17,18].
On the other hand, AML offers numerous advantages pertaining to wastewater treatment processes, significantly enhancing sustainable water reuse practices. By integrating advanced algorithms, AML facilitates reactive adjustments in operational parameters, enabling real-time optimization based on dynamic measurement data [22]. This process ensures consistent effluent quality, vital for achieving sustainable reuse standards. Partitioned algorithms within AML systems elucidate intricate relationships between diverse variables, such as pollutant concentrations and treatment efficiencies, thereby providing deeper contemplative insights for operational improvements [23,24]. Tending toward operational excellence, AML reduces chemical and energy consumption through precise control and forecasting, lowering the overall carbon footprint of treatment facilities [25,26]. Moreover, predictive maintenance enabled by AML anticipates equipment degradation before failures occur, ensuring that exterior mechanical components maintain optimal functionality over extended periods [27]. The scalability offered by automated systems allows adaptive application across various plant sizes and configurations without extensive re-engineering. Furthermore, AML could facilitate efficient resource recovery from wastewater streams, aligning with circular economy principles by identifying valuable nutrients or energy sources suitable for reuse [28]. These technologies enhance regulatory compliance through improved pollutant removal predictions, ensuring that the effluent consistently meets stringent environmental standards.
Irrigation reuse demands TDSs (Total Dissolved Solids) to be maintained at sufficiently low levels, while BOD (Biochemical Oxygen Demand) remains restricted to moderate values. Linked data streams regulate chemical dosing accordingly. Industrial settings require TSSs (Total Suspended Solids) to be minimized, with pH levels near neutral ranges. What is more, specific constraints vary across operational scenarios. Hence, AML dynamically readjusts process variables. Cooling processes benefit from reduced turbidity as measured in NTU (Nephelometric Turbidity Units), and the total hardness should align with recommended guidelines. Additionally, continuous real-time feedback refines those setpoints. Likewise, aquifer recharge necessitates nitrate levels to be within acceptably safe ranges, alongside the near-absence of coliform organisms [13,29]. Different treatment environments prompt specialized calibration.
AML has also been adapted to wastewater treatment, significantly reducing chemical and energy consumption through precise control and forecasting [30,31]. Confined within operational boundaries, AML facilitates the modulation of chemical dosing and energy input, responding dynamically to real-time variations. Forming hierarchical control structures—such as two-layer systems comprising distinct parts for trajectory generation and tracking—enables consistent reductions in energy usage, with savings of up to 7.86% demonstrated experimentally [32]. Correlative analysis using advanced neural network models, including centripetal multi-modal VAR-LSTM architectures, further refines carbon footprint predictions, optimizing operational efficiency through precise resource allocation [31,33].
Fifty distinct sets were simulated with varying load intensities. The predicted effluent turbidity was 3.2 NTU, achieving 93% removal. Notably, AML results showed consistent TSSs below 8 mg/L, with 96% reliability.
The frequencies of pH fluctuations remained within ± 0.2 for 87% of trials. The role of ML algorithms was vaguely described, yet they secured a 0.91 F1 score, further indicating robust classification features. Among tested sequences, the valuation of reaction stability was primary, verified using 2.3% standard deviation. Additional energy consumption dropped by 14%. Nonetheless, sensor-based feedback improved chemical usage efficiency, specifically reducing polydispersant demands by 11%, indicating minimal membrane fouling at day 30. This procedure balanced partial nitrification rates under variable hydraulic loading. Minimal effluent coliform was detected below 2 MPN/100 mL.
Specifically, Table 1 reports stable TSSs below 10 mg/L. Among the tested runs, pH variations remained within acceptable frequencies. Nonetheless, notable, the membrane fouling incidence rose under higher loads. The valuation of ML-driven classification revealed stable morphological features. A moderate shift occurs here, imposing new phrasing and indicating improved performances at medium loads, with turbidity values near 3.2 NTU, signifying primary reliability. Additional benefits included steady nitrification, suggesting feasible adaptation.

1.1. Previous Work

A vision for sustainable water reuse has been established through Biomimicry-Inspired Automated ML Fit-for-Purpose Wastewater Treatment. Multiple microbial collectives were integrated with syntrophic bacterial clusters to mimic ruminant digestion, facilitating methane yield increments of nearly 85% via real-time process optimization [34]. Interfacing with reinforcement learning was adopted to alter aeration protocols based on feedback signals, enabling energy consumption reductions by approximately 25%, with effluent ammonia consistently below 0.5 mg/L. An intermittent dynamic was created through dissolved oxygen measurements together with nitrogen tracking, replicating self-regulating loops found in natural water bodies [35]. Predictive methods were constructed to identify fouling events in membrane bioreactors, providing early warnings 48 h beforehand, with accuracy reaching 92%. Forecasting was performed by supervised learning trained on historical observations, prompting proactive backwashing that extended the membrane’s lifespan by almost 30%. These mechanisms represent evolutionary progress within the scope of advanced water treatment. The elementary discussion pertaining to biomimetic scheduling and chemical usage reduction currently lacks distinct connections to actual wastewater treatment processes. Clarification regarding practical implementation within real-world operational properties is necessary. Abecedarian descriptions of biomimetic methods need explicit correlations with treatment processes to establish bimodal applicability. Integrating biomimetic principles into existing infrastructure presents notable challenges requiring detailed adaptation strategies [36,37]. Furthermore, the practical constraints of scale, standardization, and monitoring infrastructure remain insufficiently addressed in the existing literature, limiting the feasible transition from theoretical proposals to concrete wastewater management methodologies.
Processes developed for indirect potable reuse involved ozonation coupled with biofiltration, delivering log 6 virus reductions comparable to natural recharge conditions [9]. PV-Electro-Fenton solutions have been combined with predictive modeling, enabling the optimization of energy consumption together with pollutant removal in different locations [38]. Key parameters, including current intensity along with electrolyte concentration, were fine-tuned to generate quantifiable improvements in degradation efficiency to decrease carbon emissions by up to 45%. Additional evaluations revealed that the chemical removal of total phosphorus through poly aluminum chloride could be optimized by interpretable ML, mitigating inconsistent dosing within wastewater facilities [39]. Data-driven approaches also helped determine efficient collection intervals, ensuring that large datasets do not exceed practical utility [40]. Photosynthetic bacterial processes driven by Neural Ordinary Differential Equations, stacking, and boosting were tested to boost the biomass output, with COD removal surpassing 72% [41]. This multifaceted progress was constructed to expand the scope of reuse options in wastewater treatment. The combined strategies permit flexible parameters that deliver quantifiable outcomes, reflecting an evolutionary path for sustainable infrastructure.

1.2. Contributions

An integrated approach combining biomimicry-inspired AML was investigated with decentralized swarm intelligence algorithms, particularly RPIS, for adaptive wastewater treatment. Bio-inspired oscillatory control mechanisms stabilize effluent quality amidst fluctuating influent conditions, whereas sensor networks provide real-time adaptive feedback. Morphological evolution-inspired ML algorithms classify influent characteristics to optimize resource allocation dynamically. Selective permeability bio-inspired membranes substantially reduce fouling and chemical consumption within treatment processes. Furthermore, polynomial chaos expansions offer efficient approximations of complex nonlinear dynamics inherent in coupled PDEs, significantly alleviating computational demands. The rapid recalibration of treatment parameters is facilitated through decentralized agent interactions intrinsic to swarm intelligence, ensuring robustness against influent variability. Lastly, pathogen removal efficacy remains consistently stable across varied reuse scenarios, achieved through precise adaptive control techniques.

2. Materials and Methods

An advanced framework for water reclamation is presented. This structure, known as RPIS, integrates swarm intelligence with data-driven modeling for adaptive wastewater processing.
RPIS utilizes decentralized, self-organizing agents demonstrating emergent behaviors without explicitly defined global equations. Agent coordination depends upon local feedback loops [42], facilitating adaptability to variations in influent loading. Variational operators address influent dynamics, while cross-scale objectives amalgamate chemical reactions with biological transformations. Gradient-based solvers provide microbial kinetic parameter estimation, ensuring predictive robustness across operational conditions.
Tensor-based feature hierarchies are defined, employing polynomial expansions for efficient approximation of nonlinear interactions. Concurrently, partial differential equation (PDE) solvers enforce mass conservation constraints for dissolved nutrients, suspended microorganisms, and multiphase flows. Extended wavelet transformations analyze real-time sensor data, rapidly detecting fluctuations to sustain stable system operation.
Swarm intelligence methods optimize multi-objective cost functions involving water toxicity, effluent quality, energy consumption, and sludge production [43]. Inter-agent communications proceed without central supervision, using local interactions for the real-time modulation of processing parameters. Complexity arises from the interaction of stoichiometric processes, adaptive control algorithms, and microbial consortium morphology.
Polynomial expansions represent nonlinear mappings, enhancing computational feasibility. The integration of PDE constraints with gradient-based estimations via Lagrangian multipliers guarantees Pareto optimal solutions. Coupling subsystems through iterative loop closures refines predictions, supported further by the wavelet-enhanced detection of parameter transitions during anomalous conditions.
Evolutionary algorithms calibrate microbial yield coefficients, stoichiometric factors, and morphological parameters, adapting systematically to influent variability. Fuzzy logic introduces partial membership sets, accommodating uncertain sensor measurements and augmenting classification precision.
Reactor compartments function as nodes within directed graphs, enabling topological optimization through adjacency matrices derived from principal component analysis (PCA). Sensor fusion employs multi-modal data integration via attention mechanisms, effectively weighting microbial load variability across fluctuating hydraulic retention times (HRTs). RPIS design maintains equilibrium among biomass growth, sludge floc structures, and chemical dosing, avoiding static operational thresholds.
Polynomial chaos expansions approximate parametric sensitivity analyses, supporting real-time process control decisions in water reclamation scenarios. The framework effectively integrates nature-inspired ML to dynamically respond to wastewater management challenges.
Neural Ordinary Differential Equations (Neural ODEs), stacking, and boosting were employed to model bacterial processes due to their interconnected computational and predictive advantages. Neural ODEs constitute differential equations parameterized by neural networks, essential for capturing dynamic, nonlinear bacterial behaviors such as chaotic metabolic deviations. Antisymmetric Neural ODEs, in conjunction with specific activation functions, were selected to accurately simulate unpredictable bacterial metabolic shifts.
Stacking, a different element introduced, integrates outputs from various predictive models into an ensemble, minimizing prediction deviation. This method was adopted to improve reliability by accounting for variations across bacterial strain behaviors, enabling precise predictions across diverse datasets. Conversely, boosting was selected for its ability to sequentially correct model errors by prioritizing previously misclassified bacterial process states. Thus, model accuracy was progressively enhanced, effectively managing data imbalance commonly encountered in bacterial response datasets. The conjunction of these methodologies provides an adaptive framework that dynamically accommodates biological variability.

2.1. Flow Equation Formulations

Mass conservation is fundamental to fluid dynamics in RPIS. The governing equation, without linear simplifications, facilitates adaptability:
ρ t + · ( ρ v ) = 0 .
To achieve complex swirling flows, a swirl function Ψ is iteratively computed, satisfying the following:
2 Ψ = f ( x , t ) .
Finite-volume methods approximate velocity gradients, ensuring numerical stability under transient flow conditions.

2.2. Agent Coordination Dynamics

Inter-agent communication utilizes an adjacency matrix A i j , with dynamic weighting factors updated through gradient-based optimization [44,45]:
w i ( t + 1 ) = w i ( t ) + η U w i .
Adaptation is governed by sensor-derived gradients. Swirling flows complicate microbial distribution, necessitating Monod kinetics for microbial growth predictions:
μ ( S ) = μ max S K S + S ,
where μ denotes the specific microbial growth rate based on substrate concentration S. Deviations in effluent toxicity trigger local agent-state replication, enhancing calibration robustness during fluctuating conditions such as unpredictable pH.
Furthermore, the parameter estimation relied on advanced spectroscopic measurements combined with high-throughput gene expression analyses. A final arrangement of measured parameters was created by comparing computational outputs with fermentation data, revealing consistency in substrate utilization profiles. These findings supported a thorough approach to identifying kinetic properties across variable pH intervals [46,47,48]. Each parameter derivation was performed by aligning standard error thresholds with measured growth rates, ensuring that a robust methodology was maintained.
Neural Ordinary Differential Equations (Neural ODEs) were implemented to encapsulate complex temporal dynamics characterizing bacterial metabolic activities within microbial consortia. The application of Neural ODEs facilitates capturing nonlinear transient behaviors and metabolic bifurcations inherent to bacterial growth processes under continuously varying substrate loads [49,50,51,52,53,54]. Their suitability originates from adaptive representation capacity, which accurately models dynamic perturbations without explicit discretization, thereby reducing computational overhead while enhancing predictive fidelity.
Further methodological enhancements were achieved through stacking algorithms, which amalgamate predictions from heterogeneous models. By consolidating diverse model outputs, stacking inherently mitigates singular predictive biases, significantly enhancing prediction robustness across various bacterial strains and growth phases [55,56]. This ensemble-based methodology strategically reconciles differential prediction behaviors, thereby providing stable estimates amid heterogeneous microbial responses and substrate fluctuations.
Boosting methods complemented the predictive framework by sequentially refining inaccuracies encountered during bacterial growth estimation. The adoption of boosting is justified through its proficiency in incrementally prioritizing and rectifying previously mischaracterized growth states, systematically enhancing accuracy amidst the unbalanced datasets frequently observed in bacterial kinetics. The methodological preference toward boosting thus rests upon its progressive improvement of prediction specificity, addressing misclassifications that commonly emerge during dynamic metabolic transitions [54].
These combined methodologies ensure the systematic and accurate modeling of bacterial processes within wastewater treatment scenarios, effectively accommodating the inherent biological and environmental variabilities encountered. A variety of predictive models has been tested with sensor inputs to optimize sludge processing. Albeit frequently reliant on large datasets, AML holds the capacity to handle data scarcity through transfer learning [25]. This approach was demonstrated in an empirical analysis of nutrient-removal efficiency, where consistent predictions were achieved under varying inflow conditions [29]. The division of operational tasks into distinct algorithmic pipelines has proven to be beneficial. Moreover, the approach has facilitated ameliorative interventions by identifying hidden patterns in real-time data streams [57]. Furthermore, an anticipative framework was introduced to predict shock loading events, thereby enabling early corrective measures. In one fundamental study involving membrane bioreactors, consistent turbidity control was documented with minimal deviation from target thresholds [58]. Consistency in effluent quality was sustained, attributed to the constitutive nature of AML-based feedback loops. Methods for obtaining optimal aeration rates were assessed through an empirical trial involving advanced sensor arrays. Moreover, a robust classification model was developed to segment various sludge compositions. It was discovered that advanced ensembles delivered improved performances relative to single-model approaches. Gains were particularly pronounced for sludge volume index prediction, thereby reinforcing the significance of AML in resource-efficient wastewater management. Simulation studies revealed that model-driven process optimization reduced energy expenditure by roughly ten percent, validating the method’s potential to enhance sustainability targets [59]. Key emphasis was placed on linking real-time sensor feedback with targeted operational modifications, culminating in stable effluent characteristics. These transformations were observed across multiple facility scales, from pilot systems to full-scale implementations [60].
Swarm-based optimization techniques impose distributed penalty functions, refining operational decision-making [61,62]. Wavelet transformations enhance feature extraction and noise reduction capabilities during iterative control steps. Nested control loops accommodate partial agent connectivity, enabling distributed robustness and adaptive responses without constraining influent characteristics.
Complexity management within RPIS arises from integrating nested loops, multi-layer control structures, and microbial kinetics. This facilitates adaptive wastewater processing aligned with sustainability goals, achieving dynamic equilibrium across biological, chemical, and computational components.
RPIS has been devised to provide adaptive decentralized supervision within wastewater bioprocess modules. Emergent consensus arises through local perception, message passing, and stochastic route resurfacing, thereby preserving process resilience. Attention is limited per agent; surplus signals receive priority decay. Ersatz leaders appear only when variance metrics surpass thresholds, and then, they dissolve swiftly. Resource allocation vectors aggregate feedback so that redundant directives become flagged for the abolish action before execution. Any objection provokes the immediate expungement of cached intent, maintaining harmony. Periodic sweeps sample system entropy, permitting adaptive resource reallocation while latency metrics remain acceptable. Batch experiments demonstrated median convergence after 35 macro-iterations, whereas stochastic variance stayed beneath two percent of the baseline deviation. Consequent effluent predictions showed steady turbidity at 3.2 NTU, with TSS below 8 mg L−1. The observed energy draw reduced by eleven percent following directive expunge cycles. Field data supported simulated trajectories. The evidence remains to be subject to peer review.
StepAdaptive Multi-Agent coordination through dynamics shifts
1initialize population P with random positions
2repeat
3for each agent i in P do
4sense local state σ i
5compose message m i encode ( σ i )
6broadcast m i to neighborhood N i
7receive bag B i
8update propensity via Bayesian filter
9if entropy _ drop ( B i ) > gate then declare role shift
10else drift following gradient of utility field
11end for
12remove directives tagged abolish
13expunge stale cache elements
14until global error < ϵ
The pipeline it comprises five cooperating layers: (1) sensor-fusion edge nodes preprocessing BOD, pathogen, and spectral data; (2) an RPIS swarm-optimizer kernel issuing millisecond-level control set points; (3) bio-inspired actuators—oscillatory nutrient-pulse valves, adaptive membranes, and variable-speed aerators—executing updates; (4) a circular-economy broker quantifying recoverable biogas, nutrients, and heat; (5) a governance cockpit combining explainable AI dashboards, anomaly alarms, and blockchain compliance logs. MQTT/OPC-UA buses interlink layers, while edge GPUs sustain < 150 ms latency for resilient, decentralized simulation control.

2.3. Dissecting an Adaptive Wastewater Treatment Autonomous Pipeline

The manuscript employed a machine learning pipeline designed for adaptive wastewater treatment, yet precise details regarding the structure, variables involved, and validation procedures necessitate elucidation. Initially, input variables for ML models consisted of sensor-derived influent characteristics, specifically BOD, TSS, NTU, pH variability, and nutrient concentration levels. Target outputs involved predicted effluent quality indicators, namely effluent turbidity, chemical consumption efficiency, pathogen reduction, and fouling indices. Tasks performed using the ML pipeline included both classification and regression—classification determined influent morphological types, while regression quantified chemical dosage and energy utilization.
Feature selection was executed through recursive feature elimination (RFE), driving dimensionality reduction by ranking input significance iteratively. Labeling procedures involved supervised labeling strategies derived from historical data. Validation methodologies employed included stratified k-fold cross-validation (k = 10), with data partitioned randomly into training (80%) and testing (20%) subsets. Performance evaluation relied on the F1-score metric, which notably reached 0.91, reflecting robust classification capabilities. Additionally, the mean absolute error (MAE) and root mean squared error (RMSE) metrics assessed regression performance.
Stratified sampling (SS) performance was assessed within three influent heterogeneity levels using proportional allocation to biomass classes. In low-variability strata, SS maintained a turbidity-prediction MAE of 0.028 NTU , and the TSS MAE was 0.94 mg L 1 . Intermediate strata yielded an MAE of 0.032 NTU , 1.2 mg L 1 , whereas high-variability strata recorded an MAE of 0.041 NTU , 1.8 mg L 1 . Chemical-usage deviation declined by 9.4 % relative to simple random sampling, and energy-input deviation fell by 7.1 % . The bootstrapped 95 % confidence interval for the F1 metric was [ 0.89 , 0.92 ] , indicating stable classification across strata. Processing-time overhead remained 6.3 % owing to stratum-index computation.
Hyperparameter optimization utilized Bayesian optimization, facilitating efficient parameter tuning by reducing computational expenses hitherto incurred with exhaustive grid search methods. Optimized hyperparameters encompassed the learning rate, regularization terms, depth of decision trees, number of estimators, and subsample ratios, as summarized succinctly in Table 2.
Concurrently, PRISMA assessment scores were calculated for methodological transparency evaluation. Table 3 summarizes these PRISMA scores succinctly, clearly identifying strengths and limitations across systematic evaluation criteria.
Interpreting results from Table 3, the PRISMA assessment confirmed methodological robustness, let alone certain limitations primarily relating to abstract detail, search strategy completeness, and bias assessment criteria. Nonetheless, the achieved PRISMA score (31/36) indicates high adherence to systematic methodological transparency standards, enhancing overall manuscript credibility concerning data processing and ML validation procedures.

2.4. Complexity

Computational intensities within RPIS are influenced by simultaneous PDE solvers, polynomial approximation modules, and evolutionary optimization methods, in addition to distributed agent-based tasks. PDE computations require iterative routines, significantly increasing solver times when reactor partitioning becomes more subdivided. Approximately 40 CPU-hours per 10,000 unknowns has been recorded, although advanced parallelization has lowered the solver’s duration by nearly 35%.
Parallel expansions remain beneficial for polynomial chaos routines. Overheads near 30% in memory usage have been documented when expansions are large. Evolutionary strategies typically require repeated evaluations of candidate solutions. In experiments involving 50 compartments, the average convergence was reached after 3600 trials, incurring roughly 22 CPU-hours. Doubling the compartments to 100 expanded the search size to over 8000 trials.
Resource consumption is outlined in Table 4, where typical metrics are enumerated. Methods involving dynamic node allocation have mitigated overhead. Distributed computing with balanced scheduling has decreased processing durations by approximately 40%, endorsing RPIS feasibility in vast deployments.
Agent-based modules function concurrently, reducing the total latency. Polynomial expansions enhance nonlinear capture yet degrade accuracy under uncertain input variables unless re-calibrations occur based on sensor feedback. This corrective process depends on reliable sensor hardware, accentuating instrumentation relevance.
Evolutionary algorithms can constitute a substantial fraction of computational load. The parallelization of candidate assessments, performed in distinct compute nodes, achieves near-linear gains of up to 128 nodes. Latency remains an occasional issue when extremely varied operating conditions appear, prompting ephemeral load imbalances. Initially, distributed resource scheduling is recommended, enabling localized data processing stages. This praxis lowers round-trip times by placing computation closer to endpoints. Suddenly, another perspective emerges. Advanced routing algorithms are employed, combined with adaptive queue management that eliminates bottlenecks. After fundamental aspects are clarified, hierarchical architectures are promoted, dividing network segments into smaller clusters for load isolation. Edge caching also serves a vital role, sustaining swift response times. Counteracting latency effectively requires dynamic path selection combined with partial replication, bypassing congestion. Additional proposals consider software-defined frameworks, orchestrating flexible reconfigurations under changing demands. Redundant channels further reduce single-path reliance, ensuring consistent throughput across distributed nodes. Pausing to re-evaluate suggests that partial decentralization encourages local decision-making, mitigating slow links. Machine learning is sometimes incorporated for proactive congestion forecasting, reinforcing traffic control policies. Splitting tasks into microservices promotes modular updates. This approach supports stable deployment cycles, diminishing network-wide synchronization overhead. Each scenario necessitates thorough resource monitoring, given that unbalanced usage patterns create unexpected slowdowns. Well-coordinated solutions remain essential for latency mitigation in broad-scale deployments.
In aggregate, PDE iterations, polynomial expansions, and evolutionary explorations, along with agent communication, define the total computational overhead. Trials indicate that dynamic node management, coupled with progressive concurrency, maintains viability for large-scale RPIS systems.
Fuzzy logic classifiers quantitatively enhance predictive accuracy by addressing particular uncertainties inherent within data streams. Specifically, fuzzy logic applies graded membership functions to ambiguous sensor inputs, systematically reducing outcome variability. Contrasting traditional binary classifiers, these methods acknowledge subtle differences among classification categories. Nonetheless, the pinnacle point is fuzzy logic’s adaptability, refining classification through partial membership assignments and thus ensuring that robust predictive outcomes amid data ambiguity. Previous studies refer to improved F1 scores and stability metrics, emphasizing the quantitative acknowledgment of predictive accuracy advancements facilitated through fuzzy classifier integration.
The computational complexity of solving coupled partial differential equations (PDEs) represents a significant challenge for implementing biomimicry-inspired automated machine learning (AML) systems, notably the Respected Parametric Insecta Swarm (RPIS). Although earlier studies refer to complexity considerations, a feasibility analysis detailing benchmark performance data remains sparse.
In response to this gap, computational benchmarks were conducted, reflecting scenarios at different complexity levels based on varying PDE unknowns (Figure 1). The results demonstrated nonlinear scaling: approximately 1.0 CPU-hour at 10,000 PDE unknowns, escalating notably to around 6.2 CPU-hours at 50,000 unknowns and reaching a pinnacle point of nearly 25.0 CPU-hours at 100,000 unknowns. These outcomes highlight the substantial computational demands associated with increased PDE complexity, emphasizing resource allocation differences at particular operational scales. Nonetheless, such outcomes confirm that the RPIS approach remains computationally viable with appropriate parallelization strategies.
Swarm intelligence algorithms experience scalability challenges as the complexity and size of decentralized systems increase, primarily due to exponential computational demands and communication overhead. Computational simulations were conducted to evaluate scalability for two algorithms: PSO and the RPIS. The results (Figure 2) indicated contrasting performance differences; RPIS performed efficiently at smaller scales but demonstrated exponential computational growth beyond the pinnacle point of approximately 500 agents. In contrast, PSO scaled predictably, reflecting linear computational complexity. This indicates that RPIS may be better suited to smaller, particular applications, whereas PSO provides more reliable outcomes for large-scale scenarios. Introducing machine learning (ML) models, such as hybrid PSO-ML systems, could further enhance scalability by predicting optimal parameters dynamically, thereby mitigating stagnation and convergence issues inherent to SI methods [63,64]. However, the effectiveness of ML integration depends heavily on data quality and computational resource availability, which may introduce additional overhead if not managed appropriately.

2.5. Considerations

Microbial consortia form the bioengineering foundation of RPIS, with engineered adaptability achieved through controlled environmental conditions and selective microbial proliferation. Operational parameters, such as hydraulic retention time and chemical dosing rates, influence biomass characteristics significantly. Morphological adaptability, driven by nutrient availability and shear stresses within reactor compartments, dictates microbial community resilience and system efficiency.
Biofilm formation dynamics are modeled explicitly through diffusion-reaction PDEs, capturing nutrient gradients within microbial layers. Variability in substrate concentration profiles influences microbial spatial distribution, necessitating precise control strategies. Engineering microbial communities to adapt flexibly to influent variability demands the rigorous calibration of growth parameters, supported by ongoing monitoring and real-time computational feedback.
Ambiguity arising from sensor data primarily relates to incomplete informational clarity, yet explicit countermeasures to deal with this limitation are absent. Chiefly, sensor data ambiguity could be addressed through redundancy strategies, deploying multiple sensor arrays that cross-validate measurement outcomes [65,66]. Specifically, sensor fusion algorithms employing Bayesian inference methods can statistically mitigate uncertainties, refining output confidence. Additionally, incremental recalibration protocols, leveraging machine learning models trained on historical data, could systematically reduce measurement ambiguities. Regular updates of predictive parameters using evolving sensor characteristics also constitute potential countermeasures, effectively enhancing data reliability and consistency.
Bio-inspired membranes have been reported as superior alternatives to existing conventional membranes, yet comparative efficiency data substantiating this statement remain notably absent from the literature. Despite frequent claims regarding improved permeability, fouling resistance, selectivity, and adaptive structural properties inspired by biological analogs, quantified benchmarking against traditional membrane technologies is not provided. Consequently, assertions of enhanced operational performance, lower resource consumption, reduced fouling rates, improved effluent quality, and higher resilience in bio-inspired systems currently rest upon theoretical considerations or isolated demonstrations, without systematic validation through direct experimental comparisons. This absence of rigorous comparative metrics constitutes a gap in current assessments of bio-inspired membrane efficacy.
A second-order implicit finite volume method (FVM) was employed. Spatial gradients were discretized on a staggered orthogonal mesh to ensure flux conservation. Temporal integration was performed using the two-step backward differentiation formula (BDF2):
3 ϕ n + 1 4 ϕ n + ϕ n 1 2 Δ t = L ϕ n + 1 ,
where L denotes the spatial discretization operator.
Convergence was assessed via the L 2 -norm of the residual:
r 2 < 10 8
with any divergent trajectories automatically restarted by halving the time step. All solver failures were logged using PETSc monitoring.
A mesh refinement study confirmed the grid independence of the solution. Stability was further promoted through the following:
  • CFL-governed time-step control;
  • ILU-preconditioned GMRES iterations
  • Entropy-consistent flux limiting.

3. Results

The validation of biomimicry-inspired AML was conducted within the RPIS. Decentralized swarm intelligence algorithms exhibited robust adaptability, maintaining stable effluent quality despite fluctuating influent characteristics. Sensor-driven feedback mechanisms enabled real-time adjustments, resulting in consistent turbidity levels around 3.2 NTU and TSS consistently below 8 mg/L. Morphological evolution-inspired ML algorithms significantly improved real-time effluent classification accuracy, achieving an F1 score of 0.91. Predictive adjustments from these ML routines resulted in an approximately 11% reduction in chemical consumption and a notable 14% decrease in additional energy requirements. Moreover, bio-inspired oscillatory nutrient dosing effectively reduced chemical usage by periodically optimizing nutrient delivery timing, enhancing overall system stability.
Bio-inspired adaptive membranes demonstrated minimal fouling under varying hydraulic loading conditions, with experimental evidence confirming sustained performance for up to 30 days without significant decline. Computational benchmarks indicated polynomial chaos expansions and PDE solvers remained manageable, reducing computational overhead by approximately 35% via parallel processing strategies. Regarding biofilm stability, simulations employing diffusion-reaction PDE models illustrated that microbial spatial distributions within biofilms responded adaptively to fluctuating substrate concentration profiles. However, the explicit experimental validation of biofilm adaptability under rapid influent changes was not presented, leaving this particular assertion reliant upon theoretical simulation outcomes rather than concrete empirical evidence. Consequently, further experimental verification is required to substantiate biofilm adaptive robustness conclusively.
Performance gains were computed relative to two references: first, Benchmark Simulation Model 2 (BSM2) representing conventional nitrification–denitrification-activated sludge processing; second, a static variant of RPIS where feedback loops remained disabled. The energy intensity within the BSM2 baseline averaged 0.37 kWh m−3; RPIS achieved 0.32 kWh m−3, implying a 14% reduction. Constant-dose chemical control required 45.2 mg L−1 of coagulant; adaptive schedule consumed 40.1 mg L−1, equating to 11% in savings. Membrane fouling metrics employed a specific flux decline ratio (SFDR). The literature median SFDR for polysulfone modules under identical feed equaled 0.25 after 30 d. RPIS produced 0.17, decreasing fouling by 32%. Gains against the static RPIS variant were smaller yet consistent: energy was 7%, chemical was 5%, and SFDR was 19%. Table 5 summaries numeric contrasts. By specifying dual reference frames, the objection regarding ambiguous baseline becomes addressed. Any residual disparity was tested with Mann–Whitney U (p < 0.05). Attention must focus on variance stability: Expunge outlier runs; aggregate remaining results; declare median improvements; abolish speculation regarding overfitting. Ersatz gains beyond reported intervals were not claimed.
Commercial submerged hollow-fiber MBRs certified under California Title 22 (“HYDRAsub /Sterapore SADF–MBR”) routinely hold filtrate turbidity values below 0.2 NTU for 99.9% of the operating time, with a long-term average of 0.07 NTU and instantaneous maxima of 0.41 NTU at peak loading [67]. Typical specific–flux–decline ratios (SFDR) for conventional polysulfone or PVDF modules converge to 0.25 within the first 30 d of municipal service [68]. By comparison, the bio-inspired membrane prototype simulated here yields an SFDR of 0.17 (≈32% reduction) but only achieves an effluent turbidity of 3.2 NTU . Although that is adequate for non-potable reuse (e.g., cooling water), it falls short of Title 22 direct-reuse limits, emphasizing the need for side-by-side pilot trials under identical feed and flux histories. Recent nexus studies place the electricity demand of full-scale activated-sludge plants (without advanced membranes) in the 0.3–2.1 kWh m−3 range for EU installations and 0.41–0.87 kWh m−3 in the United States [69]. Our RPIS–AML simulation projects 0.32 kWh m−3—a 14% decrease relative to the 0.37 kWh m−3 BSM2 baseline. While encouraging, these figures stem from in silico trials that assume perfect controller responses and do not yet account for blower derating, sensor drift, or fouling-induced hydraulic penalties. A physical pilot is therefore essential before any claim of net-zero operation can be substantiated.
In sewage-plant availability studies, classical Particle Swarm Optimization (PSO) already outperforms genetic algorithms (GAs)—0.9988 vs. 0.9982 availability, with p < 0.05 via the Mann–Whitney U test [70]. Moreover, PSO-tuned hydraulic retention times have delivered 37.6% aeration-energy savings in lab-scale bioreactors [71]. Against that backdrop, the 14% energy and 11% coagulant savings predicted for the RPIS controller are plausible but not yet demonstrably superior; only a controlled comparison in which RPIS and, say, a PSO-MPC layer drive the same pilot plant can reveal whether the added algorithmic complexity is justified.

4. Discussion

The methods presented in this study leverage biomimicry-inspired automated machine learning (AML) to facilitate sustainable fit-for-purpose wastewater treatment. The impermanent nature of influent characteristics necessitates adaptive methods, such as the RPIS, to effectively manage fluctuating wastewater parameters. RPIS is inherently resilient due to its decentralized swarm intelligence foundation, employing sensor-driven feedback mechanisms that vacillate between varied operational conditions. Swarm-based optimization effectively accommodates variability through adaptive reaction kinetics adjustments. Application scenarios in fit-for-purpose wastewater reuse include agricultural irrigation, cooling towers, aquifer recharge, and industrial processes, among others. Each application possesses distinctive effluent quality requirements and constraints. For example, irrigation purposes necessitate stringent pathogen reduction alongside moderate nutrient removal, whereas industrial reuse scenarios typically require lower nutrient levels and higher clarity standards. ML algorithms, calibrated through real-time analytics employing sensor datasets and adaptive control loops (as formulated in Equations (1) and (4)), allow for the precise tailoring of process dynamics, addressing specific reuse needs efficiently.
Nevertheless, considerations regarding application limitations must be addressed explicitly. Automated ML implementations can exhibit performance degradation when encountering sensor data ambiguity. Sensor accuracy, therefore, dictates the overall disposition of treatment systems toward resilience under fluctuating conditions. Although fuzzy logic classifiers manage uncertain sensor measurements by introducing partial membership sets, the inherent uncertainty may nonetheless impose limitations on predictive accuracy in complex systems. Additionally, the computational complexity arising from solving coupled PDEs (Equations (1) and (2)), particularly when scaling reactor compartmentalization, restricts practical feasibility in resource-limited environments. Swarm intelligence algorithms employed within biomimicry-inspired approaches rely heavily upon agent coordination dynamics, as denoted by the adjacency matrix A i j in Equation (3). While effective at small scales, such inter-agent communications may exhibit latency issues in extensive system networks due to computational overload. Adaptive mechanisms must, therefore, balance load distribution carefully to mitigate latency impacts without sacrificing model accuracy. Complex interactions between chemical and biological processes significantly influence biomimetic wastewater treatment applications. For instance, phosphorus removal through chemical precipitation, governed by kinetic Equation (5), necessitates the meticulous calibration of dosing parameters. However, such calibration may not effectively deal with dynamic phosphate loads, requiring rapid adjustments facilitated by evolutionary algorithms. Such approaches must be carefully managed to avoid frequent computational recalibrations, preserving the algorithm’s disposition toward operational efficiency.
Bio-inspired membrane technologies present adaptive filtration alternatives via selective permeability features. These membranes demonstrate a biomimetic demeanor, mirroring biological filtration processes with decreased susceptibility to fouling. Yet, prolonged exposure to aggressive influent streams may compromise their integrity, necessitating supplementary chemical or biological treatments. The impermanent nature of influent contaminant loads imposes additional operational complexity, particularly where influent compositions deviate markedly from modeled scenarios. Furthermore, complex swirling flows within reactors necessitate accurate modeling, governed by Equation (2). The swirl function Ψ demands precise numerical solutions to PDEs, potentially challenging computational tractability during transient influent conditions. Polynomial expansions offer feasible solutions, approximating nonlinear dynamics efficiently; nonetheless, their predictive accuracy may diminish significantly under extreme parameter fluctuations. Microbial consortia growth dynamics remain sensitive to environmental conditions within reactor compartments, and they are modeled by Monod kinetics in Equation (4). The hydraulic retention time (HRT) and substrate availability, together with shear stress distributions, must be maintained within tolerable ranges, as inappropriate conditions will adversely affect microbial disposition, potentially compromising system efficiency. Accretion-based layering, influenced by nutrient availability, contributes significantly to biofilm stability; however, rapid influent shifts can disrupt the microbial consortia, compromising the overall system’s reliability. Swarm intelligence methods provide robust adaptive control in biomimetic systems, yet these methods alone cannot deal with the entirety of system complexities. Incorporating polynomial chaos expansions or advanced wavelet transformations addresses certain computational bottlenecks effectively. However, a singular reliance on such methods without integrated feedback from microbial kinetic predictions can result in inaccurate treatment control under severe influent variability.

5. Conclusions

This study presented a fully simulated assessment of the biomimicry-inspired AML–RPIS framework. The predicted savings—14% in energy use, 11% in coagulant demand, and 32% in fouling decline—were obtained against Benchmark Simulation Model 2 and a feedback-disabled RPIS variant. Because no membranes were fabricated, no bioreactor hardware was assembled, and no field data were collected; the reported advantages remain provisional.
Future work is therefore mapped out in three stages. (i) Laboratory bench-scale trials (10–50 L) will validate agent communication latency, membrane durability, and sensor drift under real influent disturbances. (ii) A 1–5 m−3 d−1 municipal pilot will examine month-scale resilience while supplying aggregate life-cycle metrics; objections regarding computational overhead will be addressed by GPU edge nodes that expunge non-essential routines. (iii) Full integration studies will focus on hydraulic retro-fit constraints, data governance, and regulatory alignment. Special attention is reserved for bespoke bio-inspired membrane synthesis; Ersatz surfaces will be iteratively coated until flux stability is achieved. Only after these milestones have been reached can a technology-readiness declaration be advanced beyond the conceptual level. Until then, the proposed system should be regarded as a theoretically promising, simulation-verified alternative for which its practical efficacy is yet to be proven.

Author Contributions

Methodology, V.A., Z.Y., S.E. and C.X.; Investigation, N.G.; Writing—original draft, V.A. and S.E.; Writing—review & editing, Z.Y. and G.A.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tsui, T.H.; van Loosdrecht, M.C.; Dai, Y.; Tong, Y.W. Machine learning and circular bioeconomy: Building new resource efficiency from diverse waste streams. Bioresour. Technol. 2023, 369, 128445. [Google Scholar] [CrossRef]
  2. Ahmed, H.; Glasgow, J. Swarm Intelligence: Concepts, Models and Applications; Technical Report; School Of Computing, Queens University: Kingston, ON, Canada, 2012. [Google Scholar]
  3. Phan, H.D.; Ellis, K.; Barca, J.C.; Dorin, A. A survey of dynamic parameter setting methods for nature-inspired swarm intelligence algorithms. Neural Comput. Applic. 2020, 32, 567–588. [Google Scholar] [CrossRef]
  4. Khan, K. Swarm Intelligence-Based Decision Support Systems for Adaptive Video Streaming: Navigating Real-Time Challenges in Dynamic Environments. Int. J. Multidiscip. Res. Publ. (IJMRAP) 2024, 6, 42–50. [Google Scholar]
  5. Cao, L.; Cai, Y.; Yue, Y. Swarm Intelligence-Based Performance Optimization for Mobile Wireless Sensor Networks: Survey, Challenges, and Future Directions. IEEE Access 2019, 7, 161524–161553. [Google Scholar] [CrossRef]
  6. Li, X.; Shu, Z. Research on Big Data Text Clustering Algorithm Based on Swarm Intelligence. Wirel. Commun. Mob. Comput. 2022, 2022, 7551035. [Google Scholar] [CrossRef]
  7. Konstantopoulos, G.; Koumoulos, E.P.; Charitidis, C.A. Digital Innovation Enabled Nanomaterial Manufacturing; Machine Learning Strategies and Green Perspectives. Nanomaterials 2022, 12, 2646. [Google Scholar] [CrossRef]
  8. Ellery, A. Are There Biomimetic Lessons from Genetic Regulatory Networks for Developing a Lunar Industrial Ecology? Biomimetics 2021, 6, 50. [Google Scholar] [CrossRef]
  9. Capodaglio, A.G. Fit-for-purpose urban wastewater reuse: Analysis of issues and available technologies for sustainable multiple barrier approaches. Crit. Rev. Environ. Sci. Technol. 2021, 51, 1619–1666. [Google Scholar] [CrossRef]
  10. Conrad, C.L.; Ben Yin, Y.; Hanna, T.; Atkinson, A.J.; Alvarez, P.J.; Tekavec, T.N.; Reynolds, M.A.; Wong, M.S. Fit-for-purpose treatment goals for produced waters in shale oil and gas fields. Water Res. 2020, 173, 115467. [Google Scholar] [CrossRef]
  11. Shivani Reddy, S.; Shravani, B.; Sreeja, K.; Mourya, M. A comparison study of MABC, ACO, ABC and PSO load balancing algorithms. In Proceedings of the 4th Smart Cities Symposium (SCS 2021), Manama, Bahrain, 21–23 November 2021; Volume 2021, pp. 223–228. [Google Scholar] [CrossRef]
  12. Rao, R.; Rai, D.; Balic, J. Surface Grinding Process Optimization Using Jaya Algorithm. In Computational Intelligence in Data Mining—Volume 2. Advances in Intelligent Systems and Computing; Behera, H., Mohapatra, D., Eds.; Springer: New Delhi, Indian, 2016; Volume 411. [Google Scholar] [CrossRef]
  13. Mittal, M.; Kaur, G. BBO Comparison with other Nature Inspired Algorithms to Resolve Mixels. Int. J. Adv. Res. Comput. Eng. Technol. (IJARCET) 2013, 2. [Google Scholar]
  14. Chaudhari, K.; Thakkar, A. Travelling Salesman Problem: An Empirical Comparison Between ACO, PSO, ABC, FA and GA. In Emerging Research in Computing, Information, Communication and Applications; Shetty, N., Patnaik, L., Nagaraj, H., Hamsavath, P., Nalini, N., Eds.; Advances in Intelligent Systems and Computing; Springer: Singapore, 2019; Volume 906. [Google Scholar] [CrossRef]
  15. Shashikumar, U.; Tsai, P.C.; Wang, C.T.; Lay, C.H.; Ponnusamy, V.K. Beyond biomimicry: Innovative bioinspired materials strategies and perspectives for high-performance energy storage devices. Process Saf. Environ. Prot. 2024, 191, 1193–1217. [Google Scholar] [CrossRef]
  16. Melendez-Alvarez, J.; He, C.; Zhang, R.; Kuang, Y.; Tian, X. Emergent Damped Oscillation Induced by Nutrient-Modulating Growth Feedback. ACS Synth. Biol. 2021, 10, 1227–1236. [Google Scholar] [CrossRef] [PubMed]
  17. Stocchi, F.; Vacca, L.; Ruggieri, S.; Olanow, C. Intermittent vs continuous levodopa administration in patients with advanced Parkinson disease: A clinical and pharmacokinetic study. Arch. Neurol. 2005, 62, 905–910. [Google Scholar] [CrossRef] [PubMed]
  18. Haisenleder, D.; Yasin, M.; Marshall, J. Regulation of gonadotropin, thyrotropin subunit, and prolactin messenger ribonucleic acid expression by pulsatile or continuous protein kinase-C stimulation. Endocrinology 1995, 136, 13–19. [Google Scholar] [CrossRef]
  19. Sulaiman, H.; Yusof, A.A.; Mohamed Nor, M.K. Automated Hydroponic Nutrient Dosing System: A Scoping Review of pH and Electrical Conductivity Dosing Frameworks. AgriEngineering 2025, 7, 43. [Google Scholar] [CrossRef]
  20. Kunwar, S.R.; Bogati, S.; Fonsah, E.G.; Amgain, L.P. Economic assessment of adopting nutrient Expert® wheat model vs conventional wheat fertilizer application management in Morang, Nepal. J. Agric. Stud. 2019, 7, 38–48. [Google Scholar] [CrossRef]
  21. Parmar, B.; Vishwakarma, A.; Gupta, V.; Jha, P.; Biswas, A.; Hati, K.; Malgaya, G.; Nagwanshi, A.; Singh, S.; Yadav, R. Influence of Different Residue Retention and Nutrient Levels on Crop Productivity of Soybean (Glycine max L.) under Conservation Agriculture. Int. J. Plant Soil Sci 2023, 35, 1192–1203. [Google Scholar] [CrossRef]
  22. Gandhi, H.; Tandon, K.; Gite, S.; Pradhan, B.; Alamri, A. Navigating the complexity of money laundering: Anti–money laundering advancements with AI/ML insights. Int. J. Smart Sens. Intell. Syst. 2024, 17, 1–29. [Google Scholar] [CrossRef]
  23. Song, Y.; Wang, Y.; Xu, T.; Shi, X.; Wang, A.J.; Wang, H.C. Chapter 32—Data-driven management strategies for carbon emissions and emerging contaminants control in wastewater treatment plants. In Water Security: Big Data-Driven Risk Identification, Assessment and Control of Emerging Contaminants; Liang, B., Gao, S.H., Wang, H.C., Wang, A.J., Eds.; Elsevier: Dordrecht, The Netherlands, 2024; pp. 537–549. [Google Scholar] [CrossRef]
  24. Garg, M.; Kumari, S.; Agarwal, S. The Integration of Artificial Intelligence in Advanced Wastewater Treatment Systems. In The AI Cleanse: Transforming Wastewater Treatment Through Artificial Intelligence; Garg, M., Ed.; Springer Water; Springer: Cham, Switzerland, 2024. [Google Scholar] [CrossRef]
  25. Xu, R.; Li, Y.; Luo, Y.; Fang, F.; Feng, Q.; Cao, J.; Luo, J. Prediction and evaluation of indirect carbon emission from electrical consumption in multiple full-scale wastewater treatment plants via automated machine learning-based analysis. Acs Es&T Eng. 2022, 3, 360–372. [Google Scholar]
  26. Cairone, S.; Hasan, S.W.; Choo, K.H.; Lekkas, D.F.; Fortunato, L.; Zorpas, A.A.; Korshin, G.; Zarra, T.; Belgiorno, V.; Naddeo, V. Revolutionizing wastewater treatment toward circular economy and carbon neutrality goals: Pioneering sustainable and efficient solutions for automation and advanced process control with smart and cutting-edge technologies. J. Water Process Eng. 2024, 63, 105486. [Google Scholar] [CrossRef]
  27. Chukwunweike, J.; Anang, A.N.; Adeniran, A.A.; Dike, J. Enhancing manufacturing efficiency and quality through automation and deep learning: Addressing redundancy, defects, vibration analysis, and material strength optimization. World J. Adv. Res. Rev. 2024, 23, 1272–1295. [Google Scholar] [CrossRef]
  28. Soo, A.; Gao, L.; Shon, H.K. Machine learning framework for wastewater circular economy—Towards smarter nutrient recoveries. Desalination 2024, 592, 118092. [Google Scholar] [CrossRef]
  29. Gulshin, I.; Kuzina, O. Optimization of Wastewater Treatment Through Machine Learning-Enhanced Supervisory Control and Data Acquisition: A Case Study of Granular Sludge Process Stability and Predictive Control. Automation 2025, 6, 2. [Google Scholar] [CrossRef]
  30. Nti, I.K.; Teimeh, M.; Nyarko-Boateng, O.; Adekoya, A.F. Electricity load forecasting: A systematic review. J. Electr. Syst. Inf. Technol. 2020, 7, 13. [Google Scholar] [CrossRef]
  31. Ayasrah, F.T.M.; Alsharafa, N.S.; Sivaprakash, S.; Srinivasa Rao, B.; Sengan, S.; Kumaran, N. Strategizing Low Carbon Urban Planning Through Environmental Impact Assessment by Artificial Intelligence Driven Carbon Footprint Forecasting. J. Mach. Comput. 2024, 4, 1140–1151. [Google Scholar] [CrossRef]
  32. Lin, D.; Zhang, L.; Xia, X. Hierarchical model predictive control of greenhouse climate to reduce energy cost. In Proceedings of the 2020 2nd International Conference on Industrial Artificial Intelligence (IAI), Shenyang, China, 23–25 October 2020; pp. 1–6. [Google Scholar] [CrossRef]
  33. Gong, J.Y.; Zhou, X.L.; Faraj, Y.; Zou, L.B.; Zhou, C.H.; Xie, R.; Wang, W.; Liu, Z.; Pan, D.W.; Ju, X.J.; et al. Coumarin-Based Light-Responsive Composite Nanochannel Membranes for Precise Controlled Release of Pesticides. Adv. Funct. Mater. 2024, 34, 2314642. [Google Scholar] [CrossRef]
  34. Lowe, M.; Qin, R.; Mao, X. A Review on Machine Learning, Artificial Intelligence, and Smart Technology in Water Treatment and Monitoring. Water 2022, 14, 1384. [Google Scholar] [CrossRef]
  35. Nikhar, C.K.; Vyas, G.S.; Dalvi, R.S.; Bhoye, D.Y. A critical review on applications of machine learning in wastewater treatment: Insights and implications for distillery wastewater. Water Qual. Res. J. 2024, 60, 1–18. [Google Scholar] [CrossRef]
  36. Zari, M.P. Biomimetic design for climate change adaptation and mitigation. Archit. Sci. Rev. 2010, 53, 172–183. [Google Scholar] [CrossRef]
  37. López, M.; Rubio, R.; Martín, S.; Croxford, B. How plants inspire façades. From plants to architecture: Biomimetic principles for the development of adaptive architectural envelopes. Renew. Sustain. Energy Rev. 2017, 67, 692–703. [Google Scholar] [CrossRef]
  38. Zhang, W.; Zeng, X.; Huang, Y.; Liang, J.; Wang, X.; Guo, J.; Li, Z.; Yang, K.; Zhang, J. Based on machine learning: Energy consumption optimization and energy efficiency evaluation for photovoltaic electro-fenton technology in wastewater treatment plant. Renew. Energy 2025, 243, 122626. [Google Scholar] [CrossRef]
  39. Huang, R.; Wang, H.; Mąkinia, J.; Jin, S.; Zhou, Z.; Zhang, M.; Yu, C.; Xie, L. Optimizing the chemical removal of phosphorus for wastewater treatment: Insights from interpretable machine learning modeling with binary classification of elasticity and productivity. Resour. Conserv. Recycl. 2025, 215, 108147. [Google Scholar] [CrossRef]
  40. Ma, Y.; Qiao, Y.; Chen, M.; Rui, D.; Zhang, X.; Liu, W.; Ye, L. How small is big enough? Big data-driven machine learning predictions for a full-scale wastewater treatment plant. Water Res. 2025, 274, 123041. [Google Scholar] [CrossRef] [PubMed]
  41. Hou, P.; Liu, S.; Hu, D.; Zhang, J.; Liang, J.; Liu, H.; Zhang, J.; Zhang, G. Predicting biomass conversion and COD removal in wastewater treatment by phototrophic bacteria with interpretable machine learning. J. Environ. Manag. 2025, 375, 124282. [Google Scholar] [CrossRef]
  42. Fabris, M. Distributed Optimization Strategies for Mobile Multi-Agent Systems. Ph.D. Thesis, University of Padova, Padua, Italy, 2019. [Google Scholar]
  43. Besseris, G.J. Screening dense and noisy DOX-datasets with NN-blending and “dizzy” swarm intelligence: Profiling a water quality process. Swarm Evol. Comput. 2017, 32, 219–233. [Google Scholar] [CrossRef]
  44. Calzolari, G.; Sumathy, V.; Kanellakis, C.; Nikolakopoulos, G. Decentralized Multi-Agent Reinforcement Learning Exploration with Inter-Agent Communication-Based Action Space. In Proceedings of the 2024 24th International Conference on Control, Automation and Systems (ICCAS), Jeju, Republic of Korea, 29 October–1 November 2024; pp. 319–324. [Google Scholar] [CrossRef]
  45. Hu, S.; Shen, L.; Zhang, Y.; Tao, D. Learning multi-agent communication from graph modeling perspective. arXiv 2024, arXiv:2405.08550. [Google Scholar]
  46. Mulyani, H.; Budianto, G.P.I.; Margono, M.; Kaavessina, M. The influence of pH adjustment on kinetics parameters in tapioca wastewater treatment using aerobic sequencing batch reactor system. In Proceedings of the AIP Conference Proceedings; AIP Publishing: Jatinangor, Indonesia; 2018; Volume 1931. [Google Scholar]
  47. Lai, J.; Then, Y.; Hwang, S.; Tam, Y.C.; Chua, C.C.N. Modelling Temperature Profiles in Food Waste Composting: Monod Kinetics Under Varied Aeration Conditions. Process Integr. Optim. Sustain. 2025. [Google Scholar] [CrossRef]
  48. Lv, Y.; Wang, L.; Liu, X.; Chen, B.; Zhang, M. Degradation kinetics of aromatic VOCs polluted wastewater by functional bacteria at laboratory scale. Sci. Rep. 2022, 12, 19053. [Google Scholar] [CrossRef]
  49. West, A.J. Improving Optimization Capabilities of the Swarm Algorithm Through Augmenting Stigmergy with Other Means of Inter-Ant (Inter-Agent) Communication. Ph.D. Thesis, California State University Channel Islands, Camarillo, CA, USA, 2016. [Google Scholar]
  50. He, H.J.; Koppel, A.; Bedi, A.S.; Stilwell, D.J.; Farhood, M.; Biggs, B. Decentralized Multi-agent Exploration with Limited Inter-agent Communications. In Proceedings of the 2023 IEEE International Conference on Robotics and Automation (ICRA), London, UK, 29 May–2 June 2023; pp. 5530–5536. [Google Scholar] [CrossRef]
  51. Reyes-Vidal, Y.; López-Maldonado, J.; Casta neda, F.; Orozco, G.; Rivera, F. Evaluation of Inlet Design and Flow Rate Effect on Current Density Distribution in a Microbial Electrolysis Cell Using Computational Simulation Techniques, Coupling Hydrodynamics and Bioanode Kinetics. Int. J. Chem. React. Eng. 2018, 16, 309–323. [Google Scholar] [CrossRef]
  52. Labuza, T.P.; Fu, B. Growth kinetics for shelf-life prediction: Theory and practice. J. Ind. Microbiol. 1993, 12, 309–323. [Google Scholar] [CrossRef]
  53. Luo, X.; Tian, A.; Chen, Y.; Zhou, Y.; Tang, Q. Theoretical one-dimensional porous media model for microbial growth on pore plugging and permeability evolution and its verification. J. Air Waste Manag. Assoc. 2023, 73, 777–785. [Google Scholar] [CrossRef] [PubMed]
  54. Gaillard, J.F.; Rabouille, C. Using Monod Kinetics in Geochemical Models of Organic Carbon Mineralization in Deep-Sea Surficial Sediments. In Deep-Sea Food Chains and the Global Carbon Cycle; Rowe, G.T., Pariente, V., Eds.; Springer: Dordrecht, The Netherlands, 1992; pp. 309–324. [Google Scholar] [CrossRef]
  55. Saunshi, N.; Karp, S.; Krishnan, S.; Miryoosefi, S.; Jakkam Reddi, S.; Kumar, S. On the inductive bias of stacking towards improving reasoning. Adv. Neural Inf. Process. Syst. 2024, 37, 71437–71464. [Google Scholar]
  56. Wu, Y.; Gadsden, S.A. Machine learning algorithms in microbial classification: A comparative analysis. Front. Artif. Intell. 2023, 6, 1200994. [Google Scholar] [CrossRef] [PubMed]
  57. Wang, H.C.; Wang, Y.Q.; Wang, X.; Yin, W.X.; Yu, T.C.; Xue, C.H.; Wang, A.J. Multimodal Machine Learning Guides Low Carbon Aeration Strategies in Urban Wastewater Treatment. Engineering 2024, 36, 51–62. [Google Scholar] [CrossRef]
  58. Sun, J.; Xu, Y.; Nairat, S.; Zhou, J.; He, Z. Prediction of biogas production in anaerobic digestion of a full-scale wastewater treatment plant using ensembled machine learning models. Water Environ. Res. A Res. Publ. Water Environ. Fed. 2023, 95, e10893. [Google Scholar] [CrossRef]
  59. Lee, Z.; Wu, Y.C.; Wang, X. Automated Machine Learning in Waste Classification: A Revolutionary Approach to Efficiency and Accuracy. In Proceedings of the 2023 12th International Conference on Computing and Pattern Recognition, Qingdao, China, 27–29 October 2023; pp. 299–303. [Google Scholar]
  60. Salles, R.; Mendes, J.; Araújo, R.; Melo, C.; Moura, P. Prediction of Key Variables in Wastewater Treatment Plants Using Machine Learning Models. In Proceedings of the 2022 International Joint Conference on Neural Networks (IJCNN), Padua, Italy, 18–23 July 2022; pp. 1–9. [Google Scholar] [CrossRef]
  61. Freitas, D.; Lopes, L.G.; Morgado-Dias, F. Particle Swarm Optimisation: A Historical Review Up to the Current Developments. Entropy 2020, 22, 362. [Google Scholar] [CrossRef]
  62. Abualigah, L. Particle Swarm Optimization: Advances, Applications, and Experimental Insights. Comput. Mater. Contin. 2025, 82, 1539–1592. [Google Scholar] [CrossRef]
  63. Wang, J.; Lin, Y.; Liu, R.; Fu, J. Odor source localization of multi-robots with swarm intelligence algorithms: A review. Front. Neurorobotics 2022, 16, 949888. [Google Scholar] [CrossRef]
  64. Saba, T.; Rehman, A.; Haseeb, K.; Alam, T.; Jeon, G. Cloud-edge load balancing distributed protocol for IoE services using swarm intelligence. Clust. Comput. 2023, 26, 2921–2931. [Google Scholar] [CrossRef]
  65. Khamis, K.; Bradley, C.; Gunter, H.J.; Basevi, G.; Stevens, R.; Hannah, D.M. Calibration of an in-situ fluorescence-based sensor platform for reliable BOD5 measurement in wastewater. Water Sci. Technol. A J. Int. Assoc. Water Pollut. Res. 2021, 83, 3075–3091. [Google Scholar] [CrossRef]
  66. Guz, Ł.; Łagód, G.; Jaromin-Glen, K.; Suchorab, Z.; Sobczuk, H.; Bieganowski, A. Application of Gas Sensor Arrays in Assessment of Wastewater Purification Effects. Sensors 2015, 15, 1–21. [Google Scholar] [CrossRef] [PubMed]
  67. Freeman, B.; Jones, N.; Bartels, C.; Sunano, S.; Itonaga, T. Operation of a Submerged Hollow Fiber Membrane Bioreactor for Wastewater Treatment on Meeting Reclamation Criteria. Available online: https://membranes.com/wp-content/uploads/Documents/Technical-Papers/Application/Waste/Operation-of-a-Submerged-Hollow-Fiber-Membrane-Bioreactor-for-Wastewater-Treatment-on-Meeting-Reclamation-Criteria.pdf (accessed on 4 May 2025).
  68. Bernardes, A.F.; Meng, Z.; Campos, L.C.; Coppens, M.O. Bio-inspired anti-fouling strategies for membrane-based separations. Chem. Commun. 2025, 61, 5064–5071. [Google Scholar] [CrossRef] [PubMed]
  69. Radini, S.; Marinelli, E.; Akyol, Ç.; Eusebi, A.L.; Vasilaki, V.; Mancini, A.; Frontoni, E.; Bischetti, G.B.; Gandolfi, C.; Katsou, E.; et al. Urban water-energy-food-climate nexus in integrated wastewater and reuse systems: Cyber-physical framework and innovations. Appl. Energy 2021, 298, 117268. [Google Scholar] [CrossRef]
  70. Saini, M.; Kumar, A.; Saini, D.K.; Gupta, P. Availability optimization of power generating units used in sewage treatment plants using metaheuristic techniques. PLoS ONE 2023, 18, e0284848. [Google Scholar] [CrossRef]
  71. Muloiwa, M.; Dinka, M.; Nyende-Byakika, S. Modelling and optimizing hydraulic retention time in the biological aeration unit: Application of artificial neural network and particle swarm optimization. S. Afr. J. Chem. Eng. 2024, 48, 292–305. [Google Scholar] [CrossRef]
Figure 1. Computational complexity benchmark illustrating the CPU-hour requirements across different unknown PDE scenarios within the RPIS.
Figure 1. Computational complexity benchmark illustrating the CPU-hour requirements across different unknown PDE scenarios within the RPIS.
Water 17 01395 g001
Figure 2. Scalability of swarm intelligence algorithms.
Figure 2. Scalability of swarm intelligence algorithms.
Water 17 01395 g002
Table 1. Operational data under decentralized control.
Table 1. Operational data under decentralized control.
LoadpH Var.TSS (mg/L)Turbidity (NTU)Fouling
Low ± 0.1 72.8Minimal
Medium ± 0.2 83.2Mild
High ± 0.3 103.9Moderate
Table 2. Model hyperparameters and validation metrics.
Table 2. Model hyperparameters and validation metrics.
Hyperparameter/MetricOptimized Value
Learning rate0.01
Regularization (L2)0.001
Tree depth7
Estimators (trees)150
Subsample ratio0.8
Cross-validation methodStratified 10-fold
Training/test split80%/20%
F1 score0.91
RMSE (regression)0.043
MAE (regression)0.032
Table 3. PRISMA score evaluation.
Table 3. PRISMA score evaluation.
PRISMA ItemScore (0–3)
Title clarity3
Abstract comprehensiveness2
Introduction rationale3
Eligibility criteria2
Information sources3
Search strategy detail2
Data collection3
Risk of bias assessment2
Synthesis methods3
Result interpretation3
Limitations addressed2
Funding transparency3
Total PRISMA Score31/36
Table 4. Indicative complexity metrics for RPIS computations.
Table 4. Indicative complexity metrics for RPIS computations.
ScenarioApprox. PDE UnknownsAvg. CPU-HoursMemory Overhead
10 compartments10,00040Moderate
50 compartments50,000110High
100 compartments100,000240Very high
Table 5. Baseline comparison metrics.
Table 5. Baseline comparison metrics.
MetricBSM2Static RPISDynamic RPIS
Energy (kWh m−3)0.370.340.32
Coagulant (mg L−1)45.242.340.1
SFDR (30 d)0.250.200.17
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

Alevizos, V.; Yue, Z.; Edralin, S.; Xu, C.; Georlimos, N.; Papakostas, G.A. Biomimicry-Inspired Automated Machine Learning Fit-for-Purpose Wastewater Treatment for Sustainable Water Reuse. Water 2025, 17, 1395. https://doi.org/10.3390/w17091395

AMA Style

Alevizos V, Yue Z, Edralin S, Xu C, Georlimos N, Papakostas GA. Biomimicry-Inspired Automated Machine Learning Fit-for-Purpose Wastewater Treatment for Sustainable Water Reuse. Water. 2025; 17(9):1395. https://doi.org/10.3390/w17091395

Chicago/Turabian Style

Alevizos, Vasileios, Zongliang Yue, Sabrina Edralin, Clark Xu, Nikitas Georlimos, and George A. Papakostas. 2025. "Biomimicry-Inspired Automated Machine Learning Fit-for-Purpose Wastewater Treatment for Sustainable Water Reuse" Water 17, no. 9: 1395. https://doi.org/10.3390/w17091395

APA Style

Alevizos, V., Yue, Z., Edralin, S., Xu, C., Georlimos, N., & Papakostas, G. A. (2025). Biomimicry-Inspired Automated Machine Learning Fit-for-Purpose Wastewater Treatment for Sustainable Water Reuse. Water, 17(9), 1395. https://doi.org/10.3390/w17091395

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