Next Article in Journal
Navigating the Aerosolized Frontier: A Comprehensive Review of Bioaerosol Research Post-COVID-19
Previous Article in Journal
Emissions and Atmospheric Dry and Wet Deposition of Trace Metals from Natural and Anthropogenic Sources in Mainland China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mechanisms and Applications of Nature-Based Solutions for Stormwater Control in the Context of Climate Change: A Review

by
Chuanhao Sun
1,
Qiuyi Rao
2,
Biyi Chen
2,
Xin Liu
3,
Rana Muhammad Adnan Ikram
1,
Jianjun Li
1,2,
Mo Wang
1,2,* and
Dongqing Zhang
4,*
1
College of Architecture and Urban Planning, Guangzhou University, Guangzhou 510006, China
2
Architectural Design and Research Institute, Guangzhou University, Guangzhou 510091, China
3
College of Horticulture and Landscape Architecture, Southwest University, Chongqing 400715, China
4
Guangdong Provincial Key Laboratory of Petrochemical Pollution Processes and Control, School of Environmental Science and Engineering, Guangdong University of Petrochemical Technology, Maoming 525000, China
*
Authors to whom correspondence should be addressed.
Atmosphere 2024, 15(4), 403; https://doi.org/10.3390/atmos15040403
Submission received: 29 February 2024 / Revised: 15 March 2024 / Accepted: 21 March 2024 / Published: 25 March 2024
(This article belongs to the Special Issue Nature-Based Countermeasures in Atmospheric and Climate Research)

Abstract

:
Nature-based solutions (NbSs) are considered to form an innovative stormwater management approach that has living resolutions grounded in natural processes and structures. NbSs offer many other environmental benefits over traditional grey infrastructure, including reduced air pollution and climate change mitigation. This review predominantly centers on the hydrological aspect of NbSs and furnishes a condensed summary of the collective understanding about NbSs as an alternatives for stormwater management. In this study, which employed the CIMO (Context, Intervention, Mechanism, Outcome) framework, a corpus of 187 NbS-related publications (2000–2023) extracted from the Web of Science database were used, and we expounded upon the origins, objectives, and significance of NbSs in urban runoff and climate change, and the operational mechanisms of NbSs (including green roofs, permeable pavements, bioretention systems, and constructed wetlands), which are widely used in urban stormwater management, were also discussed. Additionally, the efficacy of NbSs in improving stormwater quality and quantity is discussed in depth in this study. In particular, the critical role of NbSs in reducing nutrients such as TSS, TN, TP, and COD and heavy metal pollutants such as Fe, Cu, Pb, and Zn is emphasized. Finally, the main barriers encountered in the promotion and application of NbSs in different countries and regions, including financial, technological and physical, regulatory, and public awareness, are listed, and future directions for improving and strategizing NbS implementation are proposed. This review gathered knowledge from diverse sources to provide an overview of NbSs, enhancing the comprehension of their mechanisms and applications. It underscores specific areas requiring future research attention.

1. Introduction

The burgeoning expansion of global urban landscapes, in conjunction with the multifarious impacts of climate change, is precipitating significant alterations in urban hydrological dynamics [1]. There have been variations in land use/land cover (LULC) during the urbanization process, where natural surfaces such as cropland, grassland, and waters have been largely replaced by impermeable surfaces like buildings and roads [2,3,4]. This fundamental transformation alters the surface properties of urban areas, resulting in an increase in the imperviousness and runoff coefficient of urban surfaces. The rising imperviousness leads to increased surface runoff, reduced rainwater infiltration, diminished groundwater re-charging, and deteriorating water quality [5]. Simultaneously, natural surfaces and impermeable surfaces exhibit significant differences in the properties of light absorption, heat capacity, and rainfall runoff coefficients [6]. This transition affects the energy budget and water budget of urban areas, leading to changes in the local climate of urban areas, such as abnormal variations in surface temperature and the intensity and duration of heavy precipitation events [2,7]. This has likewise led to the urban heat island (UHI) and convective rainfall initiation generation [7]. Concurrently, climatic fluctuations are inducing changes in precipitation patterns. Over the past few centuries, there have been major changes in climate behavior at global and regional scales, altering the transport patterns and availability of water [8]. The increase in average temperatures due to global warming has accelerated the hydrological cycle, which has had a direct impact on rainfall, particularly in terms of the number and duration of extreme rainfall events [9]. In addition, in the context of climate change, precipitation changes have taken place not only regarding amount, but also distribution and state [10]. Climate change may lead to changes in the distribution and seasonality of rainfall, and, according to previous studies, the proportion of seasonal and snowfall rainfall has a non-negligible impact on the average annual net flow [11]. This confluence of factors amplifies the risk of urban flooding, posing substantial threats to human settlements, infrastructural resilience, and ecological systems [12,13,14].
In response to these hydrological challenges, various nations have embarked on the implementation of sustainable approaches and technological advancements in urban stormwater management. Prominent examples include the Best Management Practices (BMPs) in the United States, Active, Beautiful, Clean (ABC) Waters in Singapore, and the Sponge City initiative in China [15,16]. In European urban centers, nature-based solutions (NbSs) have garnered increasing attention as a strategy to combat urban issues like climate change, urban decay, and aging infrastructure. NbSs are characterized as strategies that utilize natural elements to address various challenges such as climate change, food security, water resources, and disaster risk management [17]. These approaches involve conserving and utilizing biodiversity sustainably. They include green roofs, bioretention systems, and the construction of wetlands that serve to reduce surface runoff and improve water quality and air quality. Another NbS is the increased provision of urban green spaces such as parks and street trees to ameliorate high temperatures in cities [18]. Whatever the NbS measures, they will be inspired and supported by nature. For example, permeable pavements mimic the natural surface water cycle by allowing rainwater and other precipitation to infiltrate through its surface layers to nourish the groundwater. In bioretention, plants carry out the processes of rainwater retention, uptake, and conversion, and natural surfaces carry out the processes of infiltration. These processes are carried out through a series of chemical, biological, and physical processes that mimic the natural ecology [19].
There has been a global surge in research validating the efficacy of NbSs in improving water quality and reducing stormwater volume [20,21]. For instance, Liu et al. [22] examined the impacts of various NbS approaches at the watershed level, and their study found that different levels and combinations of NBS practices reduced runoff volume by 0 to 26.47%, TN by 0.30 to 34.20%, TP by 0.27 to 47.41%, TSS by 0.33 to 53.59%, Pb by 0.30 to 60.98%, BOD by 0 to 26.70%, and COD by 0 to 27.52%. Thiagarajan et al. [23] explored the comprehensive benefits of NbSs in residential areas and noted that NbSs has the capacity to capture 56 billion liters of stormwater annually if all residential properties used NbSs. The result found by Versini et al. [24] shows that a combination of several NbSs can reduce the runoff volume about 90%. While the ecological services and effectiveness of NbSs are well documented, there remain gaps, particularly in the optimization of their hydrological benefits [25]. Despite the numerous advantages of NbSs, their widespread integration faces challenges, notably the lack of comprehensive guidelines for their design, implementation, and maintenance [26], along with prevalent misconceptions about high maintenance costs and complex application, which hinder their broader acceptance [27,28].
Acknowledging the critical importance of NbSs, this review aims to synthesize the existing body of literature on NbSs in the context of urban stormwater management, highlighting its advantages, challenges, and potential solutions. Specifically, this review seeks to (1) elucidate the origins, objectives, significance, and fundamental processing mechanisms of NbSs in the midst of urban and climatic transformations; (2) critically analyze the effectiveness of several NbS practices (specifically green roofs and bioretention) in reducing surface runoff and improving water quality (the results show that NbS practices can significantly reduce surface runoff, peak flow, and remove pollutants from runoff, including TN, TP, TSS, COD, and heavy metals); and (3) provide examples of the main barriers to NbS implementation in different countries or regions and propose comprehensive strategies for their sustainable integration.

2. Methods

This investigation meticulously appraises the hydrological efficacy of NbSs, primarily focusing on studies that elucidate runoff mitigation and water quality enhancement. To ensure methodological robustness, this review employs a structured research paradigm, enabling an exhaustive reconnaissance of pertinent works in the literature and a meticulous scrutiny of relevant studies. The analytical framework employed is the Context, Intervention, Mechanism, Outcome (CIMO) model, acclaimed for its analytical acumen in systematic literature reviews [29]. Within this paradigm:
  • Context delineates the specific environmental or situational backdrop of the study;
  • Intervention denotes the specific NbS or practice under examination;
  • Mechanism investigates the causative linkages between the intervention and its ensuing effects;
  • Outcome encapsulates the resultant effects or consequences engendered by the intervention, as driven by the identified mechanism [30].
The central research inquiry addressed herein is as follows: “In the milieu of urbanization and climatic fluctuations, how do NbS interventions modulate urban hydrological outcomes via their intrinsic mechanisms?”
Operationalizing the CIMO framework involved pinpointing key terminologies such as “Nature-based solutions”, “low impact development”, “runoff”, “rainwater”, “urban drainage”, “hydrological processes”, “flood management”, “best management practices”, “sponge city”, “sustainable urban drainage systems”, and “water-sensitive urban design”. These terminologies were assimilated into an extensive search protocol, employing Boolean operators (OR within categories and AND between categories) for scrutinizing titles, abstracts, and the core content of seminal papers. The initial screening of articles was predicated on criteria such as article genre (e.g., empirical studies, reviews), publication recency (prioritizing the last decade), and linguistic medium (English). Further refinement of article selection emphasized the caliber and pertinence of the literature [31], with exclusion criteria disqualifying studies that did not directly address NbSs in urban hydrological contexts or those devoid of empirical data.
A systematic analysis of NbS research from 2000 to 2023 was carried out using the core collection database of Web of Science. The principal data repository was the Web of Science database, selected for its comprehensive scope and scientific rigor [32,33]. To augment the scope of the literature search, additional databases such as Scopus and Google Scholar were also perused. The selected articles were subjected to an in-depth content analysis, focusing on the categorization and delineation of NbS practices and their respective benefits. The synthesis approach amalgamated both qualitative and quantitative methodologies, facilitating a nuanced comprehension of NbS applications and outcomes. This hybrid methodology permits an all-encompassing exploration of the multifarious aspects of NbS efficiency in urban stormwater management. The section culminates with an acknowledgment of potential methodological constraints, such as the reliance on published academic literature, potentially excluding grey literature or unpublished studies, and the concentration on English-language articles, which may inadvertently neglect significant research published in other languages.

3. Review Results and Discussion

3.1. Literature Search Results

An advanced query within the SCI–EXPANDED (Science Citation Index Expanded) subset of the Web of Science Core Collection from 2000 to 2023 yielded 1907 documents using key terms such as “Nature-based solution”, “Best management practice”, “Water-sensitive urban design”, “Sponge city”, “Green infrastructure”, “Low impact development”, and SUDS in conjunction with hydrological variables like runoff, rainfall, drainage, and flood. This initial corpus was meticulously refined by excluding non-article document types such as reviews and meeting abstracts, which may not encapsulate a comprehensive NbS practice process. This filtration resulted in 161 articles being excluded. A subsequent thematic culling excluded disparate fields such as computer science and metallurgy, leaving 548 studies in more relevant domains. A closer examination of abstracts and titles further distilled the collection, focusing on studies directly investigating NbSs’ impact on urban hydrology, culminating in 187 studies being selected for in-depth analysis (Figure 1).

3.2. Overview of Nature-Based Solutions

3.2.1. Historical Trajectory of Nature-Based Solutions

The genesis of NbSs is anchored in the interplay between biodiversity and human welfare, initially surfacing in the ecosystem services discourse of the 1970s [34]. The late 20th century witnessed a pivotal shift in conservation strategies, transitioning from an anthropocentric focus to a more holistic human–nature synergy [35]. This transition marked the evolution from passive nature beneficiaries to active ecosystem stewards. The European Commission’s recent definition encapsulates NbSs as multifaceted, cost-effective approaches that bolster resilience while yielding environmental, social, and economic dividends [35]. This conceptual evolution has been paralleled by a diversification in methodologies, although terminologies vary across different stormwater control frameworks [36,37].

3.2.2. Intervention of Nature-Based Solution Implementations

NbSs are predicated on harnessing natural systems and processes to foster functionally equivalent hydrologic landscapes, eschewing traditional infrastructural approaches. This paradigm emphasizes natural processes like evapotranspiration and infiltration, integrating botanical, geomaterial, and bioengineering elements to modulate stormwater dynamics [38]. Empirical evidence substantiates NbSs’ role in addressing urban hydrological challenges, including runoff modulation and flood mitigation, while offering cost-effective alternatives to conventional methods [39,40]. Additionally, NbSs’ assimilation into urban landscapes augments ecological vitality, offering a spectrum of ecosystem services [41].

3.2.3. Mechanistic Underpinnings of NbSs for Stormwater and Contaminant Regulation

Deciphering NbSs’ mechanics is essential for optimized stormwater management. NbSs orchestrate a suite of processes—such as soil infiltration and vegetative absorption—to effectively mitigate stormwater runoff and modulate peak flows. Specifically, in the case of bioretention, for example, when stormwater runoff enters the bioretention system, it will first pass through the vegetation layer, where the vegetation will absorb and retain some of the water. Then, the residual water flow will enter the soil layer, and through the porous structure and water absorption capacity of the soil, the water will gradually penetrate deep into the soil until it reaches the storage layer for the temporary storage of stormwater. The stormwater is finally either discharged downstream through underdrain pipes or infiltrated into the subsoil below the soil bed [19]. During this process, some of the water may overflow the bioretention system due to soil saturation or unfavorable topography. While some water will be retained in depressions or between vegetation, this water may be absorbed by plants or evaporate directly. Through plant transpiration, water is returned to the atmosphere as water vapor. In the case of green roofs, similar to bioretention, rainfall is intercepted through a vegetative layer, allowed to infiltrate into a substrate layer, and then stored in the storage layer and finally drained through a drainage layer. Concurrently, these systems facilitate pollutant filtration and biotransformation through synergistic interactions among plant species and microbial consortia [42]. This mechanistic insight underscores the importance of strategic species and matrix selection to maximize hydrological efficiency within the NbS framework.

4. Leveraging NbS Techniques for Urban Stormwater Management

NbSs have been empirically validated as potent tools for stormwater management and the enhancement of urban hydrological systems. This review highlights four seminal NbS implementations, each chosen based on a comprehensive set of criteria: Green Roofs, Permeable Pavement Systems, Bioretention Systems, and Constructed Wetlands.
Green Roofs: In the face of escalating urbanization, the proliferation of built structures within cityscapes presents a unique opportunity. The integration of green roofs on these structures offers multifarious benefits. These include the mitigation of surface runoff, the enhancement of urban greenery, the moderation of indoor temperatures, and subsequent energy conservation. The convergence of these factors underscores the criticality of research into green roof technology.
Permeable Pavement Systems: These systems are distinguished by their versatility and minimal land use requirements, negating the need for additional land acquisition [43]. Their adaptability across varied urban contexts and cost-effectiveness compared to traditional pavement materials underscore their practical significance.
Bioretention Systems: Designed as landscaped depressions, bioretention systems not only optimize urban hydrological functions but also integrate aesthetically with urban landscapes. These systems serve a dual purpose of water management and urban beautification.
Constructed Wetlands: Representing a larger-scale NbS intervention, constructed wetlands are paramount in addressing water eutrophication and the removal of heavy metals. They exemplify the capacity of NbSs to tackle complex environmental challenges at a substantial scale.
Collectively, these NbS implementations are instrumental in reducing runoff, thereby mitigating urban flood risks. Their hydrological benefits have been substantiated through extensive field studies and advanced modeling simulations [44], demonstrating their efficacy in urban stormwater management.

4.1. Green Roofs (GRs)

Green roofs, multifunctional components of urban ecosystems, serve as water storage matrices while providing a suite of ecosystem benefits (see Figure 2). These benefits encompass thermal regulation, air pollution mitigation, and urban hydrology enhancement. In densely populated urban areas, where terrestrial green spaces are limited, green roofs emerge as pivotal repositories of urban biodiversity and essential green spaces [45].

4.1.1. Efficacy in Stormwater Runoff Retention

Table 1 delineates various studies highlighting green roofs’ capacity to mitigate stormwater runoff and diminish peak flow. Nonetheless, their hydrological efficacy is subject to a confluence of factors: climatic conditions, geometric design, substrate properties, drainage layer depth, vegetation type, and stormwater metrics [46,47].
Substrate depth and composition are critical in optimizing stormwater retention capabilities [48,49]. Vegetation type plays a salient role, particularly under arid conditions and elevated temperatures [50]. Research by Nagase and Dunnett reveals that grasses outperform other plant types, such as forbs and sedum, in stormwater retention [51]. Carter and Rasmussen report a correlation between rainfall intensity and runoff retention, noting a higher retention rate for lighter rainfall, and that runoff is reduced by just under 90% when rainfall is less than 2.54 cm and by just under 50% when rainfall is greater than 2.54 cm [52]. Villarreal and Bengtsson identified an inverse relationship between rainfall intensity and water retention [53]. Further, Li et al. [54] expounded on the hierarchical impact of various factors on runoff retention and peak flow reduction.
Table 1. The application of green roofs for the reduction of runoff peak flow.
Table 1. The application of green roofs for the reduction of runoff peak flow.
CountryGreen Roof InformationSite CharacteristicsRunoff/Outflow Reduction (%)Peak Flow Reduction (%)Other/NotesReference
China100 cm long × 100 cm wide × 40 cm high
Substrates (10 cm in depth)
The mean annual precipitation is 587 mm81.00–87.0083.00–87.00Natural rainfall eventsZhang et al. [48]
ChinaMainly refers to the national standard for the technical specifications of green roof construction Study areas under 2.70 hectares31.40–69.8019.80–65.20In the 5-year period rainfall eventsYao et al. [55]
ItalyN/AResidential area, imperviousness of 96.0%25.90–62.8031.40–83.80The rainfall duration was assumed 30 min and the time-to-peak ratio 0.4Palermo et al. [56]
ChinaFour types of vegetation cover (Portulaca grandiflora, Sedum lineare, Festuca elata, and bare substrate)Subhumid continental monsoon climate in north temperate zone41.70–54.2050.60–59.10The heaviest rainfall event during the observation period (81.4 mm)Ge and Zhang [57]
GreeceSubstrate depth is 8 cm or 16 cm while plant species is Sedum or origanum or no vegetationN/A22.80–62.0056.90–79.10The duration of the studied rainfall events ranged between 50 min and 2640 minSoulis et al. [58]
EnglandThe test bed (3 × 1 m) comprised a sedum vegetation layer growing in 80 mm of substrateLocated in typical extensive green roof build-up0.04–99.9519.81–99.93Rain depth (mm) between 8.80 mm and 99.6 mmStovin et al. [59]

4.1.2. Efficacy in Augmenting Stormwater Quality

Table 2 provides insights into green roofs’ proficiency in reducing pollutants such as Total Suspended Solids (TSS), Chemical Oxygen Demand (COD), Total Nitrogen (TN), and Total Phosphorus (TP). Green roofs act as contaminant filters, leveraging processes like filtration, sedimentation, adsorption, plant uptake, and biodegradation [60]. However, discrepancies in effluent quality are observed, attributed to design variability, construction methods, substrate types, and environmental factors [61,62]. Studies in diverse geographical locales by Liu et al. [63] and Koc et al. [64] demonstrate significant pollutant reductions. The results of Liu et al. showed that the removal ratios of GR for TSS, COD, and TN reached 31.60%, 25.10%, and 37.8%, respectively, while Koc et al. demonstrated 41.15%, 39.73%, and 29.58%. Moreover, green roofs have shown efficacy in heavy metal sequestration from precipitation, as indicated in Table 3.
Conversely, instances of elevated pollutant concentrations, notably total phosphorus and nitrate, have been observed in green roof runoff [47]. For example, Gong et al. [81] noted a doubling of nitrate concentrations following rainfall events on green roofs. Razzaghmanesh et al. [82] reported a significant increase in nitrate concentrations in Adelaide, Australia. Such phenomena may be exacerbated by the use of phosphorus-rich fertilizers, as evidenced by Castro et al. [83], who observed a substantial increase in TP concentrations in Porto Alegre, Brazil.
While green roofs offer a multitude of ecological and hydrological benefits in urban environments, their impact on stormwater quality demands detailed examination. This necessitates comprehensive research to fully understand and optimize the multifaceted roles and benefits of green roofs in urban ecosystems.

4.2. Permeable Pavement Systems (PPSs)

Permeable Pavement Systems (PPSs) represent a forward-thinking approach to stormwater management, distinguished by their capacity to facilitate water percolation into the underlying soil layer. This unique characteristic positions PPSs as a potent tool in mitigating urban flood risks, particularly in densely urbanized areas where conventional green spaces are scarce (Figure 3) [84]. PPSs significantly contribute to urban infrastructural resilience by maintaining natural hydrological cycles, thereby enhancing groundwater recharge and reducing stormwater runoff. Given that pavements account for a substantial portion (20–40%) of urban areas, the transition from traditional impervious surfaces to permeable alternatives marks a critical shift towards sustainable urban water management [85].

4.2.1. Efficacy in Stormwater Runoff Retention

PPSs demonstrate a variable retention capacity for rainwater, typically between 16 and 66%, resulting in marked reductions in drainage pressure, hydraulic flow rates, and the frequency of stormwater influx [86]. Empirical studies underscore this efficacy: Kumar et al. [87] observed that PPSs in Illinois significantly exceeded regional median rainfall infiltration rates. Computational models by Liu et al. [14] in Beijing’s Haidian District indicated the peak flows were reduced 37.9–35.7%, respectively, under 1-, 2-, 5-, and 10-year storm events, attributed to the conversion of impervious surfaces to permeable materials. Park et al. [88], in Busan, South Korea, further demonstrated the varying efficacy of PPS types in reducing total and peak flows. The peak flow decreased by 9.1%, 10.7%, 5.9%, and 15.8%, and the total outflow decreased by 3.6%, 3.1%, 1.4%, and 16.3%, respectively. Wang et al. [89] highlighted PPSs’ heightened performance during less intense, shorter-duration rainfall, although its effectiveness diminished under more severe storm conditions.

4.2.2. Efficacy in Augmenting Stormwater Quality

PPSs inherently excel in pollutant sequestration, utilizing mechanisms such as interception, filtration, sedimentation, nutrient transformation, and microbial degradation. This is largely due to their porous structure [90,91]. Removal rates within the PPS matrix are notable, with TSS and COD extraction ranging broadly, and significant reductions in heavy metals such as Cd, Cu, Pb, and Zn. Liu et al. [92] correlated pollutant removal with factors like rainfall intensity and gravel layer size. The results showed that for permeable paving, the removal rates of TSS, Cd and Cu were less affected by rainfall intensity, with removal rates ranging from 93.76 to 98.66%, 92.84 to 95.97%, and 90.49 to 98.26%, respectively. A comprehensive study by Mahmoud et al. [93] in Texas, USA, reported average reductions of 76% for 23 samples, 56% for 19 samples, and 12% for 11 samples in TSS, BOD, and Escherichia coli. However, the variability in PPSs’ pollutant removal efficiency is significant, as evidenced by Wang et al. [94], who noted fluctuating removal rates for TP and TN; the TP and TN removal efficiencies of the systems can vary from 6 to 68% and 5 to 99%, respectively. A major challenge for PPSs is substrate clogging, primarily due to TSS accumulation during rainfall events [95]. Substrate composition, particularly concrete with smaller aggregates and higher air voids, has been shown to offer enhanced purification capabilities [96].

4.3. Bioretention Systems (BRs)

Bioretention systems have been acknowledged as a critical on-site intervention for urban stormwater management, principally facilitating permeation and evaporation to replicate pre-development hydrological conditions (Figure 4) [20]. These systems are adeptly integrated into urban landscapes, offering a sustainable alternative to conventional stormwater infrastructure, particularly in densely populated urban areas [97].

4.3.1. Efficacy in Stormwater Runoff Retention

The proficiency of bioretention systems in stormwater runoff retention is noteworthy, with peak discharge rate reductions ranging from 40% to an impressive 99% [98]. Table 4 delineates the effectiveness of bioretention cells in reducing stormwater runoff and attenuating peak flow. Investigations by Lucke and Nichols in Caloundra, Australia, revealed variable reductions in runoff volume and peak outflow [99]. They conducted four experiments on each of the three bioretention systems, and the results of the experiments showed that the percentage reduction in peak outflow after treatment with the three bioretention systems ranged from 79.5% to 93.6%. The role of infiltration and evapotranspiration is emphasized by Li et al. [100] of Maryland, who showed that approximately 20–50% more of the runoff entering bioretention was lost to infiltration and evapotranspiration. Winston et al. [101] reported runoff reductions in northeast Ohio in areas with low-permeability soil: runoff and peak flows were reduced by 36.00–59.00% and 24.00–96.00%, respectively. Hydraulic conductivity plays a pivotal role, with higher conductivity substrates facilitating extended interactions between the substrate and pollutants [102]. However, excessively high hydraulic conductivity could undermine the comprehensive treatment of contaminants [103]. The use of various substrates like biochar and zeolite has been explored to enhance specific characteristics such as hydraulic conductivity and plant growth [104]. Substrate depth also significantly influences hydrological performance, with deeper substrates yielding superior retention compared to shallower ones [100]. The role of flora in improving hydrological and hydraulic metrics is undeniable, with indigenous vegetation often preferred for its resilience [105].

4.3.2. Efficacy in Augmenting Stormwater Quality

Bioretention systems are highly effective in mitigating stormwater pollutants. Table 5 presents their capacity to attenuate TSS, COD, TN, and TP. Research by Shrestha et al. [110] demonstrated significant reductions in these pollutants, TSS, TN, and TP were reduced by 91.00–97.00%, 38.00–57.00%, and 86.00–94.00%, respectively. Jhonson et al. showed that bioretention removed 90.00%, 92.50%, 86.40%, and 93.50% for TSS, COD, TN, and TP, respectively. Substrate adsorption, filtration, and sedimentation play key roles in TSS removal, while the abatement of heavy metals depends on fillers and processes like precipitation and ion exchange [111,112]. Gülbaz et al. [113] showed varied metal removal rates due to different cation exchange capacities. The order of metal removal percentages was found to be Pb > Cu > Zn. Calculated delay factors ranged from 5 to 910 for Zn, 20 to 3600 for Cu, and 100 to 27,000 for Pb, with turf having the highest delay factor and gravel having the lowest. Substrate depth is a critical factor, affecting the interactions between the substrate, stormwater, and plants, thus enhancing contaminant removal [114].
Traditional bioretention systems often show limited nitrogen removal efficiency due to aerobic conditions within substrates. Innovative approaches, such as creating anaerobic environments at the base, have significantly improved NO3 and TN removal [122]. The addition of carbon sources like wood chips to the submerged zone has proven effective in reducing nitrate leaching [31]. Plants play a crucial role in the removal of organic matter, nitrogen, and phosphorus, although their effectiveness varies with the type of pollutant [126]. Plants with longer roots are typically more efficient in nitrogen and phosphorus removal [127].

4.4. Constructed Wetlands (CWs)

Constructed Wetlands (CWs) are engineered ecosystems designed to mimic the functions and features of natural wetlands. They are proficient in eliminating a range of contaminants through the synergistic action of phytoremediation, facilitated by aquatic plants, and in bioremediation, conducted by microorganisms in the soil and rhizosphere, as illustrated in Figure 5 [128]. CWs are categorized based on their flow regimes into free water surface (FWS), subsurface flow (SSF)—which includes horizontal (HSSF) and vertical (VSSF) subsurface flow—and the emerging floating treatment wetlands (FTWs). FTWs are particularly adept at coping with the hydrological fluctuations of storm events, due to their unique floating vegetation matrices [129]. As a solution to the technical challenges of traditional stormwater management, especially in fluctuating hydrological conditions, FTWs can be seamlessly integrated into existing water bodies, negating the need for additional land use [130].

4.4.1. Efficacy in Stormwater Runoff Retention

While CWs are renowned for pollutant removal, their role in stormwater runoff attenuation is less emphasized due to their primary design focus on wastewater treatment [131]. Research into their potential as stormwater volume regulators is limited. Urban CWs face challenges such as land availability constraints and, in the case of FWS CWs, a minimal risk of pathogenic transmission. However, studies highlight their effectiveness in stormwater control. A North Carolina wetland demonstrated significant reductions in runoff volumes and peak flows by 54% and 80%, respectively [132]. Liu et al. [131] reported average reductions of 29.6% in runoff volume and 98.7% in peak flow across various rainfall events. Innovative approaches integrating CWs with urban drainage systems have also been explored. Rizzo et al. [133] developed a model evaluating the hydrological benefits of CWs in tandem with combined sewer overflow (CSO) systems, observing peak flow reductions up to 95.4%.

4.4.2. Efficacy in Augmenting Stormwater Quality

CWs harness a complex interplay of physical, chemical, and biological processes, including volatilization, mineralization, and biological degradation [128]. An investigation into a North Carolina stormwater wetland revealed substantial reductions of 42%, 36%, 47%, and 49% for NH4+-N, TN, TP, and TSS [132]. Chunbo et al. [134] found that optimizing the carbon-to-nitrogen ratio within CWs can significantly enhance NH4+-N, NO3-N, TN, and TP removal; removal rates were 19.0–75.3%, 63.6–96.1%, 61.94–74.4%, and 75.0–98.8%, respectively. FTWs, due to their increased biomass interaction interfaces, amplify pollutant removal efficacy by enhancing plant–microbe symbiosis and facilitating nutrient assimilation [135]. Studies in Singapore and New Zealand have demonstrated the potential of FTWs in removing TN and TP, as well as in assimilating heavy metals [136]. However, Maxwell et al. [137] caution that FTW enhancements in stormwater ponds may not always yield optimal water quality improvements.

4.5. Combination NbS Practices

In the evolving landscape of urban hydrology and water quality improvement, most evaluations of NbSs have traditionally focused on single-strategy implementations. However, recent research indicates a significant synergistic enhancement when multiple NbS approaches are integrated. This enhanced efficacy arises from the comprehensive application and mutual reinforcement provided by combined systems [138]. For instance, a singular NbS approach resulted in an 18.62% reduction in runoff and a 16.61% decrease in peak flow. Conversely, a composite Green Stormwater Infrastructure approach, incorporating infiltration trenches, rain barrels, and bioretention systems, led to a more substantial reduction of 42.95% in runoff and 31.17% in peak flow [138]. Echoing these findings, Liu et al. [22] demonstrated that integrated NbS solutions notably outperformed individual strategies in reducing runoff volume and pollutant load.
Delving into computational analysis, Luan et al. [139] explored the impact of a strategic combination of diverse NbS tools, including a concave greenbelt with a significant 50% concavity ratio, permeable pavements, bioretention cells, and vegetative swales. This comprehensive approach markedly surpassed the effectiveness of isolated methods, achieving an impressive peak discharge reduction of 55.7% across various rainfall events. However, it is crucial to recognize the variability and site-specific characteristics in NbS implementation. This necessitates the tailored customization of NbS combinations, ensuring that they are optimally aligned with the unique environmental contexts they are designed to enhance.

5. Barriers and Strategies for NbS Implementation

5.1. Barriers and Challenges

GSI has demonstrated great ecological benefits in urban hydrological management and has become an integral part of sustainable urban planning [140]. In the last two decades, there has been an upsurge in research on NbSs due to the increasing intensity and frequency of urban flooding events triggered by urbanization and climate change, which have caused substantial economic losses to human societies. These catastrophic experiences have prompted the public and stakeholders to explore NbSs as a sustainable solution. Despite the burgeoning recognition of NbSs in stormwater management, their adoption is often eclipsed by conventional Grey Infrastructure for Rainwater and Effluent (GREI) due to multifarious impediments. Variability in NbS applications and developmental paradigms across nations are significant, particularly in developing countries, where gaps exist in professional expertise, financial resources, and other essential areas. Additionally, varying climatic and pedological conditions impede the straightforward transplantation of NbS methodologies across regions. This discord between theoretical frameworks and practical application has notably hindered global NbS deployment. Predominant barriers to NbS implementation include financial constraints, technical and physical limitations, regulatory challenges, and public awareness issues.

5.1.1. Financial Constraints

While NbSs employ economically viable materials for micro-scale urban interventions, large-scale deployments demand substantial capital investments, constituting a significant economic hurdle [141]. Current research into cost–benefit analyses of NbSs suggests that although lifecycle costs are lower than traditional GREI, initial expenditures and ongoing maintenance costs are comparatively higher [89]. Presently, governmental funding is the primary financial source for NbS projects, with private investment remaining limited due to disproportionate returns. Moreover, ancillary benefits such as social welfare enhancement and land value appreciation are often underappreciated, thus deterring governmental financial backing.

5.1.2. Technical and Physical Limitations

The dearth of robust technical support impedes the widespread adoption of NbSs, leading to disparities in implementation success. Despite the availability of technical guidelines, a lack of expertise and skilled labor across design, construction, and maintenance phases poses significant challenges [142]. Data limitations in areas like hydrological characteristics and construction methodologies obstruct precise computational modeling, thereby undermining public confidence in NbSs [143]. Furthermore, diverse interpretations of NbSs, efficacy uncertainties, and the absence of localized benchmarks hinder progress [144].

5.1.3. Regulatory Barriers

While NbSs emerged in the 20th century, only a few countries have developed comprehensive legal frameworks and institutional mechanisms for their implementation. For example, Guangzhou, China, has integrated the ‘sponge city’ concept into urban planning, but such initiatives are not universally observed. Legislative restrictions and a lack of intersectoral collaboration further constrain NbS expansion. In addition, in places such as the USA, where conflicts between local ordinances and private property rights are clear, conflicts can also complicate the attribution of responsibility for maintenance when NBS areas are located on private property [19].

5.1.4. Public Awareness Challenges

Public unfamiliarity with NbSs’ transformative potential in urban hydrological management and landscape enhancement remains a significant barrier [141]. For example, in American cities, the prevailing belief among residents is that the current price of municipal water is exceedingly low. Consequently, they view the adoption of new water harvesting techniques as lacking cost-effectiveness. This perception leads to reluctance in supporting new NbS practices [145]. Governmental entities play a critical role in advocating for NbSs, yet their efforts are often limited, underscoring the need for active public engagement and advocacy to foster broader acceptance and implementation.

5.2. Future Directions for Improving and Strategizing NbS Implementation

The advancement of NbSs is primarily driven by the urgent challenges faced by developing nations. Overcoming the barriers to NbS deployment necessitates the integrated efforts of various stakeholders, including government entities, the public, and the scientific community. A thorough assessment of NbSs, integrating social, economic, and environmental dimensions, is vital to highlight their benefits, particularly in the context of rapid urbanization and significant climatic shifts. The following strategies are proposed to address the previously identified challenges and offer guidance to urban planners and researchers.

5.2.1. Financial Strategy Enhancement

Governments should broaden and increase funding channels to support the ongoing operation and maintenance of NbSs. Collaborations between local authorities, the business sector, and private citizens are crucial for promoting NbS adoption (Shafique and Kim, 2017) [28]. Fiscal incentives and legislative support can help alleviate financial barriers, enhancing public–private partnerships. For example, China’s Ministry of Finance issued a notification in 2014 to provide financial subsidies of RMB 80 per m2 for three-star-level green buildings, while in Nanjing, Jiangsu Province, China, the design fee of the district is only RMB 20 per m2 [146]. A financial subsidy can effectively stimulate the use of NbSs by enterprises. Furthermore, in-depth investigations into the economic feasibility of NbSs are essential to tap into their potential financial benefits.

5.2.2. Theoretical Framework Development

Educational institutions are key in advancing technical knowledge of NbSs and providing operational and maintenance guidance. Collaborative educational initiatives with the corporate sector can produce a workforce skilled in NbSs. The development of technical guides tailored to regional climatic and geological specifics, focusing on aspects like NbS design, material selection, and plant species, is necessary. Comprehensive maintenance plans are crucial for maintaining NbS effectiveness. Augmented research, including cost-efficiency analyses, can propel NbS advancements and enrich their ecological benefits. Cross-disciplinary collaborations can reveal innovative, cost-effective NbS approaches.

5.2.3. Legislative and Regulatory Advancements

Addressing regulatory barriers requires careful policy recalibration and the creation of appropriate legal frameworks. An equitable legislative structure or institution can promote inter-sectoral collaboration and NbS implementation. Authoritative bodies should develop and enforce NbS-specific regulations, integrating them into urban infrastructure planning. For example, Australia has established an intergovernmental committee to provide guidance on the implementation of WSUD [147]. Establishing minimum standards for post-development NbS management, stormwater discharge, groundwater protection, and nonpoint source pollution mitigation is crucial [143]. Oversight agencies or existing entities should monitor NbS development to ensure the sustainable integration into urban transformation.

5.2.4. Public Advocacy and Awareness

Enhancing the public awareness of NbSs can facilitate governmental support and mitigate economic challenges. Educating property owners about environmental and infrastructural aspects can encourage the adoption of NbS elements like green roofs and rainwater harvesting systems in existing buildings. The successful implementation of NbSs depends on diverse commitments, encompassing financial investments, coordinated efforts, and sustained dedication. Dedicated periods for specialized education, public awareness, and rigorous research are necessary. With prudent management and continuous refinement, these challenges—technical, legislative, financial, and socio-cultural—can be overcome, strengthening the resilience and efficacy of NbSs.

6. Conclusions

Emerging as a pivotal paradigm in urban development, NbS practices have witnessed escalating global commendation, anchoring extensive empirical research and pragmatic applications. Intrinsically designed as a sophisticated strategy for land utilization, NbSs exhibit an exceptional proficiency in orchestrating stormwater runoff management, enhancing aquatic quality, and fortifying environmental preservation. This proficiency is accentuated by their reliance on decentralized, micro-scale interventions, which draw inspiration from ecologically harmonious designs. Through the lens of the CIMO logical framework, this review endeavored to dissect an intricate question: “In the backdrop of urbanization and climatic shifts, how do the mechanisms underpinning NbS practices influence urban hydrological outcomes?”
A meticulous analysis of international research, emphasizing the mitigation of surface runoff and the elevation of water quality, demystified a discernible variability in the operational success of NbSs. This variability can be ascribed to a spectrum of localized environmental dynamics and distinct geographical contexts. The text delineates four salient impediments constraining the universal adoption of NbSs: fiscal limitations, intricate technical prerequisites, regulatory encumbrances, and perceptual voids in the public sphere. Concurrently, it accentuates strategies meticulously crafted to navigate these impediments, emphasizing the criticality of tailoring NbS design standards to resonate with indigenous environmental nuances and galvanizing comprehensive research endeavors in the NbS domain.
To actualize a resilient and efficacious urban stormwater management framework, intensified synergies among a diverse array of stakeholders—inclusive of academic researchers, urban planning mavens, and policy connoisseurs—are indispensable. Such a confluence of expertise and vision promises to curate and promulgate NbS strategies that are both innovative and apt for mitigating urban inundation challenges.

Author Contributions

Conceptualization, M.W.; methodology, M.W. and C.S.; software, C.S.; validation, B.C. and J.L.; formal analysis, M.W. and D.Z.; investigation, C.S.; resources, M.W.; data curation, C.S.; writing—original draft preparation, C.S.; writing—review and editing, M.W. and R.M.A.I.; visualization, C.S. and X.L.; supervision, M.W. and Q.R.; funding acquisition, M.W. and D.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Guangdong Basic and Applied Basic Research Foundation, China, grant number 2023A1515030158, grant number 2023A1515012130; the Guangzhou Sci. Technol. Program, China, grant number 202201010431; the Maoming Sci. Technol. Program, China, grant number 2021S0054; Guangzhou University, RC2023008 and the Graduate Student Innovation Ability Training Funding Program of Guangzhou University, grant number 2023GDJC.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

The study did not involve humans.

Data Availability Statement

The study did not report any publicly archived datasets.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Do, T.A.T.; Do, A.N.T.; Tran, H.D. Quantifying the spatial pattern of urban expansion trends in the period 1987–2022 and identifying areas at risk of flooding due to the impact of urbanization in Lao Cai city. Ecol. Inform. 2022, 72, 101912. [Google Scholar] [CrossRef]
  2. Lin, L.; Gao, T.; Luo, M.; Ge, E.; Yang, Y.; Liu, Z.; Zhao, Y.; Ning, G. Contribution of urbanization to the changes in extreme climate events in urban agglomerations across China. Sci. Total Environ. 2020, 744, 140264. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, H.; Zou, L.; Xia, J.; Chen, T.; Wang, F. Impact assessment of climate change and urbanization on the nonstationarity of extreme precipitation: A case study in an urban agglomeration in the middle reaches of the Yangtze river. Sustain. Cities Soc. 2022, 85, 104038. [Google Scholar] [CrossRef]
  4. Shahid, M.; Cong, Z.; Zhang, D. Understanding the impacts of climate change and human activities on streamflow: A case study of the Soan River basin, Pakistan. Theor. Appl. Climatol. 2018, 134. [Google Scholar] [CrossRef]
  5. Chen, J.; Theller, L.; Gitau, M.W.; Engel, B.A.; Harbor, J.M. Urbanization impacts on surface runoff of the contiguous United States. J. Environ. Manag. 2017, 187, 470–481. [Google Scholar] [CrossRef]
  6. Dissanayake, D. Land Use Change and Its Impacts on Land Surface Temperature in Galle City, Sri Lanka. Climate 2020, 8, 65. [Google Scholar] [CrossRef]
  7. Xihui, G.; Zhang, Q.; Singh, V.; Changqing, S.; Peng, S.; Li, J. Potential contributions of climate change and urbanization to precipitation trends across China at national, regional and local scales. Int. J. Climatol. 2019, 39, 2998–3012. [Google Scholar] [CrossRef]
  8. Seyoum, M.; Gan, T.Y. Possible impact of climate change on future extreme precipitation of the Oldman, Bow and Red Deer River Basins of Alberta. Int. J. Climatol. 2015, 36, 208–224. [Google Scholar] [CrossRef]
  9. Shahid, M.; Rahman, K. Identifying the Annual and Seasonal Trends of Hydrological and Climatic Variables in the Indus Basin Pakistan. Asia-Pac. J. Atmos. Sci. 2020, 57, 191–205. [Google Scholar] [CrossRef]
  10. Cong, Z.; Shahid, M.; Zhang, D.; Lei, H.; Yang, D. Attribution of runoff change in the alpine basin: A case study of the Heihe Upstream Basin, China. Hydrol. Sci. J. 2017, 62, 1013–1028. [Google Scholar] [CrossRef]
  11. Feng, X.; Vico, G.; Porporato, A. On the effects of seasonality on soil water balance and plant growth. Water Resour. Res. 2012, 48, W05543. [Google Scholar] [CrossRef]
  12. Tao, W.; Bays, J.; Meyer, D.; Smardon, R.; Levy, Z.F. Constructed Wetlands for Treatment of Combined Sewer Overflow in the US: A Review of Design Challenges and Application Status. Water 2014, 2014, 3362–3385. [Google Scholar] [CrossRef]
  13. Gimenez-Maranges, M.; Breuste, J.; Hof, A. Sustainable Drainage Systems for transitioning to sustainable urban flood management in the European Union: A review. J. Clean. Prod. 2020, 255, 120191. [Google Scholar] [CrossRef]
  14. Liu, W.; Chen, W.; Peng, C. Assessing the effectiveness of green infrastructures on urban flooding reduction: A community scale study. Ecol. Model. 2014, 291, 6–14. [Google Scholar] [CrossRef]
  15. Davis, A. Green Engineering Principles Promote Low-Impact Development. Environ. Sci. Technol. 2005, 39, 338A–344A. [Google Scholar] [CrossRef]
  16. Ren, N.-Q.; Wang, Q.; Wang, Q.; Huang, H.; Wang, X.-H. Upgrading to urban water system 3.0 through sponge city construction. Front. Environ. Sci. Eng. 2017, 11, 9. [Google Scholar] [CrossRef]
  17. Kabisch, N.; Frantzeskaki, N.; Pauleit, S.; Naumann, S.; Davis, M.; Artmann, M.; Haase, D.; Knapp, S.; Korn, H.; Stadler, J.; et al. Nature-based solutions to climate change mitigation and adaptation in urban areas: Perspectives on indicators, knowledge gaps, barriers, and opportunities for action. Ecol. Soc. 2016, 21, 39. [Google Scholar] [CrossRef]
  18. Bowler, D.; Buyung, A.; Knight, T.; Pullin, A. Urban greening to cool towns and cities: A systematic review of the empirical evidence. Landsc. Urban Plan. 2010, 97, 147–155. [Google Scholar] [CrossRef]
  19. Li, C.; Peng, C.; Chiang, P.-C.; Cai, Y.; Wang, X.; Yang, Z. Mechanisms and applications of green infrastructure practices for stormwater control: A review. J. Hydrol. 2019, 568, 626–637. [Google Scholar] [CrossRef]
  20. Li, L.; Yang, J.; Davis, A.; Liu, Y. Dissolved Inorganic Nitrogen Behavior and Fate in Bioretention Systems: Role of Vegetation and Saturated Zones. J. Environ. Eng. 2019, 145, 04019074. [Google Scholar] [CrossRef]
  21. Zhuang, Y.; Zhang, L.; Du, Y.; Chen, G. Current patterns and future perspectives of best management practices research: A bibliometric analysis. J. Soil Water Conserv. 2016, 71, 98A–104A. [Google Scholar] [CrossRef]
  22. Liu, Y.; Bralts, V.; Engel, B. Evaluating the effectiveness of management practices on hydrology and water quality at watershed scale with a rainfall-runoff model. Sci. Total Environ. 2015, 511, 298–308. [Google Scholar] [CrossRef] [PubMed]
  23. Thiagarajan, M.; Newman, G.; Van Zandt, S. The Projected Impact of a Neighborhood-Scaled Green-Infrastructure Retrofit. Sustainability 2018, 10, 3665. [Google Scholar] [CrossRef] [PubMed]
  24. Versini, P.A.; Kotelnikova, N.; Poulhes, A.; Tchiguirinskaia, I.; Schertzer, D.; Leurent, F. A distributed modelling approach to assess the use of Blue and Green Infrastructures to fulfil stormwater management requirements. Landsc. Urban Plan. 2018, 173, 60–63. [Google Scholar] [CrossRef]
  25. Miller, J.D.; Vesuviano, G.; Wallbank, J.R.; Fletcher, D.H.; Jones, L. Hydrological assessment of urban Nature-Based Solutions for urban planning using Ecosystem Service toolkit applications. Landsc. Urban Plan. 2023, 234, 104737. [Google Scholar] [CrossRef]
  26. Liu, T.; Lawluvy, Y.; Shi, Y.; Yap, P.S. Low Impact Development (LID) Practices: A Review on Recent Developments, Challenges and Prospects. Water Air Soil Pollut. 2021, 232, 344. [Google Scholar] [CrossRef]
  27. Brown, H.; Bos, D.; Walsh, C.J.; Fletcher, T.; RossRakesh, S. More than money: How multiple factors influence householder participation in at-source stormwater management. J. Environ. Plan. Manag. 2014, 59, 79–97. [Google Scholar] [CrossRef]
  28. Shafique, M.; Kim, R. Retrofitting the Low Impact Development Practices into Developed Urban areas Including Barriers and Potential Solution. Open Geosci. 2017, 9, 240–254. [Google Scholar] [CrossRef]
  29. Denyer, D.; Tranfield, D. Producing a systematic review. In The Sage Handbook of Organizational Research Methods; Sage Publications Ltd.: Thousand Oaks, CA, USA, 2009; pp. 671–689. [Google Scholar]
  30. Colicchia, C.; Strozzi, F. Supply chain risk management: A new methodology for a systematic literature review. Supply Chain. Manag. Int. J. 2012, 17, 403–418. [Google Scholar] [CrossRef]
  31. Zhang, K.; Chui, T.F.M. Linking hydrological and bioecological benefits of green infrastructures across spatial scales—A literature review. Sci. Total Environ. 2018, 648, 1219–1231. [Google Scholar] [CrossRef]
  32. Olawumi, T.; Chan, D.D. A scientometric review of global research on sustainability and sustainable development. J. Clean. Prod. 2018, 183, 231–250. [Google Scholar] [CrossRef]
  33. Thomé, A.M.; Ceryno, P.; Scavarda, A.; Remmen, A. Sustainable infrastructure: A review and a research agenda. J. Environ. Manag. 2016, 184, 143–156. [Google Scholar] [CrossRef] [PubMed]
  34. Gómez-Baggethun, E.; de Groot, R.; Lomas, P.L.; Montes, C. The history of ecosystem services in economic theory and practice: From early notions to markets and payment schemes. Ecol. Econ. 2010, 69, 1209–1218. [Google Scholar] [CrossRef]
  35. Cohen-Shacham, E.; Andrade, A.; Dalton, J.; Dudley, N.; Jones, M.; Kumar, C.; Maginnis, S.; Maynard, S.; Nelson, C.R.; Renaud, F.G.; et al. Core principles for successfully implementing and upscaling Nature-based Solutions. Environ. Sci. Policy 2019, 98, 20–29. [Google Scholar] [CrossRef]
  36. Fletcher, T.D.; Shuster, W.; Hunt, W.F.; Ashley, R.; Butler, D.; Arthur, S.; Trowsdale, S.; Barraud, S.; Semadeni-Davies, A.; Bertrand-Krajewski, J.L.; et al. SUDS, LID, BMPs, WSUD and more—The evolution and application of terminology surrounding urban drainage. Urban Water J. 2015, 12, 525–542. [Google Scholar] [CrossRef]
  37. Vogel, J.; Moore, T.; Coffman, R.; Rodie, S.; Hutchinson, S.; McDonough, K.; McLemore, A.; McMaine, J. Critical Review of Technical Questions Facing Low Impact Development and Green Infrastructure: A Perspective from the Great Plains. Water Environ. Res. 2015, 87, 849–862. [Google Scholar] [CrossRef]
  38. Mwangi, J.; Shisanya, C.; Gathenya, J.; Namirembe, S.; Moriasi, D. A modeling approach to evaluate the impact of conservation practices on water and sediment yield in Sasumua Watershed, Kenya. J. Soil Water Conserv. 2015, 70, 75–90. [Google Scholar] [CrossRef]
  39. Roseen, R.; Janeski, T.; Simpson, M.; Houle, J.; Gunderson, J.; Ballestero, T. Economic and Adaptation Benefits of Low Impact Development. In Proceedings of the 2011 Low Impact Development Conference, Philadelphia, PA, USA, 25–28 September 2011. [Google Scholar]
  40. Xu, Z.; Dong, X.; Zhao, Y.; Du, P. Enhancing resilience of urban stormwater systems: Cost-effectiveness analysis of structural characteristics. Urban Water J. 2021, 18, 850–859. [Google Scholar] [CrossRef]
  41. Addo-Bankas, O.; Zhao, Y.; Gomes, A.; Stefanakis, A. Challenges of Urban Artificial Landscape Water Bodies: Treatment Techniques and Restoration Strategies towards Ecosystem Services Enhancement. Processes 2022, 10, 2486. [Google Scholar] [CrossRef]
  42. Grebel, J.; Mohanty, S.; Torkelson, A.; Boehm, A.; Higgins, C.; Maxwell, R.; Nelson, K.; Sedlak, D. Engineered Infiltration Systems for Urban Stormwater Reclamation. Environ. Eng. Sci. 2013, 30, 437–454. [Google Scholar] [CrossRef]
  43. Hung, A.; Li, L.; Swei, O. Evaluation of permeable highway pavements via an integrated life-cycle model. J. Clean. Prod. 2021, 314, 128043. [Google Scholar] [CrossRef]
  44. Baek, S.; Ligaray, M.; Pachepsky, Y.; Chun, J.A.; Yoon, K.-S.; Park, Y.; Cho, K. Assessment of a green roof practice using the coupled SWMM and HYDRUS models. J. Environ. Manag. 2020, 261, 109920. [Google Scholar] [CrossRef] [PubMed]
  45. Hong, W.Y.; Guo, R.Z.; Tang, H. Potential assessment and implementation strategy for roof greening in highly urbanized areas: A case study in Shenzhen, China. Cities 2019, 95, 102468. [Google Scholar] [CrossRef]
  46. Speak, A.; Rothwell, J.; Lindley, S.; Smith, C. Metal and nutrient dynamics on an aged intensive green roof. Environ. Pollut. 2014, 184, 33–43. [Google Scholar] [CrossRef] [PubMed]
  47. Alim, M.A.; Rahman, A.; Tao, P.Z.; Garner, B.; Griffith, R.; Liebman, M. Green roof as an effective tool for sustainable urban development: An Australian perspective in relation to stormwater and building energy management. J. Clean. Prod. 2022, 362, 132561. [Google Scholar] [CrossRef]
  48. Zhang, S.; Lin, Z.; Sunxun, Z.; Ge, D. Stormwater retention and detention performance of green roofs with different substrates: Observational data and hydrological simulations. J. Environ. Manag. 2021, 291, 112682. [Google Scholar] [CrossRef] [PubMed]
  49. Kazemi, F.; Mohorko, R. Review on the Roles and Effects of Growing Media on Plant Performance in Green Roofs in World Climates. Urban For. Urban Green. 2017, 23, 13–26. [Google Scholar] [CrossRef]
  50. Dunnett, N.; Nagase, A.; Hallam, A. The dynamics of planted and colonising species on a green roof over six growing seasons 2001–2006: Influence of substrate depth. Urban Ecosyst. 2008, 11, 373–384. [Google Scholar] [CrossRef]
  51. Nagase, A.; Dunnett, N. Amount of water runoff from different vegetation types on extensive green roofs: Effects of plant species, diversity and plant structure. Landsc. Urban Plann. 2012, 104, 356–363. [Google Scholar] [CrossRef]
  52. Carter, T.; Rasmussen, T. Hydrologic behavior of vegetated roofs. J. Am. Water Resour. Assoc. 2006, 42, 1261–1274. [Google Scholar] [CrossRef]
  53. Villarreal, E.; Bengtsson, L. Response of a Sedum green-roof to individual rain events. Ecol. Eng. 2005, 25, 1–7. [Google Scholar] [CrossRef]
  54. Li, Y.; Liu, J. Green roofs in the humid subtropics: The role of environmental and design factors on stormwater retention and peak reduction. Sci. Total Environ. 2022, 858, 159710. [Google Scholar] [CrossRef] [PubMed]
  55. Yao, L.; Wu, Z.; Wang, Y.; Sun, S.; Wei, W.; Xu, Y. Does the spatial location of green roofs affects runoff mitigation in small urbanized catchments? J. Environ. Manag. 2020, 268, 110707. [Google Scholar] [CrossRef] [PubMed]
  56. Palermo, S.A.; Talarico, V.; Turco, M. On the LID systems effectiveness for urban stormwater management: Case study in Southern Italy. IOP Conf. Ser. Earth Environ. Sci. 2020, 410, 012012. [Google Scholar] [CrossRef]
  57. Ge; Zhang, S.-H. Impacts of Vegetation on Hydrological Performances of Green Roofs under Different Rainfall Conditions. Environ. Sci. 2018, 39, 5015–5023. [Google Scholar] [CrossRef]
  58. Soulis, K.; Valiantzas, J.; Ntoulas, N.; Kargas, G.; Nektarios, P. Simulation of green roof runoff under different substrate depths and vegetation covers by coupling a simple conceptual and a physically based hydrological model. J. Environ. Manag. 2017, 200, 434–445. [Google Scholar] [CrossRef] [PubMed]
  59. Stovin, V.; Vesuviano, G.; Kasmin, H. The hydrologic performance of a green roof test bed under UK climatic conditions. J. Hydrol. 2012, 414, 148–161. [Google Scholar] [CrossRef]
  60. Todorov, D.; Driscoll, C.; Todorova, S.; Montesdeoca, M. Water quality function of an extensive vegetated roof. Sci. Total Environ. 2018, 625, 928–939. [Google Scholar] [CrossRef] [PubMed]
  61. Kuppusamy, V.; Joshi, U.; Balasubramanian, R. A Field Study to Evaluate Runoff Quality from Green Roofs. Water Res. 2012, 46, 1337–1345. [Google Scholar] [CrossRef]
  62. Kuppusamy, V.; Joshi, U. Can green roof act as a sink for contaminants? A methodological study to evaluate runoff quality from green roofs. Environ. Pollut. 2014, 194, 121–129. [Google Scholar] [CrossRef]
  63. Liu, R.; Stanford, R.; Deng, Y.; Liu, D.; Liu, Y.; Yu, S. The influence of extensive green roofs on rainwater runoff quality: A field-scale study in southwest China. Environ. Sci. Pollut. Res. 2020, 27, 12932–12941. [Google Scholar] [CrossRef]
  64. Koc, K.; Ekmekcioğlu, Ö.; Özger, M. An integrated framework for the comprehensive evaluation of low impact development strategies. J. Environ. Manag. 2021, 294, 113023. [Google Scholar] [CrossRef] [PubMed]
  65. Thomaidi, V.; Petousi, I.; Kotsia, D.; Kalogerakis, N.; Fountoulakis, M. Use of green roofs for greywater treatment: Role of substrate, depth, plants, and recirculation. Sci. Total Environ. 2021, 807, 151004. [Google Scholar] [CrossRef] [PubMed]
  66. Zhang, J.; Xu, C.; Fu, D.; Liu, W. Improvement of Pollutant Controlling Performance by Modified Aggregate Structure in Extensive Green Roof Substrate Layer. Environ. Sci. Water Res. Technol. 2022, 8, 1709–1718. [Google Scholar] [CrossRef]
  67. Dutta, A.; Sanchez Torres, A.; Vojinovic, Z. Evaluation of Pollutant Removal Efficiency by Small-Scale Nature-Based Solutions Focusing on Bio-Retention Cells, Vegetative Swale and Porous Pavement. Water 2021, 13, 2361. [Google Scholar] [CrossRef]
  68. Zhang, W.; Dai, J.; Che, W.; Sun, H. Runoff quality and quantity after usage of layered or mixed substrate green roofs: A laboratory study. Desalin. Water Treat. 2020, 178, 396–404. [Google Scholar] [CrossRef]
  69. Jeon, J.; Hong, J.; Jeon, M.; Shin, D.; Kim, L.-H. Assessment of hydrologic and environmental performances of green roof system for improving urban water circulation. Desalin. Water Treat. 2019, 161, 14–20. [Google Scholar] [CrossRef]
  70. Chai, H.; Tang, Y.; Su, X.; Wang, W.; Lu, H.; Shao, Z.; He, Q. Annual variation patterns of the effluent water quality from a green roof and the overall impacts of its structure. Environ. Sci. Pollut. Res. 2018, 25, 30170–30179. [Google Scholar] [CrossRef] [PubMed]
  71. Zhang, W.; Zhong, X.; Che, W.; Sun, H.; Zhang, H. A laboratory study to determine the use of polluted river sediment as a substrate for extensive green roofs. Water Sci. Technol. 2018, 78, 2247–2255. [Google Scholar] [CrossRef]
  72. Chandrasekaran, R.; Smith, C.; Memon, F.; Philip, L. Removal of chemical and microbial contaminants from greywater using a novel constructed wetland: GROW. Ecol. Eng. 2017, 106, 55–65. [Google Scholar] [CrossRef]
  73. Bui, X.-T.; Nguyen, T.; Phan, V.; Hien, V.; Dan, N.; Koottatep, T. Performance of wetland roof with Melampodium paludosum treating septic tank effluent. Desalin. Water Treat. 2013, 52, 1070–1076. [Google Scholar] [CrossRef]
  74. Zhou, S.J.; Ren, B.Z.; Deng, R.J. An Experimental Study on Green Water-storing Roof’s Removal Efficiency of Pollutants in Rainwater Runoff. In Proceedings of the Beijing International Environmental Technology Conference, Beijing, China, 16–19 October 2009; pp. 287–293. [Google Scholar]
  75. Guo, J.; Zhang, Y.; Che, S. Performance analysis and experimental study on rainfall water purification with an extensive green roof matrix layer in Shanghai, China. Water Sci. Technol. 2017, 77, wst2017582. [Google Scholar] [CrossRef]
  76. Kuppusamy, V.; Joshi, U. Application of seaweed as substrate additive in green roofs: Enhancement of water retention and sorption capacity. Landsc. Urban Plann. 2015, 143, 25–32. [Google Scholar] [CrossRef]
  77. Kuppusamy, V.; Raja, F. Design and development of green roof substrate to improve runoff water quality: Plant growth experiments and adsorption. Water Res. 2014, 63, 94–101. [Google Scholar] [CrossRef]
  78. Seidl, M.; Gromaire, M.-C.; Saad, M.; Gouvello, B. Effect of substrate depth and rain-event history on the pollutant abatement of green roofs. Environ. Pollut. 2013, 183, 195–203. [Google Scholar] [CrossRef]
  79. Gregoire, B.G.; Clausen, J.C. Effect of a modular extensive green roof on stormwater runoff and water quality. Ecol. Eng. 2011, 37, 963–969. [Google Scholar] [CrossRef]
  80. Sarah, A.; Ebbs, S.; Battaglia, L.; Retzlaff, B. Heavy metals in leachate from simulated green roof systems. Ecol. Eng. 2011, 37, 1709–1717. [Google Scholar] [CrossRef]
  81. Gong, Y.; Zhang, X.; Li, H.; Zhang, X.; He, S.; Miao, Y. A comparison of the growth status, rainfall retention and purification effects of four green roof plant species. J. Environ. Manag. 2021, 278, 111451. [Google Scholar] [CrossRef]
  82. Razzaghmanesh, M.; Beecham, S.; Kazemi, F. Impact of green roofs on stormwater quality in a South Australian urban environment. Sci. Total Environ. 2013, 470–471, 651–659. [Google Scholar] [CrossRef]
  83. Castro, N.; Goldenfum, J.; Lopes da Silveira, A.L.; DallAgnol, A.; Loebens, L.; Demarco, C.; Leandro, D.; Nadaleti, W.; Quadro, M. The analysis of green roof’s runoff volumes and its water quality in an experimental study in Porto Alegre, Southern Brazil. Environ. Sci. Pollut. Res. 2020, 27, 9520–9534. [Google Scholar] [CrossRef]
  84. Imran, H.M.; Akib, S.; Karim, M. Permeable pavement and stormwater management systems: A review. Environ. Technol. 2013, 34, 2649–2656. [Google Scholar] [CrossRef]
  85. Wu, J.; Thompson, J. Quantifying impervious surface changes using time series planimetric data from 1940 to 2011 in four central Iowa cities, U.S.A. Landsc. Urban Plan. 2013, 120, 34–47. [Google Scholar] [CrossRef]
  86. Hernández-Crespo, C.; Fernández-Gonzalvo, M.; Martín Monerris, M.; Andrés-Doménech, I. Influence of rainfall intensity and pollution build-up levels on water quality and quantity response of permeable pavements. Sci. Total Environ. 2019, 684, 303–313. [Google Scholar] [CrossRef]
  87. Kumar, K.; Kozak, J.; Hundal, L.; Cox, A.; Zhang, H.; Granato, T. In-situ infiltration performance of different permeable pavements in a employee used parking lot—A four-year study. J. Environ. Manag. 2016, 167, 8–14. [Google Scholar] [CrossRef]
  88. Park, J.; Park, J.; Cheon, J.; Lee, J.; Shin, H. Analysis of Infiltrating Water Characteristics of Permeable Pavements in a Parking Lot at Full Scale. Water 2020, 12, 2081. [Google Scholar] [CrossRef]
  89. Wang, M.; Zhang, D.-Q.; Cheng, Y.; Tan, S. Assessing performance of porous pavements and bioretention cells for stormwater management in response to probable climatic changes. J. Environ. Manag. 2019, 243, 157–167. [Google Scholar] [CrossRef]
  90. Pezzaniti, D.; Beecham, S.; Kandasamy, J. Stormwater Treatment Using Permeable Pavements. Water Manag. 2012, 165, 161–170. [Google Scholar] [CrossRef]
  91. Drake, J.; Bradford, A.; Marsalek, J. Review of environmental performance of permeable pavement systems: State of the knowledge. Water Qual. Res. J. Can. 2013, 48, 2013. [Google Scholar] [CrossRef]
  92. Liu, J.; Yan, H.; Liao, Z.; Zhang, K.; Schmidt, A.; Tao, T. Laboratory analysis on the surface runoff pollution reduction performance of permeable pavements. Sci. Total Environ. 2019, 691, 1–8. [Google Scholar] [CrossRef]
  93. Mahmoud, A.; Alam, T.; Sanchez, A.; Guerrero, J.; Oraby, T.; Ibrahim, E.; Jones, K. Stormwater Runoff Quality and Quantity from Permeable and Traditional Pavements in Semiarid South Texas. J. Environ. Eng. 2020, 146, 15. [Google Scholar] [CrossRef]
  94. Wang, J.; Meng, Q.; Zou, Y.; Qi, Q.; Tan, K.; Santamouris, M.; He, B.-J. Performance synergism of pervious pavement on stormwater management and urban heat island mitigation: A review of its benefits, key parameters, and co-benefits approach. Water Res. 2022, 221, 118755. [Google Scholar] [CrossRef] [PubMed]
  95. Debnath, B.; Sarkar, P. Clogging in Pervious Concrete Pavement Made with Non-conventional Aggregates: Performance Evaluation and Rehabilitation Technique. Arab. J. Sci. Eng. 2021, 46, 10381–10396. [Google Scholar] [CrossRef]
  96. Park, S.-B.; Tia, M. An experimental study on the water-purification properties of porous concrete. Cem. Concr. Res. 2004, 34, 177–184. [Google Scholar] [CrossRef]
  97. Wang, M.; Zhang, D.-Q.; Li, Y.; Hou, Q.; Yu, Y.; Qi, J.; Fu, W.; Dong, J.; Cheng, Y. Effect of a Submerged Zone and Carbon Source on Nutrient and Metal Removal for Stormwater by Bioretention Cells. Water 2018, 10, 1629. [Google Scholar] [CrossRef]
  98. Davis, A.; Hunt, W.; Traver, R.; Clar, M. Bioretention Technology: Overview of Current Practice and Future Needs. J. Environ. Eng. 2009, 135, 109–117. [Google Scholar] [CrossRef]
  99. Lucke, T.; Nichols, P. The Pollution Removal and Stormwater Reduction Performance of Street-side Bioretention Basins after Ten Years in Operation. Sci. Total Environ. 2015, 536, 784–792. [Google Scholar] [CrossRef] [PubMed]
  100. Li, H.; Sharkey, L.; Asce, M.; Hunt, W.; Davis, A.; Asce, F. Mitigation of Impervious Surface Hydrology Using Bioretention in North Carolina and Maryland. J. Hydrol. Eng. 2009, 14, 407–415. [Google Scholar] [CrossRef]
  101. Winston, R.; Dorsey, J.; Hunt, W. Quantifying volume reduction and peak flow mitigation for three bioretention cells in clay soils in northeast Ohio. Sci. Total Environ. 2016, 553, 83–95. [Google Scholar] [CrossRef] [PubMed]
  102. Kluge, B.; Markert, A.; Facklam, M.; Sommer, H.; Kaiser, M.; Pallasch, M.; Wessolek, G. Metal accumulation and hydraulic performance of bioretention systems after long-term operation. J. Soils Sediments 2018, 18, 431–441. [Google Scholar] [CrossRef]
  103. Liu, J.; Sample, D.; Bell, C.; Guan, Y. Review and Research Needs of Bioretention Used for the Treatment of Urban Stormwater. Water 2014, 6, 1069–1099. [Google Scholar] [CrossRef]
  104. Kuppusamy, V.; Ranga Suba, P. Dracaena marginata biofilter: Design of growth substrate and treatment of stormwater runoff. Environ. Technol. 2015, 37, 1101–1109. [Google Scholar] [CrossRef]
  105. Skorobogatov, A.; He, J.; Chu, A.; Valeo, C.; van Duin, B. The impact of media, plants and their interactions on bioretention performance: A review. Sci. Total Environ. 2020, 715, 136918. [Google Scholar] [CrossRef] [PubMed]
  106. Yang, N.; Du, W.; Chen, L.; Shen, Z.; Chang, C.-C.; Ma, Y. Prioritizing the soil and filler layers of a bioretention system by considering multiple hydrological effects. J. Hydrol. 2021, 603, 127008. [Google Scholar] [CrossRef]
  107. Mahmoud, A.; Alam, T.; Rahman, M.Y.A.; Sanchez, A.; Guerrero, J.; Jones, K. Evaluation of Field-Scale Stormwater Bioretention Structure Flow and Pollutant Load Reductions in a Semi-Arid Coastal Climate. Ecol. Eng. 2019, 142, 100007. [Google Scholar] [CrossRef]
  108. Jia, H.; Wang, X.; Ti, C.; Zhai, Y.; Field, R.; Tafuri, A.; Cai, H.; Yu, S. Field monitoring of a LID-BMP treatment train system in China. Environ. Monit. Assess. 2015, 187, 4595. [Google Scholar] [CrossRef] [PubMed]
  109. Brown, R.; Hunt, W. Improving Bioretention/Biofiltration Performance with Restorative Maintenance. Water Sci. Technol. 2012, 65, 361–367. [Google Scholar] [CrossRef] [PubMed]
  110. Shrestha, P. Effects of different soil media, vegetation, and hydrologic treatments on nutrient and sediment removal in roadside bioretention systems. Ecol. Eng. 2018, 112, 116–131. [Google Scholar] [CrossRef]
  111. Hunt, W.; Davis, A.; Traver, R. Meeting Hydrologic and Water Quality Goals through Targeted Bioretention Design. J. Environ. Eng. 2012, 138, 698–707. [Google Scholar] [CrossRef]
  112. Blecken, G.; Marsalek, J.; Viklander, M. Laboratory Study of Stormwater Biofiltration in Low Temperatures: Total and Dissolved Metal Removals and Fates. Water Air Soil Pollut. 2011, 219, 303–317. [Google Scholar] [CrossRef]
  113. Gülbaz, S.; Kazezyılmaz-Alhan, C.; Copty, N. Evaluation of Heavy Metal Removal Capacity of Bioretention Systems. Water Air Soil Pollut. 2015, 226, 376. [Google Scholar] [CrossRef]
  114. Osman, M.; Wan Yusof, K.; Takaijudin, H.; Goh, H.; Abdul Malek, M.; Azizan, N.; Ab Ghani, A.; Abdurrasheed, A. A Review of Nitrogen Removal for Urban Stormwater Runoff in Bioretention System. Sustainability 2019, 11, 5415. [Google Scholar] [CrossRef]
  115. Jhonson, P.; Goh, H.; Chan, D.; Juiani, S.F.; Zakaria, N. Potential of bioretention plants in treating urban runoff polluted with greywater under tropical climate. Environ. Sci. Pollut. Res. 2022, 30, 24562–24574. [Google Scholar] [CrossRef]
  116. Zhang, H.; Zhang, X.; Liu, J.; Zhang, L.; Li, G.; Zhang, Z.; Gong, Y.; Li, H.; Li, J. Coal gangue modified bioretention system for runoff pollutants removal and the biological characteristics. J. Environ. Manag. 2022, 314, 115044. [Google Scholar] [CrossRef]
  117. Chen, Y.; Shao, Z.; Kong, Z.; Gu, L.; Fang, J.; Chai, H. Study of pyrite based autotrophic denitrification system for low-carbon source stormwater treatment. J. Water Process Eng. 2020, 37, 101414. [Google Scholar] [CrossRef]
  118. Yang, F.; Fu, D.; Liu, S.; Zevenbergen, C.; Singh, R. Hydrologic and Pollutant Removal Performance of Media Layers in Bioretention. Water 2020, 12, 921. [Google Scholar] [CrossRef]
  119. Shrestha, P.; Faulkner, J.; Kokkinos, J.; Hurley, S. Influence of low-phosphorus compost and vegetation in bioretention for nutrient and sediment control in runoff from a dairy farm production area. Ecol. Eng. 2020, 150, 105821. [Google Scholar] [CrossRef]
  120. Zhang, L.; Ye, Z.; Shibata, S. Assessment of Rain Garden Effects for the Management of Urban Storm Runoff in Japan. Sustainability 2020, 12, 9982. [Google Scholar] [CrossRef]
  121. Xiong, J.; Ren, S.; He, Y.; Bai, X.; Wang, J.; Dzakpasu, M. Bioretention cell incorporating Fe-biochar and saturated zones for enhanced stormwater runoff treatment. Chemosphere 2019, 237, 124424. [Google Scholar] [CrossRef]
  122. Qiu, F.; Zhao, S.; Zhao, D.; Wang, J.; Fu, K. Enhanced Nutrients Removal in Bioretention Systems Modified with Water Treatment Residual and Internal Water Storage Zone. Environ. Sci. Water Res. Technol. 2019, 5, 993–1003. [Google Scholar] [CrossRef]
  123. Johnson, J.; Hunt, W. A Retrospective Comparison of Water Quality Treatment in a Bioretention Cell 16 Years Following Initial Analysis. Sustainability 2019, 11, 1945. [Google Scholar] [CrossRef]
  124. Braswell, A.; Anderson, A.; Hunt, W. Hydrologic and Water Quality Evaluation of a Permeable Pavement and Biofiltration Device in Series. Water 2018, 10, 33. [Google Scholar] [CrossRef]
  125. Nichols, P.; Lucke, T.; Drapper, D. Field and Evaluation Methods Used to Test the Performance of a Stormceptor® Class 1 Stormwater Treatment Device in Australia. Sustainability 2015, 7, 16311–16323. [Google Scholar] [CrossRef]
  126. Dagenais, D.; Brisson, J.; Fletcher, T. The role of plants in bioretention systems; does the science underpin current guidance? Ecol Eng. 2018, 120, 532–545. [Google Scholar] [CrossRef]
  127. Read, J.; Fletcher, T.; Wevill, T.; Deletic, A. Plant Traits that Enhance Pollutant Removal from Stormwater in Biofiltration Systems. Int. J. Phytoremediat. 2010, 12, 34–53. [Google Scholar] [CrossRef]
  128. Vymazal, J. Removal of Nutrients in Various Types of Constructed Wetlands. Sci. Total Environ. 2007, 380, 48–65. [Google Scholar] [CrossRef]
  129. Headley, T.; Tanner, C. Constructed Wetlands with Floating Emergent Macrophytes: An Innovative Stormwater Treatment Technology. Crit. Rev. Environ. Sci. Technol. 2011, 42, 2261–2310. [Google Scholar] [CrossRef]
  130. Sharma, R.; Vymazal, J.; Malaviya, P. Application of floating treatment wetlands for stormwater runoff: A critical review of the recent developments with emphasis on heavy metals and nutrient removal. Sci. Total Environ. 2021, 777, 146044. [Google Scholar] [CrossRef]
  131. Liu, A.; Egodawatta, P.; Goonetilleke, A. Ranking Three Water Sensitive Urban Design (WSUD) Practices Based on Hydraulic and Water Quality Treatment Performance: Implications for Effective Stormwater Treatment Design. Water 2022, 14, 1296. [Google Scholar] [CrossRef]
  132. Lenhart, H.; Hunt, W. Evaluating Four Storm-Water Performance Metrics with a North Carolina Coastal Plain Storm-Water Wetland. J. Environ. Eng. ASCE 2010, 137, 155–162. [Google Scholar] [CrossRef]
  133. Rizzo, A.; Bresciani, R.; Masi, F.; Boano, F.; Revelli, R.; Ridolfi, L. Flood reduction as an ecosystem service of constructed wetlands for combined sewer overflow. J. Hydrol. 2018, 560, 150–159. [Google Scholar] [CrossRef]
  134. Chunbo, Y.; Zhao, F.; Zhao, X.; Zhao, Y. Woodchips as sustained-release carbon source to enhance the nitrogen transformation of low C/N wastewater in a baffle subsurface flow constructed wetland. Chem. Eng. J. 2020, 392, 124840. [Google Scholar] [CrossRef]
  135. Tanner, C.; Headley, T. Components of floating emergent macrophyte treatment wetlands influencing removal of stormwater pollutants. Ecol. Eng. 2011, 37, 474–486. [Google Scholar] [CrossRef]
  136. Chua, L.; Tan, S.; Sim, C.H.; Goyal, M. Treatment of baseflow from an urban catchment by a floating wetland system. Ecol. Eng. 2012, 49, 170–180. [Google Scholar] [CrossRef]
  137. Maxwell, B.; Winter, D.; Birgand, F. Floating treatment wetland retrofit in a stormwater wet pond provides limited water quality improvements. Ecol. Eng. 2020, 149, 105784. [Google Scholar] [CrossRef]
  138. Hua, P.; Yang, W.; Qi, X.; Jiang, S.; Xie, J.; Xianyong, G.; Li, H.; Zhang, J.; Krebs, P. Evaluating the effect of urban flooding reduction strategies in response to design rainfall and low impact development. J. Clean. Prod. 2020, 242, 118515. [Google Scholar] [CrossRef]
  139. Luan, Q.; Fu, X.; Song, C.; Wang, H.; Liu, J.; Wang, Y. Runoff Effect Evaluation of LID through SWMM in Typical Mountainous, Low-Lying Urban Areas: A Case Study in China. Water 2017, 9, 439. [Google Scholar] [CrossRef]
  140. Wang, M.; Sun, C.; Zhang, D. Opportunities and challenges in green stormwater infrastructure (GSI): A comprehensive and bibliometric review of ecosystem services from 2000 to 2021. Environ. Res. 2023, 236, 116701. [Google Scholar] [CrossRef]
  141. Nguyen, T.T.; Ngo, H.; Guo, W.; Ren, N.-Q.; Li, G.; Ding, J.; Liang, H. Implementation of a specific urban water management—Sponge City. Sci. Total Environ. 2018, 652, 147–162. [Google Scholar] [CrossRef]
  142. Kim, J.-H.; Kim, H.; Demarie, F. Facilitators and Barriers of Applying Low Impact Development Practices in Urban Development. Water Resour. Manag. 2017, 31, 3795–3808. [Google Scholar] [CrossRef]
  143. Li, H.; Ding, L.; Ren, M.; Li, C.; Wang, H. Sponge City Construction in China: A Survey of the Challenges and Opportunities. Water 2017, 9, 594. [Google Scholar] [CrossRef]
  144. Ortega, A.; Rodríguez Sánchez, J.; Bharati, L. Building flood-resilient cities by promoting SUDS adoption: A multi-sector analysis of barriers and benefits in Bogotá, Colombia. Int. J. Disaster Risk Reduct. 2023, 88, 103621. [Google Scholar] [CrossRef]
  145. Labadie, K. Identifying Barriers to Low Impact Development and Green Infrastructure in the Albuquerque Area. Master’s Thesis, The University of New Mexico, Albuquerque, NM, USA, 2011. [Google Scholar]
  146. Xu, C.Q.; Tang, T.; Jia, H.F.; Xu, M.; Xu, T.; Liu, Z.J.; Long, Y.; Zhang, R.R. Benefits of coupled green and grey infrastructure systems: Evidence based on analytic hierarchy process and life cycle costing. Resour. Conserv. Recycl. 2019, 151, 104478. [Google Scholar] [CrossRef]
  147. Roy, A.H.; Wenger, S.J.; Fletcher, T.D.; Walsh, C.J.; Ladson, A.R.; Shuster, W.D.; Thurston, H.W.; Brown, R.R. Impediments and solutions to sustainable, watershed-scale urban stormwater management: Lessons from Australia and the United States. Environ. Manag. 2008, 42, 344–359. [Google Scholar] [CrossRef] [PubMed]
Figure 1. A step-by-step process for filtering relevant articles from search results.
Figure 1. A step-by-step process for filtering relevant articles from search results.
Atmosphere 15 00403 g001
Figure 2. Mechanism diagram of green roof in providing regulation services for hydrological process.
Figure 2. Mechanism diagram of green roof in providing regulation services for hydrological process.
Atmosphere 15 00403 g002
Figure 3. Mechanism diagram of permeable pavement providing regulation services for hydrological processes.
Figure 3. Mechanism diagram of permeable pavement providing regulation services for hydrological processes.
Atmosphere 15 00403 g003
Figure 4. Mechanism diagram of bioretention in in providing regulation services in hydrological processes.
Figure 4. Mechanism diagram of bioretention in in providing regulation services in hydrological processes.
Atmosphere 15 00403 g004
Figure 5. Mechanism diagram of Constructed Wetlands, providing regulation services for hydrological processes.
Figure 5. Mechanism diagram of Constructed Wetlands, providing regulation services for hydrological processes.
Atmosphere 15 00403 g005
Table 2. The attenuation of TSS, COD, TN, and TP in green roofs.
Table 2. The attenuation of TSS, COD, TN, and TP in green roofs.
CountryRunoff SourceScaleFilter MediaPollutant Removal Efficiency (%)Reference
TSSCODTNTP
GreeceNatural rainfall eventsFieldVermiculite93.0091.0087.00N/AThomaidi et al. [65]
ChinaSimulated rainfall eventsLaboratoryPeat soil, vermiculite and polyaluminum chloride (PAC)N/AN/A6.0484.33Zhang et al. [66]
TurkeySimulated rainfall eventsLaboratoryN/A41.1539.7329.5832.26Koc et al. [64]
The NetherlandsSimulated rainfall eventsLaboratoryN/A22.00N/A19.0020.00Dutta et al. [67]
ChinaSimulated rainfall eventsLaboratoryCombined substrate44.77N/A19.6045.51Zhang et al. [68]
Chinanatural rainfall eventsFieldCommercial substrate31.6025.1037.80N/ALiu et al. [63]
Republic of Koreanatural rainfall eventsFieldN/A77.00N/A57.0053.00Jeon et al. [69]
Chinanatural rainfall eventsFieldPerlite or recycled bricks37.85N/A14.5212.93Chai et al. [70]
ChinaSimulated rainfall eventsLaboratoryPeat soilN/A30.0042.0047.00Zhang et al. [71]
IndiaSimulated rainfall eventsLaboratorySand, brick bats, and gravel85.00–90.0088.0088.00–99.0092.00Chandrasekaran et al. [72]
VietnamNatural rainfall eventsFieldSoil, sand, crushed stone, and gravel 64.30–73.1077.00–78.0088.00–91.0072.00–78.00Bui et al. [73]
ChinaSimulated rainfall eventsLaboratoryN/A80.00–90.0050.00–70.0050.00–70.0040.00–70.00Zhou et al. [74]
Table 3. The attenuation of heavy metals in green roofs.
Table 3. The attenuation of heavy metals in green roofs.
CountryFilter MediaRunoff SourceRemoval Efficiency of Heavy Metal (%)Reference
CuZnPbCd
ChinaLoam, perlite, pure cocopeat, and sodium polyacrylateSimulated rainfall eventsN/A94.5598.84N/AGuo et al. [75]
IndiaPerlite, vermiculite, sand, crushed brick, cocopeat, and T. conoidesSimulated rainfall events95.5096.6098.3097.80Kuppusamy and Joshi [76]
IndiaPerlite, crushed brick, and sandSimulated rainfall events99.2097.4099.9099.90Kuppusamy and Raja [77]
FranceCommercial substrateNatural rainfall events87.00–90.0070.00–98.00N/AN/ASeidl et al. [78]
USACommercial substrateNatural rainfall events50.0065.80Nearly 100N/AGregoire et al. [79]
USAAn expanded clay mixed with pine barkSimulated rainfall events94.0065.0280.46N/ASarah et al. [80]
Table 4. The application of bioretention systems for the reduction of runoff and peak flow.
Table 4. The application of bioretention systems for the reduction of runoff and peak flow.
CountryBioretention System InformationSite CharacteristicsRunoff/Outflow Reduction (%)Peak Flow Reduction (%) Other/NotesReference
ChinaA 10 cm aquifer layer, 5 cm mulch layer, 30 cm soil medium layer, 40 cm filler layer, and 15 cm gravel layer were set by geotextilesThe mean annual precipitation is 587 mm14.00–78.009.00–91.00In the 2-year, 30-year, and 100-year return periodYang et al. [106]
USA(250 ft) linear bioretention cellLocated in a highly impermeable area with a total area of 8494 square meters80.10–98.20N/A45 storm events were observed, ranging from 1.8 to 49.5 mmMahmoud et al. [107]
USAReferred to design guidance in the Ohio Rainwater and Land Development Manual0.36 ha, 77.1% impervious catchment36.00–59.0024.00–96.001-year design rainfall intensitiesWinston et al. [101]
AustraliaConsisted of a filter medium (usually sandy), underlaid by a gravel drainage layerThe BRSs were located directly adjacent to the roadway32.70–84.3079.50–93.60Natural rainfall eventsLucke and Nichols. [99]
GreeceHas a depth of 0.95 m, and media contained a 0.35 m depth of gravel and a 0.4 m depth of soil/planting soil.Received stormwater runoff from a playground with an area of 7672 m247.00–80.0050.00–84.00A total of 19 natural rainfall events were monitoredJia et al. [108]
USAThe BRS employed two different media depths (0.6 and 0.9 m)N/A63.00–89.0084.00–95.00Rain depth (mm) between 8.80 and 99.6 mmBrown and Hunt. [109]
Table 5. The attenuation of TSS, COD, TN, and TP in bioretention systems.
Table 5. The attenuation of TSS, COD, TN, and TP in bioretention systems.
CountryRunoff Source ScaleFilter MediaPollutant Removal Efficiency (%)Reference
TSSCODTNTP
MalaysiaN/ALaboratorySand, topsoil, and compost90.0092.5086.4093.50Jhonson et al. [115]
ChinaSimulated rainfall eventsLaboratoryCoal gangue (CG)N/A33.00–86.0030.00–70.0094.00–99.00Zhang et al. [116]
ChinaSimulated rainfall eventsLaboratoryPyrite and zeoliteN/AN/A89.3081.60Chen et al. [117]
ChinaSimulated rainfall eventsLaboratoryTraditional substrate: sandN/A86.0071.8068.00Yang et al. [118]
USANatural rainfall eventsFieldCompost83.00–96.00N/A17.30–38.5080.00–92.00Shrestha et al. [119]
JapanSimulated rainfall eventsLaboratoryN/A13.00–15.5012.90–16.1712.83–17.3414.03–19.07Zhang et al. [120]
ChinaSimulated rainfall eventsLaboratoryBiochar31.6078.5–94.682.30–97.0057.36–93.70Xiong et al. [121]
ChinaSimulated rainfall eventsLaboratorySandy loam92.00–97.0064.00–95.0075.00>99.00Qiu et al. [122]
USANatural rainfall eventsFieldSandyN/AN/A72.0079.00Johnson and Hunt [123]
USANatural rainfall eventsFieldSand, compost, and pure sand91.00–97.00N/A38.00–57.0086.00–94.00Shrestha et al. [110]
USANatural rainfall eventsFieldN/A96.00N/A42.0075.00Braswell et al. [124]
AustraliaNatural rainfall eventsFieldN/A83.00N/A23.0011.00Nichols et al. [125]
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

Sun, C.; Rao, Q.; Chen, B.; Liu, X.; Adnan Ikram, R.M.; Li, J.; Wang, M.; Zhang, D. Mechanisms and Applications of Nature-Based Solutions for Stormwater Control in the Context of Climate Change: A Review. Atmosphere 2024, 15, 403. https://doi.org/10.3390/atmos15040403

AMA Style

Sun C, Rao Q, Chen B, Liu X, Adnan Ikram RM, Li J, Wang M, Zhang D. Mechanisms and Applications of Nature-Based Solutions for Stormwater Control in the Context of Climate Change: A Review. Atmosphere. 2024; 15(4):403. https://doi.org/10.3390/atmos15040403

Chicago/Turabian Style

Sun, Chuanhao, Qiuyi Rao, Biyi Chen, Xin Liu, Rana Muhammad Adnan Ikram, Jianjun Li, Mo Wang, and Dongqing Zhang. 2024. "Mechanisms and Applications of Nature-Based Solutions for Stormwater Control in the Context of Climate Change: A Review" Atmosphere 15, no. 4: 403. https://doi.org/10.3390/atmos15040403

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