Next Article in Journal
Developing Students Well-Being and Engagement in Higher Education during COVID-19—A Case Study of Web-Based Learning in Finland
Next Article in Special Issue
Utilization of Food Waste for the Development of Composite Bread
Previous Article in Journal
Impact of Management and Reverse Logistics on Recycling in a War Scenario
Previous Article in Special Issue
Recovery of Sugar and Nutrients from Algae and Colocasia esculenta (Taro) Leaves Using Chemical Hydrolysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Decision Framework for Designing Sustainable Wastewater-Based Resource Recovery Schemes

1
Department of Mechanical Engineering, University of South Florida, Tampa, FL 33620, USA
2
Department of Civil & Environmental Engineering, University of South Florida, Tampa, FL 33620, USA
3
Department of Civil Engineering, University of Kentucky, Lexington, KY 40506, USA
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(4), 3839; https://doi.org/10.3390/su15043839
Submission received: 21 December 2022 / Revised: 30 January 2023 / Accepted: 11 February 2023 / Published: 20 February 2023
(This article belongs to the Collection Waste Utilization and Resource Recovery)

Abstract

:
The availability of sufficient water supply is a challenge many municipalities have faced in recent decades and a challenge that is expected to intensify with time. While several choices remain for selecting alternatives to freshwater sources, water reclamation offers an opportunity for sustainable resource recovery. Nonetheless, tradeoffs exist in the selection of the most sustainable technology for recovering resources from wastewater when long-term impacts are taken into consideration. This article investigates the factors influencing the environmental and economic impacts of resource recovery technologies through the analysis of life cycle environmental and economic impact case studies. Key characteristics were extracted from life cycle assessment and life cycle cost case studies to evaluate the factors influencing the sustainability of the resource recovery systems. The specific design parameters include the type of resources to be recovered, technology utilized, scale of implementation, location, and end users. The design of sustainable resource recovery systems was found to be largely driven by scale, location (e.g., as it pertains to the energy mix and water quality restrictions), and the scope of the system considered. From this analysis, a decision framework for resource recovery-oriented wastewater management was developed and then applied to an existing case study to demonstrate its usability.

1. Introduction

Water scarcity has affected 2.3 billion people around the globe [1], which poses a grand challenge to achieve Sustainable Development Goal (SDG) 6 “ensure availability and sustainable management of water and sanitation for all”. It also presents critical stressors on our communities as water is needed to maintain human health, agriculture, and many industrial processes. Our communities have been further challenged by the increased intensity of weather events exacerbated by climate change [2,3]. Self-sufficiency can offer empowerment for members of a community [4,5,6] and plays an even more critical role during disaster situations [7,8]. Fortunately, wastewater serves as a valuable resource that can provide water, energy, and nutrients [9,10], amongst other resources such as precious metals [11,12], to our communities. The true cost of resource recovery, however, must be considered to avoid shifting impacts from one impact category to another.
To evaluate the environmental and economic impacts of a technology, Life Cycle Assessment (LCA) and Life Cycle Cost Analysis (LCCA) are often employed. By applying these techniques during the design phase, a multi-criteria impact assessment can be conducted to inform strategies for environmental impact and potential cost savings over the system’s lifetime. The life cycle impacts of water and wastewater systems have been studied extensively, as evidenced by previously published review articles [13,14,15,16]. The methods implemented for evaluating the sustainability of resource recovery systems have been investigated in prior reviews [14,15]. Additionally, frameworks have been developed to guide decision making as it relates to resource extraction from sanitation systems [17] and the design of centralized water reclamation systems in highly urban environments [18].
The previous studies show that the economic and environmental impacts of resource recovery systems are highly driven by the scale of implementation [15,19,20], selection of treatment technology and treatment train [15,19], location and topography [20], and end uses [21,22]. Several previous studies have focused on different aspects of the design for wastewater-based resource recovery systems (e.g., treatment techniques and system scales), with the purpose of providing guidance for municipalities. However, lack of a systematic decision framework that can be used by a variety of stakeholders to plan for the most sustainable resource recovery scheme in their water service area is evident.
Given the trends in the environmental and economic impacts of wastewater-based resource recovery systems [10], the objective of this article is to provide a decision framework for sustainable resource recovery as stakeholders choose how to recover resources from their wastewater systems and decide where to send those resources to. This framework will serve to facilitate the decision-making process for the design of sustainable resource recovery systems and can be applied by researchers and municipalities alike. The article seeks answers to the following research questions: (1) What factors are driving the greenhouse gas emissions and costs of resource recovery systems? and (2) How can prior case studies inform the decision-making process? An analysis of existing life cycle case studies is conducted to support the development of this framework. The scope of this research considers water, energy, and nutrient recovery from domestic wastewater.

2. Materials and Methods

Life cycle environmental and economic impact case studies were analyzed in this research for wastewater-based resource recovery systems. In all, 21 water reclamation, 26 energy recovery, and 29 nutrient recovery life cycle assessments (LCAs) were analyzed (see the Appendix A for the list of articles). The analysis of the environmental impact assessments focused on greenhouse gas (GHG) emissions as this impact category was consistently reported across the LCAs. For the life cycle cost (LCC) case studies, 16 water reclamation, 10 energy recovery, and 10 nutrient recovery case studies were assessed. Key characteristics were extracted from these data samples to conduct the statistical analysis as described in the sections that follow.

2.1. Resource Recovery Processes

2.1.1. Water Reclamation

The characteristics that were extracted from the water reclamation case studies included the scale of implementation, the location of the system, the technology used for resource recovery, the end use of the reclaimed water, water quality parameters, the life cycle phases considered in the assessment, and the physical scope of the study (e.g., treatment and distribution). A variety of wastewater treatment trains were used within the case studies and they were categorized by secondary and tertiary treatment processes (see Figure 1a). Activated sludge (AS), absorption/bio-oxidation (AB), oxidation ditch (OD), anaerobic/anoxic/oxic (A2O), membrane bioreactor (MBR), upflow anaerobic sludge blanket (UASB), and septic tank (ST) are treatment technologies based on suspended growth. Rotating biological contactor (RBC) is a type of attached growth technology. Wastewater stabilization ponds and vertical flow constructed wetlands (VFCW) are considered lagoon-based technologies. These aforementioned treatment technologies are categorized as secondary biological treatment (2ndBio). Tertiary treatment technology is applied after secondary treatment (2ndBio) to obtain a better effluent quality for water reuse. Enhanced nitrogen and/or phosphorus removal are tertiary biological treatment processes (3rdBio). Flocculation, coagulation, adsorption (e.g., granular activated carbon [GAC]), and filtration (e.g., sand filtration [SF] and media filtration [MF]) are categorized as tertiary physical and/or chemical treatment (3rdPC). Reverse osmosis (RO), ultrafiltration (UF), and microfiltration (MF) are categorized as membrane filtration (3rdM).

2.1.2. Energy Recovery

Energy, including electricity and heat, can be recovered from sludge treatment through anaerobic digestion (AD), incineration, and composting, and from wastewater through a heat exchanger, hydropower, or pressure exchanger (Figure 1b). The sludge treatment process may include a volume reduction process (V) and stabilization (S) before transferring the sludge for its end use. Gravity thickening, dynamic thickening, thermal drying, sludge drying beds, centrifuge dewatering, filter press dewatering, and belt filter press dewatering are available technologies for volume reduction. Anaerobic digestion, as a reliable stabilization process for biological treatment of the produced sludge, is utilized in most of the articles analyzed. Composting, pasteurization, incineration, combustion, gasification, pyrolysis, and landfilling are the selected processes of ultimate sludge utilization. The sludge utilization step was categorized as either composting, incineration, pyrolysis, or landfilling.

2.1.3. Nutrient Recovery

Nutrient recovery processes include the treatment of the source (e.g., wastewater, urine, sludge) and the ultimate utilization of sludge (Figure 1c). Fertilizer, fertigation, and land reclamation material are common products of nutrient recovery. Nutrient recovery processes were grouped into different sources, including urine, greywater, sludge liquor, side stream, sludge, and a combination of urine, feces, and greywater. For urine as a nutrient recovery source, the available treatment trains are: (1) urine source separation (USS); (2) chemical treatment (Chem), membrane treatment (M), or drying (Dry); and (3) fertilizer or fertigation. Nutrients in greywater can be recovered by a physical treatment process and used for fertigation. For the combination source of urine, feces, and greywater, two recovery processes are considered: (1) USS and fertigation, or (2) AD and fertilizer. Sludge liquor contains an abundance of nutrients and can be recovered as a fertilizer through chemical treatment (Chem) and dewatering (Dew) processes. The side stream of a wastewater treatment plant is a good source for fertigation. AD + fertilizer and Dry + fertilizer are two processes to recover nutrients in wastewater and sludge. Nutrients in sludge can be recovered through four processes of: (1) sludge treatment wetland (STW) + fertilizer; (2) STW + Dew + fertilizer; (3) AD + incineration + land reclamation (LR); and (4) Dry + incineration + LR.

2.2. Statistical Analysis

The effluent quality of reclaimed water for different categories of reuse (non-potable reuse [NPR], indirect potable reuse [IPR], or direct potable reuse [DPR]) must meet water quality regulations. Therefore, the effluent quality and removal rate are considered in the correlation analysis. The representative water quality parameters consist of biological oxygen demand (BOD), chemical oxygen demand (COD), total nitrogen (TN), total phosphorus (TP), and total suspended solids (TSS). The influent and effluent water quality data are obtained from the literature identified in Table A1 of the Appendix A, and the removal rates are calculated quantities.
A bivariate correlation analysis was conducted in Minitab 19 with the life cycle case study data. More specifically, the correlation analysis included the evaluation of relationships between GHG emissions, scale, cost, treatment trains, effluent water quality parameters, removal rates, the type of sewage system, country, and national electrical greenhouse gas emissions. The absolute value of the correlation coefficient, r, closer to 1.0 shows a stronger correlation between variables. The outcomes of the correlation analysis with a p-value less than or equal to 0.05 are highlighted in Table A2 of the Appendix A.
Descriptive statistics were analyzed for sub-sets of GHG emission data within each resource category (water, energy, and nutrients). As the data were found to not be normally distributed, statistically significant differences in the reported environmental impacts were evaluated using the Mann–Whitney U test. Samples of GHG emission data were compared with respect to the case study location, treatment train, reuse type, life cycle phases, and system scope. The U statistic was calculated for pairwise comparisons (see Equations (1) and (2)) where R1 and R2 is the sum of the ranks for each data sample, and n1 and n2 represent the number of samples. The smaller U value (Equation (3)) was compared to the critical U value when the data samples were less than 20. When the number of samples (n1 and n2) exceeded 20, the null hypothesis was rejected for z-scores (see Equation (4)) that were less than −1.96 or greater than 1.96 (a two-tailed test).
U 1 = R 1 n 1 × ( n 1 + 1 ) 2  
U 2 = R 2 n 2 × ( n 2 + 1 ) 2
U = min ( U 1 ,   U 2 )
z = U n 1 × n 2 2 n 1 × n 2 × ( n 1 + n 2 + 1 12 )

3. Factors Influencing Life Cycle Impacts

This section reports on the outcomes of the correlation analysis to identify the extent to which specific factors influenced the life cycle GHG emissions and LCC of wastewater-based resource recovery systems.

3.1. Correlation Analysis—Water Reclamation

The factors considered for water reclamation systems include the target end use (NPR, IPR, or DPR), the water quality of the influent and effluent, the wastewater treatment process (in most cases, several options can achieve a similar effluent water quality), the redundancy in the treatment train, the design capacity of the treatment system, and the operating scale [15,19,20]. Factors that influence the pumping energy for reclaimed water distribution include the flowrate, the size of the service area for the end users, and the topography of the region [19,20,23]. However, since these parameters were rarely reported consistently in the articles reviewed, this was outside of the scope of this manuscript.

3.1.1. Treatment Train

Specific capital and/or O&M cost (USD 2019/m3) had a high positive correlation to GHG emissions (kg CO2-eq/m3) (r = 0.592, p = 0.026). This demonstrates that systems with high costs tended to result in high GHG emissions, which may be attributed, in part, to the high energy consumption of wastewater treatment technology and distribution pumps [19]. In contrast, specific costs and GHG emissions are both negatively correlated to scale (r = −0.535 and r = −0.453, respectively). Accordingly, large scale systems tend to cost less and produce less GHG emissions on a per unit basis, and thus benefit from economies of scale [15].

3.1.2. Water Quality

The target effluent water quality and removal rate aids in the determination of feasible treatment trains for water reclamation systems. Figure 2 shows the influent and effluent water quality parameters treated by various treatment trains. BOD, COD, TN, and TP were selected as they are representative of key characteristics of the wastewater influent and effluent. Only four treatment trains are discussed due to the limited availability of water quality information. The difference in the influent water quality could be due to variations in the composition of the water consumers (e.g., residential, industrial, or utilities) and the characteristics of the communities across the case studies [21,22,24]. Minor variations are observed in the effluent water quality (see Figure 2); TN and TP in the effluent is predominantly less than 10 mg/L, which aligns with many state regulations for water reuse [25]. Other differences may be attributed to variations in the influent water quality, reuse requirements, and the treatment applied.
The water quality parameters were found to have a strong association to the treatment train selected. For example, although the effluent BOD and COD were generally similar for the processes achieving tertiary treatment, biological processes (2ndBio + 3rdBio) tended to be selected to treat higher COD influent (r = 0.628 and p = 0.002). Additionally, the selection of this treatment process (2ndBio + 3rdBio) was correlated with the TN removal rate (r = 0.518 and p = 0.019). Figure 2 shows that this treatment train was used to treat higher levels of TN in the influent, with an average of approximately 60 mg/L, and provided the best effluent quality among the treatment trains. For total phosphorus management, the trains with tertiary treatment (2ndBio + 3rdM, 2ndBio + 3rdPC) had lower effluent TP overall. Significant variation was observed for effluent TP for tertiary membrane treatment (2ndBio + 3rdM), and thus no correlation between TP and the treatment technologies was identified for the case studies analyzed. When the treatment systems are classified as centralized and decentralized systems, effluent TP was found to be higher for decentralized systems (r = 0.722 and p = 0.004) relative to centralized systems (r = −0.722 and p = 0.004). Table A2 summarizes the outcomes of the correlation analysis in the Appendix A.

3.2. Correlation Analysis—Energy Recovery

Various processes are implemented for wastewater-based energy recovery (e.g., anaerobic digestion, incineration, thermal energy recovery, and hydropower generation). Given this diversity, some factors are more specific to the technology, although the processing rate has been found to influence the resources consumed for energy recovery for most of these systems [15].
Cost (USD 2019/MJ) was found to be negatively correlated to the rate of biosolids processing (r = −0.839 and p = 0), demonstrating the influence of economies of scale. Cost was also negatively correlated to recovery via anaerobic digestion and landfilling (AD + landfilling, r = −0.331 and p = 0), indicating unit costs tended to be lower when this type of energy recovery system was implemented. GHG emissions and AD + landfilling were also correlated, but to a lesser degree (r = 0.274 and p = 0.015), which demonstrates that the treatment technology was not the only factor influencing the GHG emissions.
Other factors that are expected to influence the sustainability of energy recovery systems include climate and topography for thermal energy recovery systems and hydropower generation systems, respectively. For example, Ravichandran et al. [26] evaluated the influence of local conditions on the sustainability of drain water heat recovery systems (DWHRSs) and found it was environmentally and economically beneficial to implement DWHRSs in cold/very cold climates. Similarly, hydropower generation is known to be a function of headloss (influenced by topography and flow rate); however, only a few studies have been conducted [27,28] on the sustainability of wastewater-based hydropower generation systems.

3.3. Correlation Analysis—Nutrient Recovery

For nutrient recovery, the source being processed (i.e., urine, wastewater and sludge, biosolids) highly influenced treatment technology selection (urine treated by urine source separation (USS) and chemical treatment for fertilizer production, r = 0.896, p = 0). Treatment technologies (USS and AD for fertigation and fertilizer) were also correlated with scale (m3/day) (r = 0.539, p = 0). No correlation was identified for cost; however, GHG emissions were found to be correlated to the source being processed. Specifically, GHG emissions are positively correlated to nutrient recovery from wastewater and sludge (r = 0.28 and p = 0.001), while emissions are negatively correlated to urine as the source for nutrient recovery (r = −0.194 and p = 0.024). This indicates that more GHGs are released when trying to recover nutrients once it has reached an offsite treatment facility, relative to trying to recover nutrients from the source (urine). This finding is in alignment with work from Ishii and Boyer [29] and Landry and Boyer [30] who investigated the life cycle impacts of recovering nutrients via USS and struvite precipitation. A summary of feasible recovery technologies (for water, energy and nutrients) at varied scales can be found in Diaz-Elsayed et al. [10].

4. Informing the Selection of Treatment Technologies

4.1. Life Cycle Costs

The average Specific Net Present Value (SNPV) [19] is calculated using Equations (5) and (6) where NPV represents the net present value of the system, CFt the cashflows for time period t, i the discount rate, n the lifespan of the system, and Pt the resources recovered during time period t. Descriptive statistics of the SNPV for the resource recovery systems are presented in Table 1. For water reclamation, biological (2ndBio + 3rdBio) and physical/chemical processes (2ndBio + 3rdBio + 3rdPC) are the least expensive relative to other treatment trains. When membrane treatment is applied, a significantly higher SNPV is attained as 2ndBio + 3rdM and 2ndBio + 3rdM + 3rdPC had an average cost of $2.68 and $7.23 USD 2019/m3, respectively. One benefit to consider for selecting a membrane treatment technology is the improved effluent water quality that can be achieved as discussed in Section 3.1.2.
N P V = 0 n C F t ( 1 + i ) t
S N P V = N P V 1 n 0 n P t d t
For energy recovery from wastewater and sludge, anaerobic digestion with landfilling (AD + landfilling) has the largest range of SNPV, as well as the highest standard deviation, which may be a result of differences in the transportation and labor fees for different locations. The lowest SNPV for AD + landfilling from sludge is due to greater revenue from the recovered product (e.g., electricity and district heating) than spent costs. When sludge is used as the source for resource recovery most treatment technologies resulted in a negative SNPV, which shows the benefit from recovered products. For energy recovery from wastewater as a source, hydropower generation results in a lower SNPV than heat exchangers.
Based on the case studies analyzed, nutrient recovery from urine requires a higher expense than most recovery technologies from sludge. For sludge, anaerobic digestion for fertilizer application (AD + fertilizer) is the most expensive technology with a mean SNPV of 24.46 USD 2019/kg P-eq. This is a result of high energy consumption for pelletization and transportation fees [31,32]. The large standard deviation of AD + fertilizer from sludge is due to the difference in rates for electricity, transportation, construction material, and labor across countries, including Sweden [33], China [32,34], Italy [31], and Japan [35]. Drying, incineration, and land reclamation (Dry + Inc. + LR) provided the lowest SNPV with a mean value of −4.74 USD 2019/kg P-eq. However, only one datum point is available for this technology, which introduces uncertainty for its performance.
In summary, most energy recovery technologies provide a higher revenue than water and nutrient recovery technologies considering the mean SNPV. While treatment of wastewater to some extent is required prior to releasing it back to the environment, a decision can be made on implementation of additional treatment trains to recover more resources (e.g., energy and nutrients). Once the decision is made about which resources to recover and which sources to recover from, the technology can be selected with consideration of the average SNPV shown in Table 1.

4.2. Life Cycle Greenhouse Gas Emissions

This section seeks to identify how the GHG emissions of wastewater-based resource recovery systems are influenced by varied conditions in the case studies evaluated. The correlation analysis revealed a correlation between the GHG emissions of water reclamation case studies with cost, scale, and effluent (BOD and TP) and influent (TN) water quality parameters (see Table A2 in the Appendix A). While GHG emissions and costs were positively correlated (r: 0.588, p: 0.027), economies of scale were verified as increasing the scale of implementation reduced GHG emissions (r: −0.452, p: 0.035). These findings confirm prior findings [15], and reiterate the ability for larger systems to more efficiently consume resources and lower the environmental “cost” relative to smaller systems.
GHG emissions were evaluated relative to the following factors expected to influence the environmental impact of the water reclamation systems (see Figure 3a–c): location, treatment train, reuse type, life cycle phases, and system scope. Statistically significant differences in the outcomes were identified as follows:
  • Case studies in Spain resulted in lower GHG emissions than those in the USA (mean of 0.34 kg CO2-eq/m3 vs. 1.21 kg CO2-eq/m3), which aligns with the reduced dependency on fossil fuels for the Spanish electrical energy mix (0.179 kg CO2-eq/MJ for Spain [36,37,38], vs. ~0.40 kg CO2-eq/MJ for the USA [36,38,39,40];
  • The inclusion of the construction phase in the LCAs resulted in significantly higher impacts relative to only considering the O&M plus disposal phases (1.62 kg CO2-eq/m3 and 0.61 kg CO2-eq/m3, respectively);
  • In comparing the scope of the water reclamation system, the inclusion of the collection stage resulted in a statistically significant increase in GHG emissions relative to the consideration of only the treatment and water reuse stages (mean: 0.90 kg CO2-eq/m3 vs. 0.33 kg CO2-eq/m3);
  • No statistically significant difference was observed in GHG emissions relative to the reuse type or treatment train. This may be attributed to the dominance of the previously identified factors (i.e., energy mix, life cycle phases, and system scope).
Few LCAs for wastewater-based energy recovery were available, so a comparative analysis was conducted that evaluated data from the following scenarios: (1) natural gas production from anaerobic digestion and landfilling in the USA that considered the construction, O&M, disposal, and recovery life cycle phases for resource recovery from wastewater and biosolids; and (2) a thermal heat exchanger in the United Kingdom that considered the O&M, disposal, and recovery life cycle phases for resource recovery from wastewater. Both sets of scenarios were small-scale case studies, with the former varying from 0.189 to 37.85 m3/day and the latter ranging from 3.12 to 12.5 m3/day. The USA scenarios had higher variability (standard deviation: 0.03 vs. 0.004 kg CO2-eq/MJ of energy recovered), but resulted in lower GHG emissions on average (0.002 vs. 0.08 kg CO2-eq/MJ) (see Figure 3d). This suggests that high recovery efficiency with reduced environmental impacts can be realized when targeting energy recovery from biosolids, which aligns with its implementation across many WWTPs [10,41].
In a similar vein, the nutrient recovery case studies suggested that, on average, lower GHG emissions resulted from nutrient recovery from biosolids (0.97 kg CO2-eq/kg P-eq) relative to urine (5.50 kg CO2-eq/kg P-eq), as shown in Figure 3e. Moreover, the GHG emissions for nitrogen fertilizer tended to result in lower impacts relative to phosphorus fertilizer (mean of 2.09 vs. 5.79 kg CO2-eq/kg P-eq). The null hypothesis could not be rejected for the comparative analyses of nutrient recovery case studies relative to the life cycle phases, location, degree of centralization, or treatment train with the data acquired.

5. A Framework for Resource Recovery

The framework serves to aid the decision-making process for the design of wastewater-based resource recovery systems. It was developed with guidance from practitioners from the Hillsborough County Florida Public Utilities [42], and is summarized in Figure 4.
Step 1: Determine the quantity of reclaimed water that is needed or will be needed according to the projected population for the time span of interest. Source separation is feasible for household or building level recovery (see Figure 5). Existing sub-urban and urban communities typically have sewer connections already in place for treatment at a WWTP, so water reclamation can occur during the conveyance or treatment stages. Since the efficiency of resource recovery varies with scale [10], the wastewater flow rate should be estimated; for reference, about 0.42 m3/capita-day of wastewater is generated in the United States [43]. If water will not be reclaimed, continue to Step 4.
Step 2: Identify the target effluent water quality as determined by the potential end use/end users of the reclaimed water. At a WWTP, three forms of water reclamation are feasible: NPR, IPR, and DPR. Guidelines and local regulations for target effluent water quality for each type are available [10,25].
Step 3: Determine the set of wastewater treatment technologies that are feasible considering the target effluent water quality, treatment scale, and local ordinances. Section 3.1 introduced water reclamation technologies in the context of the effluent water quality achieved in prior life cycle studies. Additional local considerations, such as land availability and proximity to the residential areas, can also be accounted for.
Step 4: Determine if energy and/or nutrients can be recovered for the scope under consideration, e.g., onsite or a WWTP (see Figure 5). End uses for onsite recovery will likely be driven by the need for a particular resource and/or major end users (e.g., local farms or green space for fertilizer). For offsite treatment of combined wastewater flows, energy and/or nutrients can be recovered from conveyance, wastewater treatment, and/or biosolids processing. While biosolids are produced at a rate of ~24.3 kg dry solids/PE-year [44], co-digestion (e.g., with yard or food waste) can be considered when only small quantities are available.
Step 5: Identify local considerations or other constraints that would restrict the recovery of energy and/or nutrients, e.g., local restrictions on the quality and quantity of biosolids for land application during agricultural off-seasons, space constraints, or restrictions on WWTP GHG emissions. Determine the feasible processes for energy and/or nutrient recovery considering these constraints.
Step 6: Determine the life cycle costs of the resource recovery systems as options are prioritized. Costs are influenced by several factors including flow rate, local rates (e.g., for energy, construction, material, and labor), climate, distance to the end users, etc. Readers can refer to Section 4.1 where the LCC of alternative systems were presented.
Step 7: Determine the environmental impact of the resource recovery system. Life cycle GHG emissions are emphasized as they are reported most often in LCAs [10]. Section 4.2 can be referenced for developing these estimates. Considering the prominence of the energy mix in the environmental impacts of wastewater treatment systems, low-emission energy sources should be utilized when feasible.
Step 8: Determine the final configuration of the resource recovery system.

6. Application of the Framework to An Existing Case Study

A step-by-step assessment was conducted through a case study to demonstrate the applicability of the developed framework in finding sustainable solutions for implementation of wastewater-based resource recovery schemes. Accordingly, Ryaverket Wastewater Treatment Facility located at Norra Fågelrovägen 3 in Göteborg (Sweden) was selected, and the results were compared to the findings from the Lundin et al. [33] study that evaluated alternatives for sludge handling at the WWTP. The treatment facility has a design capacity of 91 MGD, providing service for 617,781 habitants in the water service area [33]. The service area, which is approximately 172.9 mi2, has a population density of 3573 capita/mi2 and a population growth rate of 1.19% as of 2021 [45]. The current treatment facility consists of pre-precipitation with Ferrous Sulphate followed by Activated Sludge (2nd Bio) and Nitrogen removal. The produced biosolids are sent to digestors for sludge handling and energy recovery via biogas production. The digestate is used for soil improvement and land reclamation. Currently, the effluent from the treatment plant is being released to the North Sea (environmental reuse).
Lundin et al. [33] evaluated four alternatives for sludge handling at the treatment facility. The alternatives consisted of agricultural use, co-incineration with waste, incineration combined with phosphorus recovery (Bio-Con), and fractionation with phosphorus recovery (Cambi-KREPRO). According to their study, there are also some limitations that need to be considered and addressed while designing for a wastewater-based resource recovery system in the service area. The major limitations consist of: (1) limitation on land application of dewatered digestate due to emerging legislation; (2) concerns regarding the presence of pathogen and harmful chemicals in the produced biosolids; (3) limitation on the capacity of the current incineration system in the city; (4) low phosphorus recovery in the current system; and (5) low demand for dewatered digestate from the farmers. Anaerobic digestion of biosolids followed by incineration combined with phosphorus recovery (Bio-Con), land application, and increasing the capacity of incinerators was the proposed solution in Lundin et al. study [33].

6.1. Framework Application

The developed framework was applied to a design for a resource recovery scheme in the selected treatment facility and its corresponding service area. The assessments associated with each step in the developed framework are presented in this section, and the outcomes from the application study have been summarized in Figure 6.
Step 1: The sewer collection system is currently implemented in the water service area. The service area has a population of 579,281 as of 2019, with an estimated growth rate of 1.19% as of 2021. The current treatment system has a design capacity of 91 MGD, which can provide service for 780,000 habitants in the service area. Since the collection and treatment system is currently implemented and the design capacity of the system can provide service for the residents until 2044 (with a 1.19% population growth rate), larger scale resource recovery systems would be more feasible options for this case. The current system recovers energy in the form of electricity and heat, and no water reclamation scenario (i.e., NPR, IPR, and DPR) has been implemented in the service area.
Step 2: Since the current treatment system consists of pre-precipitation with Ferrous Sulphate followed by an Activated Sludge system and Nitrogen removal, using the reclaimed water for NPR purposes seems to be the more feasible scenario due to the effluent water quality from the current plant. The treatment system is located within the residential area of the city of Gothenburg, Sweden, which makes the distance between the generation of reclaimed water and the residents relatively short. Moreover, the plant is located near an industrial site (within 1 mile of the WWTP). Consideration of these local conditions associated with the treatment facility make Industrial Reuse and Distributed Urban Reuse suitable scenarios for the water service area.
Step 3: To meet the water quality requirements for industrial reuse, the current treatment train (2ndBio) can be extended by implementation of a micro-filtration (MF), followed by UV disinfection (i.e., 2ndBio + 3rdM + UV). The design capacity for this expansion depends on the demand for the reclaimed water, with a maximum of the current system’s design capacity (91 MGD). These modifications also make the effluent water quality suitable for urban reuse scenarios. Hence, the excess reclaimed water (e.g., during the low-demand industrial seasons) can be sent to the residential areas for urban reuse purposes. For industrial reuse, the current system can also be enhanced by implementation of hardness removal, if it is necessary for specific types of industrial reuse options (e.g., for boilers or cooling water that require lower water hardness) at the industrial site (i.e., 2ndBio + 3rdM + 3rd PC). Considering the topography of the current WWTP, expansion of the current facility seems to be a feasible option. If the current treatment system was not in place for this water service area, due to the proximity of the plant’s location to the residential areas and the limitations regarding land application of biosolids in the service area, a membrane bioreactor (MBR) followed by biological nutrient removal and UV disinfection (2ndBio or 3rdBio) would have made a good treatment train for this case. This not only reduces the land requirements for implementation of the treatment facility, but also reduces the volume of the produced sludge and eliminates the higher costs, energy requirements, and GHG emissions associated with more aggressive filtration processes.
Step 4: Since the current treatment system is operating at a large-scale capacity and the collection system is currently in place, according to the correlation analysis in this study, centralized resource recovery would be an economic alternative with lower environmental impacts for this water service area. Recovering energy (in the form of electricity and heat) through digestion of biosolids, which is currently implemented at the WWTP, makes a good solution to reduce the costs and the environmental impacts of the wastewater treatment system. The recovered energy (electricity + heat) can be used onsite for the operation of the treatment facility. To recover the desired amount of nutrients (especially for phosphorus recovery), due to the concerns for pathogen contents of the biosolid, struvite precipitation would be a feasible solution to recover N and P and increase the demand for the land application of the material from farmers in the area. Moreover, the volume of the remaining biosolids can be further reduced by implementation of a centrifuge system to make it more feasible to send the final biosolids to the city’s incinerators. Alternatively, a thickening belt system can be implemented as a more economical solution with lower GHG emissions if lower levels of volume reduction are desired. The final product can be sent to the city’s incinerators for further energy recovery. As the study by Lundin et al. [33] also confirms, co-incineration of biosolids with waste in the incinerators produces the highest amount of energy for the biosolid handling system in the city (approximately 2300 KWh per dry ton of produced sludge). The alternatives would be Incineration combined with phosphorus recovery (Bio-Con) and Fractionation with phosphorus recovery (Cambi-KREPRO). As it was also mentioned by Lundin et al. [33], increasing the operation capacity of the incineration system is also a good solution that decreases the overall environmental impacts of the wastewater system in the area.
Step 5: One local consideration in this service area is the lower demand for biosolids from the farmers in the area. New legislation is also restricting the use of produced biosolids for land applications. Hence, the land application and agricultural use do not seem feasible scenarios for recovering nutrients from the biosolid. Alternatively, struvite precipitation not only increases the demand for the product from the farmers, but also decreases the risk of pathogens, one of the concerns associated with the use of produced biosolids in the area. Moreover, limited capacity of the incinerators makes it challenging to send a higher volume of the sludge to the incineration facility. Further dewatering of the remaining biosolid (along with increasing the capacity of incinerators) makes it more feasible for this type of energy recovery. Since the collection system is currently in place and the city is not located in a very cold region, according to this study, implementation of a thermal energy recovery system does not seem economically and environmentally sustainable. The relatively flat topography of the service area also makes energy recovery through hydropower infeasible.
Step 6: According to the correlation analysis in this study, centralized recovery technologies are more economically feasible for this service area. As the study conducted by Lundin et al. [33] also shows, co-incineration of biosolids with waste has the highest implementation and operation costs; however, it produces the highest amount of energy, which further reduces the overall environmental impacts of the design.
Step 7: According to the correlation analysis in this study, centralized recovery technologies are more environmentally friendly alternatives for this service area. Moreover, co-incineration of biosolids with waste decreases the GHG emissions associated with the system.
Step 8: The final suggested resource recovery scheme for the selected treatment facility is as below.
  • Implementation of additional treatment technologies (3rdM and UV) to the current treatment train at the WWTP (with a capacity that depends on the demand, with a maximum of the current system’s design capacity);
  • Sending the reclaimed water to the industrial site located next to the plant for industrial purposes (NPR), and the excess reclaimed water to the residential areas for urban reuse purposes during the lower-demand industrial seasons;
  • Digestion of the produced sludge to recover energy in the form of biogas (electricity + heat);
  • Implementation of a thickening belt system to further dewater the remaining biosolids at the plant;
  • Implementation of a struvite precipitation system for nutrient recovery from the centrate (filtrate);
  • Sending the remaining dewatered biosolids to the city’s incineration system for further energy recovery;
  • Sending the struvite from the struvite precipitation process and the remaining ash from the incineration system to farmers for land application.

6.2. Case Study Validation

As the outcomes of the assessment show, the developed framework can be successfully applied to propose sustainable solutions for resource recovery from wastewater in the studied area, given its local conditions and operational limitations. The detailed results of the case study assessment are provided in Table A3 of the Appendix A. The proposed treatment train for water reclamation and industrial reuse consists of pre-precipitation followed by activated sludge, microfiltration, and UV disinfection (i.e., 2ndBio + 3rdM), which has a mean SNPV of 2.68 USD/m3 (with a minimum of 0.08 USD/m3 and a maximum of 10.37 USD/m3). For sludge handling, the proposed scenario consists of anaerobic digestion of biosolids for energy recovery followed by a thickening belt filter to dewater the digestate. The centrate from the dewatering system is sent to struvite precipitation for nutrient recovery, and the thickened digestate is sent to the incinerators for additional energy recovery. The remaining ash from incineration system can also be land applied as a fertilizer.
The proposed alternatives for energy recovery (i.e., AD + Incineration) has a mean SNPV of −0.008 USD/MJ (with a minimum of −0.070 USD/MJ and a maximum of 0.028 USD/MJ), and the proposed scenario for nutrient recovery, excluding the struvite precipitation, (i.e., AD + Incineration + LR) has a mean SNPV of 0.18 USD/kg P-eq (with a minimum of 0.16 USD/kg P-eq and a maximum of 0.21 USD/kg P-eq). A study conducted by Ishii and Boyer [29] also shows that the revenue from struvite precipitation exceeds the costs associated with its operation (e.g., MgO inputs, Na3PO4 inputs, and energy requirements), if USD 0.57/kg dry weight is considered as the price of the produced struvite.
The global warming potential (GWP) for the proposed water reclamation system would be ~1 kg CO2-eq/m3, and for the proposed energy recovery system would be ~1.289 kg CO2-eq/MJ. The GHG emissions for the proposed nutrient recovery system would be between ~2.27 kg CO2-eq/kg P [46] (struvite precipitation) and ~221.85 kg CO2-eq/kg P-eq (AD + Incineration + LR). Moreover, a study conducted by Linderholm et al. [46] reviews the LCA associated with the operation of a struvite precipitation system in Sweden. Results of the assessment show that operation of a struvite precipitation system in Sweden has a GWP of ~2.27 kg CO2-eq/kg P, which is significantly lower than recovering P from minerals or from ash [46]. The GWP for the scenario proposed in Lundin et al. [33] study (i.e., AD + Incineration + LR) would be approximately 221.85 kg CO2-eq/kg P-eq for a large-scale system. Considering the significant decrease in the digestate volume, the overall GWP of the proposed scenario in this study would be significantly lower than the GWP (~221.85 kg CO2-eq/kg P-eq) associated with the scenario proposed by Lundin et al. [33]. Additionally, the proposed solution in this study addresses all the limitations in the study area (e.g., emerging legislations that limit land application of biosolids, concerns regarding the presence of pathogens in the produced biosolids, limited incinerator capacity, low phosphorus recovery in the current system, and low demand for digestate), while capable of recovering notably more valuable resources from the produced wastewater in the service area. These considerations further improve the sustainability of the proposed solution, when compared to the scenarios that are introduced in the previous studies.

7. Conclusions

A framework for integrated wastewater management has been presented in conjunction with a case study application. The design of sustainable resource recovery systems was found to be largely driven by the scale of implementation, the location (e.g., as it pertains to the energy mix and water quality restrictions), and the scope of the system considered. Specific costs and GHG emissions were both negatively correlated to scale, which suggests that large scale systems tend to cost less and produce less GHG emissions on a per unit basis—thus, benefiting from economies of scale. Some data sets of the impact assessments had large variations, but nonetheless highlighted resource recovery systems that could achieve comparable or lower impacts. For example, most energy recovery technologies provided a higher revenue than water and nutrient recovery technologies, which highlights an opportunity for a sustained investment in a technology. Future research is recommended to incorporate social impacts to the framework and to embed multi-objective optimization in the decision-making process for the simultaneous recovery of multiple resources across scales.

Author Contributions

Conceptualization, N.D.-E. and Q.Z.; Methodology, N.D.-E. and Q.Z.; Validation, N.D.-E., J.H. and N.R.; Formal Analysis, N.D.-E., J.H. and N.R.; Investigation: N.D.-E., J.H. and N.R.; Writing—Original Draft Preparation, N.D.-E., J.H. and N.R.; Writing—Review and Editing, N.D.-E., J.H., N.R. and Q.Z.; Supervision: N.D.-E. and Q.Z.; Project Administration, Q.Z., Funding Acquisition, Q.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Science Foundation Faculty Early Career Development (CAREER) grant of the United States (No. 1454559). Any opinions, findings, and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the National Science Foundation.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author, Q.Z., upon reasonable request.

Acknowledgments

The authors would like to thank Luke Mulford and Gita Iranipour from Hillsborough County for their input during the framework development process.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Appendix A.1. Case studies Included in the Correlation Analysis

Table A1. The journal articles which include a life cycle assessment of water, energy, and nutrient recovery are included in this review [19,27,29,30,31,32,33,34,35,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100]. A cost analysis is conducted in the articles with an asterisk (*), and (^) shows articles that appear in multiple columns (multiple types of resources were recovered). In addition, references [26,28,100,101,102,103,104,105] focused on cost analysis discussion.
Table A1. The journal articles which include a life cycle assessment of water, energy, and nutrient recovery are included in this review [19,27,29,30,31,32,33,34,35,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100]. A cost analysis is conducted in the articles with an asterisk (*), and (^) shows articles that appear in multiple columns (multiple types of resources were recovered). In addition, references [26,28,100,101,102,103,104,105] focused on cost analysis discussion.
References
Water[19,47,48,49,50,51,52,53,58,59,61,62,63,66], [54,55,56,60,64,65] *, [57] *^
Energy[48,62,70,71,72,74,76,77,78,80,82,83], [27,33,49,55,68,69,73,75,79,81,84] *, [50,53,67] ^
Nutrient[32,35,46,72,73,80,85,86,87,88,89,91,92,93,94,95,96,97,98,99], [29,30,31,34,90] *, [67,70] ^, [33,57] *^
Cost analysis[26,28,100,101,102,103,104,105]

Appendix A.2. Correlation Analysis Table

Table A2. The table shows the screened result of paired correlation analysis with p-value larger than 0.05, and coefficient higher than 0.5 or lower than −0.5. Abbreviations—2nd Bio: secondary biological treatment; 3rd PC: tertiary physical and/or chemical treatment; 3rd Bio: tertiary biological treatment; AD: anaerobic digestion; BOD: biological oxygen demand; Chem: chemical treatment; COD: chemical oxygen demand; eff: effluent; GHG: greenhouse gasses; in: influent; LR: land reclamation; M: membrane filtration; N: number of data points used in the analysis; r: correlation coefficient; RR: removal rate; STW: sludge treatment wetland; TD: thermal drying; TN: total nitrogen; TP: total phosphorus; TSS: total suspended solid; USS: urine source separation; V: volume reduction treatment.
Table A2. The table shows the screened result of paired correlation analysis with p-value larger than 0.05, and coefficient higher than 0.5 or lower than −0.5. Abbreviations—2nd Bio: secondary biological treatment; 3rd PC: tertiary physical and/or chemical treatment; 3rd Bio: tertiary biological treatment; AD: anaerobic digestion; BOD: biological oxygen demand; Chem: chemical treatment; COD: chemical oxygen demand; eff: effluent; GHG: greenhouse gasses; in: influent; LR: land reclamation; M: membrane filtration; N: number of data points used in the analysis; r: correlation coefficient; RR: removal rate; STW: sludge treatment wetland; TD: thermal drying; TN: total nitrogen; TP: total phosphorus; TSS: total suspended solid; USS: urine source separation; V: volume reduction treatment.
Sample 1Sample 2Nrp-Value
Water
TSS inTP eff7−0.9990
TP effTP RR13−0.9940
COD effTP RR12−0.9840
COD inBOD in130.9790
COD effTSS RR60.9690.001
TP RRCOD RR120.9660
TP effTSS RR7−0.9650
TN inTSS RR7−0.9580.001
BOD inTSS RR70.9580.001
TSS inTSS RR70.9580.001
COD effCOD RR20−0.9570
TP effCOD RR12−0.9480
TN effTN RR20−0.9480
TP inBOD in130.8630
TN inTP eff140.8570
TP inBOD RR120.8550
BOD effBOD RR12−0.8340.001
BOD inCOD eff130.7840.002
BOD inTP RR13−0.7510.003
BOD RR2nd Bio + 3rd PC12−0.7420.006
TP effCentralized14−0.7220.004
TP effDecentralized140.7220.004
COD inTP eff130.7190.006
BOD inBOD RR120.7110.01
COD RRcost (USD2019/m3)13−0.6940.008
COD effcost (USD2019/m3)130.6890.009
BOD eff2nd Bio + 3rd PC200.6740.001
TN inCOD in210.6590.001
COD inTP RR13−0.6550.015
TP inCOD in210.6430.002
BOD inCOD RR12−0.640.025
COD inBOD RR120.6340.027
COD in2nd Bio + 3rd Bio210.6280.002
cost (USD2019/m3)2nd Bio + 3rd Bio14−0.6250.017
TN RRCOD RR190.6070.006
cost (USD2019/m3)GHG (kg CO2-eq/m3)140.5920.026
cost (USD2019/m3)Decentralized140.5860.028
TP effBOD RR120.5860.045
cost (USD2019/m3)Centralized14−0.5860.028
TN RRDecentralized20−0.5810.007
TN RRCentralized200.5810.007
TP inTN in210.5750.006
TN inCOD RR200.5570.011
COD RR2nd Bio + 3rd Bio200.5560.011
TN eff2nd Bio + 3rd Bio21−0.5510.01
BOD effscale (m3/day)200.5470.013
cost (USD2019/m3)scale (m3/day)14−0.5350.049
COD effTN RR19−0.5310.019
TN RR2nd Bio + 3rd Bio200.5180.019
TN eff2nd Bio + 3rd PC210.4840.026
TN inDecentralized220.4760.025
TN inCentralized22−0.4760.025
TN effCOD RR20−0.4720.036
BOD effGHG (kg CO2-eq/m3)20−0.460.041
GHG (kg CO2-eq/m3)scale (m3/day)22−0.4530.034
COD in2nd Bio + 3rd M21−0.4410.045
COD inTN eff21−0.4340.049
Energy
m3/daysludge (tons/day)80.9510
cost (USD2019/MJ)sludge (tons/day)14−0.8390
wastewaterheat exchanger1740.7010
wastewatersludge174−0.6570
wastewater + sludgeAD + landfilling1740.4890
wastewaterhydropower1740.5050
wastewaterAD + composting174−0.4610
sludgeheat exchanger174−0.460
wastewaterwastewater + sludge174−0.4240
sludgewastewater + sludge174−0.4040
hydropowerscale (m3/day)1160.3960
sludgeAD + incineration1740.3740
wastewater + sludgeAD + composting1740.3520
sludgehydropower174−0.3320
cost (USD2019/MJ)AD + landfilling112−0.3310
heat exchangerAD + composting174−0.3230
wastewaterAD + landfilling174−0.3160
wastewater + sludgeheat exchanger174−0.2970
sludgeV + incineration1740.280
cost (USD2019/MJ)wastewater + sludge112−0.2740.003
GHG (kg CO2-eq/MJ)V + incineration1270.2740.002
wastewaterAD + incineration174−0.2460.001
heat exchangerhydropower174−0.2440.001
sludgeTD + AD + pyrolysis1740.2390.002
hydropowerAD + composting174−0.2330.002
heat exchangerAD + landfilling174−0.2210.003
wastewaterV + incineration174−0.2160.004
GHG (kg CO2-eq/MJ)AD + landfilling127−0.2150.015
wastewater + sludgehydropower174−0.2140.005
AD + landfillingAD + composting174−0.2110.005
sludgeAD + pyrolysis1740.1940.01
sludgeTD + pyrolysis1740.1940.01
heat exchangerscale (m3/day)116−0.1840.049
heat exchangerAD + incineration174−0.1720.023
sludgeAD + composting1740.1730.023
AD + incinerationAD + composting174−0.1640.03
hydropowerAD + landfilling174−0.1590.036
wastewaterTD + AD + pyrolysis174−0.1570.039
wastewater + sludgeAD + incineration174−0.1510.047
heat exchangerV + incineration174−0.1510.047
Nutrient
urineUSS + Chem + fertilizer1350.8960
sludgeurine135−0.7270
sludgeUSS + Chem + fertilizer135−0.6520
urineAD + fertilizer135−0.6090
sludgeAD + fertilizer1350.5680
USS + fertigation + AD + fertilizerscale (m3/day)1310.5390
urine + faeces + greywaterscale (m3/day)1310.5390
AD + fertilizerUSS + Chem + fertilizer135−0.5450
wastewater + sludgedrying + fertilizer1350.3410
sludgewastewater + sludge135−0.3120
AD + incineration + LRsludge(tons/day)570.3430.009
GHG (kg CO2-eq/kg P-eq)wastewater + sludge1350.280.001
sludge(tons/day)scale (m3/day)570.2750.038
urineUSS + M + fertigation1350.2650.002
wastewater + sludgeAD + fertilizer1350.250.003
GHG (kg CO2-eq/kg P-eq)USS + Chem + fertilizer135−0.2160.012
AD + fertilizerscale (m3/day)131−0.2110.016
drying + incineration + LRAD + fertilizer135−0.1990.021
STW + dewatering + fertilizerAD + fertilizer135−0.1990.021
GHG (kg CO2-eq/kg P-eq)urine135−0.1940.024
sludgeurine + faeces + greywater135−0.1920.025
sludgeUSS + fertigation + AD + fertilizer135−0.1920.025
sludgeUSS + M + fertigation135−0.1920.025
wastewater + sludgeurine135−0.1870.03
urineUSS + drying + fertilizer1350.1860.031
AD + incineration + LRAD + fertilizer135−0.1810.036
sludgeAD + incineration + LR1350.1780.039
sludgescale (m3/day)131−0.1720.049

Appendix A.3. Application of the Framework to Existing Studies

The Ryaverket Wastewater Treatment Facility located at Norra Fågelrovägen 3 in Göteborg, Sweden, was selected to evaluate the applicability of the developed framework in the sustainable design of recovery systems. The treatment facility has a design capacity of 91 MGD, providing service for 617,781 habitants in the water service area [70]. The service area, which is approximately 172.9 mi2, has a population density of 3573 capita/mi2 and a population growth rate of 1.19% as of 2021 [45].
The current treatment facility consists of pre-precipitation with Ferrous Sulphate followed by Activated Sludge (2nd Bio) and Nitrogen removal. The produced biosolids are sent to digestors for sludge handling and energy recovery via biogas production. The digestate is used for soil improvement and land reclamation. Currently, the effluent from the treatment plant is being released to the North Sea (environmental reuse).
Lundin et al. [74] evaluated four alternatives for sludge handling at the treatment facility. The alternatives consisted of agricultural use, co-incineration with waste, incineration combined with phosphorus recovery (Bio-Con), and fractionation with phosphorus recovery (Cambi-KREPRO). According to Lundin et al.’s [74] study, there are also some limitations that need to be considered and addressed while designing for a wastewater-based resource recovery system in the service area. The major limitations consist of:
  • Limitation on land application of dewatered digestate due to emerging legislation;
  • Concerns regarding the presence of pathogen and harmful chemicals in the produced biosolids;
  • Limitation on the capacity of the current incineration system in the city;
  • Low P recovery in the current system;
  • Low demand for dewatered digestate from the farmers.
Table A3. Results of the case study assessment for to apply and validate the developed framework. Abbreviations—MGD: million gallons per day; WWTP: wastewater treatment plant; NPR: non-potable reuse; IPR: indirect potable reuse; DPR: direct potable reuse; Bio: biological; M: membrane; UV: ultraviolet; MF: microfiltration; PC: physical chemical; CAS: conventional activated sludge; GHG: greenhouse gas; KWh: kilowatt hour; LCC: life cycle cost; LCA: life cycle assessment.
Table A3. Results of the case study assessment for to apply and validate the developed framework. Abbreviations—MGD: million gallons per day; WWTP: wastewater treatment plant; NPR: non-potable reuse; IPR: indirect potable reuse; DPR: direct potable reuse; Bio: biological; M: membrane; UV: ultraviolet; MF: microfiltration; PC: physical chemical; CAS: conventional activated sludge; GHG: greenhouse gas; KWh: kilowatt hour; LCC: life cycle cost; LCA: life cycle assessment.
StepAssessment
Step 1: Determine the quantity of reclaimed water that is needed or will be needed according to the projected population in the considered time span for the design. What is the scope of resource recovery, i.e., onsite recovery (household or building level) or larger scales such as a wastewater treatment plant? If planning is occurring at the household or building level, source separation would be feasible to separate wastewater streams (see Figure 3b). Sub-urban and urban communities typically have sewer connections already in place for treatment at a wastewater treatment facility, so resource recovery can occur during conveyance, wastewater treatment, or biosolids processing. Opportunities for resource recovery vary by scale8, so the wastewater flow rate should be estimated. For reference, about 0.42 m3/capita-day of wastewater is generated in the United States [19]. If water will not be reclaimed, continue to Step 4.The sewer collection system is currently implemented in the water service area. The service area has a population of 579,281 as of 2019, with an estimated growth rate of 1.19% as of 2021. The current treatment system has a design capacity of 91 MGD, which can provide service for 780,000 habitants in the service area. Since the collection and treatment system is currently implemented and the design capacity of the system can provide service for the residents until 2044 (with 1.19% population growth rate), larger scale resource recovery systems would be more feasible options for this case. The current system recovers energy in the form of electricity and heat, and no water reclamation scenario (i.e., NPR, IPR, and DPR) has been implemented in the service area.
Step 2: Identify the target effluent water quality as determined by the potential end use/end users of the reclaimed water. At a WWTP, three forms of water reclamation are feasible: non-potable reuse, indirect potable reuse (IPR), and direct potable reuse (DPR). Guidelines and local regulations for target effluent water quality for each type are available [8,20].Since the current treatment system consists of pre-precipitation with Ferrous Sulphate followed by an Activated Sludge system and Nitrogen removal, using the reclaimed water for NPR purposes seem to be more feasible scenarios due to the effluent water quality from the current plant. The treatment system is located within the residential area in the city of Gothenburg, Sweden, which makes the distance between generation of reclaimed water and the residents relatively short. Moreover, the plant is located near an industrial site (within 1 mile of the WWTP). Consideration of these local conditions associated with the treatment facility make Industrial Reuse and Distributed Urban Reuse suitable scenarios for the water service area.
Step 3: With consideration of the target effluent water quality, the scale of treatment, and local ordinances, determine the set of wastewater treatment technologies that can feasibly be applied. Readers can refer back to Section 3.1, which discussed the technologies that have been implemented for water reclamation in prior life cycle studies and the effluent water quality achieved for the technologies used in the studies. Other considerations may also be taken into account in this step, such as restrictions on implementation of specific treatment technologies. For instance, some treatment technologies such as CAS require larger area for implementation. For urban areas with limitations on land availability, this treatment technology may not be a feasible option. Moreover, other restrictions such as proximity to the residential areas may also limit selection and implementation of treatment technologies with higher level of odor issues (e.g., anaerobic treatment techniques).To meet the water quality requirements for industrial reuse, the current treatment train (2ndBio) can be extended by implementation of a micro-filtration (MF), followed by UV disinfection (i.e., 2ndBio + 3rdM + UV). The design capacity for this expansion depends on the demand for the reclaim water, with the maximum of current system’s design capacity (91 MGD). These modifications also make the effluent water quality suitable for urban reuse scenarios. Hence, the excess reclaimed water (e.g., during the low-demand industrial seasons) can be sent to the residential areas for urban reuse purposes. For industrial reuse, the current system can also be enhanced by implementation of hardness removal, if it is necessary for specific type of industrial reuse options (e.g., using for boilers or cooling water that require lower hardness in the water) in the industrial site (i.e., 2ndBio + 3rdM + 3rd PC). Considering the topography of the current WWTP, expansion of the current facility seems to be a feasible option.If the current treatment system was not in place for this water service area, due to the proximity of the plant’s location to the residential areas and the limitations regarding land application of biosolids in the service area, membrane bioreactor (MBR) followed by biological nutrients removal and UV disinfection (2ndBio or 3rdBio) would have made a good treatment train for this case. This not only reduces the land requirements for implementation of the treatment facility, but also reduced the volume of the produced sludge and eliminates the higher costs, energy requirements, and GHG emissions associated with more aggressive filtration processes.
Step 4: Determine if energy and/or nutrients can be recovered for the scope under consideration (i.e., onsite vs. a WWTP). Figure 3b presents options for the recovery of energy and nutrients as well (adapted from Diaz-Elsayed et al. [8]).For onsite recovery, consider the potential end use applications of energy and nutrients: fertilizer, fertigation, and thermal energy recovery. Fertilizer can also be transported offsite for local use if there is not an immediate need onsite. The end use of the recovered resource will likely be driven by the need for a particular resource and/or the major end users available to consume the recovered resource (e.g., local farms or green space for fertilizer).For combined wastewater flows that are treated offsite (e.g., in a sub-urban or urban community), several stages can be considered for energy and nutrient recovery:
  • Conveyance: Hydropower can be generated during conveyance and used directly for pumping stations. Additionally, thermal energy can be recovered before or after arriving at the WWTP and used as a heat source for the community;
  • Wastewater Treatment: If nutrients remain in the reclaimed water, then fertigation is feasible. Additionally, thermal energy can be recovered for district heating (if it was not done so prior to treatment).
Biosolids Processing: During this stage, energy and/or nutrients can be recovered via anaerobic digestion, composting, or combustion processes (see Figure 3b). The rate of biosolids production for a community can be approximated at 24.3 kg dry solids/PE-year if data are not readily available [21]. When the rate of biosolids generation is relatively small, co-digestion with other waste (e.g., yard or food waste) can be considered.
Since the current treatment system is operating on a large-scale capacity and the collection system is currently in place, according to the correlation analysis in this study, centralized resource recovery would be an economic alternative with lower environmental impacts for this water service area. Recovering energy (in form of electricity and heat) through digestion of biosolids, which is currently implemented at the WWTP, makes a good solution to reduce the costs and the environmental impacts of the wastewater treatment system. The recovered energy (electricity + heat) can be used onsite for the operation treatment facility. To recover the desired amount of nutrients (especially for phosphorus recovery), due to the concerns on pathogen contents of the biosolid, struvite precipitation would be a feasible solution to recover N and P and increase the demand for the land application of the material from the farmers in the area. Moreover, volume of the remaining biosolids can be further reduced by implementation of a centrifuge system to make it more feasible to send the final biosolids to the city’s incinerators. Alternatively, a thickening belt system can be implemented as a more economical solution with lower GHG emissions if lower levels of volume reduction are desired. The final product can be sent to the city’s incinerators for further energy recovery. As the study by Lundin et al. (2004) also confirms, co-incineration of biosolids with waste in the incinerators produces the highest among of energy recovery for the biosolid handling system in the city (approximately 2300 KWh per dry ton of produced sludge). The alternatives would be Incineration combined with phosphorus recovery (Bio-Con) and Fractionation with phosphorus recovery (Cambi-KREPRO). As it was also mentioned by Lundin et al. (2004), increasing the operation capacity of incineration system is also a good solution that decreases the overall environmental impacts of the wastewater system in the area.
Step 5: Are there local considerations or other constraints to account for that would restrict the recovery of energy and/or nutrients? For example, are there restrictions on the quality and quantity (especially during agricultural off-seasons) of biosolids to be reused for land application or restrictions on emissions from the incineration process. If so, what are the set of feasible options for resource recovery?One local consideration in this service area is the lower demand for biosolids from the farmers in the area. New legislations are also restricting the use of produced biosolids for land applications. Hence, the land application and agricultural use do not seem feasible scenarios for recovering nutrients from the biosolid. Alternatively, struvite precipitation not only increases the demand for the product from the farmers, but also decreases the risk of pathogens, one of the concerns associated with the use of produced biosolids in the area. Moreover, limited capacity of the incinerators makes it challenging to send a higher volume of the sludge to the incineration facility. Further dewatering of the remaining biosolid (along with increasing the capacity of incinerators) makes it more feasible for this type of energy recovery. Since the collection system is currently in place and the city is not located in a very cold region, according to this study, implementation of a thermal energy recovery system does not seem economically and environmentally sustainable. The relatively flat topography of the service area also makes energy recovery through hydropower infeasible.
Step 6: Determine the life cycle costs (LCC) of the resource recovery systems as options are prioritized. The LCC can be influenced by a variety of factors including the flow rate, local costs (e.g., energy, construction, material, and labor), climate, the distance to the end users, etc. Readers can refer back to Section 4.1 where the LCC of alternative resource recovery systems were presented.According to the correlation analysis in this study, centralized recovery technologies are more economically feasible for this service area. As the study conducted by Lundin et al. (2004) also shows, co-incineration of biosolids with waste has the highest implementation and operation costs; however, it produces the highest amount of energy, which also further reduces the overall environmental impacts of the design.
Step 7: Determine the environmental impact of the resource recovery system. Life cycle GHG emissions are reported most often in LCAs [8]; a comparison of the GHG emissions for resource recovery systems were presented in Section 4.2 for reference.According to the correlation analysis in this study, centralized recovery technologies are more environmentally friendly alternatives for this service area. Moreover, co-incineration of biosolids with waste decreases the GHG emissions associated with the system.
Step 8: Determine the final configuration of the resource recovery system.
  • Implementation of additional treatment technologies to add (3rdM and UV) to the current treatment train at the WWTP (with a capacity that depends on the demand, with a maximum of the current system’s design capacity).
  • Sending the reclaimed water to the industrial site located next to the plant for industrial purposes (NPR), and the excess reclaimed water to the residential areas for urban reuse purposes during the lower-demand industrial seasons.
  • Digestion of the produced sludge to recover energy in the form of biogas (electricity + heat).
  • Implementation of a thickening belt system to further dewater the remaining biosolids at the plant.
  • Implementation of a struvite precipitation system for nutrient recovery from the centrate (filtrate).
  • Sending the remaining dewatered biosolids to the city’s incineration system for further energy recovery.
  • Sending the struvite from the struvite precipitation process and the remaining ash from incineration system to the farmers for land applications.

References

  1. UN Water. Summary Progress: Update 2021: SDG 6—Water and Sanitation for All. 2021. Available online: https://www.unwater.org/sites/default/files/app/uploads/2021/07/SDG-6-Summary-Progress-Update-2021_Version-July-2021.pdf (accessed on 20 December 2022).
  2. Goss, M.; Swain, D.L.; Abatzoglou, J.T.; Sarhadi, A.; Kolden, C.A.; Williams, A.P.; Diffenbaugh, N.S. Climate change is increasing the likelihood of extreme autumn wildfire conditions across California. Environ. Res. Lett. 2020, 15, 094016. [Google Scholar] [CrossRef] [Green Version]
  3. Knutson, T.R.; McBride, J.L.; Chan, J.; Emanuel, K.; Holland, G.; Landsea, C.; Held, I.; Kossin, J.P.; Srivastava, A.K.; Sugi, M. Tropical cyclones and climate change. Nat. Geosci. 2010, 33 3, 157–163. [Google Scholar] [CrossRef] [Green Version]
  4. Craig, G. Towards the Measurement of Empowerment: The Evaluation of Community Development. J. Community Dev. Soc. 2002, 33, 124–146. [Google Scholar] [CrossRef]
  5. Pigg, K.E. Three Faces of Empowerment: Expanding the Theory of Empowerment in Community Development. J. Community Dev. Soc. 2002, 33, 107–123. [Google Scholar] [CrossRef]
  6. Pinkett, R.; O’Bryant, R. Building Community, Empowerment and Self-sufficiency. Inf. Commun. Soc. 2003, 6, 187–210. [Google Scholar] [CrossRef]
  7. Buheji, M.; Vovk Korže, A.; Eidan, S.; Abdulkareem, T.A.; Perepelkin, N.; Mavric, B.; Preis, J.; Bartula, M.; Ahmed, D.; Buheji, A.; et al. Optimising Pandemic Response through Self-Sufficiency—A Review Paper. Am. J. Econ. 2020, 10, 277–283. [Google Scholar] [CrossRef]
  8. Sioen, G.; Sekiyama, M.; Terada, T.; Yokohari, M. Post-Disaster Food and Nutrition from Urban Agriculture: A Self-Sufficiency Analysis of Nerima Ward, Tokyo. Int. J. Environ. Res. Public Health 2017, 14, 748. [Google Scholar] [CrossRef] [Green Version]
  9. Diaz-Elsayed, N.; Mo, W.; Zhang, Q. The Sustainability Dimensions of Resource Recovery from “Wastewater.”. In Resource Recovery from Wastewater: Toward Sustainability, 1st ed.; Gude, V.G., Ed.; Apple Academic Press Inc.: New York, USA, 2022; pp. 24–66. [Google Scholar]
  10. Diaz-Elsayed, N.; Rezaei, N.; Guo, T.; Mohebbi, S.; Zhang, Q. Wastewater-based resource recovery technologies across scale: A review. Resour. Conserv. Recycl. 2019, 145, 94–112. [Google Scholar] [CrossRef]
  11. Gwak, G.; Kim, D.I.; Hong, S. New industrial application of forward osmosis (FO): Precious metal recovery from printed circuit board (PCB) plant wastewater. J. Memb. Sci. 2018, 552, 234–242. [Google Scholar] [CrossRef]
  12. Wu, D.; Lu, D.; Sun, F.; Zhou, Y. Process optimization for simultaneous antibiotic removal and precious metal recovery in an energy neutral process. Sci. Total Environ. 2019, 695, 133914. [Google Scholar] [CrossRef]
  13. Byrne, D.M.; Lohman, H.A.C.; Cook, S.M.; Peters, G.M.; Guest, J.S. Life cycle assessment (LCA) of urban water infrastructure: Emerging approaches to balance objectives and inform comprehensive decision-making. Environ. Sci. Water Res. Technol. 2017, 3, 1002–1014. [Google Scholar] [CrossRef]
  14. Corominas, L.; Byrne, D.; Guest, J.S.; Hospido, A.; Roux, P.; Shaw, A.; Short, M.D. The application of life cycle assessment (LCA) to wastewater treatment: A best practice guide and critical review. Water Res. 2020, 184, 116058. [Google Scholar] [CrossRef] [PubMed]
  15. Diaz-Elsayed, N.; Rezaei, N.; Ndiaye, A.; Zhang, Q. Trends in the environmental and economic sustainability of wastewater-based resource recovery: A review. J. Clean. Prod. 2020, 265, 121598. [Google Scholar] [CrossRef]
  16. Loubet, P.; Roux, P.; Loiseau, E.; Bellon-Maurel, V. Life cycle assessments of urban water systems: A comparative analysis of selected peer-reviewed literature. Water Res. 2014, 67, 187–202. [Google Scholar] [CrossRef] [PubMed]
  17. Trimmer, J.T.; Miller, D.C.; Byrne, D.M.; Lohman, H.A.C.; Banadda, N.; Baylis, K.; Cook, S.M.; Cusick, R.D.; Jjuuko, F.; Margenot, A.J.; et al. Re-Envisioning Sanitation As a Human-Derived Resource System. Environ. Sci. Technol. 2020, 54, 10446–10459. [Google Scholar] [CrossRef] [PubMed]
  18. Chen, Z.; Wu, Q.; Wu, G.; Hu, H.-Y. Centralized water reuse system with multiple applications in urban areas: Lessons from China’s experience. Resour. Concerv. Recycl. 2017, 117, 125–136. [Google Scholar] [CrossRef] [Green Version]
  19. Rezaei, N.; Diaz-Elsayed, N.; Mohebbi, S.; Xie, X.; Zhang, Q. A multi-criteria sustainability assessment of water reuse applications: A case study in Lakeland, Florida. Environ. Sci. Water Res. Technol. 2019, 5, 102–118. [Google Scholar] [CrossRef]
  20. Rezaei, N.; Sierra-Altamiranda, A.; Diaz-Elsayed, N.; Charkhgard, H.; Zhang, Q. A multi-objective optimization model for decision support in water reclamation system planning. J. Clean. Prod. 2019, 240, 118227. [Google Scholar] [CrossRef]
  21. Brent, D.A.; Cook, J.H.; Olsen, S. Social Comparisons, Household Water Use, and Participation in Utility Conservation Programs: Evidence from Three Randomized Trials. J. Assoc. Environ. Resour. Econ. 2015, 2, 597–627. [Google Scholar] [CrossRef]
  22. Sharvelle, S.; Dozier, A.; Arabi, M.; Reichel, B. A geospatially-enabled web tool for urban water demand forecasting and assessment of alternative urban water management strategies. Environ. Model. Softw. 2017, 97, 213–228. [Google Scholar] [CrossRef]
  23. Kavvada, O.; Horvath, A.; Stokes-Draut, J.R.; Hendrickson, T.P.; Eisenstein, W.A.; Nelson, K.L. Assessing Location and Scale of Urban Nonpotable Water Reuse Systems for Life-Cycle Energy Consumption and Greenhouse Gas Emissions. Environ. Sci. Technol. 2016, 50, 13184–13194. [Google Scholar] [CrossRef]
  24. Haque, M.M.; Egodawatta, P.; Rahman, A.; Goonetilleke, A. Assessing the significance of climate and community factors on urban water demand. Int. J. Sustain. Built Environ. 2015, 4, 222–230. [Google Scholar] [CrossRef]
  25. US EPA. 2012 Guidelines for Water Reuse; US EPA: Washington, DC, USA, 2012. [Google Scholar]
  26. Ravichandran, A.; Diaz-Elsayed, N.; Thomas, S.; Zhang, Q. An Assessment of the Influence of Local Conditions on the Economic and Environmental Sustainability of Drain Water Heat Recovery Systems. J. Clean. Prod. 2020, 279, 123589. [Google Scholar] [CrossRef]
  27. Chae, K.J.; Kang, J. Estimating the energy independence of a municipal wastewater treatment plant incorporating green energy resources. Energy Convers. Manag. 2013, 75, 664–672. [Google Scholar] [CrossRef]
  28. Power, C.; McNabola, A.; Coughlan, P. Development of an evaluation method for hydropower energy recovery in wastewater treatment plants: Case studies in Ireland and the UK. Sustain. Energy Technol. Assess. 2014, 7, 166–177. [Google Scholar] [CrossRef]
  29. Ishii, S.K.L.; Boyer, T.H. Life cycle comparison of centralized wastewater treatment and urine source separation with struvite precipitation: Focus on urine nutrient management. Water Res. 2015, 79, 88–103. [Google Scholar] [CrossRef]
  30. Landry, K.A.; Boyer, T.H. Life cycle assessment and costing of urine source separation: Focus on nonsteroidal anti-inflammatory drug removal. Water Res. 2016, 105, 487–495. [Google Scholar] [CrossRef] [Green Version]
  31. Longo, S.; Frison, N.; Renzi, D.; Fatone, F.; Hospido, A. Is SCENA a good approach for side-stream integrated treatment from an environmental and economic point of view? Water Res. 2017, 125, 478–489. [Google Scholar] [CrossRef] [Green Version]
  32. Murray, A.; Horvath, A.; Nelson, K.L. Hybrid Life-Cycle Environmental and Cost Inventory of Sewage Sludge Treatment and End-Use Scenarios: A Case Study from China. Environ. Sci. Technol. 2008, 42, 3163–3169. [Google Scholar] [CrossRef] [Green Version]
  33. Lundin, M.; Olofsson, M.; Pettersson, G.; Zetterlund, H. Environmental and economic assessment of sewage sludge handling options. Resour. Conserv. Recycl. 2004, 41, 255–278. [Google Scholar] [CrossRef]
  34. Shi, Y.; Zhou, L.; Xu, Y.; Zhou, H.; Shi, L. Life cycle cost and environmental assessment for resource-oriented toilet systems. J. Clean. Prod. 2018, 196, 1187–1197. [Google Scholar] [CrossRef]
  35. Hong, J.; Hong, J.; Otaki, M.; Jolliet, O. Environmental and economic life cycle assessment for sewage sludge treatment processes in Japan. Waste Manag. 2009, 29, 696–703. [Google Scholar] [CrossRef]
  36. IGES. Greenhouse Gas (GHG) Emission Database, Version [11.0]. Institute for Global Environmental Strategies (IGES) [WWW Document]. 2021. Available online: https://pub.iges.or.jp/pub/iges-ghg-emissions-database (accessed on 21 October 2021).
  37. Our World in Data. Per Capita Electricity Generation. 2011. Available online: https://ourworldindata.org/grapher/per-capita-electricity-consumption?time=2011 (accessed on 21 October 2021).
  38. The World Bank Group. World Development Indicators. Available online: https://databank.worldbank.org/reports.aspx?source=2&series=SP.POP.TOTL&country= (accessed on 31 July 2021).
  39. Our World in Data. Per Capita Electricity Generation. 2015. Available online: https://ourworldindata.org/grapher/per-capita-electricity-consumption?time=2015 (accessed on 21 October 2021).
  40. Our World in Data. Per Capita Electricity Generation. 2019. Available online: https://ourworldindata.org/grapher/per-capita-electricity-consumption?time=2019 (accessed on 21 October 2021).
  41. Gu, Y.; Li, Y.; Li, X.; Luo, P.; Wang, H.; Robinson, Z.P.; Wang, X.; Wu, J.; Li, F. The feasibility and challenges of energy self-sufficient wastewater treatment plants. Appl. Energy 2017, 204, 1463–1475. [Google Scholar] [CrossRef] [Green Version]
  42. Hillsborough County Public Utilities; Tampa, FL, USA. Personal Communication, 2020.
  43. Metcalf & Eddy; Tchobanoglous, G.; Stensel, H.D.; Tsuchihashi, R.; Burton, F.L. Wastewater Engineering: Treatment and Resource Recovery, 5th ed.; McGraw-Hill Education: New York, NY, USA, 2014. [Google Scholar]
  44. Kelessidis, A.; Stasinakis, A.S. Comparative study of the methods used for treatment and final disposal of sewage sludge in European countries. Waste Manag. 2012, 32, 1186–1195. [Google Scholar] [CrossRef] [PubMed]
  45. World Population Review. Available online: https://worldpopulationreview.com/world-cities/gothenburg-population (accessed on 14 June 2021).
  46. Linderholm, K.; Tillman, A.-M.; Mattsson, J.E. Life cycle assessment of phosphorus alternatives for Swedish agriculture. Resour. Conserv. Recycl. 2012, 66, 27–39. [Google Scholar] [CrossRef]
  47. Muñoz, I.; Rodríguez, A.; Rosal, R.; Fernández-Alba, A.R. Life Cycle Assessment of urban wastewater reuse with ozonation as tertiary treatment. A focus on toxicity-related impacts. Sci. Total Environ. 2009, 407, 1245–1256. [Google Scholar] [CrossRef]
  48. Cornejo, P.K.; Zhang, Q.; Mihelcic, J.R. How does scale of implementation impact the environmental sustainability of wastewater treatment integrated with resource recovery? Environ. Sci. Technol. 2016, 50, 6680–6689. [Google Scholar] [CrossRef] [PubMed]
  49. Hendrickson, T.P.; Nguyen, M.T.; Sukardi, M.; Miot, A.; Horvath, A.; Nelson, K.L. Life-Cycle Energy Use and Greenhouse Gas Emissions of a Building-Scale Wastewater Treatment and Nonpotable Reuse System. Environ. Sci. Technol. 2015, 49, 10303–10311. [Google Scholar] [CrossRef] [PubMed]
  50. Cornejo, P.K.; Zhang, Q.; Mihelcic, J.R. Quantifying benefits of resource recovery from sanitation provision in a developing world setting. J. Environ. Manag. 2013, 131, 7–15. [Google Scholar] [CrossRef]
  51. Meneses, M.; Pasqualino, J.C.; Castells, F. Environmental assessment of urban wastewater reuse: Treatment alternatives and applications. Chemosphere 2010, 81, 266–272. [Google Scholar] [CrossRef] [PubMed]
  52. Pasqualino, J.C.; Meneses, M.; Castells, F. Life Cycle Assessment of Urban Wastewater Reclamation and Reuse Alternatives. J. Ind. Ecol. 2011, 15, 49–63. [Google Scholar] [CrossRef]
  53. Shiu, H.-Y.; Lee, M.; Chiueh, P.-T. Water reclamation and sludge recycling scenarios for sustainable resource management in a wastewater treatment plant in Kinmen islands, Taiwan. J. Clean. Prod. 2017, 152, 369–378. [Google Scholar] [CrossRef]
  54. García-Montoya, M.; Sengupta, D.; Nápoles-Rivera, F.; Ponce-Ortega, J.M.; El-Halwagi, M.M. Environmental and economic analysis for the optimal reuse of water in a residential complex. J. Clean. Prod. 2016, 130, 82–91. [Google Scholar] [CrossRef]
  55. Cashman, S.; Ma, X.; Mosley, J.; Garland, J.; Crone, B.; Xue, X. Energy and greenhouse gas life cycle assessment and cost analysis of aerobic and anaerobic membrane bioreactor systems: Influence of scale, population density, climate, and methane recovery. Bioresour. Technol. 2018, 254, 56–66. [Google Scholar] [CrossRef] [PubMed]
  56. Stokes, J.; Horvath, A. Life cycle energy assessment of alternative water supply systems. Int. J. Life Cycle Assess 2006, 11, 335–343. [Google Scholar] [CrossRef]
  57. Lam, L.; Kurisu, K.; Hanaki, K. Comparative environmental impacts of source-separation systems for domestic wastewater management in rural China. J. Clean. Prod. 2015, 104, 185–198. [Google Scholar] [CrossRef]
  58. Muñoz, I.; Milà-I-Canals, L.; Fernández-Alba, A.R. Life Cycle Assessment of Water Supply Plans in Mediterranean Spain: The Ebro River Transfer Versus the AGUA Programme. J. Ind. Ecol. 2010, 14, 902–918. [Google Scholar] [CrossRef]
  59. Pintilie, L.; Torres, C.M.; Teodosiu, C.; Castells, F. Urban wastewater reclamation for industrial reuse: An LCA case study. J. Clean. Prod. 2016, 139, 1–14. [Google Scholar] [CrossRef]
  60. Rodriguez-Garcia, G.; Molinos-Senante, M.; Hospido, A.; Hernández-Sancho, F.; Moreira, M.T.; Feijoo, G. Environmental and economic profile of six typologies of wastewater treatment plants. Water Res. 2011, 45, 5997–6010. [Google Scholar] [CrossRef]
  61. Li, Y.; Xiong, W.; Zhang, W.; Wang, C.; Wang, P. Life cycle assessment of water supply alternatives in water-receiving areas of the South-to-North Water Diversion Project in China. Water Res. 2016, 89, 9–19. [Google Scholar] [CrossRef]
  62. Holloway, R.W.; Miller-Robbie, L.; Patel, M.; Stokes, J.R.; Munakata-Marr, J.; Dadakis, J.; Cath, T.Y. Life-cycle assessment of two potable water reuse technologies: MF/RO/UV-AOP treatment and hybrid osmotic membrane bioreactors. J. Memb. Sci. 2016, 507, 165–178. [Google Scholar] [CrossRef] [Green Version]
  63. Tangsubkul, N.; Beavis, P.; Moore, S.J.; Lundie, S.; Waite, T.D. Life cycle assessment of water recycling technology. Water Resour. Manag. 2005, 19, 521–537. [Google Scholar] [CrossRef]
  64. Opher, T.; Friedler, E. Comparative LCA of decentralized wastewater treatment alternatives for non-potable urban reuse. J. Environ. Manag. 2016, 182, 464–476. [Google Scholar] [CrossRef] [PubMed]
  65. Theregowda, R.; Vidic, R.; Dzombak, D.A.; Landis, A.E. Life cycle impact analysis of tertiary treatment alternatives to treat secondary municipal wastewater for reuse in cooling systems. Environ. Prog. Sustain. Energy 2015, 34, 178–187. [Google Scholar] [CrossRef]
  66. Shehabi, A.; Stokes, J.R.; Horvath, A. Energy and air emission implications of a decentralized wastewater system. Environ. Res. Lett. 2012, 7, 024007. [Google Scholar] [CrossRef]
  67. Tillman, A.-M.; Svingby, M.; Lundström, H. Life cycle assessment of municipal waste water systems. Int. J. Life Cycle Assess. 1998, 3, 145–157. [Google Scholar] [CrossRef]
  68. Thibodeau, C.; Monette, F.; Glaus, M. Comparison of development scenarios of a black water source-separation sanitation system using life cycle assessment and environmental life cycle costing. Resour. Conserv. Recycl. 2014, 92, 38–54. [Google Scholar] [CrossRef]
  69. Schoen, M.E.; Xue, X.; Wood, A.; Hawkins, T.R.; Garland, J.; Ashbolt, N.J. Cost, energy, global warming, eutrophication and local human health impacts of community water and sanitation service options. Water Res. 2017, 109, 186–195. [Google Scholar] [CrossRef]
  70. Hospido, A.; Moreira, M.T.; Martín, M.; Rigola, M.; Feijoo, G. Environmental Evaluation of Different Treatment Processes for Sludge from Urban Wastewater Treatments: Anaerobic Digestion versus Thermal Processes. Int. J. Life Cycle Assess. 2005, 10, 336–345. [Google Scholar] [CrossRef]
  71. Mills, N.; Pearce, P.; Farrow, J.; Thorpe, R.B.; Kirkby, N.F. Environmental & economic life cycle assessment of current & future sewage sludge to energy technologies. Waste Manag. 2014, 34, 185–195. [Google Scholar]
  72. Sadhukhan, J. Distributed and micro-generation from biogas and agricultural application of sewage sludge: Comparative environmental performance analysis using life cycle approaches. Appl. Energy 2014, 122, 196–206. [Google Scholar] [CrossRef] [Green Version]
  73. Tomei, M.C.; Bertanza, G.; Canato, M.; Heimersson, S.; Laera, G.; Svanström, M. Techno-economic and environmental assessment of upgrading alternatives for sludge stabilization in municipal wastewater treatment plants. J. Clean. Prod. 2016, 112, 3106–3115. [Google Scholar] [CrossRef]
  74. Alyaseri, I.; Zhou, J. Towards better environmental performance of wastewater sludge treatment using endpoint approach in LCA methodology. Heliyon 2017, 3, e00268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Li, H.; Jin, C.; Zhang, Z.; O’Hara, I.; Mundree, S. Environmental and economic life cycle assessment of energy recovery from sewage sludge through different anaerobic digestion pathways. Energy 2017, 126, 649–657. [Google Scholar] [CrossRef]
  76. Lombardi, L.; Nocita, C.; Bettazzi, E.; Fibbi, D.; Carnevale, E. Environmental comparison of alternative treatments for sewage sludge: An Italian case study. Waste Manag. 2017, 69, 365–376. [Google Scholar] [CrossRef] [PubMed]
  77. Li, H.; Feng, K. Life cycle assessment of the environmental impacts and energy efficiency of an integration of sludge anaerobic digestion and pyrolysis. J. Clean. Prod. 2018, 195, 476–485. [Google Scholar] [CrossRef]
  78. Wang, N.-Y.; Shih, C.-H.; Chiueh, P.-T.; Huang, Y.-F. Environmental Effects of Sewage Sludge Carbonization and Other Treatment Alternatives. Energies 2013, 6, 871–883. [Google Scholar] [CrossRef] [Green Version]
  79. Xu, C.; Chen, W.; Hong, J. Life-cycle environmental and economic assessment of sewage sludge treatment in China. J. Clean. Prod. 2014, 67, 79–87. [Google Scholar] [CrossRef]
  80. Miller-Robbie, L.; Ulrich, B.A.; Ramey, D.F.; Spencer, K.S.; Herzog, S.P.; Cath, T.Y.; Stokes, J.R.; Higgins, C.P. Life cycle energy and greenhouse gas assessment of the co-production of biosolids and biochar for land application. J. Clean. Prod. 2015, 91, 118–127. [Google Scholar] [CrossRef]
  81. Sills, D.L.; Wade, V.L.; DiStefano, T.D. Comparative Life Cycle and Technoeconomic Assessment for Energy Recovery from Dilute Wastewater. Environ. Eng. Sci. 2016, 33, 861–872. [Google Scholar] [CrossRef]
  82. Chai, C.; Zhang, D.; Yu, Y.; Wong, M.S. Carbon Footprint Analyses of Mainstream Wastewater Treatment Technologies under Different Sludge Treatment Scenarios in China. Water 2015, 7, 918–938. [Google Scholar] [CrossRef]
  83. Liu, Q.; Jiang, P.; Zhao, J.; Zhang, B.; Bian, H.; Qian, G. Life cycle assessment of an industrial symbiosis based on energy recovery from dried sludge and used oil. J. Clean. Prod. 2011, 19, 1700–1708. [Google Scholar] [CrossRef]
  84. Spriet, J.; McNabola, A. Decentralized drain water heat recovery from commercial kitchen in the hospitality sector. Energy Build. 2019, 194, 247–259. [Google Scholar] [CrossRef]
  85. Maurer, M.; Schwegler, P.; Larsen, T.A. Nutrients in urine: Energetic aspects of removal and recovery. Water Sci. Technol. 2003, 48, 37–46. [Google Scholar] [CrossRef] [PubMed]
  86. Tidåker, P.; Mattsson, B.; Jönsson, H. Environmental impact of wheat production using human urine and mineral fertilisers—A scenario study. J. Clean. Prod. 2007, 15, 52–62. [Google Scholar] [CrossRef]
  87. Remy, C.; Jekel, M. Sustainable wastewater management: Life cycle assessment of conventional and source-separating urban sanitation systems. Water Sci. Technol. 2008, 58, 1555–1562. [Google Scholar] [CrossRef]
  88. Sablayrolles, C.; Gabrielle, B.; Montrejaud-Vignoles, M. Life Cycle Assessment of Biosolids Land Application and Evaluation of the Factors Impacting Human Toxicity through Plant Uptake. J. Ind. Ecol. 2010, 14, 231–241. [Google Scholar] [CrossRef] [Green Version]
  89. Gilbert, P.; Thornley, P.; Riche, A.B. The influence of organic and inorganic fertiliser application rates on UK biomass crop sustainability. Biomass Bioenergy 2011, 35, 1170–1181. [Google Scholar] [CrossRef]
  90. Uggetti, E.; Ferrer, I.; Molist, J.; García, J. Technical, economic and environmental assessment of sludge treatment wetlands. Water Res. 2011, 45, 573–582. [Google Scholar] [CrossRef]
  91. Hospido, A.; Sanchez, I.; Rodriguez-Garcia, G.; Iglesias, A.; Buntner, D.; Reif, R.; Moreira, M.T.; Feijoo, G. Are all membrane reactors equal from an environmental point of view? Desalination 2012, 285, 263–270. [Google Scholar] [CrossRef]
  92. Bisinella de Faria, A.B.; Spérandio, M.; Ahmadi, A.; Tiruta-Barna, L. Evaluation of new alternatives in wastewater treatment plants based on dynamic modelling and life cycle assessment (DM-LCA). Water Res. 2015, 84, 99–111. [Google Scholar] [CrossRef] [PubMed]
  93. Bradford-Hartke, Z.; Lane, J.; Lant, P.; Leslie, G. Environmental Benefits and Burdens of Phosphorus Recovery from Municipal Wastewater. Environ. Sci. Technol. 2015, 49, 8611–8622. [Google Scholar] [CrossRef] [PubMed]
  94. Alvarez-Gaitan, J.P.; Short, M.D.; Lundie, S.; Stuetz, R. Towards a comprehensive greenhouse gas emissions inventory for biosolids. Water Res. 2016, 96, 299–307. [Google Scholar] [CrossRef] [PubMed]
  95. Fang, L.L.; Valverde-Pérez, B.; Damgaard, A.; Plósz, B.G.; Rygaard, M. Life cycle assessment as development and decision support tool for wastewater resource recovery technology. Water Res. 2016, 88, 538–549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Kulak, M.; Shah, N.; Sawant, N.; Unger, N.; King, H. Technology choices in scaling up sanitation can significantly affect greenhouse gas emissions and the fertiliser gap in India. J. Water Sanit. Hyg. Dev. 2017, 7, 466–476. [Google Scholar] [CrossRef]
  97. Mbaya, A.M.K.; Dai, J.; Chen, G.-H. Potential benefits and environmental life cycle assessment of equipping buildings in dense cities for struvite production from source-separated human urine. J. Clean. Prod. 2017, 143, 288–302. [Google Scholar] [CrossRef]
  98. Brown, S.; Beecher, N.; Carpenter, A. Calculator tool for determining greenhouse gas emissions for biosolids processing and end use. Environ. Sci. Technol. 2010, 44, 9509–9515. [Google Scholar] [CrossRef]
  99. Lundie, S.; Peters, G.M.; Beavis, P.C. Life cycle assessment for sustainable metropolitan water systems planning. Environ. Sci. Technol. 2004, 38, 3465–3473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Molinos-Senante, M.; Hernández-Sancho, F.; Sala-Garrido, R. Cost–benefit analysis of water-reuse projects for environmental purposes: A case study for Spanish wastewater treatment plants. J. Environ. Manag. 2011, 92, 3091–3097. [Google Scholar] [CrossRef]
  101. Herman, J.G.; Scruggs, C.E.; Thomson, B.M. The costs of direct and indirect potable water reuse in a medium-sized arid inland community. J. Water Process Eng. 2017, 19, 239–247. [Google Scholar] [CrossRef]
  102. Guo, T.; Englehardt, J.; Wu, T. Review of cost versus scale: Water and wastewater treatment and reuse processes. Water Sci. Technol. 2014, 69, 223–234. [Google Scholar] [CrossRef] [PubMed]
  103. Lam, C.M.; Leng, L.; Chen, P.C.; Lee, P.H.; Hsu, S.C. Eco-efficiency analysis of non-potable water systems in domestic buildings. Appl. Energy 2017, 202, 293–307. [Google Scholar] [CrossRef]
  104. Ip, K.; She, K.; Adeyeye, K. Life-cycle impacts of shower waste heat recovery: Case study of an installation at a university sport facility in the UK. Environ. Sci. Pollut. Res. 2018, 25, 19247–19258. [Google Scholar] [CrossRef] [Green Version]
  105. Bertrand, A.; Aggoune, R.; Maréchal, F. In-building waste water heat recovery: An urban-scale method for the characterisation of water streams and the assessment of energy savings and costs. Appl. Energy 2017, 192, 110–125. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The treatment and/or resource recovery processes considered for (a) water reclamation; (b) energy recovery; and (c) nutrient recovery.
Figure 1. The treatment and/or resource recovery processes considered for (a) water reclamation; (b) energy recovery; and (c) nutrient recovery.
Sustainability 15 03839 g001
Figure 2. Water quality of the influent and effluent relative to the treatment technologies: (a) biological oxygen demand (BOD), (b) chemical oxygen demand (COD), (c) total nitrogen (TN), and (d) total phosphorus (TP).
Figure 2. Water quality of the influent and effluent relative to the treatment technologies: (a) biological oxygen demand (BOD), (b) chemical oxygen demand (COD), (c) total nitrogen (TN), and (d) total phosphorus (TP).
Sustainability 15 03839 g002
Figure 3. GHG emissions per volume of reclaimed water with reference to (a) case study locations; (b) treatment stages; and (c) life cycle phases; (d) GHG emissions per MJ of energy recovered for the USA and UK scenarios; and GHG emissions for nutrient recovery case studies with reference to (e) the nutrient source and (f) reuse type. Abbreviations—3T: Tertiary Treatment; C: Collection; CONS: Construction Phase; CTD: Collection, Treatment, and Distribution; D: Distribution; DIS: Disposal Phase; N: Nitrogen; O&M: Operation and Maintance Phase; P: Phosphorus; T: Treatment; TD: Treatment and Distribution; UK: United Kingdom; USA: United States of America.
Figure 3. GHG emissions per volume of reclaimed water with reference to (a) case study locations; (b) treatment stages; and (c) life cycle phases; (d) GHG emissions per MJ of energy recovered for the USA and UK scenarios; and GHG emissions for nutrient recovery case studies with reference to (e) the nutrient source and (f) reuse type. Abbreviations—3T: Tertiary Treatment; C: Collection; CONS: Construction Phase; CTD: Collection, Treatment, and Distribution; D: Distribution; DIS: Disposal Phase; N: Nitrogen; O&M: Operation and Maintance Phase; P: Phosphorus; T: Treatment; TD: Treatment and Distribution; UK: United Kingdom; USA: United States of America.
Sustainability 15 03839 g003
Figure 4. The decision framework for wastewater-based resource recovery. It is recommended to reference Section 3.1 for Steps 2 and 3, Figure 5 for Step 4, Section 4.1 for Step 6, and Section 4.2 for Step 7. Abbreviations—GHG: Greenhouse Gas.
Figure 4. The decision framework for wastewater-based resource recovery. It is recommended to reference Section 3.1 for Steps 2 and 3, Figure 5 for Step 4, Section 4.1 for Step 6, and Section 4.2 for Step 7. Abbreviations—GHG: Greenhouse Gas.
Sustainability 15 03839 g004
Figure 5. The potential options for resource recovery for onsite and offsite (more centralized) applications. The resources denoted with an asterisk (*) can be diluted. Abbreviations—DPR: Direct Potable Reuse; Heat RS: Heat Recovery System; IPR: Indirect Potable Reuse; NPR: Non-Potable Reuse; USS: urine source separation; WW: Wastewater; WWSHP: wastewater source heat pump; WWT: Wastewater Treatment.
Figure 5. The potential options for resource recovery for onsite and offsite (more centralized) applications. The resources denoted with an asterisk (*) can be diluted. Abbreviations—DPR: Direct Potable Reuse; Heat RS: Heat Recovery System; IPR: Indirect Potable Reuse; NPR: Non-Potable Reuse; USS: urine source separation; WW: Wastewater; WWSHP: wastewater source heat pump; WWT: Wastewater Treatment.
Sustainability 15 03839 g005
Figure 6. Summary of the case study assessment for validation/applicability evaluation of the developed framework. Abbreviations—AD: anaerobic digestion; Bio: biological; GWP: global warming potential; LCA: life cycle assessment; LCC: life cycle costs; LR: land reclamation; M: membrane; MF: microfiltration; MGD: million gallons per day; NPR: non-potable reuse; P: phosphorous; PC: physical chemical SNPV: specific net present value; UV: ultraviolet.
Figure 6. Summary of the case study assessment for validation/applicability evaluation of the developed framework. Abbreviations—AD: anaerobic digestion; Bio: biological; GWP: global warming potential; LCA: life cycle assessment; LCC: life cycle costs; LR: land reclamation; M: membrane; MF: microfiltration; MGD: million gallons per day; NPR: non-potable reuse; P: phosphorous; PC: physical chemical SNPV: specific net present value; UV: ultraviolet.
Sustainability 15 03839 g006
Table 1. The Specific Net Present Value (SNPV) for resource recovery processes converted to USD 2019 per unit of resource recovered.
Table 1. The Specific Net Present Value (SNPV) for resource recovery processes converted to USD 2019 per unit of resource recovered.
Scope of RecoveryRecovery Type or Treatment ProcessMean SNPVStandard DeviationMaximum SNPVMinimum SNPVNo. of Samples
WaterSecondary TreatmentTertiary Treatment[USD/m3][USD/m3][USD/m3][USD/m3][-]
WastewaterBioN/A2.903.2912.230.0539
BioBio + PC0.190.050.220.133
BioPC0.620.511.470.1112
BioBio0.210.050.320.138
BioM2.684.2710.370.0812
BioM + PC7.230.837.786.283
EnergyRecovery or Treatment[USD/MJ][USD/MJ][USD/MJ][USD/MJ][-]
WastewaterHydropower0.400.200.750.1025
Heat exchanger0.0060.0030.0080.0042
Wastewater and sludgeAD + composting0.680.812.070.055
AD + landfilling24.4640.48124.00.0913
SludgeAD + composting0.090.190.43−0.3415
AD + incineration−0.008 a0.040.03−0.076
AD + landfilling0.0060.0030.0080.0042
AD + pyrolysis−0.085n/a−0.085−0.0851
V + composting0.20n/a0.200.201
V + incineration0.005n/a0.0050.0051
V + landfilling0.1260.0340.150.102
NutrientsTreatmentEnd Use[USD/kg P-eq][USD/kg P-eq][USD/kg P-eq][USD/kg P-eq][-]
UrineUSS + ChemFertilizer29.3746.82139.320.8114
USS + MFertigation12.4212.3824.401.644
SludgeADFertilizer24.8361.22195.35−21.4417
STWFertilizer6.470.797.205.664
STW + DewFertilizer6.422.5510.683.795
AD + Inc.LR0.180.030.210.163
Dry + Inc.LR−4.74N/A−4.74−4.741
a Negative values indicate income generated or greater resource recovery relative to consumption. Abbreviations—AD: anaerobic digestion; Bio: biological; Chem: chemical; Dew: dewatering; Dry: drying; LR: land reclamation; PC: physical and/or chemical; M: membrane; STW: sludge treatment wetland; USS: urine source separation; V: Volume reduction process.
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

Diaz-Elsayed, N.; Hua, J.; Rezaei, N.; Zhang, Q. A Decision Framework for Designing Sustainable Wastewater-Based Resource Recovery Schemes. Sustainability 2023, 15, 3839. https://doi.org/10.3390/su15043839

AMA Style

Diaz-Elsayed N, Hua J, Rezaei N, Zhang Q. A Decision Framework for Designing Sustainable Wastewater-Based Resource Recovery Schemes. Sustainability. 2023; 15(4):3839. https://doi.org/10.3390/su15043839

Chicago/Turabian Style

Diaz-Elsayed, Nancy, Jiayi Hua, Nader Rezaei, and Qiong Zhang. 2023. "A Decision Framework for Designing Sustainable Wastewater-Based Resource Recovery Schemes" Sustainability 15, no. 4: 3839. https://doi.org/10.3390/su15043839

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