Next Article in Journal
Legislation on Disability and Employment: To What Extent Are Employment Rights Guaranteed for Persons with Disabilities?
Next Article in Special Issue
Does Individuals’ Perception of Wastewater Pollution Decrease Their Self-Rated Health? Evidence from China
Previous Article in Journal
Occupational Mortality Matrix: A Tool for Epidemiological Assessment of Work-Related Risk Based on Current Data Sources
Previous Article in Special Issue
Highly Efficient Removal of Nitrate and Phosphate to Control Eutrophication by the Dielectrophoresis-Assisted Adsorption Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Partial Nitrification and Enhanced Biological Phosphorus Removal in a Sequencing Batch Reactor Treating High-Strength Wastewater

1
Xi’an Modern Chemistry Research Institute, Xi’an 710065, China
2
Department of Environmental Science and Engineering, School of Energy and Power Engineering, Xi’an Jiaotong University, Xi’an 710049, China
3
Department of Municipal and Environmental Engineering, Xi’an University of Technology, Xi’an 710048, China
4
State Key Laboratory of Eco-Hydraulics in Northwest Arid Region, Xi’an University of Technology, Xi’an 710048, China
5
Shaanxi Key Laboratory of Water Resources and Environment, Xi’an University of Technology, Xi’an 710048, China
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2022, 19(9), 5653; https://doi.org/10.3390/ijerph19095653
Submission received: 29 March 2022 / Revised: 30 April 2022 / Accepted: 2 May 2022 / Published: 6 May 2022
(This article belongs to the Special Issue Wastewater Pollution and Control)

Abstract

:
Complex and high levels of various pollutants in high-strength wastewaters hinder efficient and stable biological nutrient removal. In this study, the changes in pollutant removal performance and microbial community structure in a laboratory-scale anaerobic/aerobic sequencing batch reactor (SBR) treating simulated pre-fermented high-strength wastewater were investigated under different influent loading conditions. The results showed that when the influent chemical oxygen demand (COD), total nitrogen (TN), and orthophosphate (PO43−-P) concentrations in the SBR increased to 983, 56, and 20 mg/L, respectively, the COD removal efficiency was maintained above 85%, the TN removal efficiency was 64.5%, and the PO43−-P removal efficiency increased from 78.3% to 97.5%. Partial nitrification with simultaneous accumulation of ammonia (NH4+-N) and nitrite (NO2-N) was observed, which may be related to the effect of high influent load on ammonia- and nitrite-oxidising bacteria. The biological phosphorus removal activity was higher when propionate was used as the carbon source instead of acetate. The relative abundance of glycogen accumulating organisms (GAOs) increased significantly with the increase in organic load, while Tetrasphaera was the consistently dominant polyphosphate accumulating organism (PAO) in the reactor. Under high organic loading conditions, there was no significant PAO–GAO competition in the reactor, thus the phosphorus removal performance was not affected.

1. Introduction

High-strength wastewaters, mainly originating from livestock and poultry farming and food processing, have higher pollutant concentrations and ecological risks than conventional domestic wastewater [1]. The chemical oxygen demand (COD) concentration in high-strength wastewater can be as high as tens of thousands of milligrams per litre, total nitrogen (TN) levels can reach 800–23,000 mg/L, and total phosphorus (TP) levels can reach 50–230 mg/L [2,3,4,5]. If discharged directly into receiving water bodies without treatment, such large amounts of organic matter and nutrients in the wastewater will consume dissolved oxygen (DO) and cause eutrophication, promoting algal growth. This can lead to the death of aquatic organisms and the deterioration of the water environment [6]. Therefore, the treatment of high-strength wastewater is an urgent matter. Anaerobic digestion (AD) is one of the preferred methods, as it can economically and effectively treat highly concentrated organic wastewater, promoting carbon neutrality through energy recovery [7,8,9]. However, AD does not address the problem of excess nitrogen and phosphorus in high-strength wastewater. The removal of phosphorus from wastewater, in particular, is less studied, making the AD effluent still at risk of causing eutrophication in the receiving water body [4,10].
The enhanced biological phosphorus removal (EBPR) process has been widely used to remove phosphorus from domestic wastewater. This process relies on the enrichment of polyphosphate accumulating organisms (PAOs) in activated sludge [11]. Under anaerobic conditions, PAOs assimilate volatile fatty acids (VFAs), such as acetic acid and propionate acid, and break down stored polyphosphates and glycogen to generate energy and reducing power, whereas under aerobic conditions, they take up an excessive amount of phosphorus, thus achieving phosphorus removal from the wastewater [12]. In addition to PAOs, the main microorganisms in the EBPR system are glycogen accumulating organisms (GAOs), which behave similar to PAOs but do not have the ability to store polyphosphate [13]. Therefore, GAOs compete with PAOs for carbon sources during the anaerobic phase of the EBPR system but are unable to remove phosphate from the wastewater. Deterioration of the EBPR system is also attributed to an increase in the abundance of GAOs. Previous studies on EBPR systems have generally focused on optimally tuning the phosphate removal performance under relatively low organic loading conditions to give PAOs a competitive advantage over GAOs [14,15,16,17], while studies on medium-to-high-strength wastewaters containing high concentrations of organic matter (>400 mg/L) and phosphorus have been very limited [18].
EBPR process can be used to treat high-strength wastewater with different influent COD/P ratios (25:1 to 10:1), and the phosphorus removal efficiency could be maintained above 70% [19,20]. The phosphorus removal efficiency in EBPR systems treating wastewater containing high concentrations of phosphorus (30–280 mg P/L), such as dairy and manure wastewater, could be 60–90% [21,22]. However, these studies lacked a comprehensive analysis of the changes in the nitrogen and phosphorus removal performance, microbial activity and community structure in the systems. Previous studies showed that an excessively high influent COD/P ratio (>50:1) in the EBPR process treating low-strength wastewater promotes the proliferation of GAOs, which in turn affects the EBPR performance [12]. Randall and Chapin found that high influent carbon source concentrations (>740 mg COD/L) reduced the phosphorus removal stability and EBPR activity, and attributed to the fact that high organic loads favoured the growth of non-PAO and led to PAO being screened out of the system [23]. However, it is still unknown whether a carbon source competition between PAOs and GAOs occurs in high-strength wastewaters wherein various types of available carbon sources are sufficient. Therefore, it is necessary to evaluate the relationship between different functional microorganisms and the metabolic activity of different carbon sources in the EBPR process treating high-strength wastewater to ensure efficient and stable system performance.
In this study, a laboratory-scale anaerobic/aerobic sequencing batch reactor (A/O-SBR) was constructed to treat pre-fermented high-strength wastewater. The main objectives of our study were to (1) investigate the pollutant removal performance, microbial activity, and community structure in the reactor under different influent loading conditions; (2) evaluate the impact of organic load on nitrogen and phosphorus removal activity/populations, and (3) reveal whether the carbon source competition among different functionally relevant microorganisms occurs in high-strength wastewater. The outcome will provide support for the design and optimisation of the biological treatment of high-strength wastewater.

2. Materials and Methods

2.1. Reactor Setup and Operation

An SBR with a working volume of 4.62 L (Figure S1) was constructed in the laboratory. The seed sludge was taken from the No.4 wastewater treatment plant (WWTP) in Xi’an, Shaanxi Province, washed three times with tap water, and aerated before inoculating in the reactor. The reactor was operated under ambient temperature (20 ± 5 °C) without pH control. The variations in pH in the reactor (median: 7.4; range: 7.0–8.5) were possibly related to the protein hydrolysis during the experiment, yet the values are still in the suitable range for PAO activity [13,24], and the previously reported range for typical livestock wastewater (6.8–8.9) [25]. The cycle time was 8 h (anaerobic: 2.5 h; aerobic: 4.5 h; settling and draining: 1 h), and the sludge retention time (SRT) was ~14 d. In each cycle, 2.41 L of synthetic wastewater was pumped into the SBR, resulting in an exchange ratio of 0.52 and a hydraulic retention time (HRT) of 15.3 h.
The pollutant components and concentrations in the synthetic pre-fermented high-strength wastewater were defined based on the real wastewater from a local livestock farm in Xi’an, Shaanxi Province, which are also within the reported range of high-strength dairy and manure wastewater (Table S1) [26,27,28]. A mixture of complex (casein acid hydrolysate) and simple organic matter (sodium acetate and sodium propionate) in a COD ratio of 2:7:7 was used as the carbon source. Ammonium chloride and monopotassium phosphate were used as the inorganic nitrogen and phosphorus sources, respectively. During the start-up period (60 days), the influent COD (CODinf) concentration was gradually increased from ~200 to ~330 mg/L, while the influent ammonia (NH4+-Ninf) and orthophosphate (PO43−-Pinf) concentration was ~20 and ~8 mg/L, respectively. During Phases I, II, and III, the CODinf concentrations were ~400, ~700, and ~1000 mg/L; the NH4+-Ninf concentrations were ~20, ~35, and ~50 mg/L; and the PO43−-Pinf concentrations were ~8, ~15, and ~20 mg/L, respectively, resulting in a COD/N ratio of ~20:1 and a COD/P ratio of ~50:1 (Table 1). The concentrations of other trace elements (Table S2) [29] were kept constant during the experiment. To evaluate the pollutant removal performance during experiments, the influent and effluent concentrations of COD, TN, NH4+-N, nitrate (NO3-N), nitrite (NO2-N), TP, and PO43-P were regularly monitored.

2.2. Typical Cycle Study and Biological Phosphorus Removal Batch Tests

Samples were taken at different time points during one operating cycle of the reactor to analyse the changes in COD, TN, NH4+-N, NO3-N, NO2-N, and PO43−-P, in order to determine the pollutant removal kinetics and activities of the reactor during a typical cycle.
To assess the EBPR activity of the sludge, anaerobic/aerobic batch experiments were conducted as described previously [11]. Briefly, sludge samples at the end of the aerobic period in the reactor were taken and washed three times with a washing solution without carbon and phosphorus sources [29]. Allyl-N-thiourea (ATU) was added into the mixed liquor to inhibit nitrification, and the air was pumped in for 1 h of pre-aeration [30]. Then, after the residual organic matter had been consumed, nitrogen gas was pumped in to bring down the DO level to <0.1 mg/L to attain anaerobic conditions. Then, carbon (acetate or propionate) and phosphorus sources were added to the COD concentration of 100 mg/L and the PO43−-P concentration of 10 mg/L for 1 h of anaerobic condition. Thereafter, nitrogen gas was stopped and the air was pumped to obtain the aerobic condition for 2 h. During the test, pH was manually maintained at 7.0 ± 0.1 by the addition of NaOH or HCl. Temperature were maintained at 20 °C. Samples were periodically collected throughout the test and filtered through 0.45 μm filter membranes to determine COD and PO43−-P concentration. Sludge samples collected at the beginning and end of the test were collected to measure sludge concentration (i.e., mixed liquid suspended solids (MLSS) and mixed liquids volatile suspended solids (MLVSS)). The specific kinetic rate, such as anaerobic P release rate, substrate uptake rate, and aerobic P uptake rate, is expressed as the slope of the linear regression equation for the concentration–time plot (i.e., volumetric rate), dividing by MLVSS concentration. The P release–substrate uptake ratio is calculated as the mass of phosphorus released divided by the mass of substrate removed from the bulk solution.

2.3. Microbial Community Analysis

During the experiment, activated sludge samples were collected from the reactor at each phase for DNA extraction and 16S rRNA gene amplicon sequencing. Genomic DNA was extracted from each sample using the DNeasy PowerSoil Kit (QIAGEN, Inc., Hilden, Germany). Primers 338F (5′-ACTCCTACGGGAGGCAGCA-3′) and 806R (5′-GGACTACHVGGGTWTCTAAT-3′) were used for PCR amplification of the bacterial 16S rRNA gene. After the amplification, products were purified using Agencourt AMPure beads (Beckman Coulter, Indianapolis, IN, USA) and quantified using a PicoGreen dsDNA Assay Kit (Invitrogen, Carlsbad, CA, USA). Sequencing libraries were created and the Illumina MiSeq platform was used for 16S rRNA gene amplicon sequencing (Shanghai Personal Biotechnology Co., Ltd., Shanghai, China). Finally, bioinformatics analysis of the sequencing data was performed using QIIME (v1.8.0) software. The analysis of the relative abundance of the nutrient removal-related functional microorganisms was conducted by analyzing the sequencing data with the assistance of the Activated Sludge Microbial Database (MiDAS) [31] linking the taxonomy with multiple metabolic functions (e.g., nutrient removal, fermentation, etc.) [32].

2.4. Chemical Analyses

COD was determined by the potassium dichromate method. NH4+-N was determined using the Nessler reagent (Hach, Loveland, CO, USA). NO3-N and NO2-N were determined using ultraviolet spectrophotometry. PO43−-P was determined using ammonium molybdate reagent (Hach, Loveland, CO, USA). TN and TP were determined by the alkaline persulfate digestion method. Removal efficiency (%) is calculated as the difference between the pollutant (COD, NH4+-N, TN, and PO43−-P) concentration in the influent and effluent, divided by the pollutant concentration in the influent. The acetate or propionate concentration in the single carbon-feeding batch tests was determined via measuring COD in the supernatant and calculated based on the theoretical COD equivalents (i.e., 1.07 mg COD/g acetate and 1.52 mg COD/g propionate). MLSS and MLVSS were determined according to Standard Methods [33]. The modified thermal extraction method described by Domínguez et al. [34] was used to extract soluble microbial products (SMP) and extracellular polymeric substances (EPS) from the mixed liquor. The concentrations of proteins, polysaccharides, lipids, humic acids, and DNA in SMP; loosely bound EPS (LB-EPS); and tightly bound EPS (TB-EPS) were measured. A modified Lowry method [35,36] was employed to quantify proteins and humic acids. The anthrone method [37] was used to analyse polysaccharide concentration. The lipid concentration was determined using the sulfo-phospho-vanillin method [38], and DNA concentration was determined using the diphenylamine colorimetric method [39]. All samples were analysed in triplicate.

3. Results

3.1. Pollutant Removal Performance

3.1.1. COD Removal Performance

The variations in the COD removal performance in each phase during the experiment are shown in Figure 1a. In the initial start-up phase (0–16 days), fluctuations in the COD removal performance were observed. The effluent COD (CODeff) concentration was 97 ± 51 mg/L with an average removal efficiency of 47.3%. On days 17 and 41, the CODinf concentration increased from 193 ± 30 mg/L to 289 ± 20 and 333 ± 55 mg/L, respectively, and the removal performance gradually improved, which is possibly related to the consequently increased microbial growth and substrate degradation rates [40]. In Phase I, II, and III when the CODinf concentrations increased to 388 ± 25, 696 ± 27, and 983 ± 49 mg/L, respectively, the average removal efficiencies still reached 89.0%, 89.8%, and 94.5%, respectively. As the easily biodegradable matter, the VFAs (i.e., acetate and propionate) in the influent are expected to be fast degraded during the SBR cycle. While for the complex carbon source that is mainly composed of proteins and amino acids (i.e., casein acid hydrolysate) and commonly exists in dairy wastewater [41], its biodegradation would be slower and relies on the specific microorganisms capable of utilising amino acids [42]. Our results showed that the A/O-SBR was effective in degrading different organic matter under high influent loading conditions. This is possibly related to the increase in active biomass, as the sludge concentration increased from ~5 to ~7 g/L during the experiment.

3.1.2. Nitrogen Removal Performance

During the start-up phase, the average NH4+-N removal efficiency of the SBR was 93.1%, indicating that the reactor achieved good nitrogen removal performance (Figure 1b). NH4+-N removal was also consistently good in Phase I and II, with removal efficiencies of 97.2% and 98.2%, respectively. As the NH4+-Ninf concentration increased to 49.9 ± 1.3 mg/L in Phase III, the effluent NH4+-N (NH4+-Neff) concentration increased obviously to 5.1 ± 1.8 mg/L, with a declined average removal efficiency of 89.8%. This may be related to the excessive organic matters in influent that were not completely degraded during the anaerobic phase (Figure 2). The residual organic matter could promote the proliferation of other heterotrophic organisms (OHOs) during the aerobic phase, leading to intense competition between nitrifying and heterotrophic bacteria for oxygen and space [43]. The organic loading condition would also lead to inhibited nitrification due to the inactivation of enzymes in the nitrification process [44]. Meanwhile, the increased NH4+-N load in Phase III may exceed the removal capacity of nitrifying bacteria in the reactor, thus leading to an increase in the NH4+-Neff concentration. The denitrification performance of the SBR gradually increased under the three different loading conditions (Figure 1c), with the average removal efficiencies of 51.8%, 63.5%, and 64.5%, respectively, which should be attributed to the increased organic loads providing sufficient electron donors to denitrifiers [45].

3.1.3. Phosphorus Removal Performance

The biological phosphorus removal performance in SBR during the experiment is shown in Figure 1d. The average PO43−-P removal efficiency at the early stage of the start-up phase (days 1–16) was only 42.0%, and the effluent PO43−-P (PO43−-Peff) concentration was 4.7 ± 1.3 mg/L with large fluctuations. This may be related to the relatively long sludge age (~25 d). In the middle stage of the start-up phase (days 17–40), the SRT of the reactor was reduced to ~14 d by increasing the amount of waste sludge, which lead to ascending phosphorus removal performance. During the subsequent phases, as the influent loads elevate, the average PO43−-P removal efficiencies steadily increased to 77.4%, 92.3%, and 97.5% in Phase I, II, and III, respectively. The highest and most stable PO43−-P removal performance occurred in Phase III, with an average PO43−-Peff concentration of 0.56 mg/L (Figure S2), which is superior to another study treating dairy manure wastewater (PO43−-Pinf: 51.1 ± 23.0 mg/L; PO43−-P removal efficiency: 59%) [46]. Yuan et al. [18] also obtained similar effective performance in a lab-scale SBR treating synthetic wastewater (PO43−-Pinf: 40.0 mg/L; PO43−-P removal efficiency: 99.5 ± 0.8%), yet only used simple organic matter (i.e., acetate and propionate) as carbon sources. The effective PO43−-P removal in this study should be attributed to the high influent organic load, which reduced the competition for carbon sources between PAOs and other heterotrophic bacteria (e.g., denitrifying bacteria and OHOs), making more available carbon sources to PAOs.
In addition, the amounts of proteins, polysaccharides, and humic acids in the EPS of the activated sludge also largely increased with the increase in the CODinf load (Figure S3). Recent studies have demonstrated that EPS plays an important role in P removal by EBPR sludge, mainly due to its large specific surface area and abundant functional groups (e.g., hydroxyl, carboxyl, sulfonate, etc.) capable of adsorbing phosphorus [47,48]. Meanwhile, EPS was considered to have a positive effect on sludge flocculation, promoting cell aggregation during the sludge granulation process [49]. Therefore, the contribution of EPS to the EBPR process treating high-strength wastewater needs further investigation.
Considering the removal performance for each pollutant, the A/O-SBR used in this study could effectively treat different levels of pre-fermented high-strength wastewater. Although the NH4+-Neff concentration increased in Phase III, it should be possible to achieve improved nitrification performance via extending HRT and reducing residual organic matter in the aerobic phase. Further optimisation studies of the A/O-SBR are warranted to enhance its pollutant removal performance and expand its application range in treating wastewaters with varying influent loads.

3.2. Microbial Activities

3.2.1. Nitrogen Removal Activity

The effects of different influent loads on the pollutant treatment process during a typical cycle of SBR (anaerobic: 2.5 h; aerobic: 4.5 h) were analysed in different phases (Figure 2). The TN concentration decreased gradually throughout the typical cycle, and the effluent TN (TNeff) concentration decreased with an increase in the concentrations of CODinf and influent TN (TNinf), indicating that the sludge had efficient denitrification capacity under high-strength influent conditions. The NH4+-N concentration decreased during the aerobic period of the typical cycle, and the NO3-N concentration increased accordingly. Along with the gradually elevated NH4+-N supply in the influent, the average specific ammonia oxidation rate (AOR) increased from 0.51 mg N/(g VSS·h) in Phase I to 0.72 mg N/(g VSS·h) in Phase III (Table S1). Notably in Phase III, a higher NH4+-N concentration (5.4 mg/L) was detected at the end of the aerobic period. This incomplete oxidation of NH4+-N should be related to the high NH4+-Ninf load and the relatively low ammonia-oxidising bacteria (AOB) activity [50]. Regarding the specific nitrite oxidation rate (NOR), the average value decreased largely from 0.61 mg N/(g VSS·h) in Phase I to 0.18 mg N/(g VSS·h) in Phase III (Table S1), which is much lower than the AOR in Phase III (0.72 mg N/(g VSS·h)). Both AOR and NOR detected in this study are much lower than the values exhibited in other activated sludge systems (Table S3), which is probably due to the pretty low AOB and NOB abundance (See Section 4.1).
Meanwhile, substantially increased NO2-N concentrations were observed at the end of the aerobic period in Phase II (5.4 mg/L) and III (10.3 mg/L) (Figure 2), with the corresponding NO2-N accumulation rates (NAR) of 36% and 68%. This indicates that the high-strength influent conditions had a more pronounced inhibition effect on nitrite-oxidising bacteria (NOB) than on AOB, resulting in the NO2-N accumulation in the effluent.

3.2.2. Phosphorus Removal Activity

EBPR characteristics were observed during the typical cycle of SBR (Figure 2). The COD concentration in the SBR decreased sharply from 376.0–1020.8 mg/L to 55.4–95.1 mg/L within the first 60 min of the anaerobic period, indicating that most of the organic matter in the influent could be rapidly degraded under the anaerobic conditions. Correspondingly, the PO43−-P concentration largely increased to 49.7–100.0 mg/L and decreased obviously to 0.1–2.3 mg/L in the subsequent aerobic period. The anaerobic P release amount (PRA) were 40.4, 61.2, and 81.2 mg/L and the aerobic P uptake amount (PUA) values were 45.6, 71.0, and 99.9 mg/L for Phase I, II, and III, respectively, indicating that the increase in organic load promoted the P release and uptake capacities of the activated sludge in all phases.
To further investigate the effects of different carbon sources on EBPR activity, the batch tests fed with acetate or propionate were conducted using sludge samples from different phases (Figures S4 and S5). All the specific kinetic rates and stoichiometric ratios were within the range observed in other EBPR systems treating conventional low-strength municipal wastewater (Table 2). The anaerobic P release to acetate uptake (P/HAc) ratio, which is an indicator of PAO activity and abundance [51], decreased from 0.64 P-mol/C-mol in Phase I to 0.38 P-mol/C-mol in Phase III (Table 2). This indicates the presence of competition between PAOs and GAOs for carbon sources, potentially leading to decreased abundance of acetate-utilising PAOs (e.g., Accumulibacter) (as shown in Section 4.3). In contrast, the anaerobic P release to propionate uptake (P/HPr) ratio increased from 0.73 P-mol/C-mol in Phase I to 0.81 P-mol/C-mol in Phase III, and the P/HPr ratio in each phase was higher than the P/HAc ratio.

3.3. Microbial Community Structure

3.3.1. Microbial Diversity

The 16S rRNA gene amplicon sequencing data were analysed to obtain operational taxonomic units (OTUs) based on clustering at a similarity level of 0.97. Alpha diversity indices, including the Chao1, ACE, Shannon, and Gini–Simpson indices, were calculated for each activated sludge sample based on the OTUs (Table 3) [57]. The Good’s coverage index was higher than 0.99 for all 4 samples, indicating the current sequences represented the majority of the bacterial community. The sludge sample in Phase III had the highest diversity index, which is probably due to the increased influent load providing sufficient nutrients for the growth of microorganisms, as well as mitigating the competition between different microorganisms to a certain extent.

3.3.2. Microbial Community Composition

The relative abundance of microorganisms at the phylum levels in different phases during the experiment is shown in Figure 3a. In the raw sludge, Proteobacteria and Bacteroidetes were the dominant groups, with the relative abundance of 26.6% and 9.6%, respectively. After running under different loading conditions, the relative abundance of Proteobacteria and Bacteroidetes increased significantly, reaching 34.4%–54.1% and 28.1%–45.8%, respectively. Proteobacteria are dominant in many activated sludge systems, and include many microorganisms associated with organic matter degradation and nutrient cycling (e.g., some known denitrifying bacteria and PAOs) [58,59]. Lawson and Strachan [60] found that certain bacteria in Bacteroidetes also play an important role in denitrification. At the genus level, Thauera, Terrimonas, and Haliangium were the dominant genera in Phase I, while Saccharimonadaceae, Defluviicoccus, Flavobacterium, Competibacter, and Tetrasphaera were dominated in Phase II and III when the influent load reached higher levels (Figure 3b). This indicates that the microbial community structure changed largely and continuously from the raw sludge after reactor operation, which is probably related to the elevated organic and nutrient loads providing selection pressures to the community.

4. Discussion

4.1. Impact of Organic Load on Nitrogen-Removal-Related Microorganisms

The changes in functional microorganisms during the experiment (Figure 4) revealed that the relative abundance of AOB (i.e., Nitrosomonas) increased from 0.02% in Phase I to 0.09% in Phase III, whereas the relative abundance of NOB (i.e., Nitrospira) decreased substantially from 0.50% in Phase I to 0.05% in Phase III, which is consistent with the results obtained in the batch test (see Section 3.2.1). This distinct shift of the nitrifying bacterial population was possibly related to the increased influent load. For the typical WWTPs treating municipal wastewater, the COD level in the aerobic zone is often low since the influent organic matters were mainly degraded in the ahead anaerobic/anoxic zone. Therefore, less attention has been paid to the effect of organic loads on nitrifying bacteria. However, the organic matter concentration is significantly higher in high-strength wastewater, and their residues in the aerobic zone may impact nitrifying bacteria [44]. It has been found that the addition of small-molecule VFAs (i.e., formic, acetic, propionic, or butyric acid) in the aerobic phase inhibits NOB activity without affecting AOB activity, whereas the addition of valeric or capric acid negatively affects both AOB and NOB activities [61]. In this study, the gradual increase in acetate/propionate-dominated organic load might lead to higher suppression and out-selection of NOB populations compared to AOB.
The novel shortcut nitrification-denitrification (SND) and partial nitrification/anammox (PN/A) processes in the mainstream system have been receiving much attention due to the great savings in oxygen/energy consumption and carbon source utilisation [62,63]. However, studies on these processes have mainly focused on their application in wastewater treatment with low concentrations of carbon sources and/or low COD/N ratios, and very limited research has been conducted on the potential application of shortcut nitrification-based processes in high-strength wastewater treatment [64,65,66]. The simultaneous accumulation of NO2-N and NH4+-N in the A/O-SBR treating high-strength wastewater (Figure 2c), as observed in this study, provides a preliminary basis for the development of a mainstream SND- or PN/A-based process, which requires further in-depth and systematic studies.

4.2. Impact of Carbon Sources on Phosphorus Removal Activity

Previous studies have shown that acetate (49%–71% of total VFA) and propionate (24%–33% of total VFA) are the two most representative VFAs in domestic wastewater [67]. The propionate uptake rate of typical GAOs (i.e., Competibacter), which compete with typical PAOs (i.e., Accumulibacter) for carbon sources, was much lower than that of acetate [68], which explains the better EBPR activity in the reactor when propionate was used as the carbon source [69,70,71]. The differential PAO activities observed in the batch tests with different carbon sources (Table 2) suggest that the sludge may contain a large number of Competibacter GAOs that primarily utilise acetate (as shown in Figure 4), which to some extent affects the uptake of acetate by Accumulibacter PAOs and thus reduces the corresponding EBPR activity. However, compared to those in other studies [29,72,73], the P/HAc and P/HPr ratios in this study were consistently high, indicating that the presence of GAOs did not have a significant negative effect on the EBPR activity. In addition, the complex substrate (casein acid hydrolysate) in the influent was also available to other PAOs (e.g., Tetrasphaera) [74,75] and thus contributed to the overall EBPR activity.

4.3. Impact of Organic Load on Phosphorus-Removal-Related Microorganisms

The treatment performance of activated sludge systems depends heavily on the coordination among different functionally relevant microorganisms, and the microbial community composition and diversity are closely related to the system stability [76,77]. For EBPR-related functional microorganisms, the relative abundance of Competibacter GAOs with a higher acetate uptake rate increased significantly from 1.1% in Phase I to 24.3% in Phase III, whereas the relative abundance of Defluviicoccus GAOs with higher propionate uptake rate increased from 0.6% to 4.6%. Under low substrate concentration conditions, the presence of GAOs inhibited the phosphorus uptake activity of PAOs due to their competition for VFAs [78], while the GAO abundance would increase with the elevated substrate concentration in the influent [51,52]. Therefore, most EBPR studies have focused on how to limit the proliferation of GAOs through various pathways. The suitable COD/P ratio for EBPR activity was found to be 15:1–25:1 in low-strength wastewater [23,51], while higher COD/P ratios (e.g., 50:1) will be detrimental to the growth of PAOs but beneficial to GAOs [12]. However in this study, when the influent organic load was elevated to a sufficiently high level with a COD/P ratio of 50:1, EBPR activity was not affected (see Section 3.2.2) despite a significant increase in the GAO abundance. It suggests that there may not be significant substrate competition between PAOs and GAOs in the systems treating high-strength wastewater with various types of carbon sources [52,79]. Notably, the relative abundance of Accumulibacter PAOs decreased from 2.1% in Phase I to 0.6% in Phase III, whereas the relative abundance of Tetrasphaera PAOs increased from 2.0% to 3.3%. Tetrasphaera is a PAO that can utilise complex carbon sources (e.g., protein), whereas Competibacter is generally considered to utilise simple carbon sources only [79]. The different carbon source utilisation capabilities of the two groups would be beneficial in mitigating the PAO–GAO competition, and therefore high EBPR activity and phosphorus removal performance could be achieved in Phase III.
Another study proved that some species of Competibacter are denitrifying GAOs (DGAOs) [80]. The VFAs produced by Tetrasphaera when fermenting complex carbon sources would potentially provide additional substrate for Competibacter. The consequently enriched Competibacter DGAOs would promote nitrogen removal performance. Therefore, the coexistence of Tetrasphaera PAOs and Competibacter DGAOs, as observed in this study, may not only avoid competition for carbon sources but also synergistically promote nitrogen and phosphorus removal, as observed in a previous study [81]. Further investigation is warranted to determine how this synergy can be applied in a continuous flow reactor treating high-strength wastewater and coupled with the shortcut nitrification-based process discussed in Section 4.1.

5. Conclusions

(1) The anaerobic/aerobic SBR could effectively treat pre-fermented high-strength wastewater at different levels. The removal efficiencies of COD, TN, and PO43−-P were 94.5%, 64.5%, and 97.5%, respectively.
(2) The NOB activity and population were severely suppressed under high-strength influent loading conditions, achieving partial nitrification with simultaneous accumulation of NH4+-N and NO2-N in the effluent. Increased organic load promoted the anaerobic PRA and aerobic PUA. EBPR activity was higher when propionate was used as the carbon source.
(3) Sufficient organic load in the high-strength wastewater obviously mitigated the competition for substrate among PAOs and GAOs. The coexistence of Tetrasphaera and Competibacter DGAOs observed in the system would enable a synergistic effect on the simultaneous nitrogen and phosphorus removal.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijerph19095653/s1, Figure S1: Schematic diagram of the SBR system; Figure S2: Cumulative frequency curves of (a) COD, (b) NH4+-N, (c) TN and (d) PO43−-P concentrations in the effluent during the experiment; Figure S3: The concentrations of (a) proteins, (b) humic acids, and (c) polysaccharides in the soluble microbial products (SMP), loosely bound (LB-EPS) and tightly bound extracellular polymeric substances (TB-EPS) of the activated sludge during the experiment; Figure S4: Profiles of PO43−-P and COD during P release and uptake batch tests fed with acetate in (a) Phase I, (b) Phase II, and (c) Phase III; Figure S5: Profiles of PO43−-P and COD during P release and uptake batch tests fed with propionate in (a) Phase I, (b) Phase II, and (c) Phase III. Table S1: Summary of COD/P and COD/N ratios in high-strength dairy and manure wastewater [18,19,26,28,46,82]; Table S2: Component and concentrations of other trace elements in the synthetic pre-fermented high-strength wastewater; Table S3: Component and concentrations of other trace elements in the synthetic pre-fermented high-strength wastewater [83,84]; Table S4: Average specific ammonia oxidation rate (AOR) and specific nitrite oxidation rate (NOR) in typical cycles of SBR reactor during the experiment.

Author Contributions

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

Funding

This research was funded by the National Natural Science Foundation of China (No. 52070156 and 51409209), the Natural Science Basic Research Plan in Shaanxi Province of China (No. 2020JM-460), and the Scientific Research Program Funded by Shaanxi Provincial Education Department (No. 17JS097).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

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

References

  1. Pan, M.; Huang, X.; Wu, G.; Hu, Y.; Yang, Y.; Zhan, X.J.W. Performance of denitrifying phosphate removal via nitrite from slaughterhouse wastewater treatment at low temperature. Water 2017, 9, 818. [Google Scholar] [CrossRef] [Green Version]
  2. Chen, C.-Y.; Kuo, E.-W.; Nagarajan, D.; Ho, S.-H.; Dong, C.-D.; Lee, D.-J.; Chang, J.-S. Cultivating Chlorella sorokiniana AK-1 with swine wastewater for simultaneous wastewater treatment and algal biomass production. Bioresour. Technol. 2020, 302, 122814. [Google Scholar] [CrossRef] [PubMed]
  3. Li, X.; Wu, S.; Yang, C.; Zeng, G. Microalgal and duckweed based constructed wetlands for swine wastewater treatment: A review. Bioresour. Technol. 2020, 123858. [Google Scholar] [CrossRef]
  4. Cheng, H.-H.; Narindri, B.; Chu, H.; Whang, L.-M. Recent advancement on biological technologies and strategies for resource recovery from swine wastewater. Bioresour. Technol. 2020, 303, 122861. [Google Scholar] [CrossRef] [PubMed]
  5. Cheng, D.; Ngo, H.; Guo, W.; Chang, S.; Nguyen, D.; Kumar, S. Microalgae biomass from swine wastewater and its conversion to bioenergy. Bioresour. Technol. 2019, 275, 109–122. [Google Scholar] [CrossRef] [PubMed]
  6. Ndambi, O.A.; Pelster, D.E.; Owino, J.O.; De Buisonje, F.; Vellinga, T. Manure management practices and policies in sub-Saharan Africa: Implications on manure quality as a fertilizer. Front. Sustain. Food S. 2019, 3, 29. [Google Scholar] [CrossRef] [Green Version]
  7. Jang, H.M.; Ha, J.H.; Kim, M.-S.; Kim, J.-O.; Kim, Y.M.; Park, J.M. Effect of increased load of high-strength food wastewater in thermophilic and mesophilic anaerobic co-digestion of waste activated sludge on bacterial community structure. Water Res. 2016, 99, 140–148. [Google Scholar] [CrossRef]
  8. Jang, H.M.; Kim, J.H.; Ha, J.H.; Park, J.M. Bacterial and methanogenic archaeal communities during the single-stage anaerobic digestion of high-strength food wastewater. Bioresour. Technol. 2014, 165, 174–182. [Google Scholar] [CrossRef]
  9. Liew, Y.X.; Chan, Y.J.; Manickam, S.; Chong, M.F.; Chong, S.; Tiong, T.J.; Lim, J.W.; Pan, G.-T. Enzymatic pretreatment to enhance anaerobic bioconversion of high strength wastewater to biogas: A review. Sci. Total Environ. 2020, 713, 136373. [Google Scholar] [CrossRef]
  10. Torres-Franco, A.; Passos, F.; Figueredo, C.; Mota, C.; Muñoz, R. Current advances in microalgae-based treatment of high-strength wastewaters: Challenges and opportunities to enhance wastewater treatment performance. Rev. Environ. Sci. Bio. 2021, 20, 209–235. [Google Scholar] [CrossRef]
  11. Wang, D.; Tooker, N.B.; Srinivasan, V.; Li, G.; Fernandez, L.A.; Schauer, P.; Menniti, A.; Maher, C.; Bott, C.B.; Dombrowski, P. Side-stream enhanced biological phosphorus removal (S2EBPR) process improves system performance-A full-scale comparative study. Water Res. 2019, 167, 115109. [Google Scholar] [CrossRef] [PubMed]
  12. Oehmen, A.; Lemos, P.C.; Carvalho, G.; Yuan, Z.; Keller, J.; Blackall, L.L.; Reis, M.A. Advances in enhanced biological phosphorus removal: From micro to macro scale. Water Res. 2007, 41, 2271–2300. [Google Scholar] [CrossRef] [PubMed]
  13. Lopez-Vazquez, C.M.; Oehmen, A.; Hooijmans, C.M.; Brdjanovic, D.; Gijzen, H.J.; Yuan, Z.; van Loosdrecht, M.C. Modeling the PAO–GAO competition: Effects of carbon source, pH and temperature. Water Res. 2009, 43, 450–462. [Google Scholar] [CrossRef] [PubMed]
  14. Rubio-Rincón, F.; Lopez-Vazquez, C.; Welles, L.; Van Loosdrecht, M.; Brdjanovic, D. Cooperation between Candidatus Competibacter and Candidatus Accumulibacter clade I, in denitrification and phosphate removal processes. Water Res. 2017, 120, 156–164. [Google Scholar] [CrossRef]
  15. Rubio-Rincón, F.; Welles, L.; Lopez-Vazquez, C.; Nierychlo, M.; Abbas, B.; Geleijnse, M.; Nielsen, P.; van Loosdrecht, M.C.; Brdjanovic, D. Long-term effects of sulphide on the enhanced biological removal of phosphorus: The symbiotic role of Thiothrix caldifontis. Water Res. 2017, 116, 53–64. [Google Scholar] [CrossRef]
  16. Torresi, E.; Tang, K.; Deng, J.; Sund, C.; Smets, B.F.; Christensson, M.; Andersen, H.R. Removal of micropollutants during biological phosphorus removal: Impact of redox conditions in MBBR. Sci. Total Environ. 2019, 663, 496–506. [Google Scholar] [CrossRef]
  17. Yang, G.; Wang, D.; Yang, Q.; Zhao, J.; Liu, Y.; Wang, Q.; Zeng, G.; Li, X.; Li, H. Effect of acetate to glycerol ratio on enhanced biological phosphorus removal. Chemosphere 2018, 196, 78–86. [Google Scholar] [CrossRef]
  18. Yuan, Z.; Kang, D.; Li, G.; Lee, J.; Han, I.; Wang, D.; Zheng, P.; Reid, M.C.; Gu, A.Z. Combined Enhanced Biological Phosphorus Removal (EBPR) and Nitrite Accumulation for Treating High-strength Wastewater. bioRxiv 2021. [Google Scholar] [CrossRef]
  19. Bickers, P.O.; Bhamidimarri, R.; Shepherd, J.; Russell, J. Biological phosphorus removal from a phosphorus-rich dairy processing wastewater. Water Sci. Technol. 2003, 48, 43–51. [Google Scholar] [CrossRef]
  20. Broughton, A.; Pratt, S.; Shilton, A. Enhanced biological phosphorus removal for high-strength wastewater with a low rbCOD: P ratio. Bioresour. Technol. 2008, 99, 1236–1241. [Google Scholar] [CrossRef]
  21. Liu, Z.h.; Pruden, A.; Ogejo, J.A.; Knowlton, K.F. Polyphosphate-and Glycogen-Accumulating Organisms in One EBPR System for Liquid Dairy Manure. Water Environ. Res. 2014, 86, 663–671. [Google Scholar] [CrossRef] [PubMed]
  22. Jena, J.; Kumar, R.; Saifuddin, M.; Dixit, A.; Das, T. Anoxic–aerobic SBR system for nitrate, phosphate and COD removal from high-strength wastewater and diversity study of microbial communities. Biochem. Eng. J. 2016, 105, 80–89. [Google Scholar] [CrossRef]
  23. Randall, C.W.; Chapin, R.W. Acetic acid inhibition of biological phosphorus removal. Water Environ. Res. 1997, 69, 955–960. [Google Scholar] [CrossRef]
  24. Filipe, C.D.; Daigger, G.T.; Grady Jr, C.L. Effects of pH on the rates of aerobic metabolism of phosphate-accumulating and glycogen-accumulating organisms. Water Environ. Res. 2001, 73, 213–222. [Google Scholar] [CrossRef]
  25. Domingues, E.; Fernandes, E.; Gomes, J.; Martins, R.C. Advanced oxidation processes perspective regarding swine wastewater treatment. Sci. Total Environ. 2021, 776, 145958. [Google Scholar] [CrossRef]
  26. Sooknah, R.D.; Wilkie, A.C. Nutrient removal by floating aquatic macrophytes cultured in anaerobically digested flushed dairy manure wastewater. Ecol. Eng. 2004, 22, 27–42. [Google Scholar] [CrossRef]
  27. USEPA. NPDES Permit Writers’ Manual for Concentrated Animal Feeding Operations; USEPA: Washington DC, USA, 2012. Available online: https://www3.epa.gov/npdes/pubs/cafo_permitmanual_entire.pdf.
  28. Zheng, T.; Li, P.; Ma, X.; Sun, X.; Wu, C.; Wang, Q.; Gao, M. Pilot-scale experiments on multilevel contact oxidation treatment of poultry farm wastewater using saran lock carriers under different operation model. J. Environ. Sci. 2019, 77, 336–345. [Google Scholar] [CrossRef]
  29. Smolders, G.J.F.; Vandermeij, J.; Vanloosdrecht, M.C.M.; Heijnen, J.J. Model of the anaerobic metabolism of the biological phosphorus removal process - stoichiometry and pH influence. Biotechnol. Bioeng. 1994, 43, 461–470. [Google Scholar] [CrossRef]
  30. Oehmen, A.; Saunders, A.M.; Vives, M.T.; Yuan, Z.; Keller, J. Competition between polyphosphate and glycogen accumulating organisms in enhanced biological phosphorus removal systems with acetate and propionate as carbon sources. J. Biotechnol. 2006, 123, 22–32. [Google Scholar] [CrossRef]
  31. The Microbial Database for Activated Sludge (MiDAS). Available online: https://www.midasfieldguide.org/guide/search (accessed on 4 March 2022).
  32. Nierychlo, M.; Andersen, K.S.; Xu, Y.; Green, N.; Jiang, C.; Albertsen, M.; Dueholm, M.S.; Nielsen, P.H. knowledge platform for activated sludge and anaerobic digesters reveals species-level microbiome composition of activated sludge. Water Res. 2020, 182, 115955. [Google Scholar] [CrossRef]
  33. Ahpa, S.; Wef, S. Methods for the Examination of Water and Wastewater; American Public Health Association: Washington DC, USA, 2005. [Google Scholar]
  34. Domínguez, L.; Rodríguez, M.; Prats, D. Effect of different extraction methods on bound EPS from MBR sludges. Part I: Influence of extraction methods over three-dimensional EEM fluorescence spectroscopy fingerprint. Desalination 2010, 261, 19–26. [Google Scholar] [CrossRef]
  35. Griebe, T.; Nielsen, P. Enzymatic activity in the activated-sludge floc matrix. Appl. Microbiolol. Biot. 1995, 43, 755–761. [Google Scholar]
  36. Shen, Y.-x.; Xiao, K.; Liang, P.; Ma, Y.-w.; Huang, X. Improvement on the modified Lowry method against interference of divalent cations in soluble protein measurement. Appl. Microbiolol. Biot. 2013, 97, 4167–4178. [Google Scholar] [CrossRef] [PubMed]
  37. Zuriaga-Agustí, E.; Bes-Piá, A.; Mendoza-Roca, J.A.; Alonso-Molina, J.L. Influence of extraction methods on proteins and carbohydrates analysis from MBR activated sludge flocs in view of improving EPS determination. Sep. Purif. Technol. 2013, 112, 1–10. [Google Scholar] [CrossRef]
  38. Mishra, S.K.; Suh, W.I.; Farooq, W.; Moon, M.; Shrivastav, A.; Park, M.S.; Yang, J.-W. Rapid quantification of microalgal lipids in aqueous medium by a simple colorimetric method. Bioresour. Technol. 2014, 155, 330–333. [Google Scholar] [CrossRef] [PubMed]
  39. Rahban, M.; Divsalar, A.; Saboury, A.A.; Golestani, A. Nanotoxicity and spectroscopy studies of silver nanoparticle: Calf thymus DNA and K562 as targets. J. Phys. Chem. C 2010, 114, 5798–5803. [Google Scholar] [CrossRef]
  40. Fonte, E.S.; Amado, A.M.; Meirelles-Pereira, F.; Esteves, F.A.; Rosado, A.S.; Farjalla, V.F. The combination of different carbon sources enhances bacterial growth efficiency in aquatic ecosystems. Microb. Ecol. 2013, 66, 871–878. [Google Scholar] [CrossRef]
  41. McAteer, P.G.; Trego, A.C.; Thorn, C.; Mahony, T.; Abram, F.; O’Flaherty, V. Reactor configuration influences microbial community structure during high-rate, low-temperature anaerobic treatment of dairy wastewater. Bioresour. Technol. 2020, 307, 123221. [Google Scholar] [CrossRef]
  42. Vidal, G.; Carvalho, A.; Mendez, R.; Lema, J. Influence of the content in fats and proteins on the anaerobic biodegradability of dairy wastewaters. Bioresour. Technol. 2000, 74, 231–239. [Google Scholar] [CrossRef]
  43. Elenter, D.; Milferstedt, K.; Zhang, W.; Hausner, M.; Morgenroth, E. Influence of detachment on substrate removal and microbial ecology in a heterotrophic/autotrophic biofilm. Water Res. 2007, 41, 4657–4671. [Google Scholar] [CrossRef]
  44. Navada, S.; Knutsen, M.F.; Bakke, I.; Vadstein, O. Nitrifying biofilms deprived of organic carbon show higher functional resilience to increases in carbon supply. Scientific Reports 2020, 10, 7121. [Google Scholar] [CrossRef]
  45. Du, R.; Cao, S.; Li, B.; Zhang, H.; Li, X.; Zhang, Q.; Peng, Y. Step-feeding organic carbon enhances high-strength nitrate and ammonia removal via DEAMOX process. Chem. Eng. J. 2019, 360, 501–510. [Google Scholar] [CrossRef]
  46. Qureshi, A.; Lo, K.V.; Mavinic, D.S.; Liao, P.H.; Koch, F.; Kelly, H. Dairy manure treatment, digestion and nutrient recovery as a phosphate fertilizer. J. Environ. Sci. Health Part B 2006, 41, 1221–1235. [Google Scholar] [CrossRef]
  47. Ding, Y.; Dai, X.; Wu, B.; Liu, Z.; Dai, L. Targeted clean extraction of phosphorus from waste activated sludge: From a new perspective of phosphorus occurrence states to an innovative approach through acidic cation exchange resin. Water Res. 2022, 215, 118190. [Google Scholar] [CrossRef] [PubMed]
  48. Li, W.-W.; Zhang, H.-L.; Sheng, G.-P.; Yu, H.-Q. Roles of extracellular polymeric substances in enhanced biological phosphorus removal process. Water Res. 2015, 86, 85–95. [Google Scholar] [CrossRef]
  49. Wang, Y.; Wang, J.; Liu, Z.; Huang, X.; Fang, F.; Guo, J.; Yan, P. Effect of EPS and its forms of aerobic granular sludge on sludge aggregation performance during granulation process based on XDLVO theory. Sci. Total Environ. 2021, 795, 148682. [Google Scholar] [CrossRef]
  50. Seuntjens, D.; Han, M.; Kerckhof, F.-M.; Boon, N.; Al-Omari, A.; Takacs, I.; Meerburg, F.; De Mulder, C.; Wett, B.; Bott, C.; et al. Pinpointing wastewater and process parameters controlling the AOB to NOB activity ratio in sewage treatment plants. Water Res. 2018, 138, 37–46. [Google Scholar] [CrossRef]
  51. Schuler, A.J.; Jenkins, D. Enhanced Biological Phosphorus Removal from Wastewater by Biomass with Different Phosphorus Contents, Part I: Experimental Results and Comparison with Metabolic Models. Water Environ. Res. 2003, 75, 485–498. [Google Scholar] [CrossRef]
  52. Gu, A.Z.; Saunders, A.; Neethling, J.; Stensel, H.; Blackall, L. Functionally relevant microorganisms to enhanced biological phosphorus removal performance at full-scale wastewater treatment plants in the United States. Water Environ. Res. 2008, 80, 688–698. [Google Scholar] [CrossRef]
  53. Onnis-Hayden, A.; Srinivasan, V.; Tooker, N.B.; Li, G.; Wang, D.; Barnard, J.L.; Bott, C.; Dombrowski, P.; Schauer, P.; Menniti, A. Survey of full-scale sidestream enhanced biological phosphorus removal (S2EBPR) systems and comparison with conventional EBPRs in North America: Process stability, kinetics, and microbial populations. Water Environ. Res. 2020, 92, 403–417. [Google Scholar] [CrossRef]
  54. Majed, N.; Gu, A.Z. Phenotypic dynamics in polyphosphate and glycogen accumulating organisms in response to varying influent C/P ratios in EBPR systems. Sci. Total Environ. 2020, 743, 140603. [Google Scholar] [CrossRef]
  55. Pijuan, M.; Baeza, J.A.; Casas, C.; Lafuente, J. Response of an EBPR population developed in an SBR with propionate to different carbon sources. Water Sci. Technol. 2004, 50, 131–138. [Google Scholar] [CrossRef]
  56. Qiu, G.; Zuniga-Montanez, R.; Law, Y.; Thi, S.S.; Nguyen, T.Q.N.; Eganathan, K.; Liu, X.; Nielsen, P.H.; Williams, R.B.; Wuertz, S. Polyphosphate-accumulating organisms in full-scale tropical wastewater treatment plants use diverse carbon sources. Water Res. 2019, 149, 496–510. [Google Scholar] [CrossRef]
  57. Kim, B.-R.; Shin, J.; Guevarra, R.B.; Lee, J.H.; Kim, D.W.; Seol, K.-H.; Lee, J.-H.; Kim, H.B.; Isaacson, R.E. Deciphering diversity indices for a better understanding of microbial communities. J. Microbiol. Biotechn. 2017, 27, 2089–2093. [Google Scholar] [CrossRef] [Green Version]
  58. Nascimento, A.L.; Souza, A.J.; Andrade, P.A.M.; Andreote, F.D.; Coscione, A.R.; Oliveira, F.C.; Regitano, J.B. Sewage sludge microbial structures and relations to their sources, treatments, and chemical attributes. Front. Microbiol. 2018, 9, 1462. [Google Scholar] [CrossRef]
  59. Nielsen, P.H.; Mielczarek, A.T.; Kragelund, C.; Nielsen, J.L.; Saunders, A.M.; Kong, Y.; Hansen, A.A.; Vollertsen, J. A conceptual ecosystem model of microbial communities in enhanced biological phosphorus removal plants. Water Res. 2010, 44, 5070–5088. [Google Scholar] [CrossRef]
  60. Lawson, C.E.; Strachan, B.J.; Hanson, N.W.; Hahn, A.S.; Hall, E.R.; Rabinowitz, B.; Mavinic, D.S.; Ramey, W.D.; Hallam, S.J. Rare taxa have potential to make metabolic contributions in enhanced biological phosphorus removal ecosystems. Environ. Microbiol. 2015, 17, 4979–4993. [Google Scholar] [CrossRef]
  61. Eilersen, A.M.; Henze, M.; Kløft, L. Effect of volatile fatty acids and trimethylamine on nitrification in activated sludge. Water Res. 1994, 28, 1329–1336. [Google Scholar] [CrossRef]
  62. Cao, Y.; van Loosdrecht, M.C.M.; Daigger, G.T. Mainstream partial nitritation–anammox in municipal wastewater treatment: Status, bottlenecks, and further studies. Appl. Microbiolol. Biot. 2017, 101, 1365–1383. [Google Scholar] [CrossRef]
  63. Lackner, S.; Gilbert, E.M.; Vlaeminck, S.E.; Joss, A.; Horn, H.; van Loosdrecht, M.C.M. Full-scale partial nitritation/anammox experiences – An application survey. Water Res. 2014, 55, 292–303. [Google Scholar] [CrossRef] [PubMed]
  64. Chang, M.; Liang, B.; Zhang, K.; Wang, Y.; Jin, D.; Zhang, Q.; Hao, L.; Zhu, T. Simultaneous shortcut nitrification and denitrification in a hybrid membrane aerated biofilms reactor (H-MBfR) for nitrogen removal from low COD/N wastewater. Water Res. 2022, 211, 118027. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, X.; Wang, C.; Wu, P.; Xia, Y.; Chen, Y.; Liu, W.; Xu, L.; Faustin, F. A novel denitrifying phosphorus removal and partial nitrification, anammox (DPR-PNA) process for advanced nutrients removal from high-strength wastewater. Chemosphere 2021, 265, 129165. [Google Scholar] [CrossRef]
  66. Zhao, J.; Wang, X.; Li, X.; Jia, S.; Peng, Y. Combining partial nitrification and post endogenous denitrification in an EBPR system for deep-level nutrient removal from low carbon/nitrogen (C/N) domestic wastewater. Chemosphere 2018, 210, 19–28. [Google Scholar] [CrossRef]
  67. Lv, X.m.; Shao, M.f.; Li, C.l.; Li, J.; Xia, X.; Liu, D.y. Bacterial diversity and community structure of denitrifying phosphorus removal sludge in strict anaerobic/anoxic systems operated with different carbon sources. J. Chem. Technol. Biotechnol. 2014, 89, 1842–1849. [Google Scholar] [CrossRef]
  68. Shen, N.; Zhou, Y. Enhanced biological phosphorus removal with different carbon sources. Appl. Microbiolol. Biot. 2016, 100, 4735–4745. [Google Scholar] [CrossRef]
  69. Vargas, M.; Guisasola, A.; Artigues, A.; Casas, C.; Baeza, J. Comparison of a nitrite-based anaerobic–anoxic EBPR system with propionate or acetate as electron donors. Process Biochem. 2011, 46, 714–720. [Google Scholar] [CrossRef]
  70. Zeng, T.; Wang, D.; Li, X.; Ding, Y.; Liao, D.; Yang, Q.; Zeng, G. Comparison between acetate and propionate as carbon sources for phosphorus removal in the aerobic/extended-idle regime. Biochem. Eng. J. 2013, 70, 151–157. [Google Scholar] [CrossRef]
  71. Zhu, R.; Wu, M.; Zhu, H.; Wang, Y.; Yang, J. Enhanced phosphorus removal by a humus soil cooperated sequencing batch reactor using acetate as carbon source. Chem. Eng. J. 2011, 166, 687–692. [Google Scholar] [CrossRef]
  72. Lanham, A.B.; Oehmen, A.; Saunders, A.M.; Carvalho, G.; Nielsen, P.H.; Reis, M.A.M. Metabolic versatility in full-scale wastewater treatment plants performing enhanced biological phosphorus removal. Water Res. 2013, 47, 7032–7041. [Google Scholar] [CrossRef]
  73. Oehmen, A.; Yuan, Z.; Blackall, L.L.; Keller, J. Comparison of acetate and propionate uptake by polyphosphate accumulating organisms and glycogen accumulating organisms. Biotechnol. Bioeng. 2005, 91, 162–168. [Google Scholar] [CrossRef]
  74. Marques, R.; Santos, J.; Nguyen, H.; Carvalho, G.; Noronha, J.P.; Nielsen, P.H.; Reis, M.A.M.; Oehmen, A. Metabolism and ecological niche of Tetrasphaera and Ca. Accumulibacter in enhanced biological phosphorus removal. Water Res. 2017, 122, 159–171. [Google Scholar] [CrossRef]
  75. Wang, D.; Li, Y.; Cope, H.A.; Li, X.; He, P.; Liu, C.; Li, G.; Rahman, S.M.; Tooker, N.B.; Bott, C.B.; et al. Intracellular polyphosphate length characterization in polyphosphate accumulating microorganisms (PAOs): Implications in PAO phenotypic diversity and enhanced biological phosphorus removal performance. Water Res. 2021, 206, 117726. [Google Scholar] [CrossRef]
  76. Gu, Y.; Wei, Y.; Xiang, Q.; Zhao, K.; Yu, X.; Zhang, X.; Li, C.; Chen, Q.; Xiao, H.; Zhang, X. C: N ratio shaped both taxonomic and functional structure of microbial communities in livestock and poultry breeding wastewater treatment reactor. Sci. Total Environ. 2019, 651, 625–633. [Google Scholar] [CrossRef]
  77. Pelissari, C.; Guivernau, M.; Viñas, M.; de Souza, S.S.; García, J.; Sezerino, P.H.; Ávila, C. Unraveling the active microbial populations involved in nitrogen utilization in a vertical subsurface flow constructed wetland treating urban wastewater. Sci. Total Environ. 2017, 584, 642–650. [Google Scholar] [CrossRef] [Green Version]
  78. Carvalho, G.; Lemos, P.C.; Oehmen, A.; Reis, M.A. Denitrifying phosphorus removal: Linking the process performance with the microbial community structure. Water Res. 2007, 41, 4383–4396. [Google Scholar] [CrossRef]
  79. Nielsen, P.H.; McIlroy, S.J.; Albertsen, M.; Nierychlo, M. Re-evaluating the microbiology of the enhanced biological phosphorus removal process. Curr. Opin. Biotech. 2019, 57, 111–118. [Google Scholar] [CrossRef]
  80. McIlroy, S.J.; Albertsen, M.; Andresen, E.K.; Saunders, A.M.; Kristiansen, R.; Stokholm-Bjerregaard, M.; Nielsen, K.L.; Nielsen, P.H. ’Candidatus Competibacter’-lineage genomes retrieved from metagenomes reveal functional metabolic diversity. ISME J. 2014, 8, 613–624. [Google Scholar] [CrossRef] [Green Version]
  81. Marques, R.; Ribera-Guardia, A.; Santos, J.; Carvalho, G.; Reis, M.A.M.; Pijuan, M.; Oehmen, A. Denitrifying capabilities of Tetrasphaera and their contribution towards nitrous oxide production in enhanced biological phosphorus removal processes. Water Res. 2018, 137, 262–272. [Google Scholar] [CrossRef]
  82. Wang, L.; Li, Y.; Chen, P.; Min, M.; Chen, Y.; Zhu, J.; Ruan, R.R. Anaerobic digested dairy manure as a nutrient supplement for cultivation of oil-rich green microalgae Chlorella sp. Bioresour. Technol. 2010, 101, 2623–2628. [Google Scholar]
  83. Li, S.; Fei, X.; Chi, Y.; Jiao, X.; Wang, L. Integrated temperature and DO effect on the lab scale A2O process: performance, kinetics and microbial community. Int. Biodeterior. Biodegrad. 2018, 133, 170–179. [Google Scholar]
  84. Yao, Q.; Peng, D.-C. Nitrite oxidizing bacteria (NOB) dominating in nitrifying community in full-scale biological nutrient removal wastewater treatment plants. Amb Express 2017, 7, 25. [Google Scholar]
Figure 1. The removal performance of (a) COD, (b) NH4+-N, (c) TN, and (d) PO43−-P during the experiment.
Figure 1. The removal performance of (a) COD, (b) NH4+-N, (c) TN, and (d) PO43−-P during the experiment.
Ijerph 19 05653 g001
Figure 2. Profiles of COD, TN, NH4+-N, NO3-N, NO2-N, and PO43−-P concentrations in a typical cycle of SBR: (a) Phase I, (b) Phase II, and (c) Phase III.
Figure 2. Profiles of COD, TN, NH4+-N, NO3-N, NO2-N, and PO43−-P concentrations in a typical cycle of SBR: (a) Phase I, (b) Phase II, and (c) Phase III.
Ijerph 19 05653 g002
Figure 3. Relative abundances of microbial community composition at the (a) phylum and (b) genus levels during the experiment.
Figure 3. Relative abundances of microbial community composition at the (a) phylum and (b) genus levels during the experiment.
Ijerph 19 05653 g003
Figure 4. Relative abundance of known functionally relevant microorganisms for nitrogen and phosphorus removal during the experiment.
Figure 4. Relative abundance of known functionally relevant microorganisms for nitrogen and phosphorus removal during the experiment.
Ijerph 19 05653 g004
Table 1. Main components of synthetic pre-fermented high-strength wastewater.
Table 1. Main components of synthetic pre-fermented high-strength wastewater.
Influent ConcentrationStart-Up PhasePhase I
Days 61–142
Phase II
Days 143–183
Phase III
Days 184–224
Days 1–16Days 17–40Days 41–60
COD (mg/L)192 ± 27289 ± 20333 ± 55388 ± 25696 ± 27983 ± 49
TN (mg/L)23.1 ± 4.322.1 ± 2.024.1± 1.023.3 ± 3.143.5 ± 7.356.1 ± 10.4
NH4+-N (mg/L)21.4 ± 0.719.8 ± 2.319.3 ± 0.919.8 ± 1.135.4 ± 2.949.9 ± 1.3
PO43−-P (mg/L)8.2 ± 1.57.6 ± 2.38.3 ± 0.48.1 ± 0.613.9 ± 1.019.2 ± 1.5
COD/N ratio9.014.617.219.619.719.7
COD/P ratio23.438.040.147.950.151.2
Table 2. Specific kinetic rates and stoichiometric ratios observed in the P release and uptake batch tests fed with different carbon sources during the experiment.
Table 2. Specific kinetic rates and stoichiometric ratios observed in the P release and uptake batch tests fed with different carbon sources during the experiment.
Carbon Source P Release Rate [mg P/(g VSS·h)]Substrate Uptake Rate [mg C/(g VSS·h)]P Uptake Rate [mg P/(g VSS·h)]P Release/Substrate Uptake Ratio (P-mol/C-mol)Reference
AcetatePhase I10.36.24.10.64This study
Phase II4.76.41.20.30This study
Phase III7.07.33.20.38This study
Full-scale sludge5.6-31.916.1-42.52.4-9.70.29-0.75[52]
Full-scale sludge2.8-5.37.7-24.90.6-2.60.16-0.54[53]
Lab-scale sludge 4.4-50.67.7-32.79.8-23.80.22-0.60[54]
PropionatePhase I9.85.23.70.73This study
Phase II7.34.71.50.60This study
Phase III6.43.0 2.60.81This study
Lab-scale sludge13.636.718.60.27[55]
Full-scale sludge---0.38-0.60[56]
Table 3. Alpha diversity indices in activated sludge samples during the experiment.
Table 3. Alpha diversity indices in activated sludge samples during the experiment.
SamplesObserved SpeciesGood’s CoveragePielou’s EvennessChao1Gini–SimpsonShannon
Raw sludge27680.9900.79230240.9949.063
Phase I30350.9950.73630900.9848.515
Phase II30700.9960.69830850.9688.090
Phase III30650.9900.72332080.9868.378
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Feng, X.; Qian, Y.; Xi, P.; Cao, R.; Qin, L.; Zhang, S.; Chai, G.; Huang, M.; Li, K.; Xiao, Y.; et al. Partial Nitrification and Enhanced Biological Phosphorus Removal in a Sequencing Batch Reactor Treating High-Strength Wastewater. Int. J. Environ. Res. Public Health 2022, 19, 5653. https://doi.org/10.3390/ijerph19095653

AMA Style

Feng X, Qian Y, Xi P, Cao R, Qin L, Zhang S, Chai G, Huang M, Li K, Xiao Y, et al. Partial Nitrification and Enhanced Biological Phosphorus Removal in a Sequencing Batch Reactor Treating High-Strength Wastewater. International Journal of Environmental Research and Public Health. 2022; 19(9):5653. https://doi.org/10.3390/ijerph19095653

Chicago/Turabian Style

Feng, Xiaojun, Yishi Qian, Peng Xi, Rui Cao, Lu Qin, Shengwei Zhang, Guodong Chai, Mengbo Huang, Kailong Li, Yi Xiao, and et al. 2022. "Partial Nitrification and Enhanced Biological Phosphorus Removal in a Sequencing Batch Reactor Treating High-Strength Wastewater" International Journal of Environmental Research and Public Health 19, no. 9: 5653. https://doi.org/10.3390/ijerph19095653

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