Next Article in Journal
Estimating Climate Change’s Impacts on the Recharge of an Ungauged Tropical Aquifer (Togolese Coastal Sedimentary Basin)
Previous Article in Journal
Experimental Study of the Injectability of Infiltration Grouting in Surface Moraine of Pulang Copper Mine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improved Performance of Sulfur-Driven Autotrophic Denitrification Process by Regulating Sulfur-Based Electron Donors

1
Hubei Key Laboratory of Mineral Resources Processing and Environment, Wuhan University of Technology, Luoshi Road 122, Wuhan 430070, China
2
School of Resources and Environmental Engineering, Wuhan University of Technology, Luoshi Road 122, Wuhan 430070, China
*
Author to whom correspondence should be addressed.
Water 2024, 16(5), 730; https://doi.org/10.3390/w16050730
Submission received: 29 January 2024 / Revised: 21 February 2024 / Accepted: 22 February 2024 / Published: 29 February 2024

Abstract

:
Sulfur-driven autotrophic denitrification (SADN) has demonstrated efficacy in nitrate (NO3) removal from the aquatic environment. However, the insolubility of elemental sulfur (S0) (maximum 5 μg/L at 25 °C) limited the NO3 removal rate. In this study, we investigated the performance of a laboratory-scale S0-packed bed reactor (S0-PBR) under various volumetric NO3 loading rates. By filling with smaller S0 particles (0.5–1 mm) and introducing chemical sulfide (30–50 mg S2−-S/L), a high NO3 removal rate (1.44 kg NO3-N/(m3·d)) was achieved, which was substantially higher than previously reported values in SADN systems. The analysis of the average specific NO3 removal rates and the half-order kinetic constants jointly confirmed that the denitrification performance was significantly enhanced by decreasing the S0 particle sizes from 10–12 mm to 1–2 mm. The smaller S0 particles with a larger specific surface area improved the mass-transfer efficiency. Dosing chemical S2− (20 mg S2−-S/L) to trigger the abiotic polysulfuration process increased the specific NO3 removal rate from 0.366 to 0.557 g NO3-N/g VSS/h and decreased the portion of removed NO3-N in the form of nitrous oxide (N2O-N) from 1.6% to 0.7% compared to the S2−-free group.

Graphical Abstract

1. Introduction

The widespread use of nitrogenous fertilizers in agriculture and inadequate wastewater treatment has significantly increased nitrate (NO3) pollution in aquatic environments [1]. Zhang et al. (2021c) [2] measured NO3 data from 71 major rivers in 30 provinces in China, revealing that the NO3 concentration in approximately 8% of rivers exceeded the World Health Organization limit of 10 mg NO3-N/L [3]. The increasing NO3 loading to coastal zones has induced a severe algae boom, leading to the formation of a “dead zone” [4]. NO3 can be converted into nitrite (NO2) or nitrosoamines in the esophagus, which easily aroused methemoglobinemia, blue-baby syndrome, carcinoma, and mutation, thereby posing a severe threat to human life and health [5,6]. Traditional physical/chemical methods (e.g., reverse osmosis, ion exchange, and electrodialysis) for NO3 removal from wastewater have drawbacks such as high operational cost, low selectivity, and the generation of secondary brine wastes [7].
Alternatively, the biological denitrification process was considered an effective approach for removing NO3. During this process, NO3 was sequentially reduced to NO2, nitric oxide (NO), nitrous oxide (N2O), and di-nitrogen (N2) [8]. Organic matters were the most commonly used electron donors to perform heterotrophic denitrification (HD), while hydrogen (H2), elemental sulfur (S0), and iron compounds were utilized by autotrophic denitrifiers [7]. In practice, both insufficient and excessive supplements of organic matter in the HD process would result in poor performance of NO3 removal and organic residue in the effluent, respectively [9]. Organic supplementation increased the cost and caused biofouling due to the high production of biomass sludge [10].
The autotrophic denitrification process can potentially replace HD because of negligible residual organics in an effluent, given that inorganic matters are utilized as electron donors [10]. Autotrophic denitrifiers exhibited lower biomass production due to the lower biomass yields of 0.4–0.57 g VSS/g NO3-N [11] than 0.8–1.2 g VSS/g NO3-N for heterotrophic denitrifiers [12]. The sophisticated hydrogen-delivering systems involved high operating and maintenance costs, which hindered the application of hydrogen-driven autotrophic denitrification [13]. Recently, sulfur-driven autotrophic denitrification (SADN) with the stoichiometry shown as follows [14] (Equation (1)) has gained increasing attention because S0 was non-toxic, readily available, and chemically stable under normal conditions and could be used “on demand” without overdosing concerns [15]. The yields (Y) of SADN were relatively low, 0.24 g COD/COD [16], resulting in substantial sludge reduction. SADN was more economical than HD, with estimated costs of $0.45/per kg·N removed versus $1.05/per kg·N removed [17]. In addition, N2O, as an intermediate product during the biological denitrification process, is a potent greenhouse gas with approximately 300 times the global warming potential of carbon dioxide (CO2) [18]. Less N2O is produced in the SADN process [19].
1.1S0 + NO3 + 0.76H2O + 0.4CO2 + 0.08NH4+ → 0.08C5H7O2N + 1.1SO42− + 0.5N2 + 1.28H+
The orthorhombic α-S80, as the only steady form of S0 under ambient conditions, was hardly soluble in water (5 μg/L, 25 °C) due to the high bond strength between S-S bonds in S80-rings and its large molecular size [20]. Owing to this problem, the bioavailability of S0 is too poor for sulfur-respiring bacteria, such as S0-oxidizing bacteria (S0OB) and S80-reducing bacteria (S0RB). Preliminary microbial hydrolysis of S0 was required as S0 was only taken up by sulfur-respiring bacteria after its solubilization [16,21,22]. The low solubility resulted in lower kinetics than the conventional HD or sulfate (SO42−) reduction process, which could be seen as the main bottleneck preventing the S0-driven bioprocess from realistic applications. Some studies demonstrated a positive relationship between the denitrification rate and factors affecting the surface area of S0, including S0 concentration [10], particle morphology [23], and size [24,25]. Additionally, the bioavailability of biogenic sulfur (Sbio0) particles is superior to chemical sulfur (Schem0) due to its micro-crystallinity structure and higher specific area [26,27].
The nucleophilic attack between sulfide (S2−) and S0 under neutral or alkaline conditions results in the cleavage of S80 rings and the formation of polysulfide (Sn2−) as detailed in Equation (2) [20]. This chemical reaction has received much attention due to the higher solubility and bioavailability of Sn2− [20,28,29]. Therefore, the polysulfide-involved SADN (PiSADN) process (Equation (3)) might be an effective method for realizing high-rate NO3 removal. However, the main challenge is how to continuously generate Sn2− in situ. Although promoting the sulfidogenic bacteria activity for S0/SO42− reduction to trigger the polysulfuration process was an option [30], organic supplementation might lead to a failure of the SADN system because the faster growth rate of heterotrophic NO3-reducing bacteria than autotrophic NO3-reducing bacteria [31]. Interestingly, a recent study by Qiu et al. (2022) [32] proposed a novel PiSADN process for S0-packed bed reactor (S0-PBR), and the polysulfuration was induced by an autotrophic biological sulfur disproportionation (SD) process (Equation (4)). It was difficult to continuously obtain the precursor biogenic S2− through the SD process because the reaction was thermodynamically unfavored under standard conditions [33]. The novel PiSADN process was only adaptable for low-strength wastewater treatment and required sufficient alkalinity supplementation and complex internal recirculation devices.
HS + n 1 8 S 8 0     S n 2 + H +
Sn2− + 6NO3 + 2H2O → Sn-52− + 5SO42− + 3N2 + 4H+
4S0 + 4H2O → SO42− + 3HS + 5H+
As stated above, soluble Sn2− remarkably enhances the bioavailability of S0 and thus facilitates the NO3 removal performance in the SADN system. However, there are many challenges in the generation pathways for precursor biogenic S2−. Dosing organic matter in the SADN system poses a risk to the stability of the microbial community. The SD process is endergonic and is easily inhibited by the presence of high NO3 loading. These concerns have hindered the development of high-rate in situ PiSADN applications. As such, we investigated the feasibility of establishing an in situ PiSADN system by adding chemical S2− directly for the treatment of high-loading wastewater. Moreover, although the literature found that S0 particle size was a key factor affecting the denitrification rate, the underlying kinetic mechanisms were not fully understood. Therefore, a laboratory-scale S0-PBR was continuously operated for 163 days under different volumetric loading rates of NO3. A laboratory-scale S0-PBR was continuously operated for 163 days under different volumetric loading rates of NO3. The specific aims of this study were to (a) demonstrate the feasibility of achieving a high NO3 removal rate in the long-term S0-PBR by introducing smaller S0 particles and chemical S2−; (b) investigate the effect of different S0 particle sizes and S2− initial concentrations on NO3 removal, NO2 accumulation, and N2O production; and (c) analyze the underlying mechanisms of optimized NO3 removal in the bioreactor. This work might facilitate a better understanding of how to achieve an efficient SADN process in the S0-PBR.

2. Materials and Methods

2.1. Bioreactor Setup and Operation

A laboratory-scale plexiglass S0-PBR (dimension: 8 cm diameter × 40 cm height) was operated under anaerobic conditions with an effective volume of 1.55 L. The outlet was set at 36 cm from the base. The S0-PBR was covered with aluminum foil to prevent the growth of phototrophic bacteria during the entire operation period. The sample port with a 0.8 cm diameter was set at a height of 30 cm, while the bottom of the S0-PBR was provided with an outlet port (2 cm diameter). Two peristaltic pumps (KCM-B146, Kamoer, Shanghai, China) were used in the S0-PBR operation, i.e., one for feeding and the other for suction.
The S0-PBR was filled with Schem0 (0.5–1 mm) and activated carbon (0.5–1 mm) particles with a volume ratio of around 2/3 and 1/3. The inoculation sludge was obtained from the aeration tank of a municipal wastewater treatment plant (Tangxun Lake wastewater treatment plant, Wuhan, China), and the total inoculum mass was approximately 5.4 g. The bioreactor was operated continuously in an up-flow mode at 25 ± 2 °C in a temperature-controlled room.
The S0-PBR was fed with synthetic wastewater, as per Qiu et al. (2020) [30]. A step-wise increase in influent volumetric loading, 0.06 kg NO3-N/(m3·d) to 1.92 kg NO3-N/(m3·d), was achieved in Stage I (days 1–127) by increasing the NO3 concentration of 20 to 400 mg NO3-N and decreasing the hydraulic retention time (HRT). The influent flow rate was increased from 3.23 mL/min to 5.17 mL/min by adjusting the feeding pump, and accordingly, the HRT decreased from 8 h to 5 h. In Stage II (days 128–151), the chemical S2− solution was provided into the S0-PBR while maintaining the same influent NO3 loading rate as the latter Stage I (days 100–127). In Stage III (days 156–163), the working conditions of the S0-PBR were identical to the latter Stage I while ceasing the supplement of chemical S2−. Sufficient NaHCO3 was added to the synthetic wastewater, acting as an alkalinity source and inorganic carbon for S0OB growth. The details of the three operational conditions are presented in Table 1.

2.2. Batch Experiments

To investigate the appropriate S0 particle size, Test I was performed in 100 mL serum bottles placed in a chamber (20 °C, 200 rpm), including four groups with different S0 sizes, i.e., 10–12 mm, 7–9 mm, 3–5 mm, and 1–2 mm, respectively. All bottles were sealed with butyl rubber stoppers and purged with N2 to obtain anaerobic conditions. The sludge was taken from the S0-PBR, and the concentration in each bottle was 0.445 g MLVSS/L. A total of 50 mg NO3-N/L and 1 g S0 particles with the above-mentioned different sizes were added. The purpose of providing excessive S0 was to avoid the impact of S0 limitation on the denitrification process. In addition, 2 g/L NaHCO3 was provided to balance the pH and support the bacterial growth. The trace elements were the same as the S0-PBR feed solution. The batch test was conducted in duplicate for 34 h, during which samples were taken at 0 h, 3 h, 6 h, 9.5 h, 12 h, 21.5 h, and 34 h to measure NO3 and NO2.
As mentioned above, insoluble S0 would be converted into soluble Sn2− in the presence of S2− at alkaline conditions. Based on this point, Test II was launched to investigate whether Sn2− could promote the SADN process. Two sets of experiments with the presence of 0 and 20 mg S/L chemical S2− were performed in different serum bottles containing 1 g S0 particles (1–2 mm) and 0.473 g/L MLVSS. Controls lacking S2− to monitor the conventional SADN process with S0 as the only electron donor. This test lasted for 32 h, during which samples were collected at 0 h, 6 h, 9.5 h, 22 h, 27.5 h, and 32 h to measure NO3, NO2, N2O, and S2− concentrations. Other operational conditions were the same as those mentioned above.

2.3. Chemical Analysis

The NO3, NO2, S2− and SO42− in the water samples were measured after filtering using disposable Millipore filter units (pore size: 0.22 μm). NO3 and NO2 were measured using an ultraviolet–visible spectrophotometer (UV5500PC, Shanghai Metash Instruments Co., Ltd., Shanghai, China). Dissolved N2O in water samples was analyzed using a gas chromatograph (7890 plus GC, Lunan Ruihong Chemical Instrument, Tengzhou, Shandong, China) fitted with an HP-Plot Molesieve column (30 m × 0.53 mm × 25 μm) and an electron capture detector (ECD). SO42− was quantified using an ion chromatograph (883 Basic IC plus, Metrohm, Switzerland) with a conductivity detector. Total dissolved sulfide (H2S, HS, and S2−) was determined using the methylene blue method [34]. The concentration of Sn2− was indicated by the dissolved zero-valent sulfur atoms in polysulfide ions, which was measured by the above UV at a wavelength of 285 nm after filtration [35,36]. pH and temperature were measured with portable meters (Multi-Parameter Meter, HQ40D, Hach, Loveland, CO, USA). MLSS and MLVSS in the sludge used in batch tests were determined according to APHA (2005) [34].

3. Results and Discussion

3.1. Optimization of the S0-PBR Performance

To enhance the S0-PBR performance, a long-term continuous-flow experiment focused on reducing the size of Schem0 particles and facilitating the formation of Sn2−. Three operational conditions were applied in the bioreactor. In Stage I, the Schem0 particles (0.5–1 mm) were used as the main filler. The volumetric loading rate of the reactor was step-wise increased to investigate the feasibility of enhancing NO3 removal capability by reducing Schem0 particle size in S0-PBR. In Stage II, on the basis of optimum Schem0 particle size, the chemical precursor S2− (30–50 mg S2−-S/L) was added to form Sn2− to accelerate the SADN process. In Stage III, the external S2− addition was completely eliminated so that the polysulfuration process was inhibited, highlighting the positive effect of Sn2− as an electron donor on the SADN process.
In Stage I (day 1–127), the influent NO3 concentration increased from 20 to 400 mg NO3-N/L, and HRT decreased from 8 h to 5 h, resulting in a step-wise increase in volumetric loading rate from 0.06 kg NO3-N/(m3·d) to 1.92 kg NO3-N/(m3·d). Of note, even with an influent NO3 loading as low as 0.06 kg NO3-N/(m3·d) in the early phase of Stage I (day 1–40), the average NO3 removal efficiency was only 89.3% (Figure 1a,b). The main reason might be attributed to the low abundance of S0OB in the inoculation sludge and its slow growth rate, which led to a start-up period of S0-PBR as long as 28 days [37]. Yang et al. (2016a) [38] mentioned that the volumetric denitrification loading rate was less than 0.1 kg NO3-N/(m3·d) when the MLVSS concentration remained below 0.3 g/L in the anoxic filter. After the adaption period, NO3 was occasionally detected in the S0-PBR effluent during days 40–99 (Figure 1a). The accumulation of functional biomass could explain this result due to the relatively strong biomass retention capacity of packed-bed reactors [39]. The S0OB could become the dominant microbial community in the S0-PBR since the absence of organic supplementation.
Given the limited electron-scavenging capability from the solid S0 interface by S0OB [40,41], the NO3 reduction rate was significantly higher than the NO2 reduction rate, leading to NO2 accumulation in the SADN system. However, NO2 was undetectable during days 1–99 (Figure 1a), even during days 95–99, corresponding with a high volumetric loading rate of 1.44 kg NO3-N/(m3·d) and high average NO3 removal efficiency of approximately 100% (Figure 1a,b). The efficient denitrification performance of the S0-PBR could be attributed to the use of smaller S0 particles (0.5–1 mm) with larger specific surface areas in the S0-PBR than those in the literature (2–16 mm) [10,32]. The higher surface areas of S0 improved the mass-transfer efficiency during S0OB utilization of S0 [10]. Consequently, the dissolution process of Schem0, considered the main rate-limiting step, was largely promoted [10,19,42].
However, during days 100–127, the average effluent NO3 concentration increased to 78 mg NO3-N/L, and the NO3 removal efficiency continuously declined to 62.7% on day 127 (Figure 1a,b). In addition, overloading of the S0-PBR was evidenced by the detection of NO2 and high-level average N2O concentration of 32 mg N2O-N/L in the effluent (Figure 1a,c). These could be seen as reliable markers of the reactor overloading in the SADN process [10,43,44]. Therefore, the maximum NO3 removal loading rate of the S0-PBR was considered as 1.44 kg NO3-N/(m3·d) during days 95–99, which was 1.88 times higher than the result obtained by Koenig et al. (2001) [24] who used bigger size of S0 particles, 2.8–5.6 mm.
Of note, the practical effluent SO42− concentration during days 1–40 and 43–60 averaged 447 mg SO42−/L and 542 mg SO42−/L (Figure 1d), respectively, which were significantly higher than the theoretical values that were 155 mg SO42−/L and 364 mg SO42−/L. It is well known that the sulfur-based autotrophic disproportionation process occurs only after NO3 is depleted in the SADN system [45,46,47], especially near the effluent side of S0-PBR [32]. Hijnen et al. (1992) [48] pointed out that the volumetric loading rate of S0-PBR should be kept above the minimum limitation, 0.22 kg NO3-N/(m3·d), to prevent the head loss caused by the SD process. It was reasonable to infer that the occurrence of the SD process at these two early periods of Stage I, due to the relatively low volumetric loading rates, 0.06–0.15 kg NO3-N/(m3·d), and the nearly complete NO3 removal in the S0-PBR (Figure 1a,b). According to Equation (4), SO42− in excess amount of theoretical production in these two periods was likely to come from the SD process. In addition, the ratios of SO42− production to NO3 removal were becoming closer to the theoretical value as the volumetric loading rate increased (Figure 1d). As NO3 loading rates were in the range of 0.76–1.92 kg NO3-N/(m3·d) during days 80–127 (Figure 1b), this ratio was almost equivalent to the theoretical value of 7.54 mg SO42−/mg NO3-N, suggesting the inhibition of high NO3 loading rate on SD process. A previous study also reported that the SD process was completely inhibited when influent NO3 loading exceeded 0.72 kg NO3-N/(m3·d) and the concentration of the sulfur-heterologous electron acceptors (e.g., NO3, NO2, and dissolved oxygen) increased to 1.1 mg/L [49].
In Stage II, 30 to 50 mg S/L chemical S2− was added into the S0-PBR in sequence while keeping the volumetric loading rate constant at about 1.92 kg NO3-N/(m3·d), same as that during days 100–127 in Stage I (Figure 1b). As a result of the addition of 30 mg S2−-S/L, the downward trend of NO3 removal efficiency was terminated and replaced by an upward trend during days 128–142, showing that the average NO3 removal efficiency was increased to 85.3% from 81.3% (Figure 1b). In addition, upon the overloading of influent NO3, the average N2O concentration of 2 mg N2O-N/L in effluent samples was far lower than that during days 100–127 (32 mg N2O-N/L) without chemical S2− addition, decreasing N2O accumulation by 93.8%. The result differs from previous points that the S2− could precipitate with soluble copper cofactors in the N2O reductase, leading to a rise in N2O production [50]. However, it was also observed in a previous study by Yang et al. (2016) [51] that the bio-poisoning chemical S2− could be instantly oxidized into Sn2− by membrane-bound sulfide-quinine reductase presented in almost S0OB, and Sn2− acting as a bioavailable electron donor could contribute to N2O reduction.
However, the average NO2 concentration increased from 6 mg NO2-N/L to 15 mg NO2-N/L during days 145–151 in Stage II, when the dosage increased to 50 mg S2−-S/L. The bioavailability of insoluble Schem0 in this S0-PBR could be greatly improved by adding a higher S2− concentration to promote the chemical polysulfuration process. The formation of Sn2− and the higher competitive capacity of the nitrate reductase for electrons than nitrite reductase explained well the severe NO2 accumulation [52,53]. Moreover, it has been reported that high-level chemical S2− exerted an inhibitory effect on nitrite reductase activity and ceased the NO2 reduction process [54,55,56]. As a result of the severe NO2 accumulation, the denitrification microorganisms in the S0-PBR could be further restrained [41,57,58] and caused an undesirable NO3 removal performance, exhibiting the average NO3 removal efficiency declined to 80.8% from 85.2% during days 128–142 (Figure 1d).
In Stage III (days 156–163), the operational conditions were identical to those during days 100–127 in Stage I, with an overloading volumetric loading rate of 1.92 kg NO3-N/(m3·d) and no external chemical S2− addition. The average NO3 removal efficiency decreased further to 42.1% (Figure 1b). Simultaneously, the NO2 accumulation was aggravated to 19 mg NO2-N/L (Figure 1a), suggesting the deterioration of SADN performance in the S0-PBR with high influent NO3 loading applied. The high volumetric loading rate completely prevented the SD process in Stage III, as evidenced by the similar practical SO42− production (1267 mg SO42−/L) and theoretical value (1327 mg SO42−/L) (Figure 1d). Therefore, the deterioration could be attributed to the lack of precursor, such as biogenic/chemical S2−, to induce a chemical polysulfuration reaction.

3.2. The Short-Term Effects of Varying S0 Particle Sizes and Chemical S2− Addition on the SADN Process

Batch experiments were categorized into four groups based on the diameters of S0 particles, i.e., 1–2 mm, 3–5 mm, 7–9 mm, and 10–12 mm, for evaluating the effect of varying particle sizes on the SADN process.
NO3 removal fastened as the S0 particle size decreased (Figure 2a). NO3 was almost completely removed at 12 h when the S0 size was smaller than 5 mm (Figure 2a). Comparatively, the residual NO3 concentrations were approximately 16 mg NO3-N/L and 12 mg NO3-N/L at 12 h in the groups with the S0 particle sizes of 10–12 and 7–9 mm, respectively. Meanwhile, as a result of a lower NO2 reduction rate and faster NO3 reduction rate, the build-up of NO2 was gradually formed in all groups (Figure 2b), which was consistent with points that the capability of electron-scavenging for nitrite reductase was weaker than nitrate reductase [59,60].
The average specific NO3 removal rates within the first 12 h in groups with S0 sizes of 10–12 mm, 7–9 mm, 3–5 mm, and 1–2 mm applied were 0.672 g NO3-N/g VSS/h, 0.678 g NO3-N/g VSS/h, 0.850 g NO3-N/g VSS/h, and 0.910 g NO3-N/g VSS/h, respectively (Figure 2c). Additionally, it has been reported that a half-order reaction model could be used to explain the kinetics of the SADN process [24]. The half-order kinetic constants in groups with S0 sizes of 10–12, 7–9, 3–5, and 1–2 mm applied were calculated to be 0.382 mg-N1/2/L1/2/h, 0.435 mg-N1/2/L1/2/h, 0.545 mg-N1/2/L1/2/h, and 0.565 mg-N1/2/L1/2/h (Figure 2c), suggesting that the reaction rate constant increased with the specific surface area of S0 [25]. The smaller S0 size with a higher specific area not only provided a larger area for biofilm growth but, more importantly, reduced the mass-transfer resistance of insoluble S0 [10,29].
Generally, it was assumed that the saturation constants Ks was as low as 0.22 mg S/L in the SADN process [24,47], indicating that the affinities between S0 and the enzymes related in S0 oxidiation, such as SDO/SOR/Hdr were strong. Given that S0 was only taken up by S0OB after its solubilization and diffusion [16,21], the mass-transfer resistance of insoluble S0 became the main rate-limiting factor in the SADN system. The specific surface area of insoluble sulfur was the key parameter affecting the population of hydrolysis bacteria attached to its surface and the dissolution kinetics [61]. A kinetic model focusing on S0 hydrolysis as a prior and rate-limiting step was proposed, where both activities of hydrolytic biomass and autotrophic denitrifying bacteria in the SADN process were considered [19,62]. The model demonstrated that the specific surface area of S0 was the dominant factor affecting the denitrification rate.
To investigate how S2− or Sn2− promoted NO3 removal, batch Test II with an initial dosage of 20 mg S2−-S/L was conducted. During 0–9.5 h, the specific NO3 removal rates and NO3 consumption slope k in the S2−-added group were 0.557 g NO3-N/g VSS/h and 0.0465 (Figure 3a,b), respectively, significantly higher than the S2−-free group of 0.366 g NO3-N/g VSS/h and 0.0364. It could be due to the fact that the lower Gibbs energy was required when S2− with the relatively high solubility served as the additional electron donor, compared with the conventional SADN process [53]. However, the quietly close NO3 removal rates (3.0 mg NO3-N/(L·d) versus 3.6 mg NO3-N/(L·d)) were found in the two groups with S0 and chemical S2− as a single electron source by Qi et al. (2023) [63]. This result indicated that the acceleration of NO3 removal in this study should be mainly attributed to the Sn2− formation instead of chemical S2− participation. In the presence of chemicals S2− and S0, the abiotic polysulfuration process was triggered (Equation (2)). As a result of the product of soluble Sn2− with higher bioavailability, the NO3 removal rate was remarkably enhanced in the S2−-added group. This result was consistent with previous studies [29,30,32] and the improvement in NO3 removal efficiency in Stage II (day128–142) in the S0-PBR.
Of note, NO2 accumulation occurred in both groups and was aggravated by S2− addition (Figure 3c). The results were similar to the long-term performance of S0-PBR in Stage II and the previous study [63,64], indicating that dosing chemical S2− could significantly improve the NO3 reduction process but rarely promote NO2 reduction process. The aggravated NO2 accumulation in the S2−-added group resulted from the imbalance rate of the NO3 and NO2 reduction process, and the imbalance could be attributed to two main reasons. Firstly, the extent of NO2 accumulation was positively correlated with the NO3 reduction rate [60]. It can also be noticed that the NO3 removal rate was faster in the S2−-added group (Figure 3a,b) due to the presence of Sn2−, which explained the severe NO2 accumulation well. Secondly, the bio-toxicity of chemical S2− to nitrite reductase [54,55] and the higher competitive capacity of the nitrate reductase for electrons both hindered the NO2 reduction process [54,55,56].
Additionally, in the S2−-free group, only 1.6% of removed NO3-N was in the form of N2O-N within 27.5 h. The amount of N2O production was much lower than in the HD process, suggesting that less N2O was produced in the SADN process [19,65]. Additionally, a further decrease in N2O production in the S2−-added group was observed even in the presence of higher NO2 accumulation, only accounting for 0.7% of removed NO3-N within 27.5 h (Figure 3c). The result was consistent with the performance of the S0-PBR (Figure 1c) in Stage II. Similarly, a linearly proportional relationship between chemical S2− concentration and N2O emissions during autotrophic denitrification was reported in the study, including the mass ratio of S2−-S:NO3-N up to 5 [51]. Yang et al. (2016b) [51] also confirmed that chemical S2− had no inhibitory effect on nitrous oxide reductase. Moreover, Sn2− was formed in the S2−-added group due to the abiotic polysulfuration process. When Sn2− participated in the N2O reduction process, higher energy than S0-oxidation was yielded [66]. These explained the lower N2O production in the S2−-added group well.

4. Conclusions

Based on the long-term performance of the S0-PBR, the NO3 removal loading rate could be significantly enhanced using smaller S0 particle fillers with a higher specific surface area. More importantly, chemical S2− supplementation improved the performance of the S0-PBR under overloading conditions. It proved the feasibility of establishing an in situ PiSADN system by adding chemical S2− directly for high-loading wastewater treatment. The conducted batch tests have clarified the kinetic dynamics between the sizes of S0 particles and the rate of denitrification. Furthermore, the responses of the SADN process to chemical S2− were also investigated. The principle findings were summarized as follows:
Utilization of smaller S0 particles (0.5–1 mm) within the S0-PBR achieved a high volumetric loading rate of 1.44 kg NO3-N/(m3·d) and a NO3 removal efficiency nearing 100%, significantly surpassing outcomes observed in S0-PBR employing larger S0 particles (2–16 mm);
The supplementation of 30 mg S2−-S/L in the S0-PBR led to an increase in NO3 removal efficiency from 81.3%% to 85.3% and facilitated a 93.8% reduction in N2O accumulation;
In the batch tests with a S0 size of 10–12, 7–9, 3–5, and 1–2 mm applied, the average specific NO3 removal rates were 0.672 g NO3-N/g VSS/h, 0.678 g NO3-N/g VSS/h, 0.850 g NO3-N/g VSS/h, and 0.910 g NO3-N/g VSS/h, respectively, while the half-order kinetic constants were 0.382 mg-N1/2/L1/2/h, 0.435 mg-N1/2/L1/2/h, 0.545 mg-N1/2/L1/2/h, and 0.565 mg-N1/2/L1/2/h, respectively;
The specific NO3 removal rates and NO3 consumption slope k in the S2−-added group were 0.557 g NO3-N/g VSS/h and 0.0465, respectively, significantly higher than S2−-free group of 0.366 g NO3-N/g VSS/h and 0.0364;
The 1.6% of removed NO3-N was in the form of N2O within 27.5 h in the S2−-free group, while only 0.7% of the removed NO3-N was produced as N2O in the S2−-added group.

Author Contributions

Formal analysis, Conceptualization, Methodology, Investigation, Writing—original draft, J.X.; Investigation, Writing—review & editing, Z.L.; Supervision, Methodology, Funding acquisition, Y.X.; Supervision, Investigation, Methodology, Writing—review & editing, C.L.; Supervision, Investigation, Conceptualization, Project administration, Funding acquisition, Writing—review & editing, L.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (No. 52100061) and the Hubei Provincial Key Research and Development Program (No. 2022BCA067).

Data Availability Statement

Data will be made available on request.

Acknowledgments

The authors are grateful for the research collaboration.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, H.; Jiang, W.; Wan, D.; Qu, J. Study of a combined heterotrophic and sulfur autotrophic denitrification technology for removal of nitrate in water. J. Hazard. Mater. 2009, 169, 23–28. [Google Scholar] [CrossRef]
  2. Zhang, X.; Zhang, Y.; Shi, P.; Bi, Z.; Shan, Z.; Ren, L. The deep challenge of nitrate pollution in river water of China. Sci. Total Environ. 2021, 770, 144674. [Google Scholar] [CrossRef] [PubMed]
  3. Danni, S.O.; Bouchaou, L.; Elmouden, A.; Brahim, Y.A.; N’da, B. Assessment of water quality and nitrate source in the Massa catchment (Morocco) using δ 15N and δ 18O tracers. Appl. Radiat. Isot. 2019, 154, 108859. [Google Scholar] [CrossRef] [PubMed]
  4. Fields, S. Global Nitrogen: Cycling out of Control. Environ. Health Perspect. 2004, 112, A556–A563. [Google Scholar] [CrossRef]
  5. Della Rocca, C.; Belgiorno, V.; Meriç, S. Overview of in-situ applicable nitrate removal processes. Desalination 2007, 204, 46–62. [Google Scholar] [CrossRef]
  6. Panda, B.; Chidambaram, S.; Snow, D.; Malakar, A.; Singh, D.K.; Ramanathan, A.L. Source apportionment and health risk assessment of nitrate in foothill aquifers of Western Ghats, South India. Ecotoxicol. Environ. Saf. 2022, 229, 113075. [Google Scholar] [CrossRef] [PubMed]
  7. Sahinkaya, E.; Dursun, N. Sulfur-oxidizing autotrophic and mixotrophic denitrification processes for drinking water treatment: Elimination of excess sulfate production and alkalinity requirement. Chemosphere 2012, 89, 144–149. [Google Scholar] [CrossRef]
  8. Zumft, W.G. Cell biology and molecular basis of denitrification. Microbiol. Mol. Biol. Rev. 1997, 61, 533–616. [Google Scholar]
  9. Oh, S.E.; Yoo, Y.B.; Young, J.C.; Kim, I.S. Effect of organics on sulfur-utilizing autotrophic denitrification under mixotrophic conditions. J. Biotechnol. 2001, 92, 1–8. [Google Scholar] [CrossRef]
  10. Sierra-Alvarez, R.; Beristain-Cardoso, R.; Salazar, M.; Gómez, J.; Razo-Flores, E.; Field, J.A. Chemolithotrophic denitrification with elemental sulfur for groundwater treatment. Water Res. 2007, 41, 1253–1262. [Google Scholar] [CrossRef]
  11. Claus, G.N.; Kutzner, H.R. Physiology and kinetics of autotrophic denitrification by Thiobacillus denitrificans. Appl. Microbiol. Biotechnol. 1985, 22, 283–288. [Google Scholar] [CrossRef]
  12. Wiesmann, U. Biological nitrogen removal from wastewater. Adv. Biochem. Eng. Biotechnol. 1994, 51, 113–154. [Google Scholar]
  13. Lee, K.-C.; Rittmann, B.E. Applying a novel autohydrogenotrophic hollow-fiber membrane biofilm reactor for denitrification of drinking water. Water Res. 2002, 36, 2040–2052. [Google Scholar] [CrossRef]
  14. Batchelor, B.; Lawrence, A.W. Autotrophic denitrification using elemental sulfur. Water Pollut. Control Fed. 1978, 50, 1986–2001. [Google Scholar]
  15. Soares, M.I.M. Denitrification of groundwater with elemental sulfur. Water Res. 2002, 36, 1392–1395. [Google Scholar] [CrossRef]
  16. Wang, Y.; Bott, C.; Nerenberg, R. Sulfur-based denitrification: Effect of biofilm development on denitrification fluxes. Water Res. 2016, 100, 184–193. [Google Scholar] [CrossRef]
  17. Cui, Y.-X.; Biswal, B.K.; Guo, G.; Deng, Y.-F.; Huang, H.; Chen, G.-H.; Wu, D. Biological nitrogen removal from wastewater using sulphur-driven autotrophic denitrification. Appl. Microbiol. Biotechnol. 2019, 103, 6023–6039. [Google Scholar] [CrossRef]
  18. Peng, L.; Ni, B.-J.; Erler, D.; Ye, L.; Yuan, Z. The effect of dissolved oxygen on N2O production by ammonia-oxidizing bacteria in an enriched nitrifying sludge. Water Res. 2014, 66, 12–21. [Google Scholar] [CrossRef]
  19. Kostrytsia, A.; Papirio, S.; Frunzo, L.; Mattei, M.R.; Porca, E.; Collins, G.; Lens, P.N.L.; Esposito, G. Elemental sulfur-based autotrophic denitrification and denitritation: Microbially catalyzed sulfur hydrolysis and nitrogen conversions. J. Environ. Manag. 2018, 211, 313–322. [Google Scholar] [CrossRef]
  20. Florentino, A.P.; Weijma, J.; Stams, A.J.M.; Sánchez-Andrea, I. Ecophysiology and Application of Acidophilic Sulfur-Reducing Microorganisms. In Biotechnology of Extremophiles; Springer: Berlin/Heidelberg, Germany, 2016; pp. 141–175. [Google Scholar]
  21. Moraes, B.S.; Foresti, E. Determination of the intrinsic kinetic parameters of sulfide-oxidizing autotrophic denitrification in differential reactors containing immobilized biomass. Bioresour. Technol. 2012, 104, 250–256. [Google Scholar] [CrossRef] [PubMed]
  22. Batchelor, B.; Lawrence, A.W. A kinetic model for autotrophic denitrification using elemental sulfur. Water Res. 1978, 12, 1075–1084. [Google Scholar] [CrossRef]
  23. Dong, H.; Sun, Y.-L.; Sun, Q.; Zhang, X.-N.; Wang, H.-C.; Wang, A.-J.; Cheng, H.-Y. Effect of sulfur particle morphology on the performance of element sulfur-based denitrification packed-bed reactor. Bioresour. Technol. 2023, 367, 128238. [Google Scholar] [CrossRef]
  24. Koenig, A.; Liu, L.H. Kinetic model of autotrophic denitrification in sulfur packed-bed reactors. Water Res. 2001, 35, 1969–1978. [Google Scholar] [CrossRef] [PubMed]
  25. Moon, H.S.; Chang, S.W.; Nam, K.; Choe, J.; Kim, J.Y. Effect of reactive media composition and co-contaminants on sulfur-based autotrophic denitrification. Environ. Pollut. 2006, 144, 802–807. [Google Scholar] [CrossRef]
  26. Florentino, A.P.; Jan, W.; Stams, A.J.M.; Sánchez-Andrea, I. Sulfur Reduction in Acid Rock Drainage Environments. Environ. Sci. Technol. 2015, 49, 11746–11755. [Google Scholar] [CrossRef] [PubMed]
  27. Ucar, D.; Yilmaz, T.; Di Capua, F.; Esposito, G.; Sahinkaya, E. Comparison of biogenic and chemical sulfur as electron donors for autotrophic denitrification in sulfur-fed membrane bioreactor (SMBR). Bioresour. Technol. 2020, 299, 122574. [Google Scholar] [CrossRef]
  28. Wang, Y.; Sabba, F.; Bott, C.; Nerenberg, R. Using Kinetics and Modeling to Predict Denitrification Fluxes in Elemental-Sulfur Based Biofilms. Biotechnol. Bioeng. 2019, 116, 2698–2709. [Google Scholar] [CrossRef]
  29. Zhang, L.; Qiu, Y.-Y.; Zhou, Y.; Chen, G.-H.; van Loosdrecht, M.C.M.; Jiang, F. Elemental sulfur as electron donor and/or acceptor: Mechanisms, applications and perspectives for biological water and wastewater treatment. Water Res. 2021, 202, 117373. [Google Scholar] [CrossRef]
  30. Qiu, Y.-Y.; Zhang, L.; Mu, X.; Li, G.; Guan, X.; Hong, J.; Jiang, F. Overlooked pathways of denitrification in a sulfur-based denitrification system with organic supplementation. Water Res. 2020, 169, 115084. [Google Scholar] [CrossRef]
  31. Zhang, L.; Zhang, C.; Hu, C.; Liu, H.; Bai, Y.; Qu, J. Sulfur-based mixotrophic denitrification corresponding to different electron donors and microbial profiling in anoxic fluidized-bed membrane bioreactors. Water Res. 2015, 85, 422–431. [Google Scholar] [CrossRef]
  32. Qiu, Y.-Y.; Gong, X.; Zhang, L.; Zhou, S.; Li, G.; Jiang, F. Achieving a Novel Polysulfide-Involved Sulfur-Based Autotrophic Denitrification Process for High-Rate Nitrogen Removal in Elemental Sulfur-Packed Bed Reactors. ACS EST Eng. 2022, 2, 1504–1513. [Google Scholar] [CrossRef]
  33. Finster, K. Microbiological disproportionation of inorganic sulfur compounds. J. Sulfur Chem. 2008, 29, 281–292. [Google Scholar] [CrossRef]
  34. American Public Health Association (APHA); American Water Works Association (AWWA); Water Environment Federation (AEF). Standard Methods for the Examination of Water and Wastewater; American Public Health Association (APHA): Washington, DC, USA, 2005. [Google Scholar]
  35. Kleinjan, W.E.; de Keizer, A.; Janssen, A.J.H. Equilibrium of the reaction between dissolved sodium sulfide and biologically produced sulfur. Colloids Surf. B Biointerfaces 2005, 43, 228–237. [Google Scholar] [CrossRef]
  36. Zhang, L.; Zhang, Z.; Sun, R.; Liang, S.; Chen, G.-H.; Jiang, F. Self-accelerating sulfur reduction via polysulfide to realize a high-rate sulfidogenic reactor for wastewater treatment. Water Res. 2018, 130, 161–167. [Google Scholar] [CrossRef]
  37. Hao, W.; Li, Q.; Liu, P.; Han, J.; Duan, R.; Liang, P. A new inoculation method of sulfur autotrophic denitrification reactor for accelerated start-up and better low-temperature adaption. Sci. Total Environ. 2022, 823, 153657. [Google Scholar] [CrossRef]
  38. Yang, W.; Lu, H.; Khanal, S.K.; Zhao, Q.; Meng, L.; Chen, G.-H. Granulation of sulfur-oxidizing bacteria for autotrophic denitrification. Water Res. 2016, 104, 507–519. [Google Scholar] [CrossRef]
  39. Ruan, Y.-J.; Luo, G.-Z.; Tan, H.-X.; Che, X.; Jiang, Y.; Sun, D.-C. Nitrate and phosphate removal in sulphur-coral stone autotrophic denitrification packed-bed reactors. J. Environ. Eng. Sci. 2013, 8, 267–276. [Google Scholar] [CrossRef]
  40. Sadeq, M.; Moe, C.L.; Attarassi, B.; Cherkaoui, I.; ElAouad, R.; Idrissi, L. Drinking water nitrate and prevalence of methemoglobinemia among infants and children aged 1–7 years in Moroccan areas. Int. J. Hyg. Environ. Health 2008, 211, 546–554. [Google Scholar] [CrossRef]
  41. Chen, F.; Li, X.; Gu, C.; Huang, Y.; Yuan, Y. Selectivity control of nitrite and nitrate with the reaction of S0 and achieved nitrite accumulation in the sulfur autotrophic denitrification process. Bioresour. Technol. 2018, 266, 211–219. [Google Scholar] [CrossRef]
  42. Vavilin, V.A.; Fernandez, B.; Palatsi, J.; Flotats, X. Hydrolysis kinetics in anaerobic degradation of particulate organic material: An overview. Waste Manag. 2008, 28, 939–951. [Google Scholar] [CrossRef] [PubMed]
  43. Flere, J.M.; Zhang, T.C. Nitrate Removal with Sulfur-Limestone Autotrophic Denitrification Processes. J. Environ. Eng. 1999, 125, 721–729. [Google Scholar] [CrossRef]
  44. Lu, H.; Chandran, K. Factors promoting emissions of nitrous oxide and nitric oxide from denitrifying sequencing batch reactors operated with methanol and ethanol as electron donors. Biotechnol. Bioeng. 2010, 106, 390–398. [Google Scholar] [CrossRef]
  45. Sahu, A.; Conneely, T.; Nüsslein, K.; Ergas, S. Biological Perchlorate Reduction in Packed Bed Reactors Using Elemental Sulfur. Environ. Sci. Technol. 2009, 43, 4466–4471. [Google Scholar] [CrossRef]
  46. Wan, D.; Li, Q.; Liu, Y.; Xiao, S.; Wang, H. Simultaneous reduction of perchlorate and nitrate in a combined heterotrophic-sulfur-autotrophic system: Secondary pollution control, pH balance and microbial community analysis. Water Res. 2019, 165, 115004. [Google Scholar] [CrossRef]
  47. Guo, G.; Li, Z.; Chen, L.; Ling, Q.; Zan, F.; Isawi, H.; Hao, T.; Ma, J.; Wang, Z.; Chen, G.H.; et al. Advances in elemental sulfur-driven bioprocesses for wastewater treatment: From metabolic study to application. Water Res. 2022, 213, 118143. [Google Scholar] [CrossRef]
  48. Hijnen, W.; Bennekom, C.A.; Mijnarends, B.J. Optimization of the sulphur-limestone filtration process for nitrate removal from groundwater. Aqua 1992, 41, 209–218. [Google Scholar]
  49. Sun, Y.-L.; Zhai, S.; Qian, Z.-M.; Yi, S.; Zhuang, W.-Q.; Cheng, H.-Y.; Zhang, X.; Wang, A.-J. Managing Microbial Sulfur Disproportionation for Optimal Sulfur Autotrophic Denitrification in a Pilot-scale Elemental Sulfur Packed-bed Reactor. Water Res. 2023, 243, 120356. [Google Scholar] [CrossRef]
  50. Richardson, D.; Felgate, H.; Watmough, N.; Thomson, A.; Baggs, E. Mitigating release of the potent greenhouse gas N2O from the nitrogen cycle—Could enzymic regulation hold the key? Trends Biotechnol. 2009, 27, 388–397. [Google Scholar] [CrossRef]
  51. Yang, W.; Zhao, Q.; Lu, H.; Ding, Z.; Meng, L.; Chen, G.-H. Sulfide-driven autotrophic denitrification significantly reduces N2O emissions. Water Res. 2016, 90, 176–184. [Google Scholar] [CrossRef]
  52. Liu, Y.; Peng, L.; Ngo, H.; Guo, W.; Wang, D.-B.; Pan, Y.; Sun, J.; Ni, B.-J. Evaluation of Nitrous Oxide Emission from Sulfide- and Sulfur-Based Autotrophic Denitrification Processes. Environ. Sci. Technol. 2016, 50, 9407–9415. [Google Scholar] [CrossRef] [PubMed]
  53. Lin, S.; Mackey, H.; Hao, T.; Guo, G.; Loosdrecht, M.; Chen, G.H. Biological Sulfur Oxidation in Wastewater Treatment: A Review of Emerging Opportunities. Water Res. 2018, 143, 399–415. [Google Scholar] [CrossRef] [PubMed]
  54. Beristain-Cardoso, R.; Sierra-Alvarez, R.; Rowlette, P.; Razo-Flores, E.; Gomez, J.; Field, J. Sulfide oxidation under chemolithoautotrophic denitrifying conditions. Biotechnol. Bioeng. 2006, 95, 1148–1157. [Google Scholar] [CrossRef]
  55. Lu, H.; Huang, H.; Yang, W.; Mackey, H.; Khanal, S.; Wu, D.; Chen, G.-H. Elucidating the stimulatory and inhibitory effects of dissolved sulfide on sulfur-oxidizing bacteria (SOB) driven autotrophic denitrification. Water Res. 2018, 133, 165–172. [Google Scholar] [CrossRef]
  56. Huang, S.; Yu, D.; Chen, G.; Wang, Y.; Tang, P.; Liu, C.; Tian, Y.; Zhang, M. Realization of nitrite accumulation in a sulfide-driven autotrophic denitrification process: Simultaneous nitrate and sulfur removal. Chemosphere 2021, 278, 130413. [Google Scholar] [CrossRef]
  57. An, S.; Tang, K.; Nemati, M. Simultaneous biodesulphurization and denitrification using an oil reservoir microbial culture: Effects of sulphide loading rate and sulphide to nitrate loading ratio. Water Res. 2010, 44, 1531–1541. [Google Scholar] [CrossRef]
  58. Zhou, Z.; Xing, C.; An, Y.; Hu, D.; Qiao, W.; Wang, L. Inhibitory effects of sulfide on nitrifying biomass in the anaerobic–anoxic–aerobic wastewater treatment process. J. Chem. Technol. Biotechnol. 2014, 89, 214–219. [Google Scholar] [CrossRef]
  59. Di Capua, F.; Pirozzi, F.; Lens, P.N.L.; Esposito, G. Electron donors for autotrophic denitrification. Chem. Eng. J. 2019, 362, 922–937. [Google Scholar] [CrossRef]
  60. Bao, H.-X.; Li, Z.-R.; Song, Z.-B.; Wang, A.-J.; Zhang, X.-N.; Qian, Z.-M.; Sun, Y.-L.; Cheng, H.-Y. Mitigating nitrite accumulation during S0-based autotrophic denitrification: Balancing nitrate-nitrite reduction rate with thiosulfate as external electron donor. Environ. Res. 2022, 204, 112016. [Google Scholar] [CrossRef]
  61. Esposito, G.; Frunzo, L.; Panico, A.; Pirozzi, F. Modelling the effect of the OLR and OFMSW particle size on the performances of an anaerobic co-digestion reactor. Process Biochem. 2011, 46, 557–565. [Google Scholar] [CrossRef]
  62. Kostrytsia, A.; Papirio, S.; Mattei, M.; Frunzo, L.; Lens, P.N.L.; Esposito, G. Sensitivity analysis for an elemental sulfur-based two-step denitrification model. Water Sci. Technol. 2018, 78, 1296–1303. [Google Scholar] [CrossRef] [PubMed]
  63. Qi, X.; Han, J.; Kou, Z.; Liang, P. Supplementary sulfide during inoculation for improved sulfur autotrophic denitrification performance and adaptation to low temperature. Sci. Total Environ. 2023, 900, 166365. [Google Scholar] [CrossRef] [PubMed]
  64. Bao, H.-X.; Wang, H.-L.; Wang, S.-T.; Sun, Y.-L.; Zhang, X.-N.; Cheng, H.-Y.; Qian, Z.-M.; Wang, A.-J. Response of sulfur-metabolizing biofilm to external sulfide in element sulfur-based denitrification packed-bed reactor. Environ. Res. 2023, 231, 116061. [Google Scholar] [CrossRef] [PubMed]
  65. Zhou, W.; Liu, X.; Dong, X.; Wang, Z.; Yuan, Y.; Wang, H.; He, S. Sulfur-based autotrophic denitrification from the micro-polluted water. J. Environ. Sci. 2016, 44, 180–188. [Google Scholar] [CrossRef] [PubMed]
  66. Pokorna, D.; Zabranska, J. Sulfur-oxidizing bacteria in environmental technology. Biotechnol. Adv. 2015, 33, 1246–1259. [Google Scholar] [CrossRef]
Figure 1. Long-term performance of the S0-PBR: NO3 and NO2 concentrations of influent and effluent (a), influent NO3 loading and NO3 removal efficiency variations (b), effluent N2O concentration in liquid (c), and theoretical and practical SO42− generation (d).
Figure 1. Long-term performance of the S0-PBR: NO3 and NO2 concentrations of influent and effluent (a), influent NO3 loading and NO3 removal efficiency variations (b), effluent N2O concentration in liquid (c), and theoretical and practical SO42− generation (d).
Water 16 00730 g001
Figure 2. Variations of NO3 removal (a), NO2 accumulation (b), specific NO3 removal rate, and half-order reaction constant (c) with varying S0 particle size applied.
Figure 2. Variations of NO3 removal (a), NO2 accumulation (b), specific NO3 removal rate, and half-order reaction constant (c) with varying S0 particle size applied.
Water 16 00730 g002
Figure 3. Variations of NO3 and S2− (a), NO3 consumptions kinetics (b), and NO2 and N2O accumulation (c) over time in batch tests with and without the addition chemical S2−.
Figure 3. Variations of NO3 and S2− (a), NO3 consumptions kinetics (b), and NO2 and N2O accumulation (c) over time in batch tests with and without the addition chemical S2−.
Water 16 00730 g003
Table 1. Operational conditions of the S0-PBR.
Table 1. Operational conditions of the S0-PBR.
StagesStage IStage IIStage III
NO3-N (mg/L)20–400400400
HRT (h)8–555
Loading
(kg NO3-N/(m3·d))
0.06–1.921.921.92
S2− (mg S/L)-30–50-
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

Xu, J.; Lu, Z.; Xu, Y.; Liang, C.; Peng, L. Improved Performance of Sulfur-Driven Autotrophic Denitrification Process by Regulating Sulfur-Based Electron Donors. Water 2024, 16, 730. https://doi.org/10.3390/w16050730

AMA Style

Xu J, Lu Z, Xu Y, Liang C, Peng L. Improved Performance of Sulfur-Driven Autotrophic Denitrification Process by Regulating Sulfur-Based Electron Donors. Water. 2024; 16(5):730. https://doi.org/10.3390/w16050730

Chicago/Turabian Style

Xu, Jiang, Zhikun Lu, Yifeng Xu, Chuanzhou Liang, and Lai Peng. 2024. "Improved Performance of Sulfur-Driven Autotrophic Denitrification Process by Regulating Sulfur-Based Electron Donors" Water 16, no. 5: 730. https://doi.org/10.3390/w16050730

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