Next Article in Journal
Assessment of Radioactivity Concentrations and Associated Radiological Health Risk in Natural Spring Mineral Bottled Drinking Water from South Africa
Previous Article in Journal
Multi-Objective Optimal Planning and Deciding of Low Impact Development for an Intending Urban Area
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photo-Catalytic Reduction of Nitrate by Ag-TiO2/Formic Acid Under Visible Light: Selectivity of Nitrogen and Mechanism

1
School of Environmental and Chemical Engineering, Shanghai University, No. 333 Nanchen Road, Shanghai 200444, China
2
Shanghai Urban Construction Maintenance Management Co., Ltd., 403 Luoshan Road, Shanghai 200135, China
*
Authors to whom correspondence should be addressed.
Water 2025, 17(2), 155; https://doi.org/10.3390/w17020155
Submission received: 27 November 2024 / Revised: 26 December 2024 / Accepted: 7 January 2025 / Published: 9 January 2025
(This article belongs to the Section Wastewater Treatment and Reuse)

Abstract

:
Ubiquitous nitrate (NO3) in groundwater sources is considered a hazard compound for human health. Photo-catalytic reduction by Ag-TiO2/formic acid/visible light represents an emerging method for NO3 removal without secondary pollution. In this contribution, the removal of NO3 by photo-catalytic reduction and the selectivity of N2 were systematically investigated under varied conditions, including concentrations of Ag-TiO2, NO3, and formic acid (HCOOH). The removal efficiency of NO3 reached 84.47%, 82.68% of which was converted to N2 under the optimal conditions: NO3 at 50 mg-N/L, Ag-TiO2 at 1.0 g/L, HCOOH at 20.05 mmol/L, and reaction time at 120 min. The removal of NO3 was enhanced mainly by CO2 rather than by photo-generated electrons or HCOO. The results of this study indicated that the production of ·CO2 by Ag-TiO2 and HCOOH under visible light catalysis can achieve efficient NO3 removal.

1. Introduction

Nitrate (NO3) pollution is widespread in groundwater due to NO3 inevitably entering the environment through various routes, including over-fertilization and the discharge of industrial wastewater, etc. [1,2]. NO3 ingestion causes great damage to health, for example, methemoglobinemia, stomach cancer, or diabetes, with great effect on the health of babies, for example, “blue baby syndrome” or cyanosis [3,4]. High N causes water eutrophication and deteriorates natural aquatic and ecological system [5]. Therefore, it is imperative to develop effective and sustainable treatment technologies to remove NO3 and minimize associated risks.
The removal of NO3 can be realized by traditional physical–chemical treatment processes and biological denitrification [6,7]. However, conventional technologies have drawbacks, including low efficiency, secondary pollution, high cost, and residual waste streams of NO3, nitrite (NO2), and ammonia (NH4+) [8,9,10]. In general, NO3 is easier to be converted to NH4+, but the production of nontoxic forms like nitrogen (N2) is preferred [11,12,13,14,15,16,17]. Increasing N2 selectivity is a significant challenge in reducing NO3 in groundwater [11,18]. Many studies have found that photocatalysis can effectively remove pollutants from water [19], and carbon dioxide anion radical (·CO2) can efficiently reduce NO3 to N2 [20]. Chen et al. achieved 60% selectivity of N2 by ·CO2 under a UV/Fe(III)–oxalate system [10]. ·CO2 with a redox potential (−1.8 V) [21] similar to ·H (−2.3 V) [22] can be produced by the action of dissolving CO2 and solvating electrons in water [23], or by the interaction of hydrogen free radical (·H) [24], sulfate radical (·SO4) [25], and hydroxyl radical (·OH) [26] with organic carboxylic acids, including formic acid (HCOOH) and oxalic acid [27]. ·OH has shown certain advantages over ·SO4 due to the production of SO42−, which can cause the secondary pollution. However, ·OH is usually generated by vacuum ultraviolet photolysis of water (λexc = 185 nm) or the activation of H2O2, which requires high energy or has storage and transportation limitations.
Efficient ·CO2 production is urgently required. Using visible light instead of ultraviolet light to produce ·CO2 has great application prospects. Combining visible light and modified TiO2 to produce ·CO2 is a promising alternative. Modified TiO2 would show visible-light activity after deposition of Ag due to the localized surface plasmon resonance and form holes (h+) [28,29,30,31,32,33]. h+ with strong oxidizing ability is inclined to react with H2O to generate ·H, which could degrade the organic pollution efficiently. Ghafourian et al. achieved a high degradation efficiency of furan formaldehyde at 80.2% using visible/Ag-TiO2 [34]. ·OH or h+ could react with formic acid substances (formate (HCOO) or HCOOH) to generate ·CO2 [27,35]. Photo-catalytic NO3 reduction by ·CO2 is energy-efficient and has no secondary pollution.
However, visible/Ag-TiO2/HCOOH is rarely used to remove NO3, and the reduction efficiency, byproducts selectivity, and application potential of NO3 remain unclear. It should be noted that the realistic product of NO3 reduction is N2 rather than NH4+. Moreover, many studies have used HCOOH as a hole scavenger to reduce NO3, but few have focused on the role of HCOOH in the system. Duan et al. reported that photogenerated electrons capture NO3 and reduce it to N2 [29]. However, Anderson et al. believed that hole-capture agents (such as formic acid) should be added to fill the holes and generate carbon dioxide free radicals to reduce NO3 to N2 [28]. Thus, the mechanism of photo-catalytic reduction of NO3 requires further study.
Given these challenges and demands, the objectives of this study are as follows: (1) to determine the optimal experimental condition to achieve the highest NO3 removal efficiently and N2 conversion efficiency and (2) to explore whether the main reducing substances are photogenerated electrons or ·CO2 and reveal the possible mechanisms of NO3 reduction by Ag-TiO2/HCOOH under visible light.

2. Materials and Methods

2.1. Materials

All chemicals were of reagent grade without further purification. Titanium dioxide (TiO2, >99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China). Silver nitrate (AgNO3, >99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China), sodium formate (HCOONa, >99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China), formic acid (HCOOH, >99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China), potassium salt (KNO3, KCl, K2SO4, KHCO3, KH2PO4, and KI, all >99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China), hydrochloric acid (HCl, 36%~38%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China), methanol (CH3OH, >99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China), acetic acid (CH3COOH, >99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China), ethylene diamine tetra acetic acid disodium (EDTA, >99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China), and sodium hydroxide (NaOH, ≥96%, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China).

2.2. Preparation and Characterization of Ag-TiO2

Ag-TiO2 was prepared by the ultraviolet light reduction deposition method [36]. The detailed preparation steps of Ag-TiO2 are given in Text S1 (Supplementary Materials).
Ag-TiO2 was characterized by X-ray diffraction (XRD), ultraviolet visible diffuse reflectance spectra (UV–vis DRS), X-ray photoelectron spectroscopy (XPS), and photoluminescence spectra (PL). The crystal structure of Ag-TiO2 was analyzed by XRD (Dandong Aolong Radiation Instrument Group Co., Ltd., Dandong, China) with the instrument parameters for Cu Ka radiation (λ = 1.54056 Å) at 20–40 kV and 10–450 mA at a scanning rate of 8°/min over the 2θ range of 5–80°. The characterization of UV–vis DRS was operated through a Japanese Shimazu U-3010 UV–vis diffusion reflectance spectrometer (Qingdao Xingzhongyi Intelligent Technology Co., Ltd., Qingdao, China) at a scanning rate of 300 nm/min and a scanning wavelength range of 200–800 nm. The surface chemical states was analyzed by XPS run on ESCALAB 250Xi spectrometer (Dens Instrument Technology (Wuxi) Co., Ltd., Wuxi, China) using the K ray of Al target as the source of radiation (1486.6 eV), at a scanning range of 1100–0 eV; the spot size was 500 μm, and the voltage and current were 15 kV and 10 mA. The separation efficiency of the photo-generated e and h+ of Ag-TiO2 was analyzed through the characterization of PL by Fluorescence spectrophotometer (F-7000, Zhejiang Lichen Instrument Technology Co., Ltd., Shaoxing, China). Detailed characterization results are given in Figures S1–S5 (Supplementary Materials).

2.3. Photocatalytic Experiments

Photo-catalytic reduction of NO3 was performed in quartz glass (100 mL) that was continuously purged with N2 to remove dissolved oxygen; the suspension was then magnetically stirred to ensure a uniform contact between Ag-TiO2 and NO3 during the whole process. Next, 1.0 g/L of Ag-TiO2 and 100 mL of KNO3 (50 mg-N/L) was put into the reactor, 20.05 mmol/L of HCOOH was added, and the obtained suspension liquid was stirred for 30 min in the dark. Then, it was irradiated by a 500 W xenon lamp, which provided a visible light source, with continuous stirring for 3 h; a quartz cold hydrazine was used to control the temperature of the quartz glass. After the reaction was completed, the reaction fluid was filtrated through a filtration membrane (0.45 μm), and the supernatant was diluted to determine the concentration of NO3, NO2, and ammonia (NH4+) by ultraviolet spectrophotometer [10,37].

2.4. Analytical Methods

The analysis of N-containing compounds was detected using a UV–visible spectrophotometer (UV-1800PC) and ion chromatography (MIC, Zurich, Switzerland). NO2, NH4+, and N2 are the main reduction products in the process of visible light catalytic reduction of NO3. The product selectivity to NO2, NH4+, and N2 was calculated according to Equations (1)–(3).
S N O 2 = C t N O 2 C 0 N O 3 C t N O 3
S N H 4 + = C t N H 4 + C 0 N O 3 C t N O 3
S N 2 = C 0 N O 3 C t N O 3 C t N O 2 C t N H 4 + C 0 N O 3 C t N O 3
where C0 is the initial concentration of the substance, and Ct is the concentration of the substance in the test solution at reaction time t [38].

3. Results and Discussion

3.1. Removal of Nitrate

In actual wastewater, the initial concentration of NO3 fluctuates over time. Therefore, it is significant and necessary to investigate the photo-catalytic performance under the different NO3 concentrations. The reduction of NO3 was evaluated under a visible light system. Figure 1a shows the removal of NO3 in different initial concentrations (15–100 mg-N/L). After 30 min of reaction, the removal efficiency of NO3 at 15 mg-N/L reached 77.63%, and almost all of the NO3 was reduced after 60 min. The removal efficiency reached 84.51% at 50 mg-N/L after 120 min. The removal efficiency of NO3 was 70.59% after 210 min when NO3 was 100 mg-N/L. The removal efficiency of NO3 was slightly decreased with increasing NO3 concentration. Although the photo-catalytic efficiency decreased with increasing initial concentration, the actual removal of NO3 from 15 mg-N/L increased to 70.59 mg-N/L. Under each NO3 concentration investigated, NO3 reduction could be fitted by Langmuir Hinshelwood kinetic model described by Equation (4).
r = dC d t = k r K a d C 1 + K a d C
where r is the removal efficiency of NO3, C is the concentration of NO3 at reaction time t, Kr is the intrinsic rate, and Kad is the adsorption equilibrium constant.
As the adsorption is weaker, and the initial concentration of NO3 is lower, KadC can be ignored. In this case, the model can be simplified into a first-order kinetic equation (Equation (5)). For the initial reaction efficiency (t = 0, C = C0), the model can be simplified as follows (Equation (6)).
r = k r K a d C = kC
ln c c 0 = kt
where C0 is the initial concentration of NO3, k is the pseudo-first-order rate constant that can be obtained by linear regression of the plot of ln(C/C0) versus t [39,40], and the fitting parameters are shown in Table S1 (Supplementary Data). As illustrated in Figure 1b, a linear relationship between k and concentration of NO3 was established. The k was decreased with increasing NO3 concentration; that is, the photo-catalytic reduction efficiency of NO3 was decreased. When the NO3 concentration gradually increases, the active sites provided by the fixed dosage of Ag-TiO2 are insufficient [41,42]. The excess NO3 was adsorbed and accumulated on the surface of Ag-TiO2, which would inhibit the adsorption of HCOO and the production of ·CO2. Meanwhile, a high initial concentration of NO3 could reduce light penetration and photon energy and lead to a lower removal efficiency [43]. In fact, NO3 levels in drinking water sources in many countries still exceed the maximum contaminant level of 50 mg-N/L [44]. Therefore, 50 mg-N/L was selected as the initial NO3 concentration through the whole experiment.

3.2. Selectivity of Nitrogen

3.2.1. Effect of Ag-TiO2 Dosage

The photo-catalyst is an important part of photo-catalytic reaction system, and a photo-catalytic reaction cannot carry on in its absence. Figure 2 illustrates the effects of Ag-TiO2 dosage on NO3 removal efficiency and product selectivity. The removal efficiency of NO3 increased from 58.79% to 78.94% and then decreased to 54.77% with the increase of Ag-TiO2 from 0.5 to 2.0 g/L, as shown in Figure 2a. N2 selectivity showed the same trend with NO3 removal efficiency, which firstly increased from 53.10% to 69.32% and then decreased to 59.46%. The opposite trend was found for NO2 selectivity: it decreased from 40.98% to 21.84% and then increased to 29.78%. The selectivity of NH4+ increased slightly with the increase of Ag-TiO2 dosage but remained at a low concentration of <10.76%. The N2 selectivity was consistent with Sowmya et al.’s observation [40].
Further, the selection intensity of the product can be expressed by multiplying the NO3 removal efficiency by the product selectivity. The N2 selection intensity was the highest when the Ag-TiO2 dosage was at 1.0 g/L, as shown in Figure 2b. The reasons are as follows. When the dosage of Ag-TiO2 increased from 0.5 to 1.0 g/L, ·CO2 generation was increased to a sufficient value, and the turbidity induced by the addition of Ag-TiO2 did not affect the visible light. However, when Ag-TiO2 was further increased from 1.0 to 2.0 g/L, the active sites were covered by Ag-TiO2 particulate, which was not conducive to the transformation of NO3 and led to NO2 accumulation [45,46]. The turbidity also affected the visible light. The production process of ·CO2 in the photo-catalytic reduction of nitrate by Ag-TiO2/HCOOH is clearly illustrated by Equations (7)–(11). Ag-TiO2 generated hot electrons and h+ under the irradiation of visible light; meanwhile, the hot electrons are injected into the adjacent semiconductor TiO2. The h+ can react with H2O to produce ·OH Equations (7)–(9) [34]. When HCOOH is used as a hole scavenger, ·OH and h+ react with formic acid to produce ·CO2 (Equations (10) and (11)) [47].
Ag + hv(visible) → Ag*
Ag + TiO2 → Ag+ (h+) + TiO2 (e)
H2O + h+ → H+ + ·OH
·OH + HCOO → ·CO2 + H2O
h+ + HCOO → H+ + ·CO2

3.2.2. Effect of Nitrate Concentrations

Figure 3a illustrates the influence of initial NO3 concentration from 15 to 50 and 100 mg-N/L on the NO3 removal efficiency and product selectivity. With initial NO3 concentrations increasing from 15 to 50 and 100 mg-N/L, the NO3 removal efficiency decreased from 100% to 84.46% and 54.98% at a Ag-TiO2 dosage of 1.0 g/L. The removal efficiency of NO3 was negatively correlated with its concentration. The low NO3 removal efficiency at 100 mg-N/L was due to the low Ag-TiO2 dosage, which could not provide sufficient h+. The excess NO3 competed with HCOO for h+, which decreased ·CO2 production and further inhibited NO3 transformation to N2, and thus, NO2 accumulated. As shown in Figure 3a, the selectivity of NO2 increased from 0.41% to 21.63%. The selectivity of N2 was 70.46%, 82.68%, and 77.84% at NO3 concentrations of 15, 50, and 100 mg-N/L. The NH4+ selectivity decreased from 29.13% to 3.36% and 0.53% (Figure 3a). At a low NO3 of 15 mg-N/L, NO2 absorbed on the surface of Ag-TiO2 and was converted to N2 or NH4+. As the NO3 concentration increased, more ·OH was generated, as shown in Equations (12) and (13). The ·OH transformed NO2 to HOONO and reduced the selectivity of N2 and NH4+ (Equations (14) and (15)) [43]. The maximum N2 selection intensity was obtained when the NO3 concentration was 50 mg-N/L, as shown in Figure 3b. Based on NO3 removal efficiency and N2 selection intensity, the optimal NO3 concentration is 50 mg-N/L.
NO3 + hv → ·NO2 + ·O
·O + H+ ↔·OH
·OH + NO2 → ·NO2 + OH
·OH + ·NO2 →HOONO

3.2.3. HCOOH Concentrations

The effect of the concentration of HCOOH on the photo-catalytic NO3 reduction is illustrated in Figure 4. Figure 4a shows the removal efficiency of NO3 increased from 47.17% to 81.21% along with the increasing dosage of HCOOH from 4.01 mmol/L to 20.05 mmol/L and reached the highest at 20.05 mmol/L (M(HCOOH:NO3) = 5.6). When the HCOOH concentration further increased to 40.11 mmol/L and 80.22 mmol/L, the removal efficiency of NO3 decreased to 56.30% and 38.28%. The highest selectivity of N2 was achieved at 70.33%.
When the HCOOH concentration is low, less ·CO2 is produced, and NO3 removal and N2 selectivity are low. Nevertheless, a high concentration of HCOOH hampers the adsorption of NO3 on the surface of Ag-TiO2 due to the competitive adsorption between negatively charged HCOO and NO3, which also inhibits NO3 removal.
NH4+ was maintained at a low selectivity of 5% at all HCOOH concentrations. The selectivity of NO2 was high at 76.60% when the HCOOH concentration was 4.01 mmol/L. NO2 decreased at first and then increased with an increase in HCOOH concentration, and N2 showed the opposite trend. Zhang et al. explained that the insufficiency of HCOOH could promote the formation of NO2 (Equations (16) and (17)) [48,49]; hence, HCOOH was easily transformed into NO2 when NO3 was insufficient. Figure 4b shows that the maximum N2 selectivity at the concentration of HCOOH was 20.05 mmol/L. In addition, considering the limit of COD proposed by the Chinese environmental protection agency, 20.05 mmol/L of NO3 concentration was selected as the optimal dosage of hole scavengers.
2NO3 + 5HCOO + 7H+ → N2 + 5CO2 + 6H2O
NO3 + HCOO + H+ → NO2 + CO2 + H2O

3.3. Mechanism of Nitrate Reduction

3.3.1. Choice of Hole Scavengers

Hole scavengers play a significant role in the photo-catalytic process. In order to maintain the electrical neutrality of materials, the generated holes need to be transferred by hole scavengers. Hole scavengers can irreversibly combine with photo-generated h+ and effectively prevent the combination of photo-generated electrons and h+, thereby promoting the efficiency of the photo-catalytic reduction of NO3. Different hole scavengers have different binding abilities with photo-generated h+ and affect the efficiency of the photo-catalytic reduction of NO3. Hence, it is very important to select the appropriate hole scavengers. In order to determine the optimal hole scavengers with high efficiency in the photo-catalytic NO3 reduction process, we compared HCOOH with its substructural analogs, including HCOONa, CH3OH, and CH3COOH.
As shown in Figure 5a, the removal efficiency of NO3 followed the order of HCOOH > HCOONa > CH3OH > CH3COOH due to the difference in initial pH (Table 1). The initial pH can affect the surface charge of TiO2. There is a zwitterionic substance titanium hydroxyl group (TiOH) on the surface of TiO2, as shown in reaction Equations (18) and (19) [50]. The isoelectric point of titanium dioxide is 6.25. When the solution pH is below 6.25, the surface titanium hydroxyl group of TiO2 will undergo protonation, and the surface of TiO2 will carry a positive charge. This is conducive to adsorbing the negatively charged ions by electrostatic adsorption.
The pH of 2.7 of the HCOOH solution was lower than the isoelectric point ≈ 6.25 of TiO2, which was conducive to adsorption of NO3. The pH of 6.63 of the HCOONa system was higher than 6.25 (isoelectric point of TiO2), which made the surface of Ag-TiO2 negatively charged. And due to electrostatic repulsion, this was not conducive to the adsorption of NO3 and HCOO, resulting in a low NO3 removal efficiency. Although the pH in both cases was lower than 6.25, the removal efficiency of NO3 was unsatisfactory with CH3OH or CH3COOH as hole scavengers. Generally, CH3OH was used as a hole scavenger, and it was oxidized to formaldehyde by holes rather than ·CO2. Moreover, the adsorption of CH3OH on the catalyst surface was extremely weak and could not capture the h+ effectively. Ag-TiO2 has a strong adsorption capacity for CH3COOH, which inhibits the adsorption of NO3 on the surface of Ag-TiO2. It was thus concluded that HCOOH is the optimal hole scavenger.
TiOH + H+ → TiOH2+  pH < 6.25
TiOH → TiO + H+   pH > 6.25
Based on previous studies, the effect of HCOOH as hole scavenger can be summarized as follows: (1) HCOOH can capture holes, reduce the combination probability of holes and electrons, and improve the yield of the photo-quanta of the catalyst; (2) HCOOH can yield an initial pH of 2.7, lower than the isoelectric point of TiO2, which is conducive to adsorbing NO3 and HCOO on the surface of TiO2 for catalysis and promoting the reaction of the photo-catalytic reduction of NO3; (3) HCOOH has reducibility (E0 (CO2/HCOO) = −0.2 eV) and can also be used as a reducing agent for NO3 reduction in theory; and (4) HCOOH can be oxidized by holes to generate ·CO2 (E0 (CO2/·CO2) = −1.8 eV) in the process of trapping holes. In the HCOOH and HCOONa photo-catalytic reduction system, the reducing substances include photo-generated electrons (E0 (e) = −0.3 V) generated by the catalyst excited by incident light and ·CO2 from HCOO (E0 (CO2/HCOO) = −0.2 eV), which contribute to efficient NO3 removal. Some scholars believe that, in the process of the photo-catalytic reduction of NO3, ·CO2 plays a major role instead of photo-generated electrons [29]. However, fewer studies have provided controversial evidence that photo-generated electrons are more important [10,28]. This study was conducted to verify whether photo-generated e or ·CO2 is predominant in the photo-catalytic reduction of nitrate by Ag-TiO2/HCOOH under visible light.

3.3.2. Effective Reducing Substances

In this study, the major reducing agent was explored by controlling the concentration of HCOOH to control the formation of ·CO2 and combining with the free radical quenching experiment to indirectly control the production path of photo-generated e, HCOO, and ·CO2.
In order to determine the effective reductants, the removal of NO3 was compared with different hole scavengers, including H2O, HCOOH, KI, and EDTA + HCOOH. As shown in Figure 5b, the NO3 removal efficiency follows the order of HCOOH > KI > EDTA + HCOOH > H2O. Different reducing substances were produced using different hole scavengers, as shown in Table 2. The reducing substances were photo-generated e, ·CO2, HCOO, and ·OH with HCOOH as the hole scavenger in the photo-catalytic reduction of NO3 by Ag-TiO2 under visible light, and the highest NO3 removal efficiency was achieved at 82.7%. When H2O was used as the hole scavenger, the reducing substance was ·OH, and NO3 removal efficiency was only 1.3%. KI is commonly used as a hole scavenger. When photo-generated e was the reducing substance, NO3 removal efficiency was only 7.2%, as show in Table 3. EDTA is a frequently used radical quencher, which can quench ·CO2 and ·OH. Compared with the HCOOH system, the main reductants in the EDTA + HCOOH system are photo-generated e and HCOO. NO3 removal efficiency was 5.2% in the EDTA + HCOOH system. Combining the reducing substances and NO3 removal efficiency under a different hole scavenger system, the main reductant is ·CO2.
On the basis of the above discussion, the mechanism of photo-catalytic reduction of NO3 by Ag-TiO2/HCOOH under visible light was revealed. (1) When Ag-TiO2 is irradiated with visible light, high light quantum yield is produced. (2) The hot electrons generated by Ag can be effectively transferred to the energy band of TiO2, contributing to a high separation between e and h. (3) h+ reacts with H2O to generate ·OH; then, h+ reacts with ·OH and HCOOH to generate ·CO2. (4) ·CO2 is captured by NO3, and NO3 is reduced to NO3 and finally to N2 and/or NH4+.

3.4. Effect of Co-Anions

In groundwater, anions like sulphate (SO42−), chloride (Cl), bicarbonate (HCO3), and phosphate (H2PO4) are present. This work investigated the influence of these co-anions on photo-catalytic NO3 reduction and selectivity of N2 under visible light.
The NO3 removal efficiency was effected by SO42−, Cl, and H2PO4, while HCO3 did not affect NO3 removal efficiency, as shown in Figure 6. The higher the concentration of SO42−, Cl, and H2PO4, the stronger the inhibition effect on NO3 reduction. This result is consistent with Jessica et al.’s work [51]. This result aligns with the surface anionization effects suggested by Sheng et al. that SO42− and H2PO4 adsorbed h+, which blocked the adsorption and degradation of HCOO [52].
Figure 7a–d show the selectivity of the reduction products (NO2, NH4+, and N2) of NO3 under different co-anions. The existence of Cl was adverse for the reduction of NO3 but was conducive to the reduction of NO3 to N2.
H2PO4 led to an increase in NO2, as shown in Figure 7c. The reason is that H2PO4 can release H+, and PO42− and PO42− compete with NO3 for electrons, and NO3 is reduced to NO2 rather N2 [40]. SO42− can act as an electron acceptor and competed withs NO3 for electrons, resulting in a decreased NO3 removal efficiency. HCO3, as the hole scavenger, can capture h+ and ·OH to form ·CO2, which improves the removal efficiency of NO3. Meanwhile, the existence of HCO3 contributes to a high buffer capacity. When the pH is greater than 6.25, TiO2 will be deprotonated, resulting in a negative charge on the TiO2 surface. Due to electrostatic repulsion, the adsorption of NO3 and HCOO on the surface of TiO2 will be hindered, resulting in the reduction of ·CO2 and reduced NO3 removal efficiency.

4. Conclusions

The photocatalytic reduction of NO3 by Ag-TiO2/HCOOH under visible light was efficient, and the reduction reaction conformed to the first-order kinetic model. The removal efficiency of NO3 firstly increased and then decreased with the increase in the concentration of Ag-TiO2, NO3, and HCOOH. The concentration of Ag-TiO2, NO3, and HCOOH affected the removal efficiency of NO3 and the selectivity of NH4+, NO2, and N2. All of them influenced the generation of ·CO2, which was confirmed in the major reductants. When the concentration of Ag-TiO2 was 1.0 g/L, the concentration of HCOOH was 20.05 mmol/L, the removal efficiency of NO3 (50 mg-N/L) was the highest at 84.47%, and N2 conversion efficiency reached 82.68% in 120 min. Except for HCO3, the existence of co-anions affected NO3 reduction, and the higher the concentration of anions, the stronger the inhibition effect. These results demonstrated that the method of producing ·CO2 by Ag-TiO2 and HCOOH could realize efficient removal of NO3 and selectivity of N2 in groundwater under visible light.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/w17020155/s1, Figure S1: Preparation steps of Ag-TiO2; Figure S2: The UV–vis DRS of Ag-TiO2 with different Ag content; Figure S3: Energy band gap of samples; Figure S4: The XRD characterization of samples; Figure S5: The XPS characterization of Ag-TiO2; Table S1: The kinetic constants and correlation coefficients.

Author Contributions

Investigation, Y.S. and J.X.; writing—original draft, Y.S.; conceptualization, H.C. and Y.X.; writing—review and editing, J.C. and X.Z.; supervision, J.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

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

Conflicts of Interest

Author Yi Xie, Jun Xia and Xiaolin Zhang was employed by the Shanghai Urban Construction Maintenance Management Co Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential con-flict of interest.

References

  1. Burow, K.R.; Nolan, B.T.; Rupert, M.G.; Dubrovsky, N.M. Nitrate in Groundwater of the United States, 1991–2003. Environ. Sci. Technol. 2010, 44, 4988–4997. [Google Scholar] [CrossRef] [PubMed]
  2. Bijay, S.; Craswell, E. Fertilizers and nitrate pollution of surface and ground water: An increasingly pervasive global problem. SN Appl. Sci. 2021, 3, 518. [Google Scholar] [CrossRef]
  3. Islam, M.; Patel, R. Synthesis and physicochemical characterization of Zn/Al chloride layered double hydroxide and evaluation of its nitrate removal efficiency. Desalination 2010, 256, 120–128. [Google Scholar] [CrossRef]
  4. Kostraba, J.N.; Gay, E.C.; Rewers, M.; Hamman, R.F. Nitrate levels in community drinking waters and risk of IDDM. An ecological analysis. Diabetes Care 1992, 15, 1505–1508. [Google Scholar] [CrossRef]
  5. Meng, S.; Ling, Y.; Yang, M.; Zhao, X.; Osman, A.I.; Al-Muhtaseb, A.A.H.; Rooney, D.W.; Yap, P.-S. Recent research progress of electrocatalytic reduction technology for nitrate wastewater: A review. J. Environ. Chem. Eng. 2023, 11, 109418. [Google Scholar] [CrossRef]
  6. Tugaoen, H.O.; Garcia-Segura, S.; Hristovski, K.; Westerhoff, P. Challenges in photocatalytic reduction of nitrate as a water treatment technology. Sci. Total Environ. 2017, 599–600, 1524–1551. [Google Scholar] [CrossRef]
  7. Tyagi, S.; Rawtani, D.; Khatri, N.; Tharmavaram, M. Strategies for Nitrate removal from aqueous environment using Nanotechnology: A Review. J. Water Process Eng. 2018, 21, 84–95. [Google Scholar] [CrossRef]
  8. Zhang, B.; Liu, Y.; Tong, S.; Zheng, M.; Zhao, Y.; Tian, C.; Liu, H.; Feng, C. Enhancement of bacterial denitrification for nitrate removal in groundwater with electrical stimulation from microbial fuel cells. J. Power Sources 2014, 268, 423–429. [Google Scholar] [CrossRef]
  9. Wehbe, N.; Jaafar, M.; Guillard, C.; Herrmann, J.-M.; Miachon, S.; Puzenat, E.; Guilhaume, N. Comparative study of photocatalytic and non-photocatalytic reduction of nitrates in water. Appl. Catal. A Gen. 2009, 368, 1–8. [Google Scholar] [CrossRef]
  10. Chen, J.; Zhang, R.; Chen, D.; Liu, J.; Chen, S. Carbon dioxide radical reducing nitrate to nitrogen gas in a UV/Fe(III)-oxalate system. J. Water Process Eng. 2021, 40, 101934. [Google Scholar] [CrossRef]
  11. Tugaoen, H.O.N.; Herckes, P.; Hristovski, K.; Westerhoff, P. Influence of ultraviolet wavelengths on kinetics and selectivity for N-gases during TiO2 photocatalytic reduction of nitrate. Appl. Catal. B Environ. 2018, 220, 597–606. [Google Scholar] [CrossRef]
  12. Wu, L.; Chen, S.; Zhou, J.; Zhang, C.; Liu, J.; Luo, J.; Song, G.; Qian, G.; Song, L.; Xia, M. Simultaneous removal of organic matter and nitrate from bio-treated leachate via iron–carbon internal micro-electrolysis. RSC Adv. 2015, 5, 68356–68360. [Google Scholar] [CrossRef]
  13. Hérissan, A.; Meichtry, J.M.; Remita, H.; Colbeau-Justin, C.; Litter, M.I. Reduction of nitrate by heterogeneous photocatalysis over pure and radiolytically modified TiO2 samples in the presence of formic acid. Catal. Today 2017, 281, 101–108. [Google Scholar] [CrossRef]
  14. Zhao, F.; Xin, J.; Yuan, M.; Wang, L.; Wang, X. A critical review of existing mechanisms and strategies to enhance N2 selectivity in groundwater nitrate reduction. Water Res. 2021, 209, 117889. [Google Scholar] [CrossRef]
  15. Xue, S.; Wen, X.; Yun, Y. Pathway and mechanism study on improvement of N2 selectivity of catalytic denitrification. J. Saudi Chem. Soc. 2023, 27, 101577. [Google Scholar] [CrossRef]
  16. Chang, Y.; Meng, J.; Hu, Y.; Qi, S.; Hu, Z.; Wu, G.; Zhou, J.; Zhan, X. Unacclimated activated sludge improved nitrate reduction and N2 selectivity in iron filling/biochar systems. Sci. Total Environ. 2024, 947, 174581. [Google Scholar] [CrossRef]
  17. Li, F.; Zhang, W.; Zhang, P.; Gong, A.; Kexun, L. Strategies of selective electroreduction of aqueous nitrate to N2 in chloride-free system: A critical review. Green Energy Environ. 2024, 9, 198–216. [Google Scholar] [CrossRef]
  18. Mertah, O.; El Hajjaji, K.; El Amrani, S.; Tanji, K.; Goncharova, I.; Kherbeche, A. Visible-light Cu/TiO2@Ag3PO4 heterostructure photocatalyst for selective nitrate reduction and antimicrobial activity. Opt. Mater. 2022, 129, 112549. [Google Scholar] [CrossRef]
  19. Pawar, M.; Nain, P.; Rani, S.; Sharma, B.; Kumar, S.; Majeed Khan, M.A. Structural, optical, photocatalytic and thermal behaviour of (Nb, Ta) co-doped WO3 nanoparticles and its application in photocatalytic degradation of MG and RhB dyes. Opt. Mater. 2024, 157, 116277. [Google Scholar] [CrossRef]
  20. Chen, L.; Zheng, H.; Li, A.; Qiu, X.; Wang, L. Lewis acid-rich SrFexTi1−xO3/TiO2 to enhance the photocatalytic reduction of nitrate to N2. J. Hazard. Mater. 2024, 473, 198–216. [Google Scholar] [CrossRef]
  21. Wardman, P. Reduction Potentials of One-Electron Couples Involving Free Radicals in Aqueous Solution. J. Phys. Chem. Ref. Data 1989, 18, 1637–1755. [Google Scholar] [CrossRef]
  22. Li, X.; Ma, J.; Liu, G.; Fang, J.; Yue, S.; Guan, Y.; Chen, L.; Liu, X. Efficient reductive dechlorination of monochloroacetic acid by sulfite/UV process. Environ. Sci. Technol. 2012, 46, 7342–7349. [Google Scholar] [CrossRef]
  23. Grills, D.C.; Lymar, S.V. Radiolytic formation of the carbon dioxide radical anion in acetonitrile revealed by transient IR spectroscopy. Phys. Chem. Chem. Phys. 2018, 20, 10011–10017. [Google Scholar] [CrossRef]
  24. Liu, X.; Zhong, J.; Fang, L.; Wang, L.; Ye, M.; Shao, Y.; Li, J.; Zhang, T. Trichloroacetic acid reduction by an advanced reduction process based on carboxyl anion radical. Chem. Eng. J. 2016, 303, 56–63. [Google Scholar] [CrossRef]
  25. Berkovic, A.M.; Gonzalez, M.C.; Russo, N.; Michelini Mdel, C.; Pis Diez, R.; Martire, D.O. Reduction of mercury(II) by the carbon dioxide radical anion: A theoretical and experimental investigation. J. Phys. Chem. A 2010, 114, 12845–12850. [Google Scholar] [CrossRef]
  26. Vilhunen, S.; Vilve, M.; Vepsalainen, M.; Sillanpaa, M. Removal of organic matter from a variety of water matrices by UV photolysis and UV/H2O2 method. J. Hazard. Mater. 2010, 179, 776–782. [Google Scholar] [CrossRef]
  27. Sá, J.; Agüera, C.A.; Gross, S.; Anderson, J.A. Photocatalytic nitrate reduction over metal modified TiO2. Appl. Catal. B Environ. 2009, 85, 192–200. [Google Scholar] [CrossRef]
  28. Clavero, C. Plasmon-induced hot-electron generation at nanoparticle/metal-oxide interfaces for photovoltaic and photocatalytic devices. Nat. Photonics 2014, 8, 95–103. [Google Scholar] [CrossRef]
  29. Duan, L.; Lin, Q.; Peng, H.; Lu, C.; Shao, C.; Wang, D.; Rao, S.; Cao, H.; Lv, W. The catalytic reduction mechanisms of metal-doped TiO2 for nitrate produced from non-thermal discharge plasma: The interfacial photogenerated electron transfer and reduction process. Appl. Catal. A Gen. 2023, 650, 118995. [Google Scholar] [CrossRef]
  30. Nain, P.; Pawar, M.; Rani, S.; Sharma, B.; Kumar, S.; Majeed Khan, M.A. (Ce, Nd) co-doped TiO2 NPs via hydrothermal route:Structural, optical, photocatalytic and thermal behavior. Mater. Sci. Eng. B 2024, 309, 117648. [Google Scholar] [CrossRef]
  31. Khan, M.A.M.; Ansari, A.A.; Choudhary, P.; Ahmed, J.; Kumar, S.; Hussain, S. Reduced graphene oxide supported Ag-loaded brookite TiO2 nanowires: Enhanced photocatalytic degradation performance and electrochemical energy storage applications. Diam. Relat. Mater. 2023, 139, 110397. [Google Scholar] [CrossRef]
  32. Khan, M.A.M.; Nain, P.; Kumar, S.; Ansari, A.A.; Ahamed, M.; Shahabuddin, M. Improved photocatalytic and electrochemical activities of (Nd3+, Yb3+) co-doped TiO2 nanoparticles synthesized by hydrothermal protocol. J. Mater. Sci. Mater. Electron. 2024, 35, 2149. [Google Scholar] [CrossRef]
  33. Xu, R.; Guan, Y.; Cao, L.; Yu, X.; Li, L.; Ma, B.; Liu, C. Boosting light harvesting and charge separation in Au nanoparticles and g-C3N4 co-modified TiO2 nanotube arrays for efficient photoelectrocatalytic reduction of nitrate in water. J. Water Process Eng. 2024, 68, 106316. [Google Scholar] [CrossRef]
  34. Ghafourian, N.; Lashanizadegan, M.; Hosseini, S.N. Ag/TiO2/EP: A low-cost and floating plasmonic photocatalyst for degrading furfural under visible light irradiation. Int. J. Environ. Sci. Technol. 2017, 14, 2721–2732. [Google Scholar] [CrossRef]
  35. Yang, W.; Wang, J.; Chen, R.; Xiao, L.; Shen, S.; Li, J.; Dong, F. Reaction mechanism and selectivity regulation of photocatalytic nitrate reduction for wastewater purification: Progress and challenges. J. Mater. Chem. A 2022, 10, 17357–17376. [Google Scholar] [CrossRef]
  36. Ko, S.; Banerjee, C.K.; Sankar, J. Photochemical synthesis and photocatalytic activity in simulated solar light of nanosized Ag doped TiO2 nanoparticle composite. Compos. Part B Eng. 2011, 42, 579–583. [Google Scholar] [CrossRef]
  37. Adewuyi, Y.G.; Sakyi, N.Y.; Arif Khan, M. Simultaneous removal of NO and SO2 from flue gas by combined heat and Fe2+ activated aqueous persulfate solutions. Chemosphere 2018, 193, 1216–1225. [Google Scholar] [CrossRef]
  38. Chen, J.; Liu, J.; Zhou, J.; Chen, D. Reductive removal of nitrate by carbon dioxide radical with high product selectivity to form N2 in a UV/H2O2/HCOOH system. J. Water Process Eng. 2020, 33, 101097. [Google Scholar] [CrossRef]
  39. Lin, B.; Sun, P.; Zhou, Y.; Jiang, S.; Gao, B.; Chen, Y. Interstratified nanohybrid assembled by alternating cationic layered double hydroxide nanosheets and anionic layered titanate nanosheets with superior photocatalytic activity. J. Hazard. Mater. 2014, 280, 156–163. [Google Scholar] [CrossRef]
  40. Sowmya, A.; Meenakshi, S. Photocatalytic reduction of nitrate over Ag–TiO2 in the presence of oxalic acid. J. Water Process Eng. 2015, 8, e23–e30. [Google Scholar] [CrossRef]
  41. Chen, H.; Shao, Y.; Xu, Z.; Wan, H.; Wan, Y.; Zheng, S.; Zhu, D. Effective catalytic reduction of Cr(VI) over TiO2 nanotube supported Pd catalysts. Appl. Catal. B Environ. 2011, 105, 255–262. [Google Scholar] [CrossRef]
  42. Nishio, J.; Tokumura, M.; Znad, H.T.; Kawase, Y. Photocatalytic decolorization of azo-dye with zinc oxide powder in an external UV light irradiation slurry photoreactor. J. Hazard. Mater. 2006, 138, 106–115. [Google Scholar] [CrossRef]
  43. Shankar, M.V.; Nelieu, S.; Kerhoas, L.; Einhorn, J. Photo-induced degradation of diuron in aqueous solution by nitrites and nitrates: Kinetics and pathways. Chemosphere 2007, 66, 767–774. [Google Scholar] [CrossRef]
  44. Ward, M.H.; Jones, R.R.; Brender, J.D.; de Kok, T.M.; Weyer, P.J.; Nolan, B.T.; Villanueva, C.M.; van Breda, S.G. Drinking Water Nitrate and Human Health: An Updated Review. Int. J. Environ. Res. Public Health 2018, 15, 1557. [Google Scholar] [CrossRef]
  45. Guo, J.; Jiang, L.; Liang, J.; Xu, W.; Yu, H.; Zhang, J.; Ye, S.; Xing, W.; Yuan, X. Photocatalytic degradation of tetracycline antibiotics using delafossite silver ferrite-based Z-scheme photocatalyst: Pathways and mechanism insight. Chemosphere 2021, 270, 128651. [Google Scholar] [CrossRef]
  46. Xu, L.; Yang, L.; Johansson, E.M.J.; Wang, Y.; Jin, P. Photocatalytic activity and mechanism of bisphenol a removal over TiO2−x/rGO nanocomposite driven by visible light. Chem. Eng. J. 2018, 350, 1043–1055. [Google Scholar] [CrossRef]
  47. Harbour, J.R.; Hair, M.L. Spin trapping of the ·CO2 radical in aqueous medium. Can. J. Chem. 1979, 57, 1150–1152. [Google Scholar] [CrossRef]
  48. Kobwittaya, K.; Sirivithayapakorn, S. Photocatalytic reduction of nitrate over TiO2 and Ag-modified TiO2. J. Saudi Chem. Soc. 2014, 18, 291–298. [Google Scholar] [CrossRef]
  49. Zhang, F.; Jin, R.; Chen, J.; Shao, C.; Gao, W.; Li, L.; Guan, N. High photocatalytic activity and selectivity for nitrogen in nitrate reduction on Ag/TiO2 catalyst with fine silver clusters. J. Catal. 2005, 232, 424–431. [Google Scholar] [CrossRef]
  50. Kormann, C. Photolysis of Chloroform and Other Organic Molecules in Aqueous TiO2 Suspensions. Env. Sci. Technol. 1991, 25, 494–500. [Google Scholar] [CrossRef]
  51. Freire, J.M.A.; Matos, M.A.F.; Abreu, D.S.; Becker, H.; Diógenes, I.C.N.; Valentini, A.; Longhinotti, E. Nitrate photocatalytic reduction on TiO2: Metal loaded, synthesis and anions effect. J. Environ. Chem. Eng. 2020, 8, 103844. [Google Scholar] [CrossRef]
  52. Sheng, H.; Li, Q.; Ma, W.; Ji, H.; Chen, C.; Zhao, J. Photocatalytic degradation of organic pollutants on surface anionized TiO2: Common effect of anions for high hole-availability by water. Appl. Catal. B Environ. 2013, 138–139, 212–218. [Google Scholar] [CrossRef]
Figure 1. The kinetics of the NO3 reduction: (a) removal efficiency of NO3; (b) k at different initial NO3 concentrations (reaction conditions: NO3 with different concentrations; HCOOH: 20.05 mmol/L; Ag-TiO2: 1.0 g/L; visible light: 120 min).
Figure 1. The kinetics of the NO3 reduction: (a) removal efficiency of NO3; (b) k at different initial NO3 concentrations (reaction conditions: NO3 with different concentrations; HCOOH: 20.05 mmol/L; Ag-TiO2: 1.0 g/L; visible light: 120 min).
Water 17 00155 g001
Figure 2. Effect of different dosages of Ag-TiO2: (a) the removal efficiency of NO3 and selectivity to N2, NH4+, and NO2; (b) the selection intensity of products (reaction conditions: NO3: 50 mg-N/L; HCOOH: 20.05 mmol/L; irradiation time: 120 min).
Figure 2. Effect of different dosages of Ag-TiO2: (a) the removal efficiency of NO3 and selectivity to N2, NH4+, and NO2; (b) the selection intensity of products (reaction conditions: NO3: 50 mg-N/L; HCOOH: 20.05 mmol/L; irradiation time: 120 min).
Water 17 00155 g002
Figure 3. Effect of different dosages of initial concentration: (a) the removal efficiency of NO3 and selectivity to N2, NH4+, and NO2; (b) the selection intensity to products (reaction conditions: NO3: 100 mL; 50 mg-N/L mole ratio of HCOOH:NO3 = 5.6:1; Ag-TiO2: 1.0 g/L; irradiation time: 120 min).
Figure 3. Effect of different dosages of initial concentration: (a) the removal efficiency of NO3 and selectivity to N2, NH4+, and NO2; (b) the selection intensity to products (reaction conditions: NO3: 100 mL; 50 mg-N/L mole ratio of HCOOH:NO3 = 5.6:1; Ag-TiO2: 1.0 g/L; irradiation time: 120 min).
Water 17 00155 g003
Figure 4. Effect of different dosages of HCOOH: (a) the removal efficiency of NO3 and selectivity to N2, NH4+, and NO2; (b) the selection intensity to products (reaction conditions: NO3: 50 mg-N/L; Ag-TiO2: 1.0 g/L; irradiation time: 120 min).
Figure 4. Effect of different dosages of HCOOH: (a) the removal efficiency of NO3 and selectivity to N2, NH4+, and NO2; (b) the selection intensity to products (reaction conditions: NO3: 50 mg-N/L; Ag-TiO2: 1.0 g/L; irradiation time: 120 min).
Water 17 00155 g004
Figure 5. The analysis of effective reductants: (a) with different hole scavengers; (b) with different reducing substances (reaction conditions: NO3: 50 mg-N/L; Ag-TiO2: 1.0 g/L; irradiation time: 120 min. (a) M (hole scavenger:NO3 = 5.6:1); (b) HCOOH: 20.05 mmol/L, KI/EDTA: 20 mmol/L).
Figure 5. The analysis of effective reductants: (a) with different hole scavengers; (b) with different reducing substances (reaction conditions: NO3: 50 mg-N/L; Ag-TiO2: 1.0 g/L; irradiation time: 120 min. (a) M (hole scavenger:NO3 = 5.6:1); (b) HCOOH: 20.05 mmol/L, KI/EDTA: 20 mmol/L).
Water 17 00155 g005
Figure 6. NO3 removal efficiency under different co-anions.
Figure 6. NO3 removal efficiency under different co-anions.
Water 17 00155 g006
Figure 7. The selectivity of the reduction products (NO2, NH4+, and N2) of nitrate under different co-anions (reaction conditions: NO3: 50 mg-N/L; HCOOH: 20.05 mmol/L; Ag-TiO2: 1.0 g/L; irradiation time: 120 min).
Figure 7. The selectivity of the reduction products (NO2, NH4+, and N2) of nitrate under different co-anions (reaction conditions: NO3: 50 mg-N/L; HCOOH: 20.05 mmol/L; Ag-TiO2: 1.0 g/L; irradiation time: 120 min).
Water 17 00155 g007
Table 1. The initial solution pH corresponding to the different hole scavengers.
Table 1. The initial solution pH corresponding to the different hole scavengers.
Hole ScavengersHCOOHHCOONaCH3OHCH3COOH
Initial pH2.76.65.23.5
Table 2. Main reducing substances in different hole scavengers.
Table 2. Main reducing substances in different hole scavengers.
Hole ScavengerMain Reducing Substances
HCOOH·CO2, photo-generated e, HCOO, ·OH
H2O·OH
KIPhoto-generated e
EDTA+HCOOHPhoto-generated e, HCOO
Table 3. The removal efficiency of NO3 with different concentrations of KI.
Table 3. The removal efficiency of NO3 with different concentrations of KI.
KI (mmol/L)Removal Efficiency of NO3 (%)
16.8
27.1
207.2
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

Shi, Y.; Xie, Y.; Xia, J.; Zhang, X.; Cheng, H.; Chen, J. Photo-Catalytic Reduction of Nitrate by Ag-TiO2/Formic Acid Under Visible Light: Selectivity of Nitrogen and Mechanism. Water 2025, 17, 155. https://doi.org/10.3390/w17020155

AMA Style

Shi Y, Xie Y, Xia J, Zhang X, Cheng H, Chen J. Photo-Catalytic Reduction of Nitrate by Ag-TiO2/Formic Acid Under Visible Light: Selectivity of Nitrogen and Mechanism. Water. 2025; 17(2):155. https://doi.org/10.3390/w17020155

Chicago/Turabian Style

Shi, Yuanyuan, Yi Xie, Jun Xia, Xiaolin Zhang, Hui Cheng, and Jialin Chen. 2025. "Photo-Catalytic Reduction of Nitrate by Ag-TiO2/Formic Acid Under Visible Light: Selectivity of Nitrogen and Mechanism" Water 17, no. 2: 155. https://doi.org/10.3390/w17020155

APA Style

Shi, Y., Xie, Y., Xia, J., Zhang, X., Cheng, H., & Chen, J. (2025). Photo-Catalytic Reduction of Nitrate by Ag-TiO2/Formic Acid Under Visible Light: Selectivity of Nitrogen and Mechanism. Water, 17(2), 155. https://doi.org/10.3390/w17020155

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