You are on page 1of 9

Chemosphere 130 (2015) 8–16

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Remediation of nitrate–nitrogen contaminated groundwater using


a pilot-scale two-layer heterotrophic–autotrophic denitrification
permeable reactive barrier with spongy iron/pine bark
Guoxin Huang a,b, Yuanying Huang c,⇑, Hongyan Hu d, Fei Liu b, Ying Zhang e, Renwei Deng a
a
China Meat Research Center, Beijing Academy of Food Sciences, Beijing 100068, China
b
Beijing Key Laboratory of Water Resources & Environmental Engineering, China University of Geosciences (Beijing), Beijing 100083, China
c
National Research Center for Geoanalysis, Beijing 100037, China
d
Hydrogeology and Engineering Geology Prospecting Institute of Heilongjiang Province, Harbin 150030, China
e
Yunnan HITECH Environmental Protection Technology Co., Ltd., Kunming 650032, China

h i g h l i g h t s

 We proposed a heterotrophic–autotrophic denitrification permeable reactive barrier.



 The barrier achieved a high NO3 –N removal efficiency (>91%) before 38 PVs.
 The packing structure over previous ones generated more H2 and CO2.
 Aerobic heterotrophic bacteria played a dominant role in oxygen depletion.
 Spongy iron ensured a robust oxygen removal in addition to form H2.

a r t i c l e i n f o a b s t r a c t

Article history: A novel two-layer heterotrophic–autotrophic denitrification (HAD) permeable reactive barrier (PRB) was
Received 17 October 2014 proposed for remediating nitrate–nitrogen contaminated groundwater in an oxygen rich environment,
Received in revised form 31 January 2015 which has a packing structure of an upstream pine bark layer and a downstream spongy iron and river
Accepted 7 February 2015
sand mixture layer. The HAD PRB involves biological deoxygenation, heterotrophic denitrification,
Available online 3 March 2015
hydrogenotrophic denitrification, and anaerobic Fe corrosion. Column and batch experiments were per-
Handling Editor: Chang-Ping Yu formed to: (1) investigate the NO 3 –N removal and inorganic geochemistry; (2) explore the nitrogen
transformation and removal mechanisms; (3) identify the hydrogenotrophic denitrification capacity;
Keywords: and (4) evaluate the HAD performance by comparison with other approaches. The results showed that
Nitrate–nitrogen the HAD PRB could maintain constant high NO 3 –N removal efficiency (>91%) before 38 pore volumes
Groundwater remediation (PVs) of operation (corresponding to 504 d), form little or even negative NO2 –N during the 45 PVs, and
Heterotrophic–autotrophic denitrification produce low NH+4–N after 10 PVs. Aerobic heterotrophic bacteria played a dominant role in oxygen deple-
(HAD) tion via aerobic respiration, providing more CO2 for hydrogenotrophic denitrification. The HAD PRB sig-
Permeable reactive barrier (PRB) nificantly relied on heterotrophic denitrification. Hydrogenotrophic denitrification removed 10–20% of
Pine bark
the initial NO
3 –N. Effluent total organic carbon decreased from 403.44 mg L
1
at PV 1 to 9.34 mg L1
Spongy iron
at PV 45. Packing structure had a noticeable effect on its denitrification.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction recorded in numerous aquifers in China (Zhang et al., 2013), Italy


(Sacchi et al., 2013), Mexico (Pastén-Zapata et al., 2014), Spain
Nitrate–nitrogen contamination of groundwater has become an (Boy-Roura et al., 2013), America (Su and Puls, 2007), Canada
environmental and public health problem worldwide (Della Rocca (Levallois et al., 1998), Portugal (Mendes and Ribeiro, 2010), Ger-
et al., 2007). High NO
3 –N concentration (>10 mg L
1
) has been many (Shomar et al., 2008), and Australia (Salvestrin and Hagare,
2009). It is worth mentioning that most NO 3 –N laden groundwater
also contains dissolved oxygen (DO) (Huang et al., 2012). Public
⇑ Corresponding author at: 26 Baiwanzhuang Street, Xicheng District, Beijing
health concerns arise from the potential reduction of NO 3 –N to
100037, China. Tel.: +86 1068999579; fax: +86 1068998605.
NO2 –N and the endogenous formation of N-nitroso compounds,
E-mail address: huangy1978@126.com (Y. Huang).

http://dx.doi.org/10.1016/j.chemosphere.2015.02.029
0045-6535/Ó 2015 Elsevier Ltd. All rights reserved.
G. Huang et al. / Chemosphere 130 (2015) 8–16 9

which can cause methemoglobinemia, hypertension, cancers, mal- two parallel columns (Columns 1 and 2) packed with spongy iron
formation, mutation, and even death (Huang et al., 2012). and pine bark. In the two studies, both single- and double-layer
Remediating NO 3 –N contaminated groundwater has been and packing structures were designed. For the single-layer structure,
continues to be an area of active research. Among the available the mixed reactive materials (spongy iron and pine bark) were
remediation approaches (Westerhoff and James, 2003; Wang packed in a reactive zone; whereas for the double-layer structure,
et al., 2009; Demiral and Gündüzoğlu, 2010), biological denitrifica- the ZVI (steel wool or spongy iron) was placed in the upstream lay-
tion, falling into heterotrophic denitrification and autotrophic er and the solid carbon (cotton or pine bark) was filled in the
denitrification, is receiving increasing attention. Heterotrophic downstream layer. Unfortunately, the packing structure of an
denitrification is widely studied and applied in the field due to high upstream solid carbon layer followed by a downstream ZVI layer
denitrifying rate and high treatment capacity (Liu et al., 2009). has been neglected, and furthermore data on the feasibility and
Even so, it has shown some shortcomings such as remnant organic effectiveness of this packing structure are rather scarce.
carbon (requiring an intensive post-treatment process) and high Here, we report the performance of a novel two-layer HAD PRB
cell yield (0.6–0.9 g cells g1 NO 3 –N) (Ergas and Reuss, 2001; to remediate NO 3 –N contaminated groundwater in an aerobic
Shrimali and Singh, 2001). Autotrophic denitrification supported environment and further develop HAD processes and PRB systems.
with hydrogen, which is known as hydrogenotrophic denitrifica- Our HAD PRB concept, mainly involving a combination of biological
tion, is particularly promising due to clean nature of hydrogen, deoxygenation, heterotrophic denitrification, hydrogenotrophic
low biomass yield, relatively low cost, no external carbon source, denitrification, and anaerobic Fe corrosion, was composed of an
and avoidance of secondary organic carbon contamination (Lee upstream pine bark layer and a downstream spongy iron and river
and Rittmann, 2002; van Rijn et al., 2006; Moon et al., 2008). Nev- sand mixture layer. The packing structure over the previous pack-
ertheless, hydrogenotrophic denitrification is limited by inade- ing ones is expected to allow the generation of more cathodic
quate carbon dioxide, low solubility (1.6 mg L1 at 20 °C), high hydrogen and carbon dioxide to support hydrogenotrophic denitri-
flammability, and high cost of hydrogen gas (Haugen et al., 2002; fication. Heterotrophic denitrification contributed to the removal
Jha and Bose, 2005; Della Rocca et al., 2006; Ghafari et al., 2008; of most NO 3 –N. Aerobic heterotrophic bacteria played a dominant
Shin and Cha, 2008). role in oxygen depletion. Spongy iron potentially ensures a robust
To take the advantages and overcome the disadvantages of sin- oxygen removal in case of temporary insufficient biological deoxy-
gle heterotrophic denitrification and autotrophic denitrification, a genation. The objectives of this study were to: (1) investigate the
combined heterotrophic–autotrophic denitrification (HAD) NO 3 –N removal and inorganic geochemistry caused by the pro-
approach was proposed by Della Rocca et al. (2006) and further posed HAD PRB; (2) explore the nitrogen transformation and clean
developed by Liu et al. (2014) and Huang et al. (2012). In the up mechanisms; (3) identify the hydrogenotrophic denitrification
HAD processes, zero-valent iron (ZVI) via chemical reduction capacity in the HAD PRB; and (4) evaluate the HAD performance
removes most oxygen to rapidly create an anaerobic environment, by comparison with other approaches.
supporting biological denitrification. Heterotrophic denitrifiers via
heterotrophic denitrification utilize organic carbons (cotton,
2. Materials and methods
methanol, and pine bark) as electron donors to biologically reduce
most NO 3 –N and meanwhile generate CO2, favoring
2.1. Materials and chemicals
hydrogenotrophic denitrification. ZVI via anaerobic Fe corrosion
forms cathodic hydrogen with the aid of hydrogenase enzymes,
Spongy iron (0.15–2.00 mm, Brunauer–Emmett–Teller area
supporting hydrogenotrophic denitrification as well.
0.49 m2 g1, micropore volume 0.0012 cm3 g1, micropore dia-
Hydrogenotrophic denitrifiers employ the carbon dioxide and
meter 101.87 Å) was obtained from Kaibiyuan Co., Beijing, China,
cathodic hydrogen to biologically reduce the remaining NO 3 –N.
which consisted of Fe0 (60.60 wt%), C (0.78 wt%), S (0.06 wt%), P
Additionally, the solid materials serve as biofilm carriers. Although
(0.04 wt%), Mn (0.29 wt%), Ni (0.02 wt%), Cr (0.02 wt%), Cu
the symbiotic, synergistic, and promotive effects of the mechan-
(0.02 wt%), and Al (0.26 wt%). Pine bark (2–11 mm, Brunauer–Em-
isms mentioned above in the HAD processes were encouraging,
mett–Teller area 0.46 m2 g1, micropore volume 0.0018 cm3 g1,
questions arise regarding capacity of hydrogenotrophic denitrifica-
micropore diameter 159.42 Å) was purchased from a local nursery
tion and performance evaluation by comparison with other
store in Beijing, China. Gravel (2–5 mm) and river sand (0.45–
denitrification approaches.
2 mm) were donated by a quarry in Beijing, China. Hydrogen gas
Passive permeable reactive barrier (PRB) is considered one of
(industrial grade, purity 99.9%) was purchased from Beijing Hua
the innovative technologies widely accepted as an alternative to
Yuan Gas Chemical Industry Co., Ltd., Beijing, China. Deionized
the conventional pump-and-treat for sustainable in situ ground-
water was used to prepare reagent solutions. Beijing tap water
water remediation (Obiri-Nyarko et al., 2014) due to more cost
was spiked with NaNO3, NaHCO3, and K2HPO4 acting as synthetic
effectiveness and lower maintenance in the long-term (Phillips,
groundwater. In addition to NO 3 –N (20–105 mg L
1
), the synthetic
2009). PRBs have been successful in removing a variety of ground-
groundwater contained the following (final concentration in
water contaminants such as NO 3 –N, petroleum hydrocarbon, and +
mg L1 except pH): NO 2 –N (<0.1), NH4–N (<0.8), F

(0.3), Cl
heavy metals (Guerin et al., 2002; Schipper et al., 2005; Zhang 2
(20.8), HCO 3 (254.2), SO4 (46.4), Na+ (142.3–281.9), K+ (11.7),
et al., 2012). Over the past decades, much research work has been
Ca2+ (48.5), Mg2+ (28.1), DO (5–9), P (3.0), and pH (7.5–8.5). Unless
done on refining site characterization techniques, developing reac-
otherwise indicated, all chemicals used were analytical reagent
tive media (or sorbents), and improving installation and design
grade as received.
(Phillips, 2009). However, only a few attempts have been made
to combine multiple remediation mechanisms and optimize pack-
ing structures. 2.2. Column experiment setup and operation
More recently, for target NO 3 –N, several column studies have
paid particular attention to HAD PRBs. For instance, Della Rocca A pilot-scale flow through HAD PRB experiment was conducted
et al. (2006) investigated the effect factors (e.g., flow rate, influent employing a plexiglas column (20.6 cm i.d., 150 cm in length)
NO 3 –N, and retention time) using four parallel columns packed (Fig. 1a and b). A lower support layer of 5 cm of gravel was packed
with steel wool and cotton (R1–R4). Liu et al. (2014) determined at the bottom of the column, and a 5 cm gravel cap was placed at
the spatial and temporal variations of water quality indices using the top. The reactive zone was composed of an upstream layer
10 G. Huang et al. / Chemosphere 130 (2015) 8–16

(b)

(a) Effluent

Legend:
150cm
Gravel

Spongy iron
125cm
Sand

Pine bark 105cm


Water seal

Intermediate sampling ports


65cm (c)
Gas tubing 1 Gas tubing 2

25cm H2

Mixed
bacteria
0cm
+
Influent Reservoir Synthetic
Feed pump groundwater

Fig. 1. (a) Schematic and (b) photograph of the continuous flow setup in the column experiment, and (c) schematic of an incubation bottle in the batch tests.

(110 cm long) packed with pine bark (7.13 kg) and a downstream as follows: (1) add a 0.45 lm filter membrane covered with mixed
layer (30 cm long) filled with a mixture of spongy iron (7.99 kg) bacteria and 500 mL tap water enriched with 22 mg L1 NO 3 –N,
and river sand (7.40 kg), providing a pore volume (PV) of 24.77 L 3.0 mg L1 K2HPO4-P, and 350 mg L1 NaHCO3 into an incubation
and a porosity of 54.28%. The river sand was used in an effort to bottle; (2) sparge the bottle with N2 through Gas tubing 2 for
assure optimal permeability, minimize potential clogging and 5 min at 200 kPa (Mousavi et al., 2012); (3) sparge the bottle with
avoid cementation during operation. Fourteen intermediate sam- H2 through Gas tubing 1 for 30 and 10 min at 50 kPa before and
pling ports were positioned along the column height, at 15, 25, after Gas tubing 2 was sealed; (4) cover the bottle with aluminum
35, 45, 55, 65, 75, 85, 95, 105, 115, 125, 135, and 145 cm from foil; (5) statically incubate the bottle in the dark at 23 ± 5 °C; (6)
the influent end. draw water samples after 3, 6, 18, 22, and 30 d; and (7) after sam-
It is important to note that the inoculum accounting for 45% of pling, readd H2 into the bottle.
the column pore volume was introduced into the column, and then Two mixed bacterial suspensions (each 200 mL) were respec-
recycled for 20 d for the microorganisms to attach onto the surface tively collected at 65 and 125 cm of the column after 40 PVs. To
of the media particles. The inoculum was enriched from a sub-sur- separate the mixed bacteria from TOC and inhibit heterotrophic
face soil taken from a pristine and humic acid rich area, and found denitrifying bacteria, each bacterial suspension was filtered
to contain heterotrophic denitrifiers (belonging to Bacillus, Clostri- through a 0.45 lm filter membrane with depleting suspended solid,
dium, Flavobacterium, Steroidobacter, and Novosphingobium), and then the filtrate was passed through a 0.2 lm filter membrane
hydrogenotrophic denitrifiers (belonging to Pseudomonas), and with retaining the mixed bacteria on the filter membrane.
aerobic heterotrophic bacteria (belonging to Adhaeribacter and
Flavisolibacter) (Liu et al., 2014).
Under the steady state, the column was loaded continuously
2.4. Analytical methods and instruments
with synthetic groundwater in an upflow mode using an adjustable
multiport peristaltic pump set (BT 100-1F drive, DG-4 pump head, +
NO 
3 –N, NO2 –N, and NH4–N were determined using a Hewlett
Baoding Longer Precision Pump Co., Ltd., Baoding, China). The col-
Packard 8453 ultraviolet–visible spectrophotometer (USA),
umn was covered with aluminum foil to shade out photosynthesiz-
employing ultraviolet spectrophotometric method at 220 and
ers (Rust et al., 2002), and operated at room temperature
275 nm, N-(1-naphthyl)-ethylenediamine dihydrochloride colori-
(23 ± 5 °C) in 9 phases (Table S1 in Supporting Information) each
metric method at 540 nm, and Nessler’s reagent colorimetric
for 5 PVs, in sequence. Water samples were collected from the
method at 410 nm, respectively. The lower detection limits for
influent, effluent, and intermediate sampling ports, and stored at +
NO 
3 –N, NO2 –N, and NH4–N were 0.08, 0.003, and 0.025 mg N L
1
,
4 °C prior to analysis for pH, DO, total organic carbon (TOC),
+ respectively. DO and water temperature were monitored using a
NO 
3 –N, NO2 –N, NH4–N, and water temperature.
Hach HQ30d DO portable meter (USA). pH was measured using a
Sartorius PB-10 digital pH meter (Germany). TOC was measured
2.3. Batch test design using a Shimadzu 3201 TOC analyzer (Japan) by measuring the dif-
ference between the total carbon (burning at 680 °C) and inorganic
Batch tests were conducted in sterile rubber capped, 1 L brown carbon (acidification with 2 M HCl). In each analysis, at least one in
glass bottles to identify the hydrogenotrophic denitrification capa- five samples was duplicated and the deviation between the two
city (Fig. 1c). The procedures of a batch incubation were introduced samples was always less than 5%.
G. Huang et al. / Chemosphere 130 (2015) 8–16 11

2.5. Data processing ating conditions during the 36–40 PVs (Fig. 2). The drop also
demonstrated the reduction of both pine bark carbon releasing
Analytical data are presented as means of triplicate measure- capacity and organic carbon available to the heterotrophic denitri-
ments. The number of PVs was expressed as the ratio of accumulated fying population (Fig. 5b). It is evident that a time-dependent
water volume (L) over time to column pore volume (L). Total nitrogen decay in the HAD PRB denitrification performance was noticeable
(TN) (mg L1) was defined as the sum of NO 1 
3 –N (mg L ), NO2 –N after 38 PVs. Volokita et al. (1996a) have also reached a similar
1 + 1 1
(mg L ), and NH4–N (mg L ). N removed (mg L ) was defined as conclusion when newspaper served as a sole carbon source and
influent concentration (mg L1)–effluent concentration (mg L1) electron donor. Afterward, removal efficiency (77–82%) tended to
(for the column experiment) or initial concentration (mg L1)–final reach a steady state and effluent NO 3 –N was kept at a level of
concentration (mg L1) at time t (d) (for the batch tests). N formed 18.23–23.66 mg L1 during the 40–45 PVs (Fig. 2), which was
(mg L1) was defined as effluent concentration (mg L1)–influent around its limitation value (20 mg L1 for drinking use) of quality
concentration (mg L1) (for the column experiment) or final concen- standard for ground water of China (State Bureau of Technical
tration (mg L1)–initial concentration (mg L1) at time t (d) (for the Supervision (SBTS), 1993).
batch tests). Removal efficiency (%) was calculated as [influent con-
centration (mg L1)  effluent concentration (mg L1)]/influent con- 3.2. Nitrogen transformation
centration (mg L1)  100% (for the column experiment) or [initial
concentration (mg L1)  final concentration (mg L1) at time t (d)]/ Fig. 3 shows the temporal changes in nitrogen.
initial concentration (mg L1)  100% (for the batch tests). Denitrifi- As was apparent from Fig. 3, TN removed was significantly low-
cation rate (g m3 d1) was calculated as flow rate (m d1)  effective er than NO 3 –N removed at the beginning of operation (up to 10
cross-sectional flow area of the column (m2)  [influent NO 3 –N PVs), which was attributed to the high NH+4–N formation. After
3
(mg L1)  effluent NO 1
3 –N (mg L )]/effective column volume (m ). 10 PVs, TN removed was nearly equal to NO 3 –N removed, which
was a proof that relatively low NH+4–N and NO 2 –N were generated
(Fig. 3). Little NO2 –N formation of <0.05 mg L
1
or even temporary
3. Results and discussion negative formation of <0.02 mg L1 (absolute value) was found
during operation over the 45 PVs (Fig. 3). This suggested that the
3.1. Nitrate–nitrogen removal activities of nitrite reductase were higher than or equal to those
of nitrate reductase, and nitrite reductase quickly adapted itself
Fig. 2 shows the temporal changes in NO 3 –N in samples from to a new environment and refreshed its activity and synthesis
the influent and effluent together with the results of its removal when operating conditions were changed. No NO 2 –N accumula-
efficiency. tion is a desirable attribute because of reducing the risk to public
As shown in Fig. 2, high NO3 –N removal efficiency (>91%) was health. High NH+4–N formation with an average of 4.37 mg L1
maintained relatively constant, and low effluent NO 3 –N was noted during the first 10 PVs mainly due to the wash-out of
(<8.90 mg L1) remained relatively stable before 38 PVs (corre- some ammonium fertilizer applied to the pine bark beforehand,
sponding to 504 d) of operation regardless of influent NO 3 –N and and subsequently low formation of <1.90 mg L1 was observed
flow rate. In addition, NO3 –N removal efficiencies were close to mainly due to the deamination of organic nitrogen from the pine
97–100% at a VNLR of 65.92 g m3 d1 (Fig. S1 of Supporting Infor- bark (Fig. 3) (DeSimone and Howes, 1998; Liu et al., 2014). No mat-
mation). These results indicated that the HAD PRB was effective ter what the pathways are for the NH+4–N generation, the sum of
and efficient in NO 3 –N removal. But removal efficiency sharply NO +
2 –N and NH4–N formed was much less than NO3 –N reduced


dropped down to 82% and effluent NO 3 –N drastically rose up to at all times (Fig. 3), indicating most of the influent NO 3 –N was
18.23 mg L1 at PV 39 (Fig. 2). The drop in removal efficiency biologically transformed into gaseous nitrogen, as expressed by
demonstrated that influent NO 3 –N (103 mg L
1
) and flow rate Eq. (1) (Fernández-Nava et al., 2010).
1
(0.22 m d ) were not the causes of the HAD PRB performance
NO3 ! NO2 ! NOðgÞ ! N2 OðgÞ ! N2 ðgÞ ð1Þ
decrease, because the column was operated under the same oper-

Influent NO3--N 23 mg L -1 48 mg L-1 103 mg L-1

Flowrate 0.15 m d-1 0.22 m d-1 0.30 m d-1 0.15 m d-1 0.22 m d-1 0.30 m d-1 0.15 m d-1 0.22 m d-1 0.30 m d-1
Phase ĉ Ċċ Č č Ď ď Đ đ
120 100
110 90
100
NO3--N percent removal (%)

80
90 In Out Removal efficiency 70
NO3--N (mg L-1)

80
70 60
60 50
50 40
40
Standard for drinkinguse (China) 30
30
20
20
10 10
0 0
0 3 6 9 12 15 18 21 24 27 30 33 36 39 42 45
Number of pore volumes

Fig. 2. Changes in NO


3 –N in the influent and effluent and its removal efficiency over time under variable operating conditions. Vertical dashed lines represent the times when
influent NO 3 –N and/or flow rate was changed.
12 G. Huang et al. / Chemosphere 130 (2015) 8–16

120
0.06
108
0.04
96 0.02
0.00
84
-0.02 1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 33 35 37 39 41 43 45
72 -0.04
N (mg L-1)

NH4+-N formed
60 NO2--N formed

NO3--N removed
48
TN removed
10
36 8
6
24 4
2
12 0
1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 33 35 37 39 41 43 45
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45
-12
Number of pore volumes

Fig. 3. Changes in nitrogen over time. Total nitrogen (TN) (mg L1) = NO
3 –N (mg L
1
) + NO 2 –N (mg L
1
) + NH+4–N (mg L1). NO3 –N or TN removed (mg L
1
) = influent
+
concentration (mg L1)  effluent concentration (mg L1). NO
2 –N or NH4–N formed (mg L
1
) = effluent concentration (mg L1)  influent concentration (mg L1).

(a) 150 150


Influent NO3--N of 23 mg L-1 (b)
Influent NO3--N of 46 mg L-1
125 Influent NO3--N of 103 mg L-1 125
Column height (cm)
Column height (cm)

100 100

75 75

50 50

25 25

0 0
0 12 24 36 48 60 72 84 96 108 120 0.0 0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.2 3.6 4.0
NO3--N (mg L-1) NO2--N (mg L-1)

150 150
(c) (d)
125 125
Column height (cm)

Column height (cm)

100 100

75 75

50 50

25 25

0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 0 1 2 3 4 5 6 7 8 9 10
NH4+-N (mg L-1) DO (mg L-1)
+
Fig. 4. Changes in (a) NO  
3 –N, (b) NO2 –N, (c) NH4–N, and (d) dissolved oxygen (DO) along the column at three influent NO3 –N concentrations of 23 mg L
1
(pore volume (PV)
11), 46 mg L1 (PV 26), and 103 mg L1 (PV 41). Arrows indicate flow direction.

3.3. Spatial nitrogen distribution The changes in NO 3 –N at the influent concentrations of 23 and
46 mg L1 were similar to each other upward through the column:
Fig. 4 shows the spatial changes in nitrogen and oxygen at three almost all the introduced NO 3 –N (P97%) was depleted within the
influent NO 3 –N concentrations of 23 mg L
1
(PV 11), 46 mg L1 (PV first 65 cm; and the resultant low NO 3 –N (<1.50 mg L
1
)
1
26), and 103 mg L (PV 41). approached to a steady state between 65 and 150 cm (Fig. 4a). In
G. Huang et al. / Chemosphere 130 (2015) 8–16 13

(a)
Influent NO3--N 23 mg L-1 48 mg L-1 103 mg L-1

Flowrate 0.15 m d-1 0.22 m d-1 0.30 m d-1 0.15 m d-1 0.22 m d-1 0.30 m d-1 0.15 m d-1 0.22 m d-1 0.30 m d-1
Phase ĉ Ċ ċ Č č Ď ď Đ đ
10 100
9 90

NO3--N percent removal (%)


8 80
7 70
DO (mg L-1)

6 60
5 50
4 In Out Removal efficiency 40
3 30
2 20
1 10
0 0

(b) 500
450 5.0
400
4.0
350
3.0
TOC (mg L-1)

300
2.0
250
1.0
200
0.0
150 0 3 6 9 12 15 18 21 24 27 30 33 36 39 42 45
100
50
0

(c) 10.0
9.5
Standard for drinkinguse (China)
9.0
8.5
8.0
pH

7.5
7.0
6.5
6.0
Standard for drinkinguse (China)
5.5
5.0
0 3 6 9 12 15 18 21 24 27 30 33 36 39 42 45
Number of pore volumes

Fig. 5. Changes in (a) dissolved oxygen (DO), (b) total organic carbon (TOC), and (c) pH in the influent and effluent over time under variable operating conditions. Vertical
dashed lines represent the times when influent NO 3 –N and/or flow rate was changed.

the case of the influent concentration of 103 mg L1, 79% of the maximum of 2.14 mg L1 was detected in the column at the three
influent NO 3 –N was eliminated throughout the column with a high influent concentrations (Fig. 4c). In the spongy iron layer, NO
3 –N
NO 3 –N of 21.44 mg L
1
at 125 cm (Fig. 4a), suggesting that the was not chemically reduced to NH+4–N by the spongy iron, as evi-
denitrifying bacteria present in the column did not have the capa- denced by a small difference in NH+4–N at 125 cm (1.95 mg L1)
city to completely degrade high-concentration NO 3 –N due to a lack and 150 cm (1.88 mg L1) at the influent concentration of
of usable TOC. No NO 2 –N accumulation occurred along the column 103 mg L1 (Fig. 4a and c). Apparently, spongy iron-based chemical
at the influent concentrations of 23 and 46 mg L1 (Fig. 4b). Howev- reduction was not a cause of NH+4–N formation.
er, NO 2 –N sharply accumulated with a peak of 1.21 mg L
1
up to DO quickly decreased in the first 25 cm and afterward slowly
25 cm, and rapidly descended thereafter to less than 0.10 mg L1 declined to lower than 0.75 mg L1 between 25 and 130 cm at
at the influent concentration of 103 mg L1 (Fig. 4b). Contrary to any influent concentration (Fig. 4d). The sharp drop in DO around
our previous batch studies where no NH+4–N was formed when a the influent confirmed that the aerobic heterotrophic bacteria
mixture of pine bark, spongy iron, and mixed bacteria was kept in played a dominant role in oxygen depletion via aerobic respiration
incubation continuously for 105 d (Liu et al., 2014), NH+4–N with a (Eq. (2)) (Della Rocca et al., 2005). In another HAD PRB column,
14 G. Huang et al. / Chemosphere 130 (2015) 8–16

(a) 5.0 100 (b) 5.0 100


NH4+-N formed NO2--N formed
4.5 90 4.5 90
NO3--N removed TN removed
Removal efficiency 4.0 80

NO 3-- N percent removal (%)


4.0 80

NO3--N percent removal (%)


3.5 70 3.5 70

3.0 60 3.0 60

N (mg L -1)
N (mg L-1)

2.5 50 2.5 50

2.0 40 2.0 40

1.5 30 1.5 30

1.0 20 1.0 20

0.5 10 0.5 10

0.0 0 0.0 0
3 6 18 22 30 3 6 18 22 30
Reaction time (d) Reaction time (d)

Fig. 6. Changes in nitrogen and NO3 –N removal efficiency over time by hydrogenotrophic denitrification with the mixed bacteria from (a) 65 and (b) 125 cm of the column,
respectively. Total nitrogen (TN) (mg L1) = NO 3 –N (mg L
1
) + NO
2 –N (mg L
1
) + NH+4–N (mg L1). NO
3 –N or TN removed (mg L
1
) = final concentration (mg L1)  initial
+
1
concentration (mg L ) at time t (d). NO 2 –N or NH4–N formed (mg L
1
) = initial concentration (mg L1)  final concentration (mg L1) at time t (d).

Table 1
Comparison of biological denitrification capacity among different approaches.

Denitrification approach Packing material System description Maximum Maximum References


removal denitrification rate
efficiency (%) (g m3 d1)
HAD Pine bark/spongy iron/sand Double-layer PRB 99c 9.5c This study
HAD Spongy iron/sand/pine bark Media mixed single-layer 99b,c 11.7b,c Liu et al. (2014)
(Column 1) PRB
HAD Spongy iron/sand/pine bark Double-layer PRB 90b,c 8.7b,c Liu et al. (2014)
(Column 2)
Heterotrophic denitrification Iron powder/sodium citrate/ Double-layer PRB 100b 7.5b Liu et al. (2013)
activated
carbon
HAD Cotton/steel wool (R4) Double-layer column 100b 277 Della Rocca et al. (2006)
reactor
a
Heterotrophic denitrification Sawdust/soil Field pilot-scale single-layer NA 1.2 Schipper and Vojvodic-
PRB Vukovic (2000)
Heterotrophic denitrification Sawdust/sand 15-year-old pilot-scale 100b 6 Robertson et al. (2008)
single-layer PRB
Heterotrophic denitrification Cotton Field pilot-scale single-layer 100 360 Soares et al. (2000)
plant
Hydrogenotrophic Steel wool Single-layer column reactor 18%b 5.0b Till et al. (1998)
denitrification
a
No data available.
b
Estimated or calculated based on data given in the article.
c
Estimated or calculated based on NO3 –N transformed into gaseous nitrogen.

spongy iron dominantly contributed to oxygen removal via chemi- By combining the results of Fig. 4a and d, it is therefore conclud-
cal reduction (Liu et al., 2014). Clearly, more CO2 was produced ed that NO3 –N depletion and DO removal were two parallel pro-
during deoxygenation in this study, then providing more inorganic cesses, that is, the presence of oxygen did not limit biological
carbon for hydrogenotrophic denitrifiers (Eq. (3)) (Lee and denitrification. An explanation is that the dense and thick biofilms
Rittmann, 2002). It should be noted that CO2 generated in around the pine bark near the influent as well as their stratified
heterotrophic denitrification (Eq. (4)) (Della Rocca et al., 2005) structures protected the denitrifying bacteria in the inner zone of
was also an inorganic carbon source. the biofilms from oxygen inhibition.
aerobic heterotrophs
2C6 H10 O2 þ 15O2 ! 12CO2 þ 10H2 O ð2Þ 3.4. Inorganic geochemistry

NO3 þ 3:03H2 þ Hþ þ 0:229CO2 Fig. 5 shows the temporal changes in DO, TOC, and pH in sam-
hydrogenotrophic denitrifiers ples from the influent and effluent.
 ! 0:0458C5 H7 O2 N þ 0:477N2
It can be seen from the results that during operation over the 45
þ 3:37H2 O ð3Þ PVs, influent DO fluctuated between 5.84 and 8.57 mg L1, whereas
the corresponding effluent value was in the range of 0.19–
heterotrophic denitrifiers
2C6 H10 O2 þ 6NO3 þ 6Hþ ! 6CO2 þ 10H2 O þ 3N2 1.63 mg L1, indicating the good DO removal efficiency (74–97%
removed) of the HAD PRB in any phase (Fig. 5a). The low effluent
ð4Þ
DO level is thought to be desirable, as high DO may not only cause
G. Huang et al. / Chemosphere 130 (2015) 8–16 15

high biomass buildup, but may also render NO 3 –N regeneration if Even though any approach that had heterotrophic denitrifica-
NH+4–N is subsequently encountered in the downstream aquifer. tion included was capable of attaining the ideal NO 3 –N removal
Despite the fact that the spongy iron did not function efficiently efficiency (P90%), the cotton supported approach presented a
in capturing oxygen (Fig. 4d), it was pretty essential to ensure a much higher denitrification rate than the other ones (including
robust oxygen removal in case of temporary insufficient biological the HAD approach in this study) (Table 1). Nevertheless, the soft
deoxygenation in the upstream pine bark layer (e.g., as a result of texture of cotton limits its in situ applicability on a large scale in
increases in toxicity levels and change in seasonal water that it will be easily and strongly compressed at a high flow rate
temperature). (Soares et al., 2000). The single steel wool mediated
Effluent TOC decreased rapidly from 403.44 mg L1 at PV 1 to hydrogenotrophic denitrification was not able to efficiently
108.46 mg L1 at PV 8, and then declined slowly to 9.34 mg L1 remove the influent NO 3 –N (18% removed) (Table 1). The differ-
at PV 45 at the influent TOC of 62.44 mg L1 (Fig. 5b), which was ences in denitrification performance probably resulted from the
consistent with the literature describing wheat straw-supported differences in hydraulic retention time, diversity of the used inocu-
heterotrophic denitrification (Aslan and Türkman, 2004). The rapid lum, denitrification mechanism, and so on. In spite of the similar
decrease during the early operational stage could be explained due packing media (spongy iron/pine bark/sand) and denitrification
to the rapid decrease in easily soluble, amorphous components of mechanisms, the HADs achieved different removal efficiencies
the cellulose and hemicellulose contained in the pine bark. Organic and denitrification rates (Table 1), demonstrating that packing
carbon excess was also found in other studies (Volokita et al., structure had a noticeable effect on HAD PRB denitrification.
1996b; Aslan and Türkman, 2004; Della Rocca et al., 2006; Liu
et al., 2014). Consequently, how to realize a consistent, slow, and
stable carbon release of cellulose-based solid substrates is worthy
of further study. 4. Conclusions
Effluent pH was 7.07–8.50 at any time point at the influent val-
ue of 7.69–8.37 (Fig. 5c). Accordingly, the effluent pH met the Based on the research findings the main conclusions were as
water quality standard (6.5–8.5 for drinking use) (SBTS, 1993). follows:
The optimum pH for biological denitrification should be kept in
the range of 7.0–9.0 (Tang et al., 2011; Liu et al., 2013). Undoubt- 1. The proposed HAD PRB was effective and efficient in NO 3 –N

edly, the pH in this study was not a cause of lowering the HAD removal. High NO 3 –N removal efficiency (>91%) was main-

PRB denitrification performance after 36 PVs (Fig. 2). tained relatively constant before 38 PVs (corresponding to
504 d). Hydrogenotrophic denitrification (removal efficiency
10–20%) played a minor role in NO 3 –N depletion compared
3.5. Hydrogenotrophic denitrification capacity with heterotrophic denitrification.
2. Most influent NO 3 –N was biologically transformed into gaseous
Fig. 6 shows the temporal changes in nitrogen and NO 3 –N nitrogen. Little NO 1
2 –N formation of <0.05 mg L or even tem-
removal efficiency in incubation bottles by hydrogenotrophic 1
porary negative formation of <0.02 mg L (absolute value) was
denitrification supported by the mixed bacteria from 65 and found during the 45 PVs. Low NH+4–N formation of <1.90 mg L1
125 cm of the column, respectively. was observed after 10 PVs caused by the deamination of organic
NO 3 –N removal efficiency of 10–20% was achieved (Fig. 6a and nitrogen from the pine bark rather than the spongy-iron based
b), indicating that hydrogenotrophic denitrification contributed to chemical reduction.
NO 3 –N removal in the incubation bottles. Besides, considering 3. Almost all the introduced NO 3 –N (P97%) was depleted within
those hydrogenotrophic denitrifying bacteria on the microbores the first 65 cm of the column at the influent NO 3 –N concentra-
of the pine bark and spongy iron media, the hydrogenotrophic tions of 23 and 46 mg L1, whereas 79% of the influent NO 3 –N
denitrification capacity may be further strengthened. Combining was eliminated throughout the column at the influent concen-
Figs. 2 and 6, hydrogenotrophic denitrification played a minor role tration of 103 mg L1.
in NO 3 –N depletion in the HAD PRB compared with heterotrophic 4. Aerobic heterotrophic bacteria played a dominant role in oxy-
denitrification. However, hydrogenotrophic denitrification was gen depletion via aerobic respiration, providing more CO2 for
able to minimize potential gas and biomass clogging, as it con- hydrogenotrophic denitrification. Spongy iron was pretty
sumed cathodic hydrogen formed by anaerobic ZVI corrosion and essential to ensure a robust oxygen removal.
carbon dioxide formed by heterotrophic denitrification (Eq. (3)), 5. Effluent pH (7.07–8.50) met the water quality standard (6.5–8.5
and produced a low biomass yield of 0.24 g cells g1 NO 3 –N, which for drinking use). Effluent DO (0.19–1.63 mg L1) was low and
was less than 0.6–0.9 g cells g1 NO 3 –N typically reported for desirable. The rapid decrease in effluent TOC from 403.44 to
heterotrophic denitrification (Ergas and Reuss, 2001). 108.46 mg L1 during the first 8 PVs was due to the rapid
Hydrogenotrophic denitrification supported with the mixed bacte- decrease in easily soluble, amorphous components of the cellu-
ria from 65 cm showed a higher removal efficiency (17–20%) than lose and hemicellulose.
that from 125 cm (10–15%) (Fig. 6a and b). This phenomenon 6. Even though the packing media and denitrification mechanisms
demonstrated that the hydrogenotrophic denitrifiers developed are similar in the HAD PRBs, packing structure had a noticeable
better in the upstream pine bark layer owing to sufficient inorganic effect on their denitrification performance.
carbon (HCO 
3 + CO2) and NO3 –N substrates. TN removed appeared
invariably a little lower than NO 3 –N removed as a result of the lit-
tle formation of NO 2 –N (60.17 mg L
1
) and NH+4–N (60.25 mg L1)
(Fig. 6a and b).
Acknowledgements

3.6. Comparison of biological denitrification performance among This study is financially supported jointly by NSF – China
different approaches (41172226), National Program of Control and Treatment of Water
Pollution (2009ZX07424-002-002), China Postdoctoral Science
Based on a literature survey, a comparison was made of denitri- Foundation – China (2013M541111) and Beijing Excellent Talents
fication performance among different approaches (Table 1). Program (2012D001055000001).
16 G. Huang et al. / Chemosphere 130 (2015) 8–16

Appendix A. Supplementary material Moon, H.S., Shin, D.Y., Nam, K., Kim, J.Y., 2008. A long-term performance test on an
autotrophic denitrification column for application as a permeable reactive
barrier. Chemosphere 73, 723–728.
Table S1 provides the operating conditions of the column in the Mousavi, S., Ibrahim, S., Aroua, M.K., 2012. Sequential nitrification and
9 phases. Fig. S1 shows the changes in removal efficiency over denitrification in a novel palm shell granular activated carbon twin-chamber
upflow bio-electrochemical reactor for treating ammonium-rich wastewater.
volumetric NO 3 –N loading rate. Supplementary data associated
Bioresour. Technol. 125, 256–266.
with this article can be found, in the online version, at http:// Obiri-Nyarko, F., Grajales-Mesa, S.J., Malina, G., 2014. An overview of permeable
dx.doi.org/10.1016/j.chemosphere.2015.02.029. reactive barriers for in situ sustainable groundwater remediation. Chemosphere
111, 243–259.
Pastén-Zapata, E., Ledesma-Ruiz, R., Harter, T., Ramírez, A.I., Mahlknecht, J., 2014.
References Assessment of sources and fate of nitrate in shallow groundwater of an
agricultural area by using a multi-tracer approach. Sci. Total Environ. 470–471,
Aslan, S., Türkman, A., 2004. Biological denitrification of drinking water using 855–864.
various natural organic solid substrates. Water Sci. Technol. 48, 489–495. Phillips, D.H., 2009. Permeable reactive barriers: a sustainable technology for
Boy-Roura, M., Menció, A., Mas-Pla, J., 2013. Temporal analysis of spring water data cleaning contaminated groundwater in developing countries. Desalination 248,
to assess nitrate inputs to groundwater in an agricultural area (Osona, NE 352–359.
Spain). Sci. Total Environ. 452–453, 433–445. Robertson, W.D., Vogan, J.L., Lombardo, P.S., 2008. Nitrate removal rates in a 15-
Della Rocca, C., Belgiorno, V., Meric, S., 2005. Cotton-supported heterotrophic year-old permeable reactive barrier treating septic system nitrate. Groundw.
denitrification of nitrate-rich drinking water with a sand filtration post- Monit. Remediat. 28, 65–72.
treatment. Water SA 31, 229–236. Rust, C.M., Aelion, C.M., Flora, J.R.V., 2002. Laboratory sand column study of
Della Rocca, C., Belgiorno, V., Meric, S., 2006. An heterotrophic/autotrophic encapsulated buffer release for potential in situ pH control. J. Contam. Hydrol.
denitrification (HAD) approach for nitrate removal from drinking water. 54, 81–98.
Process Biochem. 41, 1022–1028. Sacchi, E., Acutis, M., Bartoli, M., Brenna, S., Delconte, C.A., Laini, A., Pennisi, M.,
Della Rocca, C., Belgiorno, V., Meriç, S., 2007. Overview of in-situ applicable nitrate 2013. Origin and fate of nitrates in groundwater from the central Po plain:
removal processes. Desalination 204, 46–62. Insights from isotopic investigations. Appl. Geochem. 34, 164–180.
Demiral, H., Gündüzoğlu, G., 2010. Removal of nitrate from aqueous solutions by Salvestrin, H., Hagare, P., 2009. Removal of nitrates from groundwater in remote
activated carbon prepared from sugar beet bagasse. Bioresour. Technol. 101, indigenous settings in arid Central Australia. Desalin. Water Treat. 11, 151–156.
1675–1680. SBTS, 1993. Quality Standard for Ground Water (GB/T 14848-93). State Bureau of
DeSimone, L.A., Howes, B.L., 1998. Nitrogen transport and transformations in a Technical Supervision, Beijing.
shallow aquifer receiving wastewater discharge: a mass balance approach. Schipper, L.A., Vojvodic-Vukovic, M., 2000. Nitrate removal from groundwater and
Water Resour. Res. 34, 271–285. denitrification rates in a porous treatment wall amended with sawdust. Ecol.
Ergas, S.J., Reuss, A.F., 2001. Hydrogenotrophic denitrification of drinking water Eng. 14, 269–278.
using a hollow fibre membrane bioreactor. J. Water Supply: Res. Technol.-AQUA Schipper, L.A., Barkle, G.F., Vojvodic-Vukovic, M., 2005. Maximum rates of nitrate
50, 161–171. removal in a denitrification wall. J. Environ. Qual. 34, 1270–1276.
Fernández-Nava, Y., Marañón, E., Soons, J., Castrillón, L., 2010. Denitrification of high Shin, K.H., Cha, D.K., 2008. Microbial reduction of nitrate in the presence of
nitrate concentration wastewater using alternative carbon sources. J. Hazard. nanoscale zero-valent iron. Chemosphere 72, 257–262.
Mater. 173, 682–688. Shomar, B., Osenbrück, K., Yahya, A., 2008. Elevated nitrate levels in the
Ghafari, S., Hasan, M., Aroua, M.K., 2008. Bio-electrochemical removal of nitrate groundwater of the Gaza strip: distribution and sources. Sci. Total Environ.
from water and wastewater – a review. Bioresour. Technol. 99, 3965–3974. 398, 164–174.
Guerin, T.F., Horner, S., McGovern, T., Davey, B., 2002. An application of permeable Shrimali, M., Singh, K.P., 2001. New methods of nitrate removal from water.
reactive barrier technology to petroleum hydrocarbon contaminated Environ. Pollut. 112, 351–359.
groundwater. Water Res. 36, 15–24. Soares, M.I.M., Brenner, A., Yevzori, A., Messalem, R., Leroux, Y., Abeliovich, A., 2000.
Haugen, K.S., Semmens, M.J., Novak, P.J., 2002. A novel in situ technology for the Denitrification of groundwater: pilot-plant testing of cotton-packed bioreactor
treatment of nitrate contaminated groundwater. Water Res. 36, 3497–3506. and post-microfiltration. Water Sci. Technol. 42, 353–359.
Huang, G., Fallowfield, H., Guan, H., Liu, F., 2012. Remediation of nitrate–nitrogen Su, C., Puls, R.W., 2007. Removal of added nitrate in the single, binary, and ternary
contaminated groundwater by a heterotrophic–autotrophic denitrification systems of cotton burr compost, zerovalent iron, and sediment: implications for
(HAD) approach in an aerobic environment. Water Air Soil Pollut. 223, 4029– groundwater nitrate remediation using permeable reactive barriers.
4038. Chemosphere 67, 1653–1662.
Jha, D., Bose, P., 2005. Use of pyrite for pH control during hydrogenotrophic Tang, Y., Zhou, C., Ziv-El, M., Rittmann, B.E., 2011. A pH-control model for
denitrification using metallic iron as the ultimate electron donor. Chemosphere heterotrophic and hydrogen-based autotrophic denitrification. Water Res. 45,
61, 1020–1031. 232–240.
Lee, K., Rittmann, B.E., 2002. Applying a novel autohydrogenotrophic hollow-fiber Till, B.A., Weathers, L.J., Alvarez, P.J.J., 1998. Fe(0)-supported autotrophic
membrane biofilm reactor for denitrification of drinking water. Water Res. 36, denitrification. Environ. Sci. Technol. 32, 634–639.
2040–2052. van Rijn, J., Tal, Y., Schreier, H.J., 2006. Denitrification in recirculating systems:
Levallois, P., Thériault, M., Rouffignat, J., Tessier, S., Landry, R., Ayotte, P., Girard, M., theory and applications. Aquacult. Eng. 34, 364–376.
Gingras, S., Gauvin, D., Chiasson, C., 1998. Groundwater contamination by Volokita, M., Abeliovich, A., Soares, M.I.M., 1996a. Denitrification of groundwater
nitrates associated with intensive potato culture in Qúebec. Sci. Total Environ. using cotton as energy source. Water Sci. Technol. 34, 379–385.
217, 91–101. Volokita, M., Belkin, S., Abeliovich, A., Soares, M.I.M., 1996b. Biological
Liu, H., Jiang, W., Wan, D., Qu, J., 2009. Study of a combined heterotrophic and sulfur denitrification of drinking water using newspaper. Water Res. 30, 965–971.
autotrophic denitrification technology for removal of nitrate in water. J. Hazard. Wang, Q., Feng, C., Zhao, Y., Hao, C., 2009. Denitrification of nitrate contaminated
Mater. 169, 23–28. groundwater with a fiber-based biofilm reactor. Bioresour. Technol. 100, 2223–
Liu, S.J., Zhao, Z.Y., Li, J., Wang, J., Qi, Y., 2013. An anaerobic two-layer permeable 2227.
reactive biobarrier for the remediation of nitratecontaminated groundwater. Westerhoff, P., James, J., 2003. Nitrate removal in zero-valent iron packed columns.
Water Res. 47, 5977–5985. Water Res. 37, 1818–1830.
Liu, F., Huang, G., Fallowfield, H., Guan, H., Zhu, L., Hu, H., 2014. Study on Zhang, Y., Li, Y., Li, J., Sheng, G., Zhang, Y., Zheng, X., 2012. Enhanced Cr(VI) removal
Heterotrophic–Autotrophic Denitrification Permeable Reactive Barriers (HAD by using the mixture of pillared bentonite and zero-valent iron. Chem. Eng. J.
PRBs) for Groundwater In Situ Remediation. Springer, Germany. 85–186, 243–249.
Mendes, M.P., Ribeiro, L., 2010. Nitrate probability mapping in the northern aquifer Zhang, X., Xu, Z., Sun, X., Dong, W., Ballantine, D., 2013. Nitrate in shallow
alluvial system of the river Tagus (Portugal) using Disjunctive Kriging. Sci. Total groundwater in typical agricultural and forest ecosystems in China, 2004–2010.
Environ. 408, 1021–1034. J. Environ. Sci. 25, 1007–1014.

You might also like