Sunteți pe pagina 1din 13

Ecological Engineering 110 (2018) 54–66

Contents lists available at ScienceDirect

Ecological Engineering
journal homepage: www.elsevier.com/locate/ecoleng

Determination of major biogeochemical processes in a denitrifying MARK


woodchip bioreactor for treating mine drainage

Albin Nordström , Roger B. Herbert
Uppsala University, Department of Earth Sciences, Villavägen 16, SE-752 36, Uppsala, Sweden

A R T I C L E I N F O A B S T R A C T

Keywords: At the Kiruna iron ore mine in northern Sweden, mine drainage and process water contain elevated con-
Denitrification centrations of nitrate (NO3−) from the use of ammonium nitrate fuel oil explosives. In order to investigate the
Sulfate reduction treatment capacity of a denitrifying woodchip bioreactor technique for the removal of NO3− through deni-
DNRA trification, a bioreactor was installed at the mine site in 2015 and operated for two consecutive years. Neutral-pH
Woodchip bioreactor
mine drainage and process water containing 22 mg NO3−-N and 1132 mg SO42− (average) was passed through
Temperature
Biogeochemical processes
the bioreactor which was filled with a reactive mixture of pine woodchips and sewage sludge, at treatment
temperatures ranging between 0.8 and 17 °C. At bioreactor temperatures above ∼5 °C, NO3− removal pro-
ceeded to below detection limits (0.06 mg N L−1) without substantial production of nitrite (NO2−), ammonium
(NH4+), nitrous oxide (N2O), or methane (CH4). The relative production of NH4+ and N2O to the NO3− reduced
increased as bioreactor temperatures decreased below ∼5 °C. Based on the resultant changes in alkalinity and
pH from the production of bicarbonate (HCO3−) and carbonic acid (H2CO3), a stoichiometric mass balance
model indicated that denitrification, nitrate reduction to ammonium (DNRA), sulfate reduction, and fermen-
tation were the major biogeochemical processes controlling pH, alkalinity and nitrogen, sulfur and carbon
concentrations in the system. It is suggested that fermentation changed from being mainly butyrate producing to
acetate producing with time, triggering a decline in biogeochemical process diversity and leaving denitrification
as the sole major electron accepting process.

1. Introduction that is unavailable for uptake by most living organisms. The stoichio-
metric definition of net denitrification is exemplified in reaction (1)
Ammonium nitrate (NH4NO3) is the most common explosive used in using glucose (C6H12O6) as the carbon substrate/electron donor
the mining industry (Forsyth et al., 1995) and is highly soluble in water. 5 1 1 1
Due to spillage during storage, transport, and loading of the explosives, NO3− + C6 H12 O6 → N2 + HCO3− + H2 CO3 + H2 O
24 2 4 2 (1)
as well as the incomplete detonation of the explosives (Revey, 1996),
NH4NO3 dissolves in mine drainage and process water and is eventually Denitrification occurs under the sequential reduction of NO3−,
discharged to the environment, primarily in the form of nitrate (NO3−) NO2−, NO, and N2O to N2, and is known to be “leaky”, such that in-
(Lindeström, 2012). Excess release of NO3− to aquatic ecosystems can termediate NOx species may accumulate.
be hazardous as it may induce eutrophication leading to hypoxia, or Woodchip bioreactors have been identified as effective technologies
may be transformed into ammonium (NH4+) or ammonia (NH3). NH4+ for the removal of NO3− from contaminated water (e.g. Moorman et al.,
is more easily incorporated into biomass than NO3− (Rittmann and 2010; Robertson and Merkley, 2009; Warneke et al., 2011a). For a
McCarty, 2001), leading to increased risks of eutrophication, and may general overview of the woodchip bioreactor technology, the reader is
further be transformed into NO2−/NO3− through nitrification, an refered to Schipper et al. (2010). In terms of the reactive material in the
oxygen consuming reaction, both leading to increased risks of hypoxia bioreactor, wood products (e.g. sawdust and woodchips) have in-
in aquatic ecosystems. NH3 is additionally toxic to aquatic biota at high creasingly been used as an electron source for denitrification due to the
concentrations (EPA, 2013). high content of carbon (electron donor) available at low cost (Schipper
Denitrification provides the possibility of completely transforming et al., 2010), and due to the longevity of the material (Moorman et al.,
NO3− into nitrogen gas (N2), a comparatively inert form of nitrogen (N) 2010; Robertson and Merkley, 2009; Robertson, 2010; Warneke et al.,
2011a).


Corresponding author.
E-mail addresses: albin.nordstrom@geo.uu.se (A. Nordström), roger.herbert@geo.uu.se (R.B. Herbert).

http://dx.doi.org/10.1016/j.ecoleng.2017.09.018
Received 29 June 2017; Received in revised form 25 September 2017; Accepted 29 September 2017
0925-8574/ © 2017 Elsevier B.V. All rights reserved.
A. Nordström, R.B. Herbert Ecological Engineering 110 (2018) 54–66

In this study, a pilot-scale woodchip bioreactor was installed in the emerging at 0.9 m from W5 (in the direction of flow, see Fig. 1a). The
subarctic climate at the Kiruna iron ore mine (northern Sweden) with elevation at the top of the vertical drainage pipe was adjustable and
the purpose of removing NO3− in mine drainage and process water regulated the water level in the bioreactor. The distribution of water
originating from the use of ammonium-nitrate fuel oil explosives over the width of the bioreactor at the inlet, and the broad-crested weir
(ANFO). To the best of our knowledge, this is the first woodchip bior- near the outlet, were both efforts to minimize preferential flow and
eactor that has been tested for the removal of NO3− from mine water. reduce the advective velocity within the bioreactor; these are two fac-
The objectives of the study were to determine the predominating tors that have been shown to maximize contact between the treated
biogeochemical processes controlling pore water composition in the solution and the reactive material (Herbert et al., 2014; Nordström and
system. Furthermore, a mass balance approach was applied to nitrogen Herbert, 2017).
and sulfur in order to calculate changes in the carbonate speciation, and The bioreactor featured 20 observation points referred to as
hence pH, based on reaction stoichiometry. A1–A20. A1 and A20 was the inlet and the outlet, respectively. A2-A19
were PVC pipes attached in different configurations on the inner walls
2. Material and methods of the bioreactor, and of two different designs used for pore water
sampling. A3, A7, A11, A15, and A18 were sampled according to the
2.1. Study site BAT sampling technique (see Torstensson, 1984). The remaining PVC
pipes ended in a 10 cm screen covered in a polyester plastic net filled
This study was conducted at the Kiruna iron ore mine, operated by with medium grained sand that worked as a filter to prevent smaller
the mining company Luossavaara-Kiirunavaara Aktiebolag (LKAB), and particles from passing through the well screens.
located in the subarctic climate of northern Sweden, Kiruna (67°51′N
20°13′E). The mean annual air temperature in the region of the study 2.3. Bioreactor material
site was −3 °C for the period 1961–1990 (SMHI, 2017).
Mine drainage and process water are released through a clarifica- Previous column studies have shown that a mixture of pine wood-
tion pond into the recipient Mettä Rakkurijärvi lake that drains through chips and sewage sludge created a suitable biogeochemical environ-
the Rakkuri river system into the Kalix River (LKAB, 2016), and ulti- ment for the promotion of denitrification at temperatures relevant at
mately to the Baltic Sea. The ore and waste rock has a low sulfide the study site (see Nordström and Herbert, 2017). Decorticated pine
mineral content, and hence the pH of the site drainage is close to woodchips (porosity 0.54, estimated in the laboratory) were retrieved
neutral at 8.05 ± 0.04. The water has relatively high concentrations of from a nearby sawmill, and digested sewage sludge was shipped from
nitrogen (N) from the use of ammonium-nitrate fuel oil (ANFO) ex- the Uddebo waste water plant, Luleå, northern Sweden. A digested form
plosives (LKAB, 2016). In 2015, ∼5,7 Mm3 of mine and process water of sewage sludge was used as a source of denitrifying bacteria and was
was released into the Mettä Rakkurijärvi lake, with an average NO3− selected in preference of activated sewage sludge as it was easier to
concentration of 27.8 mg N L−1 (LKAB, 2016). transport and handle.
The bioreactor in this study was constructed near the point where Woodchips and sewage sludge were deposited in layers when filling
water was discharged from the clarification pond to the recipient. the bioreactor. In total, 2.5 m3 of digested sewage sludge and ∼210 m3
Construction took place during May and June 2015. of decorticated pine woodchips were added to the bioreactor, yielding a
sludge-woodchip ratio of 1:84 on a volume basis. The woodchips were
2.2. Bioreactor design emplaced and gently compressed using an excavator, while the sewage
sludge was deposited manually. A sewage sludge slurry (sludge:water
The bioreactor was designed as a subsurface system: a ∼1.1 m deep volume ratio ∼1:10) was poured over the woodchips in layers located
trench with trapezoidal cross-section was excavated and enclosed by 50 cm and 100 cm above the geomembrane (S1–S2; Fig. 1c). Approxi-
∼1 m high mounds of waste rock material originating from the Kiruna mately 10 Liters of slurry were added per square meter of the woodchip
iron ore mine. The total depth was 2.1 m. At the ground surface, the surface. In addition, sludge (not slurry) was dispersed on the wood chip
bioreactor was 44 m long and 6.65 m wide (see dimensions in Fig. 1a surfaces in the first three compartments (C1–C3, Fig. 1a). Note that no
and b). The bottom of the trench was covered with a 1.5 mm thick sewage sludge (or slurry) was added to compartment C6 in the bior-
impermeable geomembrane (high-density polyethylene, HDPE) serving eactor.
as an impermeable boundary and preventing the leakage of water to the The accumulated thickness of the layers of pine woodchips and
surrounding soil. sewage sludge in C1, C5, C6 and C2–C4, was ∼1.7 m and 1.2 m, re-
Five “inner walls” of plywood (W1–W5, Fig. 1a and b) were con- spectively. The mixture of pine woodchips and sewage sludge was then
structed and placed at 8.5 m intervals, demarcating six compartments covered with a geotextile as to minimize the contact with the atmo-
(C1–C6, Fig. 1a), with the purpose of forcing water flow below the sphere. In compartments C2–C4, 0.9 m of glacial till (primarily silty-
surface of the bioreactor (i.e. avoiding surface flow). The inner walls sand with detrital organic matter, but also with larger cobbles < 300
were fixated by their own weight and by the additional weight of steel mm) was placed, and compressed, on top of the geotextile using an
I-beams that were placed on top of the walls (Fig. 1b). excavator. The purpose of the till layer was primarily for forcing the
W1–W4 measured 1.2 m in height, extending from approximately flow to pass through the more permeable woodchip material and
0.1 m above the bioreactor surface to 1 m above the trench floor minimize gas exchange (O2, CO2, N2O, CH4) between the bioreactor
(Fig. 1b). W5 covered the entire cross section of the bioreactor and interior and the surrounding atmosphere. The till layer also decreased
featured a rectangular weir (Fig. 1b) with an adjustable cross-sectional the infiltration of precipitation into the permeable woodchip material,
area intended for regulation of the water table in the bioreactor. which would lead to a dilution effect. The geotextile underlying the till
Water from the clarification pond was delivered to the bioreactor prevented the till from migrating into the woodchip material and
using a submersible pump placed at the clarification pond outlet, and a thereby risking a decrease in permeability. During the placement of the
ball valve was used to adjust the input flow rate (Q) to the bioreactor glacial till, the bioreactor was saturated with water from the clarifica-
system. The inlet water from the clarification pond was spread across tion pond at a flow rate of 2 L s−1.
the width of the bioreactor using a perforated PVC drainage pipe so that
flow would be distributed as evenly as possible over the width of the 2.4. Bioreactor operation
bioreactor.
The outlet was constructed as a basal drain connected to a polyvinyl After the saturation of the bioreactor system, the pore water was
chloride (PVC) pipe extending to the surface of the bioreactor, slowly recirculated for 10 days (cf. Nordström and Herbert, 2017) to

55
A. Nordström, R.B. Herbert Ecological Engineering 110 (2018) 54–66

Fig. 1. Construction and instrumentation of


the bioreactor. The longitudinal and cross-
sectional dimensions of the bioreactor
trench are shown in (a) and (b). The di-
mensions of the five inner walls (W1-W5)
are shown in (b). The material layering is
shown in (c).

allow for the development of the microbial community. After this 2.5. Sampling and analysis
period, water was pumped to the bioreactor inlet and the outlet was
opened. The bioreactor was thereafter operated for two consecutive In this study, data collected from the inlet (A1), outlet (A20), and
field seasons. The first operational year extended from 22nd of June the observation points located along the bottom middle profile in the
2015 until the 20th of November 2015 when air temperatures dropped bioreactor (A2, A6, A10, A14, A17) were used for analysis.
to below −20 °C. The second operational year commenced on the 9th
of May 2016 when air temperatures approached 0 °C, and the bior- 2.5.1. Water sampling
eactor was operated until the 21st of October 2016. Flow was stopped Water samples were collected from the inlet and outlet approxi-
to the bioreactor during the winter months because the piping system mately two times a week. Once a month, an extended sampling took
leading to the bioreactor would have otherwise frozen. Indeed, tem- place where bioreactor pore water was also sampled. Pore water was
perature measurements in the bioreactor (see below) indicated that sampled using a peristaltic pump. Before sampling the porewater, ∼2 L
temperatures were well above 0 °C in the active regions of the bior- of water was purged from the sampling tube to ensure the collection of
eactor (1–2 m depth) during the entire period June 2015–October a representative sample (see Appelo and Postma, 2005, p.12). Samples
2016. intended for analysis of NO3−, NO2−, NH4+, pH, alkalinity, total or-
The hydraulic residence time (HRT) in the bioreactor system was ganic carbon (TOC), and other inorganic anions were collected in clean
adjusted several times during the operation of the bioreactor (see 2 L PVC containers that were refrigerated at 4° +/− 2 °C within two
Table 1) by changing the pump discharge Q. Water flow to the bior- hours of collection. The chemical analyses on the water samples were
eactor was varied between 43.2 and 109.6 m3 d−1, with 43.2 m3 d−1 performed by LKAB’s accredited laboratory services, following stan-
being the most common flow rate used. dardized procedures as presented in table S1.1 (see supplementary
materials).
Whenever sample analysis could not be guaranteed within the
Table 1 standard time for analysis, samples were sent by ground transport to the
Theoretical hydraulic residence times (HRT) calculated by dividing the estimated pore
volume with the pump flow rate (Q).
accredited laboratory ALS Scandinavia in Luleå for analysis, which in-
cluded ∼1 day of additional storage at room temperature during
Days Q [m3 day−1] Theoretical HRT [days] transport.

−10-0 14.0 Recirculation


0–45 43.2 2.6
2.5.2. Gases
46–52 0 No flow – Pump malfunction After water sampling, a 50 mL sample was injected into a pre-
53–151 59.4 1.9 sealed, non-evacuated (air-filled), 100 mL serum vial containing 1 mL
152–321 0 No flow – winter intermission ZnCl2 (50% w/v). Zinc chloride was used as a preservative and was
322–370 43.2 2.6
prepared from milli-Q water and anhydrous, reagent grade (≥98%)
371–428 48.9 2.3
429–490 109.6 1 ZnCl2 (Redi-Dri™). In total 130 dissolved gas samples were collected,
whereof 33 were technical replicates.

56
A. Nordström, R.B. Herbert Ecological Engineering 110 (2018) 54–66

The dissolved gas samples were shipped from Kiruna to Uppsala via Generally, at temperatures above 5 °C and low flow (Table 1), the in-
ground transport (4-7 days), and stored at room temperature (22 °C) for coming NO3− is removed to below the detection limit (0.06 mg N L−1).
24–72 h for equilibration. The headspace in the 100 mL serum vials was Effluent NO2− concentrations were generally lower than the influent
sampled using a 60 mL luer lock syringe equipped with a stop cock and NO2− concentrations, with the exception of days 1–7 following the
a hypodermic needle and injected into a pneumatically pre-sealed, non- start-up of the bioreactor and during the second operational year during
evacuated (air-filled), 22 mL PerkinElmer™ glass vial, so that the at- periods of lower temperature (< 5 °C) and high flow (i.e. after day
mosphere was replaced in the vial. Sampled headspace gas was ana- 429). Higher NO2− concentrations are observed in the effluent under
lyzed for N2O and CH4 on a gas chromatograph equipped with an these conditions because denitrification kinetics are either slower (i.e.
electron capture detector (Clarus 500 GC, Perkin Elmer, CT, USA). at low temperature) or the HRT is shorter than the time required for
Surface gas fluxes (N2O, CH4) were measured using the static gas complete NO2− removal (i.e. high flow).
chamber technique, with the chambers constructed with reference to the Effluent NH4+ concentrations generally surpassed influent con-
guidelines provided by Parkin and Venterea (2010). The chambers (60 L) centrations, with the average ingoing and outgoing NH4+ concentra-
were constructed in polyethylene plastic and equipped with water locks, tions ranging from below detection limit (0.015 mg N L−1) to
vent tubes, fans, and thermometers. The chamber anchors were perma- 0.4 mg N L−1. During the first operational year, NH4+ production was
nently installed, and when not in use, the anchors were covered with intermittent; concentrations were of greater magnitude than during the
geotextile. Samples were withdrawn from the headspace in the top of second year when effluent NH4+ concentrations remained compara-
chambers at 0, 15, 30, 45, and 60 min after emplacement using a 60 mL tively stable and always greater than influent concentrations (Fig. 2c).
syringe and a hypodermic needle. The collected sample was injected into a Hence, the bioreactor was a net source of low levels of NH4+.
non-evacuated (air-filled), 22 mL PerkinElmer™ glass vial and shipped and Influent and effluent sulfate concentrations were on average 1132
analyzed together with dissolved gas samples (see above). and 1083 mg SO42− L−1, respectively. Based on the difference between
influent and effluent SO42− concentrations, there was a clearly-notable
2.6. Determination of gas concentrations and surface gas emissions loss of SO42− from the system between days 25–63 when effluent
SO42− concentrations were consistently lower than influent con-
Dissolved gas concentrations were determined via headspace equi- centrations (Fig. 2d). In addition, a strong odor of H2S was noted during
librium using Henry’s law and the Bunsen coefficient which was cal- the entire first operational year, indicating sulfate-reducing conditions
culated and adjusted to the increased pressure (2 atm) in the sampling in the bioreactor:
vials for dissolved gases (see Section 2.5.2), based on Wilhelm et al. 1
(1977) and Breibarth et al. (2004). The R (R Core Team, 2017) package SO42 − + C6 H12 O6 → H2 S + 2HCO3−
3 (2)
quantchem (Komsta, 2012a, 2012b) was used produce the calibration
curves and to account for heteroscedasticity and non-linearity. Tech- SO42− reduction is also noted between days 378 and 409 the second
nical replicates showed a significant correlation with the original year, but at a lesser magnitude. Unfortunately, analyses for H2S were
samples (see Fig. S1.1, supplementary material). Surface gas fluxes not performed during the study. Since the analytical uncertainty in
were calculated using the R package HMR (see Pedersen, 2015), ac- SO42− concentrations was ± 15% (LKAB, pers. comm.), the exact level
counting for nonlinearity in the change in gas concentration in the of H2S production cannot be calculated based solely on differences in
chamber over time. This resulted in six surface gas flux measurements SO42− concentrations.
for each sampling occasion, and the average of these were used as re- Effluent pH was consistently lower than influent pH with the lowest
presentative of the surface gas flux for the entire bioreactor. pH observed following the start-up of the bioreactor (Fig. 2e), during
the first 100 days of operation. After start-up, the effluent pH increased
2.7. Determination of the N2O emission factor and stabilized with time, varying within the neutral to slightly basic pH
range. Effluent alkalinity was consistently greater than the influent al-
The N2O emission factor is here defined as the ratio of emitted N2O kalinity (on average 44.4 mg L−1) except for days 1–28 following the
to the NO3− removed. The total mass of emitted N2O-N was divided by start-up of the bioreactor when the system was alkalinity-consuming
the total mass of removed NO3−-N on a daily basis, for each of the (see Fig. 2f). Hence, the system went from alkalinity-consuming to al-
sampling occasions, yielding the N2O emission factor. The total mass of kalinity-producing at the same time as sulfate-reducing conditions were
emitted N2O-N was calculated as the sum of dissolved and surface gas initiated during the first year of operations (see Fig. 2d and f). Except
emissions of N2O. The emitted N2O-N mass in gas phase was de- for the high alkalinity observed on day 322 (890 mg L−1) following the
termined from N2O-N surface emission rates multiplied with the bior- winter intermission, maximum alkalinity concentrations were similar
eactor surface area, under the assumption that there were no gas flux for both years of operation. As the HRT was decreased to 24 h on day
from the areas of the bioreactor covered by the glacial till. The emitted 429 (Table 1), there was an immediate decrease in alkalinity because of
N2O-N mass in dissolved phase was determined by multiplying ob- the lower removal of NO3−.
served dissolved N2O-N effluent concentrations with the input flow
rate. The daily total removal of NO3−-N was calculated as the differ- 3.2. Concentration changes along profiles
ence between the inlet and outlet concentrations for each profile sam-
pling occasion multiplied with the input flow rate. Fig. 3 shows the change in NO3−, NO2−, N2O(aq), and NH4+ con-
centrations as observed in the pore water between the inlet and outlet
3. Results and discussion on eleven different occasions during the operational period of the
bioreactor. Overall, NO3− is the major ionic nitrogen species in the
3.1. Influent and effluent concentrations system, with exceptions being the relatively high NO2− and NH4+
concentrations in the beginning of the bioreactor operations (Fig. 3b, d,
Time series of influent and effluent nitrate, nitrite, ammonium, pH, and f). During the first year, NO3− removal to below the detection limit
alkalinity, TOC, and sulfate concentrations are shown in Fig. 2. was completed in close proximity to the bioreactor inlet, whereby
Fig. 2a–c shows that NO3− is the major nitrogen species in the incoming conditions became especially conducive for sulfate reduction (see
mine water. Over the two year operational period, the average NO3− above).
concentration in the influent and effluent waters were 22.0 and The sequential production of NO2− after NO3− production shows
5.4 mg N L−1, respectively, with effluent concentrations ranging be- that NO2− is the product of microbiological NO3− reduction. In the
tween the detection limit (0.06 mg N L−1) and 23.3 mg N L−1 (Fig. 2a). initial three profiles from the first year, the rate of change in NO3−

57
A. Nordström, R.B. Herbert Ecological Engineering 110 (2018) 54–66

Fig. 2. Time series of effluent (open sym-


bols) and input (full symbols) concentra-
tions of nitrate (a), nitrite (b), ammonium
(c), sulfate (d), pH (e), alkalinity (f), and
TOC (g). Note broken x-axis, separating the
2015 data from 2016 data. Daily mean
temperature in (a) as full line with no sym-
bols.

concentration (with distance from inlet) is comparatively high and Dissolved N2O concentrations are consistently low in comparison to
occurring simultaneously as peaks in NO2− concentrations are ob- the magnitude of NO3− reduction, but the concentrations increase to-
served. Elevated NO2− concentrations are often observed when glu- ward the outlet of the bioreactor, indicating that N2O is the product of
cose, a fermentable carbon substrate, is used by denitrifiers as the sequential denitrification. The observed low concentrations of dissolved
carbon substrate in comparison to less fermentable carbon substrates, N2O imply that N2O is fully reduced to N2 in the system, that N2O
e.g. acetate (e.g. Akunna et al., 1994, 1993; Wilderer et al., 1987). The leaves the system through surface outgassing, or that N2O is never
decrease in observable NO2− in the bioreactor with time (Fig. 3) could produced due to an accumulation of NO. In denitrifying woodchip
be the result of a change in the primary carbon substrate from a fer- bioreactors, N2O emissions are mainly in dissolved form (e.g. Warneke
mentable to a less fermentable carbon substrate (e.g. from glucose to et al., 2011b). Further, any major NO accumulation from denitrification
acetate) used by the denitrifiers in the system. For example, Wilderer is unlikely based on previous studies where low emissions of NO from
et al. (1987) concluded that during fermentative conditions when glu- denitrification have been observed (e.g. Betlach and Tiedje, 1981;
cose was supplied as a carbon substrate, the growth of bacteria capable McKenney et al., 1982). Therefore, except for the observed concentra-
of only reducing NO3− to NO2− was promoted, as opposed to those tions of N2O in the bioreactor, any N2O produced by sequential deni-
capable of further reduction to N2O and N2, such that the increase in trification is assumed to be fully reduced into N2.
NO2− concentrations was due to unbalanced rates of NO3− and NO2− Fig. 4 shows the change in TOC, alkalinity, and pH in the pore water
reduction (e.g. Betlach and Tiedje, 1981). between the inlet and the outlet at the same eleven occasions as the

58
A. Nordström, R.B. Herbert Ecological Engineering 110 (2018) 54–66

Fig. 3. Nitrate, nitrite, ammonium, and ni-


trous oxide concentration profiles sampled
on days 330 (a), 1 (b), 365 (c), 22 (d), 400
(e), 57 (f), 428 (g), 85 (h), 456 (i), 113 (j),
and 477 (k). Profiles in the same row were
sampled at similar temperatures in the
bioreactor. Profiles in the left panel column
in Fig. 3 are from the first operational year
(days 1–150), while the profiles in the right
panel column are from the second opera-
tional year (days 322–486); profiles in the
same row were sampled at similar tem-
peratures (i.e. the same time of the year).
Note that N2O-N was not determined for (b)
and (d).

nitrogen profiles in Fig. 3. The observed TOC concentrations in the pore 3.3. Surface gas emissions
water samples show that the TOC concentrations were generally in-
creasing in a non-linear relationship toward the outlet of the bioreactor Fig. 5a and b show the N2O(g) and CH4(g) surface emission rates
(see Fig. 4). This nonlinear increase in TOC is most evident in the determined for each occasion of profile sampling.
profiles from the first year of operations (left side graphs in Fig. 4) when The average observed N2O(g) surface emission rate for the full dura-
NO3− removal to below detection limits occurred near the inlet, leaving tion of the bioreactor operations was 1.75 ± 0.97 μg N m−2 min−1. The
sulfate reduction as the dominating electron-accepting process. Fol- emission rates appear to negatively correlate with the bioreactor tem-
lowing the bioreactor start-up (day 1), alkalinity was fully consumed in perature (see Fig. 2a), with the highest emission rates observed at the
the pore water (Fig. 4b). With time, the zone of net alkalinity con- lowest temperatures and the lowest emission rates observed at the highest
sumption gradually moved towards the outlet (Fig. 4) as the system temperatures (Fig. 5a).
became net alkalinity producing (see above). Bioreactor pH changed The largest CH4(g) surface emission rates (average 20.81 ± 34.67
simultaneously with alkalinity. The initial pH decrease in the reactor is [μg m−2 min−1]) were observed in the beginning of the bioreactor
substantial with a decrease in ∼4.5 pH units within the first 5 m of the operations (during the start-up phase). Dissolved CH4 were generally
reactor following the start-up (Fig. 4b), where after pH gradually in- low (< 1 mg L−1), and generally increased toward the outlet of the
creased in the bioreactor (Fig. 4b, d, f, h). bioreactor, and decreased with time from bioreactor start-up (see Fig.

59
A. Nordström, R.B. Herbert Ecological Engineering 110 (2018) 54–66

Fig. 4. Alkalinity and TOC concentration,


and pH, profiles sampled on days 330 (a), 1
(b), 365 (c), 22 (d), 400 (e), 57 (f), 428 (g),
85 (h), 456 (i), 113 (j), and 477 (k). Profiles
in the same row were sampled at similar
temperatures in the bioreactor. Profiles in
the left panel column in Fig. 3 are from the
first operational year (days 1–150), while
the profiles in the right panel column are
from the second operational year (days
322–486); profiles in the same row were
sampled at similar temperatures (i.e. the
same time of the year).

S2.1, supplementary material). CH4 is the product of methanogenesis, related to the magnitude of NO3− removal (Fig. 5c). Generally, when
and may occur through two general pathways that use H2 and CO2 or the NO3− removal in the bioreactor was less than 10% in the bior-
acetate as electron donors (Appelo and Postma, 2005; Muyzer and eactor, all the reduced NO3− left the bioreactor as N2O (Fig. 5c), which
Stams, 2008) as represented by reactions (3) and (4) below: coincided with increased production of NH4+ and low temperatures
(< 5 °C) in the bioreactor (see Fig. 2c and a). N2O was predominantly
CO2 + H2 → 2H2 O + CH4 (3)
leaving the system in dissolved form, with an average of 93.8 ± 3.4%
CH3 COO− + H2 O → CH4 + HCO3− (4) (n = 8) of the emitted N2O being dissolved. The largest percentage N2O
emissions were observed on days 329 and 476 where they were de-
The observed decline in methanogenesis with time in this study (cf. termined as 139 and 111%, respectively, of the reduced NO3−. More
Fig. 5b) occurs simultaneous as an increase in pH and could be related than 100% production of N2O is most likely due to measurement errors,
to a decline in CO2/H2 production from fermentation with time (see or analytical uncertainty, as the nitrate removal was very low on these
discussion below). days (see Fig. 2a).
Fig. 5c shows the N2O emission factor as a function of the total
NO3− removed in the bioreactor, calculated on a daily basis using in- 3.4. Nitrogen transformation processes
fluent and effluent concentration data (see Section 2.7). The N2O
emission factor relative to the amount of NO3− reduced was mainly According to our general understanding of nitrate removal in anoxic

60
A. Nordström, R.B. Herbert Ecological Engineering 110 (2018) 54–66

systems with high abundance of organic carbon (Burgin and Hamilton,


2007; González-Cabaleiro et al., 2015a).
The results of this current study support these previous findings: the
generally high level of NO3− removal in parallel with a relatively low
production of NO2−, N2O and NH4+, along with alkalinity production
(see above), strongly suggest that denitrification is the primary path for
nitrate removal. Overall, effluent alkalinity varied with bioreactor
temperature and HRT, which was due to the dependence of the deni-
trification kinetics on temperature and HRT (as HRT affects the time
available for reaction and thus the completeness of the reaction; see e.g.
Nordström and Herbert, 2017). This suggests that as the effluent pH and
alkalinity stabilized during the second year of operations, the sig-
nificance of denitrification as the sole major terminal electron-ac-
cepting process (TEAP) in the bioreactor increased with time. This oc-
curred primarily because of a suppression of secondary TEAPs such as
sulfate reduction and DNRA with time, which were more prominent
during the first operational year than during the second year. Never-
theless, ammonium production by DNRA is not an insignificant process.
The net stoichiometric definition of DNRA is shown in reaction (5)
using glucose as the carbon substrate/electron donor
1
NO3− + C6 H12 O6 + H2 O → NH4+ + 2HCO3−
3 (5)

In DNRA, NO3− is sequentially reduced to NO2−, and in turn to


NH4+/NH3. The DNRA pathway represents an undesirable pathway for
NO3− removal from an environmental perspective as NH4+ oxidation is
an oxygen-consuming process. The production of NH3 in the bioreactor
was insignificant as pH values were well below the pKa of NH4+ (9.3).
In sulfidic environments, or when there is a high availability of carbon
with respect to the available NOx (high C:N ratios), the significance of
DNRA in NOx cycling increases (see Burgin and Hamilton, 2007). The
type of carbon substrate may also affect the relative prevalence of
DNRA. For example, Akunna et al. (1993) observed significant pro-
duction of NH4+ relative to the amount of NO3− reduced when fer-
mentable carbon substrates (glucose and glycerol) were used in com-
parison to less fermentable carbon substrates (acetic acid, lactic acid,
methanol). It is likely that the relatively increased effluent NH4+ con-
centrations during the first year of operations (see Section 3.1) were
due more fermentable carbon substrates being used by the NOx cycling
microbial community, which is consistent with the discussion above on
Fig. 5. N2O (a) and CH4 (b) gas emissions from the bioreactor surface [μg m−2 min−1] NO2− in the profile samples (Section 3.2).
represented as an average of the gas flux determined in six spatially separated chambers. NH4+ may, in addition to DNRA, be produced by aerobic miner-
Error bars represent one standard deviation. For day 56, the average estimated CH4 alization of organic matter (MOM). The stoichiometric definition of
ranged between 99.4 ± 316.9 [μg m−2 min−1]. Figure (c) shows the percentage of re-
MOM is exemplified in reaction (6) with C5H7O2N, an empirical com-
moved total NO3− emitted as N2O determined on a daily basis. Note log scale on axes in
(c). position of a prokaryotic cell (from Rittmann and McCarty, 2001), as
the electron donor
5O2 + C5 H7 O2 N + H2 O → NH4+ + HCO3− + 4H2 CO3 (6)

systems (e.g. Burgin and Hamilton, 2007), the processes that may be re- MOM is an aerobic process where biomass oxidation results in the
sponsible for the removal of NO3− in the bioreactor system are (1) deni- release of NH4+, while DNRA is an anaerobic process that produces
trification, (2) dissimilatory nitrate reduction to ammonium/ammonia NH4+ under the simultaneous reduction of NO3−/NO2−. On a molar
(DNRA), (3) anaerobic ammonium oxidation (ANAMMOX), and/or (4) basis, the observed production of NH4+ relative to the removal of NO3−
biomass assimilation. In previous studies on woodchip bioreactors (Elgood was on average 3.8%, leaving denitrification as the main process for
et al., 2010; Greenan et al., 2009, 2006; Moorman et al., 2010; Robertson reductive removal of NO3−. However, NH4+ concentrations deviated
and Merkley, 2009; Schipper et al., 2010; Warneke et al., 2011a,b), de- considerably from the average pattern during days 1–7 following the
nitrification has been determined as the main pathway for NO3− removal start-up of the bioreactor. During that time, relatively high NH4+
based on observed N2O production, enrichment of the 15N isotope, and the concentrations were observed to occur simultaneously with compara-
generally increased abundances of functional genes pertaining to deni- tively low effluent concentrations of NO3− and NO2− (∼8–9% of re-
trification. The comparatively low production of NH4+ relative to the duced NO3−, Fig. 2a–c), suggesting DNRA. In addition, if MOM was
amount of reduced NO3− have excluded DNRA as a major process for significant in the system, it would be expected that there would be a
NO3− removal in other studies. Further, NO3− immobilization by assim- relative increase in NH4+ close to the bioreactor inlet due to the pre-
ilation in biomass has been inferred to be a minor process, estimated to be ference of O2 over NO3− as an electron acceptor (Appelo and Postma,
within the ranges of ∼2.0-3.5% of the total NO3− removed (cf. Greenan 2005, p. 439) as the oxygenized inlet water enters the bioreactor. This
et al., 2009, 2006; Schipper et al., 2010). Additionally, ANAMMOX is is, however, not observed in any of the profiles sampled (Fig. 3). The
likely not a major pathway for NO3− removal in bioreactor systems increase in NH4+ occurs closer to the outlet, which is most evident from
(Herbert et al., 2014; Schipper et al., 2010), mainly as it suppressed in Fig. 3f, as NO3− concentrations were low, suggesting that NH4+

61
A. Nordström, R.B. Herbert Ecological Engineering 110 (2018) 54–66

production was from DNRA. The sulfidic environment and the low organic carbon and its consecutive consumption by heterotrophic
NO3− concentrations in the bioreactor, during the first year of ob- electron-accepting processes (see table S3.1, Supplementary materials).
servations (Fig. 3), would have resulted in conditions favorable for This is consistent with the observations made following the start-up of
DNRA (see above). Note, however, that the NH4+ production is clearly the bioreactor when comparatively low pH, low alkalinity, and high
suppressed during the period of clear SO42− reduction between days concentrations of TOC were observed in the bioreactor effluent (see
25–63 (i.e. based on differences in influent and effluent concentrations, Fig. 2e–g). It should be noted that the initial increase in effluent TOC
see Fig. 2c). In addition to using SO42− as the electron acceptor, sulfate concentrations is commonly seen in woodchip bioreactors, and is
reducers can also reduce NO3−/NO2− to NH4+ through DNRA (see usually explained as a “flush-out” of soluble organic carbon (e.g.
Dalsgaard and Bak, 1994; López-Cortés et al., 2006). This suggests that Hoover et al., 2016). However, the observed drop in pH and alkalinity
the sulfate reducers in the system alternated their metabolism between consumption in this study suggest a fermentative signature on the TOC
sulfate reduction and DNRA. During periods of low temperature and production. Alternatively, the increase in TOC under sulfidic conditions
low NO3− removal (days 113–126 and 455–486, Fig. 2a), and following (see Section 3.2) may indicate incomplete oxidation of high molecular
the winter intermission (days 325–329), the produced NH4+ relative to weight organic carbon substrates (e.g. glucose, butyrate) to acetate by
the amount of reduced NO3− was ∼ 8-9% and ∼50%, respectively. sulfate reducers (Muyzer and Stams, 2008). However, the simultaneous
The increased relative production of NH4+ during these periods are, alkalinity consumption and pH decrease during the production of TOC
however, due to the low absolute NO3− removal at low temperatures, suggests fermentation.
rather than an increased significance of DNRA. The product spectrum (i.e. the distribution of produced species of
organic carbon substrates) during anaerobic fermentation of glucose
3.5. Carbon cycling in bioreactor (the primary carbon substrate in the bioreactor, see above) has been
observed to shift from mainly butyrate (C3H7COO−) and acetate
For natural aquatic environments in general, it may be assumed that (CH3COO−) at low pH ranges (∼4–6.5) to mainly ethanol
anaerobic degradation processes are driven by the oxidation of carbo- (CH3CH2OH), propionate (C2H5COO−), and acetate at higher pH
hydrates (Conrad, 1999). For denitrification, the carbon contained (∼6.5–8.5) (see Fang and Liu, 2002; Horiuchi et al., 2002; Temudo
within the lignocellulose material (primarily as cellulose) must first be et al., 2008, 2007). Temudo et al. (2007) additionally observed a
solubilized through hydrolysis before the compounds are available as transition from CO2/H2 to formate production during the anaerobic
energy sources for heterotrophic denitrifying microorganisms fermentation of glucose as pH increased (range 4.0–8.5). The shift in
(Glombitza et al., 2015; Muyzer and Stams, 2008). Once solubilized, product spectrum during anaerobic fermentation of glucose has further
different microbial functional groups may compete for the labile or- been confirmed in a thermodynamic modeling framework (see
ganic carbon, which may be transformed into new organic carbon González-Cabaleiro et al., 2015a). This suggests that the biogeochem-
substrates by fermentation or sulfate reduction (see Muyzer and Stams, ical changes observed in the bioreactor with time, particularly between
2008), and/or be oxidized into inorganic carbon compounds, e.g. days 31 and 63 when effluent pH increased from 6 to 7 (see Fig. 2e),
H2CO3/CO2, HCO3−, and CO32−; this leads to characteristic changes in were due to a change in the organic carbon substrates utilized by the
pH and alkalinity. The changes in pH, alkalinity and TOC in the bior- biogeochemical processes. Indeed, the observed elevated NH4+ and
eactor pore water indicate that different biogeochemical processes are NO2− concentrations in the bioreactor following the bioreactor start-up
controlling the consumption of organic carbon substrates at different could be related to the use of more fermentable carbon substrates (see
times during the two year of operations. Section 3.2), e.g. butyrate. Further, the decline in methane production
Electron donors that are not directly produced by the hydrolysis of could be related to a decline in CO2/H2 production (see Section 3.3),
the carbon sources contained in the woody biomass must be produced the primary substrates for methanogenesis, from fermentation of glu-
by secondary reactions, for example fermentation and sulfate reduction. cose.
Woody biomass mainly consists of cellulose (40–55%), hemicellulose
(24–40%), lignin (18–35%) (Sun and Cheng, 2002), and additional 4. Mass balance modeling
extraneous substances such as organic extractives and ash/inorganic
minerals (see Fan et al., 1982). Typical hydrolysates of cellulose are To test if the inferred biogeochemical processes (see Sections 3.4
reducing sugars (e.g. glucose). Hemicellulose is more easily hydrolyzed and 3.5) indeed were responsible for the observation reduction/pro-
than cellulose due to their branched structure, and hemicellulose hy- duction of NOx, NH4+, SO42−, and TOC, mass balance modeling was
drolysates are xylose (C5H10O5), acetic acid (CH3COOH), and various used to predict the observed pH and alkalinity changes in the bioreactor
isomers of glucose (C6H12O6), e.g. galactose, and mannose (Palmqvist based on stoichiometric definitions of the respective biogeochemical
and Hahn-Hägerdahl, 2000). From the above discussion, glucose is processes. Methanogenesis was excluded in the modeling as it was
assumed as the primary carbon substrate in the bioreactor system, considered insignificant in relation to the other biogeochemical pro-
originating from the woodchips. cesses in the bioreactor. The observed alkalinity was assumed to en-
The effluent alkalinity generally followed the completeness of NO3− tirely consist of HCO3−. The mass changes of NO3−, NO2−, NH4+,
removal in the system (see Fig. 2a and f); this is supported by the SO42−, and TOC per liter were calculated from the difference in influent
stoichiometry of denitrification (reaction (1)), which is interpreted as and effluent concentrations. From these, the production/consumption
the dominating process for NO3− removal (see above). However, of HCO3− and H2CO3* were calculated by propagating the calculated
during the first 80 days after the start-up of the bioreactor, alkalinity mass changes through stoichiometric definitions of the inferred bio-
and pH consistently increased, with alkalinity consumption being ob- geochemical reactions as discussed above. The modeled production of
served immediately after the start-up. Since none of the biogeochemical alkalinity (HCO3−) and carbonic acid (H2CO3*) were used to calculate
reactions that are indicated as being active in the bioreactor (i.e deni- the resulting pH based on the following equilibrium reaction with a pKa
trification, DNRA, and sulfate reduction) are alkalinity-consuming, the value of 6.3:
observed initial alkalinity consumption must have been due to some
H2 CO3* ↔ H+ + HCO3− (7)
secondary process, possibly fermentation. For example, during the
fermentation of glucose to various volatile fatty acids (e.g. valerate, In EQ. (7), H2CO3* represents the combination of carbonic acid
butyrate, propionate, acetate, and formate) there is a consumption of (H2CO3) and dissolved carbon dioxide (CO2). For the pH ranges observed
HCO3− and a production of H2CO3 (see table S3.1, Supplementary in the bioreactor effluent (5.0–7.9, Fig. 2e), the CO32− concentration was
materials). Fermentation is an alkalinity-consuming and acidifying insignificant and it is assumed that the relative concentrations of HCO3−
process as long as there is an imbalance between the production of (alkalinity) to H2CO3* controlled the pH in the bioreactor.

62
A. Nordström, R.B. Herbert Ecological Engineering 110 (2018) 54–66

4.1. Stoichiometric definitions of the biogeochemical reactions 1 1 1 3 1


SO42 − + C6 H12 O6 → HS − + H2 S + HCO3− + H2 CO3
3 2 2 2 2 (10)
4.1.1. Selection of electron donor From the above discussion, the observed TOC production was due to
The production/consumption of HCO3− and H2CO3* is dependent fermentation, wherefore the potential of acetate production from sul-
on the electron donor (e.g. glucose, acetate, sulfides) utilized by the fate reduction was disregarded. Nevertheless, the stoichiometric defi-
biogeochemical reactions (see table S4.1 in Supplementary materials). nition of incomplete glucose oxidation to acetate by sulfate reduction
Glucose was inferred as the major electron donor in the bioreactor expands into separate stoichiometric definitions of sulfate reduction
system (see above). The net stoichiometric definitions of the secondary and fermentation of glucose to acetate (see section S5, Supplementary
reactions producing electron donors from hydrolysates, and the con- materials).
secutive consumption of the secondary electron donors by e.g. deni-
trification, condenses into direct utilization of hydrolysates (provided
4.1.4. Stoichiometric definition of fermentation
balanced production and consumption of secondary electron donors).
From the above discussion, fermentation is a major process in the
This is shown in section S3 in Supplementary materials. This justifies
system following the bioreactor start-up. Observations of elevated
the use of primary hydrolysates in the stoichiometric definitions of the
NO2− and NH4+ concentrations following the bioreactor start-up sug-
biogeochemical processes operating the bioreactor. However, it does
gest that fermentable carbon substrates are used (e.g. butyrate) in
not necessarily imply that glucose is the actual carbon substrate used by
preference over less fermentable carbon substrates (e.g. acetate) (see
e.g. denitrification, merely that the net stoichiometric definition of the
Section 3.2). The molar consumption and production of HCO3− and
biogeochemical processes condenses into glucose utilization. In fact,
H2CO3, respectively, from the fermentation of glucose is dependent on
fermentative coupling may be an energetic advantage (see González-
the type of carbon substrate produced (see table S3.1, Supplementary
Cabaleiro et al., 2015b), such that the fermentation of glucose may
materials). Since the speciation of produced organic carbon in the
provide additional carbon substrates than glucose to the biogeochem-
system was not determined, the stoichiometry of fermentation was
ical processes in the bioreactor.
based on the observations by Fang and Liu (2002), Horiuchi et al.
(2002), and Temudo et al. (2008, 2007). It was hence assumed that
4.1.2. Stoichiometric definitions of denitrification and DNRA
fermentation of glucose resulted in the production of a mixture of
The production of HCO3− and H2CO3 from denitrification was
(mainly) butyrate and acetate (reaction (11)).
modeled assuming full reduction of NO and/or N2O, but with the po-
tential for NO2− accumulation. The observed reduction in NO3− was 9α 6α
C6 H12 O6 + ⎛3 − ⎞ HCO3− → ⎛ ⎞ C3 H7 COO− + (3 − 3α ) CH3 COO−
propagated step-wise in reactions (8.1)–(8.4). ⎝ 5 ⎠ ⎝ 5 ⎠
1 1 3α
NO3− + C6 H12 O6 → NO2− + H2 CO3 + ⎛3 − ⎞ H2 CO3
12 2 (8.1) ⎝ 5 ⎠ (11)

1 3 1 In reaction (11), the parameter α determines the quotient between


NO2− + C6 H12 O6 + H2 CO3 → NO + HCO3− + H2 O
24 4 2 (8.2) butyrate and acetate production from anaerobic glucose fermentation.
If α = 1, then the inferred TOC production from the anerobic fermen-
1 1 1
NO + C6 H12 O6 → N2 O + H2 CO3 tation of glucose purely results in the production of butyrate. If α = 0,
24 2 4 (8.3)
then acetate is the single organic carbon substrate produced. The ob-
1 1 1 1 served molar production of organic carbon (TOC) was related to the
N2 O + C6 H12 O6 → N2 + H2 CO3
2 24 2 4 (8.4) consumption and production of HCO3− and H2CO3, respectively, per
mol organic carbon formed as predicted from reaction (11).
The observed production of NH4+was assumed to come from the
The parameter α was randomly sampled between 0 and 1 from a
reduction of NO2− through the DNRA pathway (reaction (9)). The
uniform distribution 1000 times. The optimal values of α were selected
produced NH4+ was subtracted from the theoretical production of
based on the absolute residuals between the observed and simulated pH
NO2− (reaction (8.1)), accounting for any observable net accumula-
and alkalinities, respectively. Since different values of α were “optimal”
tion/reduction of NO2− in the bioreactor system (Fig. 2b).
with respect to the observed alkalinity and pH, the selected value is
1 1 represented as the average between these two.
NO2− + C6 H12 O6 + H2 CO3 + H2 O → NH4+ + 2HCO3−
4 2 (9)
From the observed production of NH4+,the production and con- 4.1.5. Accounting for retention time in the system and uncertainty in
sumption of HCO3− and H2CO3 was modeled using reaction (9). In concentrations
addition to the carbonate system, denitrification and DNRA involes the The influent NO3−, NO2−, NH4+, and SO42− concentrations were
acid-base pairs of NO2−/HNO2 (pKa 3.4) and NH4+/NH3 (pKa 9.3). linearly interpolated on a daily basis between days of observation. The
From the observed pH range in the bioreactor (see above), the pro- observed effluent concentrations (on day X) were then matched with
duction of HNO2 and/or NH3 should be insignificant and is therefore the influent concentration that was interpolated/observed on day X-
disregarded in the model. HRT (HRT in days). This was done in order to account for the variability
in influent concentrations, and the delay between influx and outflux to/
4.1.3. Stoichiometric definition of sulfate reduction from the system.
The production of HCO3− and H2CO3 from heterotrophic sulfate Due to the large uncertainties in the determined sulfate concentra-
reduction was modeled assuming that all the observed reduction in tions (see above), the occurrence and magnitude of sulfate reduction is
SO42− in the bioreactor could be attributed to sulfate reduction. In highly uncertain. Because of this, pH and alkalinity were first modeled
addition to the carbonate system, heterotrophic sulfate reduction in- only accounting for heterotrophic denitrification and DNRA (scenario
volves the acid-base pair HS−/H2S (pKa 7.04), which is active in the pH 1). Uncertainties in NO3−, NO2−, and NH4+ concentrations were not
range observed in the bioreactor (see above). The quotient of produced accounted for, and the produced HCO3− was considered the only
HS−/H2S should change during the course of bioreactor operations, contributor to alkalinity. Sulfate reduction was then added to the re-
being mainly H2S prior to day 63 (pH < 7) and HS− post day 63 actions in scenario 1 (scenario 2). Sulfate reduction was only allowed to
(pH > 7). However, since the effluent pH in the bioreactor on average occur whenever there was an observed reduction of 30 mg L−1 SO42−
was 7.1 which is close to the pKa of H2S, sulfate reduction was modeled between influent and effluent data. Fermentation was added to scenario
with an equal production of H2S and HS− for simplicity (reaction (10)). 2 in scenario 3.

63
A. Nordström, R.B. Herbert Ecological Engineering 110 (2018) 54–66

4.1.6. CO2 outgassing preferred carbon substrate following the start-up of the bioreactor, then
The stoichiometric definitions of the biogeochemical processes used it should not be traceable in the bioreactor effluent. With time as pH
in the modeling implicitly assumes no CO2 outgassing from the bior- increases in the bioreactor, scenario 3 predicts a step-like change in α
eactor. CO2 surface fluxes from the bioreactor were not determined in (reaction (11)) to 1, implying that fermentation almost only produces
this study, but have nevertheless been observed in previous woodchip butyrate. The simulated change in primary glucose fermentation pro-
bioreactor studies (e.g. Elgood et al., 2010; Warneke et al., 2011a). CO2 ducts from acetate to butyrate with time indicates that the preferred
outgassing would lead to a pH increase without affecting the alkalinity carbon substrate of the TEAP biogeochemical processes (e.g. deni-
(HCO3−); however, we believe that (natural) CO2 outgassing is not trification) changed from butyrate to acetate/ethanol with time. The
substantial in the bioreactor in this study as: (1) the covering till layer change in the preferred carbon substrate of e.g. denitrification is sug-
creates confining conditions; and (2) natural waters and groundwaters gested to occur in response to the relative availability of different
are often observed to be supersaturated with respect to atmospheric carbon substrates at different pH (see e.g. Temudo et al., 2008,2007),
CO2. CO2 partial pressures (pCO2) in groundwater are usually ∼10-100 which ultimately is explained by a maximization of the energy harvest
times higher than the atmosphere (Macpherson, 2009), and boreal rate in the system (see González-Cabaleiro et al., 2015a). Indeed, the
streams, rivers, and lakes are often observed to be supersaturated with metabolic energy harvest rate has been identified as a selective force of
respect to atmospheric CO2, with pCO2 ranging between microbial activities in competitive systems (see González-Cabaleiro
∼300–5500 μatm (Roehm et al., 2009; Teodoru et al., 2009), sug- et al., 2015b).
gesting slow/incomplete CO2 outgassing.

5. Conclusions
4.2. Biogeochemical evolution of the bioreactor
The woodchip bioreactor in this study was tested as a potential method
4.2.1. Simulated pH and alkalinity to remove nitrate (NO3−) from mine and process water originating from
Scenario 1 is generally predicting a pH below the observed pH in the the Kiruna iron ore mine (northern Sweden). The test period extended for
bioreactor effluent (see Fig. 6a). However, prior to day 63 the simulated the duration of two years. The results show that the bioreactor was able to
pH falls above the observed (see Fig. 6a). reduce an average NO3− concentration of 22 mg N L−1 to below the de-
Scenario 1 is generally able to predict the alkalinity changes post tection limit (0.06 mg N L−1) at temperatures above 5 °C. The (theore-
day 63 in the bioreactor effluent (see Fig. 6b). Prior to day 63 as the tical) hydraulic residence time required for this was 1.9–2.6 days. The
observed pH falls below 7, the proposed biogeochemical reactions main NO3− removing mechanism at temperatures above 5 °C was com-
predict a too high alkalinity (HCO3−) production. Post day 63, Scenario plete denitrification. As the bioreactor temperatures decreased to below
1 is able to account for approximately all of the alkalinity production 5 °C, dissimilatory nitrate reduction to ammonium (DNRA) and in-
except during the high temperature periods when the predicted HCO3− complete denitrification to nitrite (NO2−) or nitrous oxide gas (N2O) in-
is too low. During the high temperature periods, NO3− reduction was creased in significance, resulting in a substantial production of ammonium
complete in the bioreactor (see Fig. 2a), triggering sulfidic conditions in (NH4+) and N2O relative to the amount of NO3− reduced in the system.
the bioreactor. Sulfate reduction was added to scenario 1 in scenario 2, Following bioreactor start-up, sulfate reduction (as noted by H2S odor) and
and from Fig. 6b it is seen that sulfate reduction is able to compensate the production of organic carbon from fermentation were substantial.
for the deficiency in alkalinity production from scenario 1 during the Methanogenesis, resulting in the production of methane (CH4), was ob-
high temperature periods. From Fig. 6b it is also seen that scenario 2 servable yet insignificant in relation to the other biogeochemical processes
generally overpredicts alkalinity production during periods of TOC operating in the system.
production (see Fig. 2g), particulary during the first operational year With time the diversity of biogeochemical processes controlling
(Fig. 6b). This suggests that during these periods, fermentation was major redox-active species decreased: sulfate reduction and fermenta-
active. Fermentation (reaction (11)) was added to scenario 2 in scenario tion decreased in importance with time from bioreactor start-up,
3. As seen in Fig. 6a, scenario 3 is able to predict the effluent pH from leaving denitrification as the sole major terminal electron accepting
the bioreactor, under the assumption that close to all of the produced process. The decrease in biogeochemical diversity occurred simulta-
TOC is in the form of acetate (i.e. α in reaction (11) is close to 0). At the neously as the decline in NO2−, NH4+, and TOC production, and an
pH ranges observed in the bioreactor prior to day 63 (Fig. 6a), the increase and stabilization of the effluent pH and alkalinity. Based on a
studies by Fang and Liu (2002), Horiuchi et al. (2002), and Temudo stoichiometric mass balance confined by observed changes in pH and
et al. (2008, 2007) suggests a production of butyrate and acetate from alkalinity in the bioreactor, the trigger of the decline in biogeochemical
glucose fermentation. Scenario 3 suggests no net production of buty- diversity is suggested to have been a change in the carbon substrate
rate, which could imply that butyrate is consumed by the biogeo- utilized by the nitrate and sulfate-reducing communities from primarily
chemical processes in the bioreactor. Indeed, if butyrate was the butyrate to acetate, both provided by the fermentation of glucose.

Fig. 6. Modeled and observed pH (a) and alkalinity


(b) for scenario 1, 2, and 3. Denitrification and
DNRA was considered in scenario 1, sulfate reduc-
tion was added to scenario 1 in scenario 2, and fer-
mentation was added to scenario 2 in scenario 3.

64
A. Nordström, R.B. Herbert Ecological Engineering 110 (2018) 54–66

Conflicts of interest carbon substrates for denitrification of subsurface drainage water. J. Environ. Qual.
35, 824–829. http://dx.doi.org/10.2134/jeq2005.0247.
Greenan, C.M., Moorman, T.B., Parkin, T.B., Kaspar, T.C., Jaynes, D.B., 2009.
None. Denitrification in wood chip bioreactors at different water flows. J. Environ. Qual. 38,
1664–1671. http://dx.doi.org/10.2134/jeq2008.0413.
Herbert, R.B., Winbjörk, H., Hellman, M., Hallin, S., 2014. Nitrogen removal and spatial
Acknowledgements disitribution of denitrifier and anammox communities in a bioreactor for mine
drainage treatment. Water Res. 66, 350–360. http://dx.doi.org/10.1016/j.watres.
This work was supported by the Luossavaara-Kiirunavaara 2014.08.038.
Hoover, N.L., Bhandari, A., Soupir, M.L., Moorman, T.B., 2016. Woodchip denitrification
Aktiebolag (LKAB), Boliden Aktiebolag (Boliden AB), and VINNOVA, bioreactors: impact of temperature and hydraulic retention time on nitrate removal.
The Swedish Innovation Agency, under Grant 2014-01134. The work J. Environ. Qual. 45, 803–812. http://dx.doi.org/10.2134/jeq2015.03.0161.
presented herein is part of the larger project “miNing – Reduction of Horiuchi, J.I., Shimizu, T., Tada, K., Kanno, T., Kobayashi, M., 2002. Selective production
of organic acids in anaerobic acid reactor by pH control. Bioresour. Technol. 82,
nitrogen discharges in mining processes and mitigating its environ-
209–213. http://dx.doi.org/10.1016/S0960-8524(01)00195-X.
mental impact”, a collaboration between LKAB, Boliden AB, Rock Tech Komsta, L., 2012a. Chemometric and statistical evaluation of calibration curves in
Centre, the Swedish University of Agricultural Sciences, Luleå pharmaceutical analysis: a short review on trends and recommendations. J. AOAC
University of Technology, and Uppsala University. Int. 95, 669–672.
Komsta, L., 2012b. Quantchem: Quantitative Chemical Analysis: Calibration and
Special thanks are extended to Johan Johansson for assistance Evaluation of Results. R Package Version 0.13. http://CRAN.R-project.org/
during the construction of the bioreactor, Christopher Jones and package=quantchem.
Joachim Audet for discussions and help during planning and prepara- López-Cortés, A., Fardeu, M.L., Fauque, G., Joulian, C., Ollivier, B., 2006. Reclassification
of the sulfate- and nitrate-reducing bacterium Desulfovibrio vulgaris subsp. Oxamicus
tions of gas sampling, and to two anonymous reviewers for their va- as Desulfovibrio oxamicus sp. Nov., comb. Nov. Int. J. Syst. Evol. Microbiol. 69,
luable comments on the manuscript. 1495–1499. http://dx.doi.org/10.1099/ijs.0.64074-0.
LKAB, 2016. Miljörapport LKAB Kiruna [Environmental Report LKAB Kiruna 2015].
Luossavaara-Kiirunavaara AB (in Swedish).
Appendix A. Supplementary data Lindeström, L., 2012. Kväveutsläpp från gruvindustrin: Risker för miljöproblem, krav på
utsläppsbegränsningar och möjliga åtgärder [Nitrogen releases from the mining in-
Supplementary data associated with this article can be found, in the dustry: risks for environmental problems, requirements for release limits, and pos-
sible solutions]. SveMin. (in Swedish), Stockholm.
online version, at http://dx.doi.org/10.1016/j.ecoleng.2017.09.018.
Macpherson, G.L., 2009. CO2 distribution in groundwater and the impact of groundwater
extraction on the global C cycle. Chem. Geol. 264, 328–336. http://dx.doi.org/10.
References 1016/j.chemgeo.2009.03.018.
McKenney, D.J., Shuttleworth, K.F., Vriesacker, J.R., Findlay, W.I., 1982. Production and
loss of nitric oxide from denitrification in anaerobic brookston clay. Appl. Environ.
Akunna, J.C., Bizeau, C., Moletta, R., 1993. Nitrate and nitrite reductions with anaerobic Microb. 43, 534–541.
sludge using various carbon sources: glucose, glycerol, acetic acid, lactic acid and Moorman, T.B., Parkin, T.B., Kaspar, T.C., Jaynes, D.B., 2010. Denitrification activity,
methanol. Water Res. 27, 1303–1312. http://dx.doi.org/10.1016/0043-1354(93) wood loss, and N2O emissions over 9 years from a wood chip bioreactor. Ecol. Eng.
90217-6. 36, 1567–1574. http://dx.doi.org/10.1016/j.ecoleng.2010.03.012.
Akunna, J.C., Bizeau, C., Moletta, R., 1994. Nitrate reduction by anaerobic sludge using Muyzer, G., Stams, A.J.M., 2008. The ecology and biotechnology of sulphate-reducing
glucose at various nitrate concentrations: ammonification, denitrification and me- bacteria. Nat. Rev. Microbiol. 6, 441–454. http://dx.doi.org/10.1038/nrmicro1892.
thanogenic activities. Environ. Technol. 15, 41–49. http://dx.doi.org/10.1080/ Nordström, A., Herbert, R., 2017. Denitrification in a low temperature bioreactor system
09593339409385402. at two different hydraulic residence times: laboratory column studies. Environ.
Appelo, C.A.J., Postma, D., 2005. Geochemistry, Groundwater and Pollution, Second ed. Technol. 38, 1362–1375. http://dx.doi.org/10.1080/09593330.2016.1228699.
A.A. Balkema, Leiden. Palmqvist, E., Hahn-Hägerdahl, B., 2000. Fermentation of lignocellulosic hydrolysates. II:
Betlach, M.R., Tiedje, J.M., 1981. Kinetic explanation for accumulation of nitrite, nitric inhibitors and mechanisms of inhibition. Bioresour. Technol. 74, 25–33. http://dx.
oxide, and nitrous oxide during bacterial denitrification. Appl. Environ. Microb. 42, doi.org/10.1016/S0960-8524(99)00161-3.
1074–1084. Parkin, T.B., Venterea, R.T., 2010. Chamber-based trace gas flux measurements. In:
Breibarth, E., Mills, M.M., Friedrichs, G., LaRoche, J., 2004. The Bunsen gas solubility Follett, R.F. (Ed.), Sampling Protocols. United States Department of Agriculture, Fort
coefficient of ethylene as a function of temperature and salinity and its importance Collins 3.1–3.39.
for nitrogen fixation assays. Limnol. Oceanogr-Method 2, 282–288. http://dx.doi. Pedersen, A.R., 2015. HMR: Flux Estimation with Static Chamber Data. R Package Version
org/10.4319/lom.2004.2.282. 0.4.1. http://CRAN.R-project.org/package=HMR.
Burgin, A.J., Hamilton, S.K., 2007. Have we overemphasized the role of denitrification in R Core Team, 2017. R: A Language and Environment for Statistical Computing. R
aquatic ecosystems? A review of nitrate removal pathways. Front. Ecol. Environ. 5, Foundation for Statistical Computing, Vienna, Austria. https://www.R-project.org.
89–96. http://dx.doi.org/10.1890/1540-9295(2007)5[89:HWOTRO]2.0.CO;2. Revey, G., 1996. Practical methods to control explosives losses and reduce ammonia and
Conrad, R., 1999. Contribution of hydrogen to methane production and control of hy- nitrate levels in mine water. Min. Eng. Tech. Pap. 61–64.
drogen concentrations in methanogenic soils and sediments. FEMS Microbiol. Ecol. Rittmann, B.E., McCarty, P., 2001. Biotechnology: Principles and Applications. McGraw-
28, 193–202. http://dx.doi.org/10.1111/j.1574-6941.1999.tb00575.x. Hill Book Co., New York.
Dalsgaard, T., Bak, F., 1994. Nitrate reduction in a sulfate-reducing bacterium, desulfo- Robertson, W.D., Merkley, L.C., 2009. In-Stream bioreactor for agricultural nitrate
vibrio desulfuricans, isolated from rice paddy soil sulfide inhibition, kinetics, and treatment. J. Environ. Qual. 38, 230–237. http://dx.doi.org/10.2134/jeq2008.0100.
regulation. Appl. Environ. Microb. 60, 291–297. Robertson, W.D., 2010. Nitrate removal rates in woodchip media of varying age. Ecol.
EPA, 2013. Aquatic Life Ambient Water Quality Criteria for Ammonia –freshwater. Report Eng. 36, 1581–1587. http://dx.doi.org/10.1016/j.ecoleng.2010.01.008.
EPA 822-R-13-001. United States Environmental Protection Agency, Washington, DC. Roehm, C.L., Prairie, Y.T., del Giorgio, P.A., 2009. The pCO2 dynamics in lakes in the
Elgood, Z., Robertson, W.D., Schiff, S.L., Elgood, R., 2010. Nitrate removal and green- boreal region of northern Québec, Canada. Global Biogeochem. Cy. 23, 1–9. http://
house gas production in a stream-bed denitrifying bioreactor. Ecol. Eng. 36, dx.doi.org/10.1029/2008GB003297.
1575–1580. http://dx.doi.org/10.1016/j.ecoleng.2010.03.011. SMHI, 2017. Swedish Meteorological and Hydrological Institute. https://www.smhi.se/
Fan, L.T., Lee, Y.H., Gharpuray, M.M., 1982. The nature of lignocellulosics and their klimatdata/meteorologi/temperatur (Accessed 17.06.21) (in Swedish).
pretreatment for enzymatic hydrolysis. Adv. Biochem. Eng. 23, 158–187. Schipper, L.A., Robertson, W.D., Gold, A.J., Jaynes, D.B., Cameron, S.C., 2010.
Fang, H.H.P., Liu, H., 2002. Effect of pH on hydrogen production from glucose by a mixed Denitrifying bioreactors –an approach for reducing nitrate loads to receiving waters.
culture. Bioresour. Technol. 82, 87–93. http://dx.doi.org/10.1016/S0960-8524(01) Ecol. Eng. 36, 1581–1587. http://dx.doi.org/10.1016/j.ecoleng.2010.04.008.
00110-9. Sun, Y., Cheng, J., 2002. Hydrolysis of lignocellulosic materials for ethanol production: a
Forsyth, B., Cameron, A., Miller, S., 1995. Explosives and water quality. In: Proceedings of review. Bioresour. Technol. 83, 1–11. http://dx.doi.org/10.1016/S0960-8524(01)
Sudbury ’95 Mining and the Environment. Montreal, Quebec, Canada: MEND (Mine 00212-7.
Environment Neutral Drainage). pp. 795–803. Temudo, M.F., Kleerebezem, R., van Loosdrecht, M., 2007. Influence of the pH on (Open)
Glombitza, C., Jaussi, M., Røy, H., Seidenkrantz, M.S., Lomstein, B.A., Jørgensen, B.B., mixed culture fermentation of glucose: a chemostat study. Biotechnol. Bioeng. 98,
2015. Formate, acetate, and propionate as substrates for sulfate reduction in sub- 69–79. http://dx.doi.org/10.1002/bit.21412.
arctic sediments of Southwest Greenland. Front. Microb. 6, 846. http://dx.doi.org/ Temudo, M.F., Muyzer, G., Kleerebezem, R., van Loosdrecht, M.C.M., 2008. Diversity of
10.3389/fmicb.2015.00846. microbial communities in open mixed culture fermentations: impact of the pH and
González-Cabaleiro, R., Ofiţeru, I.D., Lema, J.M., Rodríguez, J., 2015a. Microbial cata- carbon source. Appl. Microbiol. Biotechnol. 80, 1121–1130. http://dx.doi.org/10.
bolic activities are naturally selected by metabolic energy harvest rate. ISME J. 9, 1007/s00253-008-1669-x.
2630–2641. http://dx.doi.org/10.1038/ismej.2015.69. Teodoru, C.R., del Giorgio, P.A., Prairie, Y.T., Camire, M., 2009. Patterns in pCO2 in
González-Cabaleiro, R., Lema, J.M., Rodríguez, J., 2015b. Metabolic energy-based mod- boreal streams and rivers of northern Quebec, Canada. Global Biogeochem. Cy.
elling explains product yielding in anaerobic mixed culture fermentations. PLoS One 23http://dx.doi.org/10.1029/2008GB003404. GB2012.
10, 1–17. http://dx.doi.org/10.1371/journal.pone.0126739. Torstensson, B.A., 1984. A new system for ground water monitoring. Ground Water
Greenan, C.M., Moorman, T.B., Kaspar, T.C., Parkin, T.B., Jaynes, D.B., 2006. Comparing Monit. Rem. 4, 131–138. http://dx.doi.org/10.1111/j.1745-6592.1984.tb00904.x.

65
A. Nordström, R.B. Herbert Ecological Engineering 110 (2018) 54–66

Warneke, S., Schipper, L.A., Matiasek, M.G., Cameron, S., Bruesewitz, D.A., McDonald, Ecol. Eng. 37, 511–522. http://dx.doi.org/10.1016/j.ecoleng.2010.12.006.
I.R., 2011a. Nitrate removal, communities of denitrifiers and adverse effects in dif- Wilderer, P.A., Jones, W.L., Dau, U., 1987. Competition in denitrification systems af-
ferent carbon substrates for use in denitrification beds. Water Res. 45, 5463–5475. fecting reduction rate and accumulation of nitrite. Water Res. 21, 239–245. http://
http://dx.doi.org/10.1016/j.watres.2011.08.007. dx.doi.org/10.1016/0043-1354(87)90056-X.
Warneke, S., Schipper, L.A., Bruesewitz, D.A., McDonald, I., Cameron, S., 2011b. Rates, Wilhelm, E., Battino, R., Wilcock, R.J., 1977. Low-pressure solubility of gases in liquid
controls and potential adverse effects on nitrate removal in a denitrification bed. water. Chem. Rev. 77, 219–261. http://dx.doi.org/10.1021/cr60306a003.

66

S-ar putea să vă placă și