You are on page 1of 11

Water Research 128 (2018) 293e303

Contents lists available at ScienceDirect

Water Research
journal homepage: www.elsevier.com/locate/watres

Fast anaerobic sludge granulation at elevated salinity


D. Sudmalis a, *, M.C. Gagliano b, R. Pei a, K. Grolle a, C.M. Plugge b, H.H.M. Rijnaarts a,
G. Zeeman a, H. Temmink a
a
Sub-department of Environmental Technology, Wageningen University and Research, Bornse Weilanden 9, 6708 WG, Wageningen, The Netherlands
b
Laboratory of Microbiology, Wageningen University and Research, Stippeneng 4, 6708WE, Wageningen, The Netherlands

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

Article history: It is commonly accepted that high salt concentrations negatively affect microbial activity in biological
Received 21 July 2017 wastewater treatment reactors such as upflow anaerobic sludge blanket (UASB) reactors. Microbial ag-
Received in revised form gregation in such reactors is equally important. It is well documented that anaerobic granules, when
28 September 2017
exposed to high salinity become weak and disintegrate, causing wash-out, operational problems and
Accepted 18 October 2017
Available online 21 October 2017
decreasing process performance. In this research, the possibility of microbial granule formation from
dispersed biomass was investigated at salinity levels of 5 and 20 g Naþ/L. High removal efficiencies of
soluble influent organics were achieved at both salinity levels and this was accompanied by fast and
Keywords:
Anaerobic sludge granulation
robust formation of microbial granules. The process was found to be stable for the entire operational
UASB period of 217 days. As far as we know this is the first time it has been demonstrated that stable granule
High salinity formation is possible at a salinity level as high as 20 g Naþ/L. Methanosaeta was identified as the
Methanosaeta dominant methanogen at both salinity levels. Streptococcus spp. and bacteria belonging to the family
Lachnospiraceae were identified as the dominant microbial population at 5 and 20 and g Naþ/L,
respectively.
© 2017 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).

1. Introduction recovery, reduced sludge production and the possibility to handle


higher volumetric loading rates if formation of granular or immo-
Saline wastewater streams are generated by several industries bilised biomass is possible (Vieira et al., 2005; Vyrides et al., 2010;
such as the fish processing (Soto et al., 1991), chemical, agro-food, Xiao and Roberts, 2010).
petroleum and leather industry (Lefebvre and Moletta, 2006). In the field of anaerobic wastewater treatment it is generally
Biodegradation of organics in such saline waste streams is reported accepted that methanogenic activity is inhibited above sodium
to be poor due to toxic effects of sodium on the bacterial cells and levels of 5 g Naþ/L (McCarty and McKinney, 1961; Patel and Roth,
Archaea causing cell lysis and death (Sleator and Hill, 2002; Vyrides 1977; Rinzema et al., 1988; Liu and Boone, 1991; De Vrieze et al.,
and Stuckey, 2009; Sochacki et al., 2011). Physical and chemical 2016b). However, sodium toxicity may depend on the composi-
treatment processes are energy intensive and more expensive tion of wastewater due to possible antagonistic and synergistic
compared to biological processes (Fakhru'l-Razi et al., 2009), hence effects caused by other ions (McCarty and McKinney, 1961).
the need to explore the limits of biological wastewater treatment Moreover the type of organic substrate generally determines the
regarding salinity and to stretch these limits where possible. type of microorganisms evolving in the reactors and thus their
Because of a lower sensitivity of aerobic biomass to salt toxicity sensitivity to sodium toxicity (Patel and Roth, 1977; Lefebvre et al.,
compared to anaerobic biomass, most research is done on aerobic 2007). Hence, the potential of anaerobic treatment of saline
treatment (Freire et al., 2001; Tellez et al., 2002, 2005; Zhao et al., wastewaters thus far may have been largely underestimated.
2006; Ozgun et al., 2013). Anaerobic biological wastewater treat- If high rate anaerobic bioreactors such as upflow anaerobic
ment has several advantages over aerobic, such as net energy sludge blanket (UASB) for treatment of wastewater are applied,
both microbial activity and formation of granular microbial ag-
gregates plays a crucial role on the process performance (O'flaherty
* Corresponding author. et al., 1997; Yu et al., 1999; Liu et al., 2009; Lu et al., 2013). After the
E-mail address: dainis.sudmalis@wur.nl (D. Sudmalis). initial formation of granules it is of equal importance to maintain a

https://doi.org/10.1016/j.watres.2017.10.038
0043-1354/© 2017 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
294 D. Sudmalis et al. / Water Research 128 (2018) 293e303

robust granular structure (McHugh et al., 2003). Microbial granules UASB, while in R3 it was left intact. The inoculum of R2, R3 and R4
of sufficient strength and weight then allow the decoupling of was mixed with powdered activated carbon (PAC) (0.1% of inoc-
hydraulic and sludge retention times, achieve high biomass con- ulum mass; 100e400 mesh, Sigma-Aldrich).
centrations and permit the use of compact reactors (O'flaherty Fig. 1 shows a schematic drawing of R4. The set e ups of R1 - R3
et al., 1997; Liu et al., 2009). (Fig. S1) had external circulation tanks while R4 was operated with
The most commonly used methods to enhance the rate of an internal circulation. The upflow velocity of 0.2e1 m/h, was
granulation are addition of multivalent cations, such as calcium controlled by peristaltic circulation pumps (Watson e Marlow
(150e300 mg/L) (Zeeuw, 1984; Mahoney et al., 1987; Hulshoff Pol, 530U, UK).
1989; Yu et al., 2001) and/or addition of a support material, such as Table 1 contains a summary of start-up and operational condi-
powdered activated carbon (PAC) (Wirtz and Dague, 1996; Yu et al., tions for each bioreactor
1999).
Only a few studies have reported on the effect of monovalent 2.2. Process performance analysis
salt stress on anaerobic granules. All of these studies were per-
formed by inoculating (intact) granular sludge and - without 2.2.1. Analytical methods
exception - showed subsequent granule disintegration, biomass Daily biogas production was monitored with mFlow gas flow
washout and loss of microbial activity (Rinzema et al., 1988; Jeison meter (Bioprocess Control Sweden AB). Biogas composition was
et al., 2008; Ismail et al., 2010; Li et al., 2014; De Vrieze et al., 2016a). measured periodically as described in Steinbusch et al. (2008).
As far as we know, formation of anaerobic granules from dispersed Volatile fatty acids (VFAs) were quantified in soluble fraction of
biomass under saline conditions has not been reported before. effluent samples on Agilent 7890B gas chromatograph equipped
In this paper granulation from dispersed biomass, accompanied with FID detector and HP-FFAP column (25 m  0.32 mm, 0.5 mm
by stable UASB process performance, at sodium concentrations as film thickness). Helium (purity 5.0) was used as the carrier gas
high as 20 g Naþ/L is demonstrated for the first time. The rate of (1.25 ml/min for first 3.5 min and 2 ml/min further). The injection
granulation was compared to granulation at a reference level of 5 g volume (split injection 1:25) was 1 mL and the injection tempera-
Naþ/L. The effect on granulation of small amounts of powdered ture 250  C. The oven temperature had the following temperature
activated carbon (PAC) was studied. Calcium was only fed at ramp: 60  C for 3 min; 21  C/min up to 140  C; 8  C/min up to
extremely low concentrations of 13 mg/L. Microscopic analyses 150  C; 1.5 min constant 150  C; 120  C/min up to 200  C; 1.25 min
coupled with fluorescent in situ hybridization (FISH) showed constant 200  C; 120  C/min up to 240  C; 3 min 240  C constant.
dominance of filamentous Methanosaeta, irrespective of the salinity The detector temperature was constantly kept at 240  C. The data
(5 or 20 g Naþ/L) at which the UASB reactors were operated. acquisition was done with Chromeleon 6.80 SR13 software.
COD measurements were performed with LCK314, LCK514 and
2. Materials and methods LCK1414 kits (HACH GMBH, Germany) after sample dilution with
milli-Q to reach chloride levels of 0.2e1.1 g/L. Differentiation be-
2.1. Experimental set eup tween total (tCOD), colloidal and supracolloidal (cCOD) and soluble
fractions (sCOD) of COD was achieved by filtration. Raw effluent
2.1.1. Growth medium samples were filtered through 4e7 mm paper filters (Whatman™,
The nutrient medium used was adapted from Vallero et al. grade 595 ½) and part of the filtrate was further passed through a
(2002) and consisted of the following salts at final concentrations pre-washed 0.45 mm cellulose acetate membrane filter (VWR® Sy-
in g/L: NH4Cl (1.02), MgSO4*7H2O (0.05), CaCl2*2H2O (0.05), ringe Filters). Soluble fraction is defined as filtrate through 0.45 mm
KH2PO4 (0.22), NaHCO3 (1.5) and in mg/L FeCl2*4H2O (1.2), HBO3 filter (Park et al., 2006). Colloidal and supracolloidal fraction is
(0.03), ZnCl2 (0.03), CuCl2*2H2O (0.02), MnCl2*4H2O (0.3), calculated from the difference between 4 and 7 mm and 0.45 mm
(NH4)6Mo7O24*4H2O (0.05), CoCl2*6H2O (1.2), NiCl2*6H2O (0.03),
EDTA (0.6), Resazurin (0.3), Na2SeO3*5H2O (0.3) and 0.216 ml/L of
36% HCl.
The COD of the influent was soluble and consisted of D-Glucose,
Na-Acetate and Tryptone (Sigma-Aldrich, Microbiologically tested,
N content 11e16%) in COD proportions of 3:2:1. The COD concen-
tration in the influent was increased in steps from 3 g/L to 12 g/L to
reach a final organic loading rate of approximately 16 g COD/m3$d.
Analytical grade NaCl (Merck Suprapur®) was used to adjust
sodium concentration in each reactor's influent for the first 3
months and afterwards was replaced by technical grade NaCl
(VWR, minimum purity 98%). Final sodium concentrations of 5 or
20 g Naþ/L include the sodium originating from sodium acetate.

2.1.2. Inoculum, set-up and operation


Four double-walled glass UASB bioreactors (R1, R2, R3 and R4)
were inoculated with 6 g VSS/L. Origin of the biomass was a full
scale mesophilic UASB treating wastewater from a styrene and
propene-oxide production plant (Shell, Moerdijk, the Netherlands).
The sludge was adapted to approximately 8 g Naþ/L for almost 10 Fig. 1. Schematic drawing of set-up for R4. 1 e influent tank at 4  C; 2 e peristaltic
years and has a specific methanogenic activity of approximately influent pump; 3 e influent sampling port; 4 e pebbles for influent distribution; 5 e
0.2e0.3 g COD/gVSS$d with acetate as carbon source (Jeison et al., double walled glass bioreactor; 6 e effluent line; 7 - effluent collection tank at 4  C; 8 -
biogas sampling port; 9 - gas flow meter; 10 e sampling ports; 11 e peristaltic
2008; Ismail, 2013). The COD in this wastewater mainly consisted recirculation pump; 12 e computer for data logging; CIRC e circulation line; pH e pH
of benzoic and acetic acid. The inoculum for R1, R2 and R4 was sensor; redox e oxidation reduction sensor; T e temperature sensor; HW - heating
dispersed by forcing it through a 125 mm sieve before adding to the bath for temperature control at 35  C.
D. Sudmalis et al. / Water Research 128 (2018) 293e303 295

Table 1
Summary of reactors operation parameters and dimensions. CODinf e influent COD concentration.

Parameters R1 R2 R3 R4

Inoculum Sieved Biomass Sieved Biomass þ 0.1% PAC Intact Biomass þ 0.1% PAC Sieved Biomass þ 0.1% PAC
Operating Temperature 35 ± 1  C 35 ± 1  C 35 ± 1  C 35 ± 1  C
Salinity 5 g Naþ/L 5 g Naþ/L 20 g Naþ/L 20 g Naþ/L
Total Reactor volume prior height increase 680 mL 730 mL 730 mL 2940 mL
Additional Gas e Solids eLiquid separator volume added on day 102 190 mL 190 mL 190 mL Not needed
Reactor volume prior sedimentation zone prior height increase 500 mL 550 mL 550 mL 1960 mL
Reactor height prior sedimentation zone 0.24 m 0.26 m 0.26 m 0.65 m
Reactor Diameter 0.052 m 0.052 m 0.052 m 0.062 m
Loading rate during start - up 1 g COD/L$d 1 g COD/L$d 1 g COD/L$d 1.4 g COD/L$d
Final loading rate after 100 days 11 g COD/L$d 11 g COD/L$d 11.5 g COD/L$d 9.11 g COD/L$d
Increase of upflow velocity from 0.2 m/h to 1 m/h Day 30 Day 30 Day 30 Day 18
Increase of CODinf from 3 g/L to 7 g/La Day 31 Day 31 Day 31 day 26
Increase of CODinf from 7 g/L to 12 g/La Day 52 Day 52 Day 52 Day 60
a
Increase of CODinf corresponds to increasing loading rates.

filtrates (Dulekgurgen et al., 2006). Total COD is measured in ho- bioreactors, singletons were measured at each time point for
mogenized effluent samples. Particulate COD (pCOD) is defined as strength characterization.
difference between 4 and 7 mm filtrate COD and total COD. The integrity coefficient IC is calculated as follows:
During the entire operation period grab samples of effluent COD  
were collected and sCOD removal efficiency was measured in SSx
ICx ¼ 1 *100
singleton per sample. 40 h composite effluent samples were SS0
collected on day 179 in R1, R2 and R3 (day 100 for R4). COD analysis
on different fractions of the samples was performed in triplicate. where ICx e the integrity coefficient after x minutes of vortexing,
SSx e the amount of solids quantified after x minutes of vortexing,
SS0 - the total amount of solids used for the test. The higher IC value
2.2.2. Sludge retention time
corresponds to stronger aggregates.
Suspended solids were measured in duplicate according to
APHA standard method Nr.2540D. 24 h composite effluent samples
to determine the sludge retention time (SRT) were collected on 2.4. Microscopy and image analysis
days 178 and 179. Sampling of the sludge from bottom and top of
the reactor columns was performed while flushing them with N2 2.4.1. Scanning electron microscopy (SEM)
gas to achieve good mixing. Samples of granules for SEM analysis were prepared as
The SRT was calculated as follows: described in Ismail et al. (2010) and imaged at an acceleration
voltage of 2 kV and beam current of 6 pA, at room temperature in a
TSSReactor TSSReactor field emission scanning electron microscope (Magellan 400, FEI
SRT ¼ ¼
TSSwasted TSSeff þ TSSdischarged Company, Oregon, USA).

where: SRT e sludge retention time (days); TSSReactor e the amount 2.4.2. Fluorescence in situ hybridization (FISH)
of solids within a bioreactor (gTSS); TSSwasted e the amount of solids Samples taken at the end of continuous experiment from each
wasted per day (gTSS/day); TSSeff - the amount of solids leaving UASB reactor sludge bed were directly fixed with 37% w/w form-
with the effluent (gTSS/day); TSSdischarged - the amount of solids aldehyde according to the procedure described in Amann et al.
discharged from the bioreactor (gTSS/day). Due to low frequency (1995). After fixation, samples were washed several times with
and low amounts of solids sampled and no intentional sludge PBS 1 and then stored at 20  C in ethanol:PBS 1 (1:1). Before
wasting from reactors, the term TSSdischarged was neglected. further FISH analysis fixed granules were gently crushed through a
0.7 mm diameter needle and diluted (Amann et al., 1995). Oligo-
2.3. Granular sludge characteristics nucleotide probes applied are listed in Table S2. Probes were
labelled with Cy3-red or Alexa488-green fluorophores. Samples
2.3.1. Particle size distribution (PSD) and strength were examined by epifluorescence microscopy (Olympus BX41,
Particle size distribution of the microbial granules was Olympus, Japan) equipped with Infinity Camera (Lumenera cor-
measured as described in Ismail et al. (2010). The measurement was poration, Canada). FISH images were modified using the ImageJ
performed in triplicate on a Mastersizer 2000 (Malvern, UK). software package (version1.37v, Wayne Rasband, National Institute
Sampling was done as described above. The volume and surface of Health, Bethesda, MD, USA).
based mean particle sizes were calculated based on total volume or
total surface of the particles, respectively. 3. Results
Strength of the microbial aggregates was measured as described
in Xiao et al. (2008). Approx. 0.12 g of suspended solids were used 3.1. Granulation
in each test. The microbial aggregates were put in 50 mL centrifuge
tubes, filled with reactor medium up to 30 mL and vortexed at 3.1.1. Size, strength and retention
2500 rpm for 2, 4 and 8 min. After each period of vortexing, the During start-up fast granulation (R1, R2, R4) and growth of
aggregates were allowed to settle for 1 min and the supernatant particle size (R3) was observed in all bioreactors, irrespective of the
was decanted to quantify the amount of fines produced as total salinity level or whether PAC was added to the inoculum. As an
suspended solids. To minimize biomass removal from running example, Fig. 2 illustrates the development of granules over time in
296 D. Sudmalis et al. / Water Research 128 (2018) 293e303

Fig. 2. Images of granular sludge formation in R4. A e day 4; B e day 23; C e day 33; D e day 43. White circle in B for guiding the reader's attention to the aggregates appearing in
the bioreactor. Reactor inoculated with 125 mm sieved biomass and operated at 20 g Naþ/L.

R4. Microbial aggregation was already observed on day 23 (Fig. 2 e resistant to shear forces compared to R1-R3. This difference may be
B), while on day 33 the majority of the sludge bed consisted of caused by scaled e up set e up of R4, which may have resulted in
granules (Fig. 2 e C). By day 43 the sludge bed was completely filled sludge exposure to different shear forces and mixing patterns of
with granular biomass (Fig. 2 e D). influent.
Particle size distribution (PSD) was followed over time and was
similar for R1 and R2. Fig. 3-A-B shows an increase of particle sizes
3.1.2. Microscopic analyses
in R1 and R2 operated at 5 g Naþ/L compared to the maximum
Microscopic analysis carried out using FESEM and Fluorescence
particle size of 125 mm of the inoculum. Fig. 3-C-D show that also in
microscopy coupled with FISH visualized the differences between
R3 and R4, operated at 20 g Naþ/L the particle size increased.
granules operated at 5 and 20 g Naþ/L. SEM images of granules
Table 2 summarizes the development of volume and surface based
showed that the dominant morphology of the microorganisms was
average particle sizes in all four bioreactors. Addition of PAC to the
filamentous (Fig. 4). In R1 and R2 operated at 5 g Naþ/L the most
inoculum did not result in a faster development of granules
abundant morphologies were coccus-shaped microorganisms
compared to the inoculum where no PAC was added (R1 compared
organized in chains (Fig. 4 e A,B,C,D). In R3, operated at 20 g Naþ/L,
to R2). Also, no observable difference in the rate of granulation was
filaments were also observed but these mainly consisted of
found when comparing the start eup of reactors with sieved
aggregated rod shaped cells (Fig. 4 e E, F). The difference in
biomass at 5 g Naþ/L (R2) and 20 g Naþ/L (R4).
morphology at 5 g Naþ/L and 20 g Naþ/L suggests that different
On Day 79 the volume based average size of granules in R3
microbial communities were “selected” in response to salinity. For
was higher than in R1 and R2. This difference is due to larger
further analysis FISH probes were used to target the microbial
particles inoculated in R3. Compared to the inoculum, the
groups commonly found in anaerobic treatment systems (Table S2)
average volume based size increased by 141.1 mm, 124.7 mm and
and in anaerobic environments.
121.4 mm in R1, R2 and R3, respectively. Thus, the rate at which
The use of ARC915 archaea domain specific probe, combined
particle size increased in R1, R2 and R3 was similar. The SRTs
with the application of specific probes for Methanosarcinales,
were 74.9, 98.6 and 70 days for reactors R1, R2 and R3, respec-
Methanosaetaceae and Methanomicrobiales (Table S2) revealed
tively. For R4 no solids balance was made, as it was merely
acetoclastic Methanosaetaceae as the dominant methanogen at
started up to study if the process of granulation from dispersed
both salinities (Fig. 5 - A, C, E and F in green), while members of the
biomass is possible at 20 g Naþ/L.
hydrogenotrophic order of Methanomicrobiales were present in
In Table 3 strength measurements of the granules, indicate that
much lower amounts in both R1 (Fig. 5-E) and R3 (Fig. 5-F) pop-
in all bioreactors aggregates of higher strength compared to the
ulations. In R1 rods and filaments appeared, whereas only few rods
inoculum developed. The strength of microbial granules in R1, R2
hybridized with the probes in R3. Within the bacterial population
and R3 was similar, implying that strength was not affected by the
in all samples taken at the end of the operational runs (day 217),
sodium concentration or by addition of PAC. Also, the strength of
Streptococcus spp. dominated in low salinity reactors (5 g Naþ/L)
granules formed from sieved biomass (R1 and R2) was not different
(Fig. 5-B), which corresponds to the twisted chain shaped micro-
from the strength of the granules that developed from intact
organism shown by SEM at the exterior of the granule (Fig. 4 A-D),
granules (R3). The integrity coefficients during strength tests in R1
while a long filamentous bacterium belonging to the family Lach-
e R3 stabilised with vortexing time while in R4 a decreasing trend
nospiraceae was the most represented in the granules grown at 20 g
remained. These outcomes suggest that granules in R4 were less
Naþ/L (Fig. 5-D). Both belong to the phylum Firmicutes and are
D. Sudmalis et al. / Water Research 128 (2018) 293e303 297

Table 3
Integrity coefficient measurements of sludge from reactors and of intact inoculum
sludge. n.d. not determined.

Reactor Intact inoculum R1 R2 R3 R4 R1 R2 R3 R4

Day 0 148 148 148 126 205 205 205 n.d


IC2, % 45.3 60.2 66.5 61.6 74.7 61.3 59.9 58.4 n.d
IC4, % 34.9 49.6 51.9 53.1 48.9 49.7 45.3 47.0 n.d
IC8, % 30 44.5 43.9 48.2 27.4 44.7 38.6 39.7 n.d

3.2. COD removal efficiency, fractionation in the effluent and


methane production

Fig. 6 shows that all reactors accomplished high removal effi-


ciencies regardless of the salinity level and COD loading rate. The
average removal efficiencies based on sCOD during the entire
operation period were 97.2 ± 2.1%, 96.6 ± 3.3%, 94.0 ± 2.2% and
92.8 ± 5.4% for R1, R2, R3 and R4, respectively, showing that the
reactors operated at 5 g Naþ/L (R1 and R2) performed slightly better
than the ones operated at 20 g Naþ/L (R3 and R4). No clear differ-
ences were observed between reactors seeded and not seeded with
PAC (R1 and R2) at any stage of the continuous operation, thus
showing no effect of additional PAC on start e up of R1 and R2.
Effluent tCOD was higher in reactors operated at 20 g Naþ/L. In
composite effluent samples at day 179 (day 100 for R4), tCOD was
257 mg/L, 330 mg/L, 769 mg/L and 1170 mg/L for R1, R2, R3 and R4,
respectively. The pCOD concentrations in the effluent were 31, 89.5,
120.8 and 222.5 mg/L for R1, R2, R3 and R4, respectively. The
fractions contributing to total effluent COD were similar for all re-
actors and are depicted in Fig. 7-A.
The detected VFAs in the effluent were acetate and propionate.
The sum of these VFAs contributed between 29.5 and 60.5% to the
effluent soluble COD (Fig. 7-B). The highest VFA fraction was found
in R3 (operated at 20 g Naþ/L) and the lowest fraction in R2
(operated at 5 g Naþ/L). These results show that between 39.5 and
70.5% of the soluble effluent COD consisted of organic material
other than VFAs.
The profiles of effluent VFAs are shown in Fig. S3. Generally,
lower effluent acetate and propionate concentrations were
measured in R1 and R2 compared to R3 and R4. However, both in
Fig. 3. Particle size distribution in time. Particle size distribution presented as % of total R3 and R4 the effluent VFA concentrations showed a downward
volume of the measured particles. The vertical lines in A, B, D represent the maximum trend, indicating an improvement of reactor performance over
particle size of inoculum for R1, R2 and R4. Samples were taken from top sampling time. The effluent acetate concentration at the end of operational
ports of the bioreactors, while mixing the reactor with nitrogen gas.
runs of R3 and R4 were 85 and 34.7 mg/L, respectively, whereas
propionate concentration was 34.8 mg/L in R3 and could not be
detected in R4.
reported as carbohydrate fermenting bacteria (Cotta and Forster,
Cumulative biogas production as well as biogas composition
2006; Milinovich et al., 2008). Apparently, salinity resulted in a
were monitored over time. Biogas composition over the whole
shift in the bacterial and hydrogenotrophic methanogens pop-
period was similar in the reactors operated at both 5 g (R1 and R2)
ulations, while the other methanogenic biomass was not affected
and 20 g Naþ/L (R3 and R4): 63.6 ± 4.6%, 65.9 ± 6.5%, 68.4 ± 4.3%
by increasing sodium concentration.
and 70 ± 2.9% for R1, R2, R3 and R4, respectively. Biogas production
No other specific group of microorganisms could be visualized
curves are presented in Fig. S4 and show continuous biogas
when applying all probes listed in Table S2.

Table 2
Development of volume and surface based average particle sizes within bioreactors.

Reactor Naþ, Inoculum Day 79 surface Day 87 surface Day 148 surface Day 205 surface Day 79 vol Day 87 vol Day 148 vol Day 205 vol
g/L Typea,b based, mm based, mm based, mm based, mm based, mm based, mm based, mm based, mm

R1 5 Sieved 96.3 e 74.4 75.8 220.7 e 303.8 329.6


R2 5 Sievedþ PAC 64.3 e 61.5 73.5 204.0 e 295.20 289.1
R3 20 Intact þ PAC 84.7 e 98.2 103 426.9 e 368.4 384.7
R4 20 Sieved þ PAC e 74.5 e e e 250.5 e e
e 8 Intact
e 8 Sieved
a
Surface and volume based average diameter of intact inoculum is 82.3 and 305.5 mm, respectively.
b
Surface and volume based average diameter of sieved inoculum is 37.9 and 79.3 mm, respectively.
298 D. Sudmalis et al. / Water Research 128 (2018) 293e303

Fig. 4. FESEM images of crushed granules from reactors. A, B e samples from R1 operated at 5 g Naþ/L; C, D e samples from R2 operated at 5 g Naþ/L and seeded with PAC upon
start-up; E, F e samples from R3 operated at 20 g Naþ/L and seeded with PAC upon start-up.

production with increasing slopes as the loading rates increase. Robust granules were formed, even though the wastewater only
contained a (background) concentration of 13 mg/L of calcium
4. Discussion during the first 3 months of operation. Multivalent cations are
known to play a crucial role in the formation of microbial granules/
4.1. Granulation and start e up period biofilms by bridging negatively charged groups of EPS (Mahoney
et al., 1987; Yu et al., 2001; Vlyssides et al., 2009; Flemming and
4.1.1. Rapid granulation and start eup at high salinity: role of Wingender, 2010), although their role with respect to effects on
cations and PAC rates of granulation is somewhat controversial. Mahoney et al.
In our UASB reactors a rapid granular sludge formation from (1987) and Yu et al. (2001) reported that dosing of 100e300 mg
dispersed biomass at 5 g Naþ/L and even at 20 g Naþ/L occurred Ca2þ/L increases the rates of granulation and biomass retention
(Figs. 2 and 3). The first granules were already observed after 23 during UASB start e up, whereas Hulshoff Pol (1989) and Guiot
days (Fig. 2). This is in contrast with Ismail et al. (2008) who et al. (1988) found no positive effect upon addition of
showed that considerable washout of (granular) biomass from 150e640 mg Ca2þ/L at non e saline conditions. Since at high Naþ
UASB reactors took place, even at salinity levels as low as 5 g Naþ/L. concentrations Ca2þ is displaced in the EPS matrix of granules,
D. Sudmalis et al. / Water Research 128 (2018) 293e303 299

Fig. 5. FISH images of samples taken from R1 (5 g Naþ/L, in A, B and E) and R3 (20 g Naþ/L, in C, D and F) at the end of the reactor runs. In A and C samples were hydribized with the
EUB338mix probe (domain Bacteria, red) and the ARC915 probe (domain Archaea - green). In B and D the EUB338mix probe (red) was applied together with specific probes. In B,
Streptococcus sp. was identified in reactor R1 samples by using Strept probe (in green). In D, filamentous bacteria of reactor R3 were further identified with the probe LAC435
(green). In E and F, ARC915 probe (green) was applied together with Methanobacteriales specific probe MG1200b, and arrows indicate archaeal cells positive to both probes. Details
on the probes are reported in Table S2. Yellow bar is 10 mm. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this
article.)

thereby weakening the granular structure (Jeison et al., 2008; faster granulation upon addition of PAC to inoculum. However, the
Ismail et al., 2010), it would be expected that at highly saline con- amount of PAC that was mixed with the inoculum in our study was
ditions additional Ca2þ dosing would be needed to promote gran- only 0.02 compared to 0.24 gPAC/gVSS in experiments of Yu et al.
ulation. Here we demonstrated that at highly saline conditions (1999). Zhou et al. (2015) demonstrated that the diameter of the
granulation rate can be fast without dosing of additional calcium, activated carbon can be a determining factor for enhancing the
probably due to gelling of EPS with sodium ions, as discussed in granulation process of sludge due to velocity field differences be-
paragraph 4.1.3. tween the flocs and activated carbon. It was concluded that 0.33 g
For all reactors a start e up period (as defined by Ghangrekar GAC/gVSS with a diameter of 0.2 mm enhances microbial granu-
et al. (1996)) of only 3 weeks was necessary at both salinity levels lation and reactor start-up. Also Yu et al. (1999) used PAC with a
and regardless of the inoculum being sieved or intact (Fig. 6 A-D). diameter of 0.2 mm, whereas in our study the diameter of PAC was
This period coincided with the appearance of the first granules 0.037e0.149 mm. Due to differences in both type and quantity of
(Fig. 2 e B) and was as fast as reported at non e saline conditions, PAC used in our study, a definitive conclusion on, why additional
with operational parameters chosen for enhanced UASB start - up PAC had no effect on process performance and granulation cannot
(Ghangrekar et al., 1996). be made.
Interestingly, during the start-up no sizable difference was
observed in process performance between R1 (no PAC added) and 4.1.2. Microbial key players
R2 (0.1% of inoculum mass PAC) and in the development of the SEM (Fig. 4) and FISH (Fig. 5) analysis show that different bac-
particle size (Fig. 3 A-B) as opposed to Yu et al. (1999), who reported terial communities developed at different salinity levels, even
300 D. Sudmalis et al. / Water Research 128 (2018) 293e303

Fig. 6. COD loading rates and sCOD removal efficiencies. VLR e volumetric loading rate. A e R1 (at 5 g Naþ/L), B e R2 (at 5 g Naþ/L), C e R3 (at 20 g Naþ/L), D e R4 (at 20 g Naþ/L).
Dotted lines indicate the periods with different influent COD concentrations.

A 100
B
257 330 769 1170
100
% of VFA COD in soluble

31 90
effluent COD fraction

121 223
% of total effluent COD

75
75
113 87 323 512
50 50

25 113 153 325 435 25

0
R1 R2 R3 R4 0
cCOD sCOD pCOD R1 R2 R3 R4

Fig. 7. 40 h composite effluent sample COD fractions collected on day 179 in R1, R2, R3 and on day 100 in R4. A e COD fractions in the effluent; B e fraction of effluent sCOD as
acetate and propionate. Numbers above bars in A show total effluent COD in mg/L, numbers next to the fractions show their concentrations in mg/L.

though the same organic substrate and inoculum were used. aggregation.
Within the archaeal population, a member of the Methanosaetaceae This is the first time that Streptococcus sp. and Lachnospiraceae
was the dominant methanogen in all reactors studied. Meanwhile, family members were identified at salinity levels of 5e20 g Naþ/L,
the hydrogenotrophic Methanomicrobiales decreased in abundance and as key population within anaerobic granules. FISH analysis
with increasing salinity. confirmed their association with Methanosaeta clusters (Fig. 5-A, C).
It has been hypothesized by Wiegant (1988), that a precursor of In the 20 g Naþ/L granules the filamentous members of the family
granulation is formation of filamentous Methanosaeta nuclei, to Lachnospiraceae and Methanosaeta clusters were uniformly
which other microbial cells can attach. The aggregation properties distributed within the granule (Fig. 4 E-F), with a clear dominance
of Methanosaeta were also found to be crucial during anaerobic of the methanogens along the granular structure (Fig. 5-C). The
digestion in several other studies (Calli et al., 2003; Angenent et al., close connection between the methanogenic and bacterial com-
2004; Li et al., 2015; Gagliano et al., 2017a). In R1 and R3 granules, munity in the granules suggests the presence of a syntrophic
Methanosaeta-like cells and filaments were always detected in large metabolism which, given the substrate composition of the syn-
clusters, forming a nucleus for aggregation, to which a bacterial thetic wastewater stream and the microbial metabolic character-
population could attach (Fig. 5). This demonstrated the possible istics, could be based on glucose fermentation coupled with
fundamental role of this acetoclastic methanogen in granulation, as methane production. The presence of Methanomicrobiales was
already discussed by Gagliano et al. (2017b). However, the presence already shown in highly saline natural environments and waste-
of the two bacteria as filaments may also have contributed to the water treatment systems (Webster et al., 2014; De Vrieze et al.,
D. Sudmalis et al. / Water Research 128 (2018) 293e303 301

2016b) and suggests that hydrogenotrophic methanogenesis was 2007; Jeison et al., 2008; Li et al., 2014; De Vrieze et al., 2016a), our
taking place, most likely coupled with either protein or glucose findings are in agreement with de Baere et al. (1984), who have
fermentation (Stams, 1994). However, FISH analysis could not reported salt tolerance of anaerobic biomass similar to the one
clarify the utilization of the proteinaceous substrate, which could demonstrated here. Potentially both the substrate and inoculum
be partially used for the EPS build-up by one of microbial (adapted to 8 g Naþ/L) used, may have influenced the positive
protagonists. outcome of this research (McCarty and McKinney, 1961; Patel and
Roth, 1977; Lefebvre et al., 2007).
4.1.3. Strength of granules From bioenergetics point of view in most prokaryotes, uptake of
The strength of the granules at low and high salinity was similar osmoprotectants from surrounding liquid is preferred over de novo
and the granules that were formed were stronger than the granules synthesis (Oren, 1999; Roebler and Müller, 2001; Gunde-Cimerman
in the inoculum (Table 3). This is in conflict with the results of et al., 2006). Since also amino acids and their derivatives can act as
others using the same inoculum (Ismail et al., 2008, 2010; Jeison osmolytes (Roberts, 2005), tryptone fed to the reactors during our
et al., 2008). Also, the high values of 70e98.6 days found for SRTs study may have acted as source of “cheap” osmoprotectants after
confirmed that robust granules were formed. The different out- undergoing hydrolysis. Also in experiments of de Baere et al. (1984),
comes between ours and the other studies may be explained by who reported salt tolerance of anaerobic biomass similar to one
different carbon sources used, a factor shown to affect chemical presented here, peptone (tryptic soy) was added to synthetic
composition of EPS (Wolfaardt et al., 1998), which can in turn affect wastewater, in contrast to other studies reporting sodium inhibi-
the gelling properties of these polymers. tion in the absence of such substances.
Both monovalent and divalent cations can form gels with micro- Use of salt adapted microorganisms as inoculum was suggested
bial polysaccharides (e.g., gellan) (Chandrasekaran and Radha, 1995) as one of the most suitable strategies to overcome sodium toxicity
and alginate-like EPS are likely to bind both calcium and sodium ions in anaerobic wastewater treatment (Vyrides, 2015). The inoculum
(Seviour et al., 2012; Stewart et al., 2014), meaning that both can play used in our study had been adapted to 8 g Naþ/L, which is
a role in granulation at high salinity. Calcium ions will form stronger considerably lower than the 20 g Naþ/L applied in R3 and R4.
gels than monovalent cations because carboxylate—Ca2þ—carbox- However, in the few studies where high methanogenic activity was
ylate interactions are stronger than those of, for example carbox- found at comparable salinities, the inoculum was adapted to similar
ylate—Naþ—water—Naþ—carboxylate (Chandrasekaran and Radha, salinity as reported here (Ismail et al., 2010; Gagliano et al., 2017b),
1995). Due to a much higher concentration of sodium over calcium demonstrating the importance of inoculum at high salinities.
in the influent wastewater of all reactors in our study, sodium is very
likely to play an important role in the process of granulation, thus 5. Conclusions
explaining the similar strength of granules measured both at 5 g Naþ/
L and 20 g Naþ/L. Fig. S5 shows that the concentration of sodium is Granulation of anaerobic biomass from dispersed inoculum is
5.75 and 22.79 g/L in granular sludge of R1 and R3, respectively, possible, even at sodium concentrations as high as 20 g Naþ/L,
whereas the concentration of calcium is 0.159 and 0.037 in the which was not reported in literature before. Moreover, calcium
granular sludge of R1 and R3. The preliminary outcomes of metal concentration as low as 13 mg/L is sufficient to form microbial
analysis within the granular sludge indeed indicate, that further granules at 20 g Naþ/L and is around ten times lower than the
investigation of the abovementioned hypothesis is needed to further traditionally reported 100e150 mg Ca2þ/L in non-saline conditions.
understand the gelling properties of EPS formed at high monovalent The granulation process at 20 g Naþ/L can be accompanied by
salt concentration. efficient soluble COD removal, however the removal efficiency is
still lower (92.8 ± 5.4% to 94.0 ± 2.2%) compared to reactors
4.2. Process performance operated with the same substrate at 5 g Naþ/L (96.6 ± 3.3% to
97.2 ± 2.1%).
4.2.1. Performance of bioreactors at 5 and 20 g Naþ/L: role of Interestingly, addition of powdered activated carbon to inoc-
substrate and inoculum biomass ulum does not guarantee a faster reactor start-up.
The reactors R1, R2 and R3 have been continuously operated for High salinity acted as a selective pressure on the dominant
more than 200 days and reactor R4 for more than 170 days with active microbial (Streptococcus spp at 5 g Naþ/L and bacterium
stable process performance (Fig. 6). This shows that with an belonging to Lachnospiraceae at 20 g Naþ/L) and hydrogenotrophic
appropriate inoculum selection, start-up procedure and operation, archaeal community (decreasing abundance of Methanomicrobiales
highly saline waste streams can be treated in UASB reactors at at elevated salinity). Interestingly, the active acetoclastic archaeal
loading rates as high as 16 g COD/m3$d. community remained similar within the range of 5e20 g Naþ/L
Even though good COD removal efficiencies were obtained at 5 (Methanosaeta as dominant methanogen at both salinities).
and 20 g Naþ/L (Fig. 6 A-D), it becomes clear that total effluent COD
concentrations around 1000 mg/L require a further polishing step. Acknowledgements
The higher effluent COD concentration at 20 g Naþ/L can be
explained by a lower methanogenic activity (Li et al., 2014; De This research is financed by the Netherlands Organisation for
Vrieze et al., 2016a), sloughing of loosely bound EPS, a higher Scientific Research (NWO), which is partly funded by the Ministry of
production of soluble EPS (Corsino et al., 2017), washout of indi- Economic Affairs, and co-financed by the Netherlands Ministry of
vidual cells and by a lower affinity of the microorganisms to the Infrastructure and Environment and partners of the Dutch Water
substrate, since different microbial communities developed at 5 Nexus consortium (project nr. STW 14300 Water Nexus 2.1). Addi-
and 20 g Naþ/L (Fig. 5). Fig. 7 shows, that colloidal and soluble tionally we would like to thank dr. ing. Marcel Giesbers from
fractions were contributing to the largest effluent COD difference Wageningen Electron Microscopy Centre for providing support and
between 5 and 20 g Naþ/L reactors. In these fractions both EPS and training on electron microscopy and the technical and analytical
microorganisms can potentially contribute to the COD content. support team of the Sub-department of Environmental Technology
Although deteriorating process performance and operational Vinnie de Wilde, Bert Willemsen, Jean Slangen, Livio Carlucci, Hans
problems were found by other authors at levels close to or below Beijleveld and Ilse Gerrits. A special thank you to Zacchariah Ross for
5 g Naþ/L (Patel and Roth, 1977; Rinzema et al., 1988; Lefebvre et al., the final proof reading and correcting the English of this manuscript.
302 D. Sudmalis et al. / Water Research 128 (2018) 293e303

Appendix A. Supplementary data granules. Biotechnol. Adv. 27 (6), 1061e1070.


Liu, Y., Boone, D.R., 1991. Effects of salinity on methanogenic decomposition. Bio-
resour. Technol. 35 (3), 271e273.
Supplementary data related to this article can be found at Lu, Y., Slater, F., Bello-Mendoza, R., Batstone, D.J., 2013. Shearing of biofilms enables
https://doi.org/10.1016/j.watres.2017.10.038. selective layer based microbial sampling and analysis. Biotechnol. Bioeng. 110
(10), 2600e2605.
Mahoney, E., Varangu, L., Cairns, W., Kosaric, N., Murray, R., 1987. The effect of
References calcium on microbial aggregation during UASB reactor start-up. Water Sci.
Technol. 19 (1e2), 249e260.
Amann, R.I., Ludwig, W., Schleifer, K.-H., 1995. Phylogenetic identification and in McCarty, P.L., McKinney, R.E.C.F.p.d.A., 1961. Salt toxicity in anaerobic digestion.
situ detection of individual microbial cells without cultivation. Microbiol. Rev. J. Water Pollut. Control Fed. 33 (4), 399e415.
59 (1), 143e169. McHugh, S., O'reilly, C., Mahony, T., Colleran, E., O'flaherty, V., 2003. Anaerobic
Angenent, L.T., Sung, S., Raskin, L., 2004. Formation of granules and Methanosaeta granular sludge bioreactor technology. Rev. Environ. Sci. Biotechnol. 2 (2e4),
fibres in an anaerobic migrating blanket reactor (AMBR). Environ. Microbiol. 6 225e245.
(4), 315e322. Milinovich, G.J., Burrell, P.C., Pollitt, C.C., Bouvet, A., Trott, D.J., 2008. Streptococcus
Calli, B., Mertoglu, B., Tas, N., Inanc, B., Yenigun, O., Ozturk, I., 2003. Investigation of henryi sp. nov. and Streptococcus caballi sp. nov., isolated from the hindgut of
variations in microbial diversity in anaerobic reactors treating landfill leachate. horses with oligofructose-induced laminitis. Int. J. Syst. Evol. Microbiol. 58 (1),
Water Sci. Technol. 48 (4), 105e112. 262e266.
Chandrasekaran, R., Radha, A., 1995. Molecular architectures and functional prop- O'flaherty, V., Lens, P., De Beer, D., Colleran, E., 1997. Effect of feed composition and
erties of gellan gum and related polysaccharides. Trends Food Sci. Technol. 6 (5), upflow velocity on aggregate characteristics in anaerobic upflow reactors. Appl.
143e148. Microbiol. Biotechnol. 47 (2), 102e107.
Corsino, S.F., Capodici, M., Torregrossa, M., Viviani, G., 2017. Physical properties and Oren, A., 1999. Bioenergetic aspects of halophilism. Microbiol. Mol. Biol. Rev. 63 (2),
Extracellular Polymeric Substances pattern of aerobic granular sludge treating 334e348.
hypersaline wastewater. Bioresour. Technol. 229, 152e159. Ozgun, H., Ersahin, M.E., Erdem, S., Atay, B., Kose, B., Kaya, R., Altinbas, M., Sayili, S.,
Cotta, M., Forster, R., 2006. The Prokaryotes. Springer, pp. 1002e1021. Hoshan, P., Atay, D., 2013. Effects of the pre-treatment alternatives on the
de Baere, L.A., Devocht, M., Van Assche, P., Verstraete, W., 1984. Influence of high treatment of oil-gas field produced water by nanofiltration and reverse osmosis
NaCl and NH4Cl salt levels on methanogenic associations. Water Res. 18 (5), membranes. J. Chem. Technol. Biotechnol. 88 (8), 1576e1583.
543e548. Park, C., Muller, C.D., Abu-Orf, M.M., Novak, J.T., 2006. The effect of wastewater
De Vrieze, J., Coma, M., Debeuckelaere, M., Van der Meeren, P., Rabaey, K., 2016a. cations on activated sludge characteristics: effects of aluminum and iron in floc.
High salinity in molasses wastewaters shifts anaerobic digestion to carboxylate Water Environ. Res. 78 (1), 31e40.
production. Water Res. 98, 293e301. Patel, G., Roth, L., 1977. Effect of sodium chloride on growth and methane pro-
De Vrieze, J., Regueiro, L., Props, R., Vilchez-Vargas, R., Ja uregui, R., Pieper, D.H., duction of methanogens. Can. J. Microbiol. 23 (7), 893e897.
Lema, J.M., Carballa, M., 2016b. Presence does not imply activity: DNA and RNA Rinzema, A., van Lier, J., Lettinga, G., 1988. Sodium inhibition of acetoclastic
patterns differ in response to salt perturbation in anaerobic digestion. Bio- methanogens in granular sludge from a UASB reactor. Enzyme Microb. Technol.
technol. Biofuels 9 (1), 244. 10 (1), 24e32.
Dulekgurgen, E., Dog ruel, S., Karahan, O., € Orhon, D., 2006. Size distribution of Roberts, M.F., 2005. Organic compatible solutes of halotolerant and halophilic mi-
wastewater COD fractions as an index for biodegradability. Water Res. 40 (2), croorganisms. Saline Syst. 1 (1), 5.
273e282. Roeßler, M., Müller, V., 2001. Osmoadaptation in bacteria and archaea: common
Fakhru’l-Razi, A., Pendashteh, A., Abdullah, L.C., Biak, D.R.A., Madaeni, S.S., principles and differences. Environ. Microbiol. 3 (12), 743e754.
Abidin, Z.Z., 2009. Review of technologies for oil and gas produced water Seviour, T., Yuan, Z., van Loosdrecht, M.C., Lin, Y., 2012. Aerobic sludge granulation: a
treatment. J. Hazard. Mater. 170 (2), 530e551. tale of two polysaccharides? Water Res. 46 (15), 4803e4813.
Flemming, H.-C., Wingender, J., 2010. The biofilm matrix. Nat. Rev. Microbiol. 8 (9), Sleator, R.D., Hill, C., 2002. Bacterial osmoadaptation: the role of osmolytes in
623e633. bacterial stress and virulence. FEMS Microbiol. Rev. 26 (1), 49e71.
Freire, D., Cammarota, M., Sant'Anna, G., 2001. Biological treatment of oil field Sochacki, K.A., Shkel, I.A., Record, M.T., Weisshaar, J.C., 2011. Protein diffusion in the
wastewater in a sequencing batch reactor. Environ. Technol. 22 (10), 1125e1135. periplasm of E. coli under osmotic stress. Biophys. J. 100 (1), 22e31.
Gagliano, M., Gallipoli, A., Rossetti, S., Braguglia, C., 2017a. Efficacy of methanogenic Soto, M., Me ndez, R., Lema, J., 1991. Biodegradability and toxicity in the anaerobic
biomass acclimation in mesophilic anaerobic digestion of ultrasound pretreated treatment of fish canning wastewaters. Environ. Technol. 12 (8), 669e677.
sludge. Environ. Technol. 1e25 (just-accepted). Stams, A.J., 1994. Metabolic interactions between anaerobic bacteria in methano-
Gagliano, M., Ismail, S., Stams, A., Plugge, C., Temmink, H., Van Lier, J., 2017b. Biofilm genic environments. Ant. Van Leeuwenhoek 66 (1e3), 271e294.
formation and granule properties in anaerobic digestion at high salinity. Water Steinbusch, K.J.J., Hamelers, H.V.M., Buisman, C.J.N., 2008. Alcohol production
Res. 121, 61e71. through volatile fatty acids reduction with hydrogen as electron donor by
Ghangrekar, M., Asolekar, S., Ranganathan, K., Joshi, S., 1996. Experience with UASB mixed cultures. Water Res. 42 (15), 4059e4066.
reactor start-up under different operating conditions. Water Sci. Technol. 34 (5), Stewart, M.B., Gray, S.R., Vasiljevic, T., Orbell, J.D., 2014. The role of poly-M and poly-
421e428. GM sequences in the metal-mediated assembly of alginate gels. Carbohydr.
Guiot, S., Gorur, S., Bourque, D., Samson, R., 1988. Metal effect on microbial aggre- Polym. 112, 486e493.
gation during upflow anaerobic sludge bed-filter (UBF) reactor start-up. Granul. Tellez, G.T., Nirmalakhandan, N., Gardea-Torresdey, J.L., 2002. Performance evalua-
Anaerob. Sludge Microbiol. Technol. 187e194. tion of an activated sludge system for removing petroleum hydrocarbons from
Gunde-Cimerman, N., Oren, A., Plemenitas, A., 2006. Adaptation to Life at High Salt oilfield produced water. Adv. Environ. Res. 6 (4), 455e470.
Concentrations in Archaea, Bacteria, and Eukarya. Springer Science & Business Tellez, G.T., Nirmalakhandan, N., Gardea-Torresdey, J.L., 2005. Kinetic evaluation of a
Media. field-scale activated sludge system for removing petroleum hydrocarbons from
Hulshoff Pol, L., 1989. The Phenomenon of Granulation of Anaerobic Sludge. oilfield-produced water. Environ. Prog. 24 (1), 96e104.
Landbouwuniversiteit te Wageningen. Vallero, M., Pol, L.H., Lens, P., Lettinga, G., 2002. Effect of high salinity on the fate of
Ismail, S., Gonzalez, P., Jeison, D., Van Lier, J., 2008. Effects of high salinity waste- methanol during the start-up of thermophilic (55 C) sulfate reducing reactors.
water on methanogenic sludge bed systems. Water Sci. Technol. 58 (10), Water Sci. Technol. 45 (10), 121e126.
1963e1970. Vieira, D.S., Servulo, E.F.C., Cammarota, M., 2005. Degradation potential and growth
Ismail, S., de La Parra, C., Temmink, H., Van Lier, J., 2010. Extracellular polymeric of anaerobic bacteria in produced water. Environ. Technol. 26 (8), 915e922.
substances (EPS) in upflow anaerobic sludge blanket (UASB) reactors operated Vlyssides, A., Barampouti, E.M., Mai, S., 2009. Influence of ferrous iron on the
under high salinity conditions. Water Res. 44 (6), 1909e1917. granularity of a UASB reactor. Chem. Eng. J. 146 (1), 49e56.
Ismail, S.B., 2013. Anaerobic Wastewater Treatment of High Salinity Wastewaters: Vyrides, I., Stuckey, D., 2009. Adaptation of anaerobic biomass to saline conditions:
Impact on Bioactivity and Biomass Retention. Wageningen University. role of compatible solutes and extracellular polysaccharides. Enzyme Microb.
Jeison, D., Rio, A.D., Lier, J.V., 2008. Impact of high saline wastewaters on anaerobic Technol. 44 (1), 46e51.
granular sludge functionalities. Water Sci. Technol. 57 (6), 815e820. Vyrides, I., Santos, H., Mingote, A., Ray, M., Stuckey, D., 2010. Are compatible solutes
Lefebvre, O., Moletta, R., 2006. Treatment of organic pollution in industrial saline compatible with biological treatment of saline wastewater? Batch and contin-
wastewater: a literature review. Water Res. 40 (20), 3671e3682. uous studies using submerged anaerobic membrane bioreactors (SAMBRs).
Lefebvre, O., Quentin, S., Torrijos, M., Godon, J.J., Delgenes, J.P., Moletta, R., 2007. Environ. Sci. Technol. 44 (19), 7437e7442.
Impact of increasing NaCl concentrations on the performance and community Vyrides, I., 2015. Environmental Microbial Biotechnology. Springer, pp. 105e117.
composition of two anaerobic reactors. Appl. Microbiol. Biotechnol. 75. Webster, G., O'sullivan, L.A., Meng, Y., Williams, A.S., Sass, A.M., Watkins, A.J.,
Li, J., Yu, L., Yu, D., Wang, D., Zhang, P., Ji, Z., 2014. Performance and granulation in an Parkes, R.J., Weightman, A.J., 2014. Archaeal community diversity and abun-
upflow anaerobic sludge blanket (UASB) reactor treating saline sulfate waste- dance changes along a natural salinity gradient in estuarine sediments. FEMS
water. Biodegradation 25 (1), 127e136. Microbiol. Ecol. 91 (2), 1e18.
Li, L., Zheng, M., Ma, H., Gong, S., Ai, G., Liu, X., Li, J., Wang, K., Dong, X., 2015. Wiegant, W., 1988. The spaghetti theory on anaerobic granular sludge formation, or
Significant performance enhancement of a UASB reactor by using acyl homo- the inevitability of granulation. Granul. Anaerob. Sludge 146e152.
serine lactones to facilitate the long filaments of Methanosaeta harundinacea Wirtz, R.A., Dague, R.R., 1996. Enhancement of granulation and start-up in the
6Ac. Appl. Microbiol. Biotechnol. 99 (15), 6471e6480. anaerobic sequencing batch reactor. Water Environ. Res. 68 (5), 883e892.
Liu, X.-W., Sheng, G.-P., Yu, H.-Q., 2009. Physicochemical characteristics of microbial Wolfaardt, G., Lawrence, J., Robarts, R., Caldwell, D., 1998. In situ characterization of
D. Sudmalis et al. / Water Research 128 (2018) 293e303 303

biofilm exopolymers involved in the accumulation of chlorinated organics. UASB reactor start-up. Water Res. 35 (4), 1052e1060.
Microb. Ecol. 35 (3e4), 213e223. Zeeuw, W., 1984. Acclimatization of Anaerobic Sludge for UASB-reactor Start-up.
Xiao, F., Yang, S., Li, X., 2008. Physical and hydrodynamic properties of aerobic Editora. Agricultural University of Wageningen.
granules produced in sequencing batch reactors. Sep. Purif. Technol. 63 (3), Zhao, X., Wang, Y., Ye, Z., Borthwick, A.G.L., Ni, J., 2006. Oil field wastewater treat-
634e641. ment in Biological Aerated Filter by immobilized microorganisms. Process
Xiao, Y., Roberts, D.J., 2010. A review of anaerobic treatment of saline wastewater. Biochem. 41 (7), 1475e1483.
Environ. Technol. 31 (8e9), 1025e1043. Zhou, J.-h., Zhao, H., Hu, M., Yu, H.-t., Xu, X.-y., Vidonish, J., Alvarez, P.J.J., Zhu, L.,
Yu, H., Tay, J., Fang, H.H., 1999. Effects of added powdered and granular activated 2015. Granular activated carbon as nucleating agent for aerobic sludge granu-
carbons on start-up performance of UASB reactors. Environ. Technol. 20 (10), lation: effect of GAC size on velocity field differences (GAC versus flocs) and
1095e1101. aggregation behavior. Bioresour. Technol. 198, 358e363.
Yu, H., Tay, J., Fang, H.H., 2001. The roles of calcium in sludge granulation during

You might also like