You are on page 1of 10

Bioresource Technology 234 (2017) 310–319

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Optimization of hydrogen dispersion in thermophilic up-flow reactors


for ex situ biogas upgrading
Ilaria Bassani a, Panagiotis G. Kougias a,⇑, Laura Treu a,b, Hugo Porté a, Stefano Campanaro c,
Irini Angelidaki a
a
Department of Environmental Engineering, Technical University of Denmark, Kgs. Lyngby DK-2800, Denmark
b
Department of Agronomy, Food, Natural Resources, Animals and Environment (DAFNAE), University of Padua, viale dell’Università 16, 35020 Legnaro (PD), Italy
c
Department of Biology, University of Padua, Via U. Bassi 58/b, 35121 Padova, Italy

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 Ex situ biogas upgrading to 96% CH4 is


achieved in thermophilic up-flow
reactors.
 The totality of H2 and CO2 is converted
to CH4 reaching 0.25 LCH4/LH2 CH4
yield.
 Higher gas recirculation and diffusers’
pore size improved gas-liquid contact.
 H2 selectively enhances
hydrogenotrophic methanogens and
syntrophic bacteria.

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

Article history: This study evaluates the efficiency of four novel up-flow reactors for ex situ biogas upgrading converting
Received 11 January 2017 externally provided CO2 and H2 to CH4, via hydrogenotrophic methanogenesis. The gases were injected
Received in revised form 6 March 2017 through stainless steel diffusers combined with alumina ceramic sponge or through alumina ceramic
Accepted 8 March 2017
membranes. Pore size, input gas loading and gas recirculation flow rate were modulated to optimize
Available online 12 March 2017
gas-liquid mass transfer, and thus methanation efficiency. Results showed that larger pore size diffusion
devices achieved the best kinetics and output-gas quality converting all the injected H2 and CO2, up to
Keywords:
3.6 L/LREACTORd H2 loading rate. Specifically, reactors’ CH4 content increased from 23 to 96% and the
Ex situ biogas upgrading
Gas-liquid mass transfer rate
CH4 yield reached 0.25 LCH4/LH2. High throughput 16S rRNA gene sequencing revealed predominance of
Ceramic sponge bacteria belonging to Anaerobaculum genus and to uncultured order MBA08. Additionally, the massive
Ceramic membrane increase of hydrogenotrophic methanogens, such as Methanothermobacter thermautotrophicus, and syn-
16S rRNA gene sequencing trophic bacteria demonstrates the selection-effect of H2 on community composition.
Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction

The reduction of greenhouse gases emissions, such as CO2, is of


great importance for a sustainable industrial development and to
move towards cleaner energy production and utilization. Conse-
⇑ Corresponding author at: Department of Environmental Engineering, Technical
quently, several technologies have been developed to remove, con-
University of Denmark, Bld 113, 2800 Lyngby, Denmark.
vert or store CO2 (Mikkelsen et al., 2010). In this context, biological
E-mail address: panak@env.dtu.dk (P.G. Kougias).

http://dx.doi.org/10.1016/j.biortech.2017.03.055
0960-8524/Ó 2017 Elsevier Ltd. All rights reserved.
I. Bassani et al. / Bioresource Technology 234 (2017) 310–319 311

carbon fixation emerged as a very promising alternative to existing as mixing speed (Kramer and Bailey, 1991; Luo and Angelidaki,
methods, offering the advantage of working in mild conditions of 2012), gas recirculation (Guiot et al., 2011) and H2 diffusion device
temperature and pressure and without the use of chemical sub- (Díaz et al., 2015; Luo and Angelidaki, 2013). Previous studies on
stances (Alitalo et al., 2015; Lee et al., 2012). Biological carbon fix- H2 assisted biogas upgrading reported kL a ranging from 1 to
ation besides capturing CO2, can also convert it into different 10⁄103 day1, depending on applied diffusion device and selected
products, such as CH4. operating conditions (i.e. mixing speed, gas recirculation flow rate)
Anaerobic digestion (AD) of biomass is an effective method for (Díaz et al., 2015; Luo et al., 2012; Luo and Angelidaki, 2013).
bioenergy production. The resulting biogas consists of 50–70% The current study aimed at maximizing H2 uptake by methano-
CH4 and 30–50% CO2 and it is commonly burned in a combined gens, by increasing the H2 gas-liquid mass transfer rate, so as to
heat and power unit (CHP) to produce electricity and heat. Nowa- improve the efficiency of ex situ biogas upgrading. This was
days, there is an increasing interest to expand the utilization achieved by designing and operating four novel up-flow configura-
potentials of biogas by a process called biogas upgrading. Through tions, in which the gas was introduced to the reactors through dif-
this process, the CH4 content in the biogas is increased by remov- ferent gas distribution devices, i.e. stainless steel diffusers and
ing or transforming the contained CO2 (and Supplementary due to ceramic sponge or ceramic membrane. Moreover, different pore
removal of other impurities, such as water, hydrogen sulfide and sizes of the distribution devices, input gas flow rate ðQ IN Þ and gas
siloxanes) (Kougias et al., 2017). Thereafter, the upgraded biogas recirculation rate ðQ RC Þ were tested to improve H2 and CO2 conver-
can be injected into the natural gas grid or can be used as fuel sion to CH4. Finally, the microbial profile at the beginning and at
for vehicles (Deng and Hägg, 2010). the end of the process was analyzed via high throughput 16S rRNA
Moreover, H2 production from renewable sources, such as wind amplicon sequencing, to elucidate the microbial community
and solar power, is gaining greater importance in the context of change driven by the applied operating conditions.
alternative energy development. In fact, in northern EU countries,
such as Denmark, >26% of the electricity produced from wind mills 2. Materials and methods
is a temporary surplus and although some of it is exported to
neighbouring countries, the potential of wind mills is not fully uti- 2.1. Inoculum characteristics and feedstock preparation
lized (Carton and Olabi, 2010; Sharman, 2005; Sovacool, 2013). H2
generation from water electrolysis using off-peak electricity sur- The digestate used as nutrient feedstock was obtained from
plus from wind and/or solar power can solve the limitations of Snertinge biogas plant (Denmark). After arrival to the laboratory,
variable wind power production and electricity storage (Levene the digestate was filtered through a 2 mm net to remove large par-
et al., 2007). Nevertheless, because of its low density, H2 presents ticles and stored at 55 °C at anaerobic conditions in 5 L bottles. The
limitations linked to its transportation and management bottles were kept in the incubator for one month in order to be
(Granovskii et al., 2006). Therefore, the opportunity of coupling completely degassed. The chemical composition and trace ele-
the H2 with the CO2 present in the biogas and their subsequent ments content of the digestate are shown in Tables S1 and S2.
conversion to CH4 is a promising effective and cheaper alternative Before usage, the digestate was diluted with water and 1 M HCl
technology to upgrade the biogas to higher CH4 content. In fact, to reduce the pH from average 8.48 ± 0.11 to 6.91 ± 0.19 and main-
although information about biological upgrading energy require- tain reactors’ pH values within the range for methanogenesis
ments is not currently available, previous studies affirm that, dur- (Weiland, 2010). The ratio of digestate, 1 M HCl and water was
ing this process, the electricity consumed for conversion and 1:0.7:0.3. The reactors were inoculated with 600 mL of undiluted
storage, with the exception of electrolysis process (4–5 kW h/ degassed digestate and 250 mL of active enriched hydrogeno-
m3 H2), is considered to be negligible, especially when compared trophic inoculum obtained from an upgrading biogas reactor
with surplus electricity (Götz et al., 2016; Jürgensen et al., 2014). (Bassani et al., 2015). The purpose of the enriched culture was to
Moreover, the heat generate from biomethanation can be used to provide active hydrogenotrophic methanogens and thus shorten
heat a biogas digester (e. g. 420 kW for a 5 MW plant) (Götz the overall adaptation period. 50 mL of fresh diluted digestate were
et al., 2016). daily replaced from the bottom of the reactor.
Simplicity and robustness of biological biogas upgrading tech-
nology is mainly attributable to the effectiveness of the heteroge- 2.2. Reactor’s setup and operation
neous microbial consortium involved in AD process. In fact,
biomethanation efficiency is strongly determined by syntrophic A detailed reactors’ description is provided in Supplementary
relations between bacteria and/or archaea co-existing in anaerobic Information (SI) (Fig. S1). Each reactor setup was composed of an
digesters’ microbial communities (Kougias et al., 2016b). During ex up-flow reactor with 850 mL working volume. Reactors were
situ biogas upgrading process, biogas and H2 are injected into an maintained at thermophilic conditions (55 ± 1 °C) by circulating
anaerobic reactor, where the CO2 contained in the biogas is cou- hot water through a water jacket around the reactor’s glass walls.
pled with H2, to be converted to CH4, by a consortium of mixed Reactors’ mixing was ensured by continuous gas recirculation. A
microbial species, via hydrogenotrophic methanogenesis (Kougias gas mixture composed of H2, CH4 and CO2 with ratio 62:23:15
et al., 2017). Previous studies on H2 assisted biogas upgrading (AGA A/S, Denmark) was continuously provided to the reactors.
showed that H2 addition affected microbial community composi- In reactor 1 (R1) and reactor 2 (R2) the gas mixture was injected
tion enhancing hydrogenotrophic methanogenesis and bacterial through 3 stainless steel diffusers with pore size 0.5 lm and
syntrophic interactions with methanogenic archaea (Bassani 2 lm, respectively. Moreover, R1 and R2 were filled with inert alu-
et al., 2015). Nevertheless, due to the fact that the microbial pro- mina ceramic sponge, to reduce H2 bubbles’ size and extend gas
cesses occur exclusively in aqueous environment, the injected retention time in the liquid increasing the contact with reactors’
gases must be dissolved in the reactor liquid phase in order to be liquid phase. Conversely, in reactor 3 (R3) and reactor 4 (R4), the
utilized by microorganisms. However, H2 gas-liquid mass transfer gas mixture was injected through an Al2O3 ceramic membrane
rate is known to be very low, strongly limiting H2 availability for with pore size 0.4 lm and 1.2 lm, respectively. The experiment
methanogens (Bassani et al., 2015, 2016; Luo et al., 2012; Luo was divided into 6 periods, where, either the gas retention time
and Angelidaki, 2012). H2 gas transfer coefficient ðkL a) is mainly (GRT) was reduced by increasing input gas flow rate ðQ IN Þ ([L/LRd]
determined and can be modulated by changing parameters such expressed as per litter reactor (LR)) or gas recirculation rate ðQ RC Þ
312 I. Bassani et al. / Bioresource Technology 234 (2017) 310–319

[L/LRh] was increased, in order to optimize CO2 and H2 conversion CH4 yield per H2 loaded [LCH4/LH2] was determined by Eq. (2):
efficiencies (gCO2 and gH2 ) (Table 1). Q RC was initially set to, on
P CH4
average, 42 times Q IN and then increased to 82 times as previously Y CH4 ¼ ð2Þ
H2used;T
described (Kougias et al., 2017).
where H2used;T represented the theoretical H2 utilization rate [LH2/LRd]
2.3. Calculations (i.e. considering that the 2% extra H2 is remained unutilized according
to the stoichiometry of hydrogenotrophic methanogenesis) and is
The residual biogas produced mainly from content of volatile calculated as described by Eq. (3):
fatty acids (VFA) in the digestate used (essentially acetate and pro-
pionate) was calculated according to the following conversion H2used;T ¼ Q H2 ;IN  Q H2 ;OUT;T ð3Þ
reactions and subtracted from output CH4 and CO2 rates ðQ CH4 ;OUT
and Q CO2 ;OUT expressed as [L/LRd]) where Q H2 ;IN represents the H2 loading rate [LH2/LRd] and Q H2 ;OUT;T
the theoretical H2 output rate [LH2/LRd].
Acetate CH3 COOH ! CH4 þ CO2 H2 utilization efficiency [%] was calculated as described by Eq. (4):
H2used
Propionate CH3 CH2 COOH þ 0:5 H2 O ! 1:75 CH4 þ 1:25 CO2 gH2 ¼  100 ð4Þ
H2used;T
Similarly, as it will be further discussed, extra H2 was found to
be coupled with a portion of extra CO2 produced from VFA degra- where H2used represented the actual H2 utilization rate [LH2/LRd], cal-
dationðQ CO2 ;VFA , [LCO2/LRd]) generating additional CH4. Therefore, culated as follows (Eq. (5)):
H2 utilized for this reaction has been added to the H2 output rate
H2used ¼ Q H2 ;IN þ Q H2 ;VFA  Q H2 ;OUT ð5Þ
(Q H2 ;OUT , [LH2/LRd]) to provide unambiguous results when defining
ex situ biogas upgrading mass balance. Indeed, a source of extra H2 where Q H2 ;VFA is the H2 utilized in VFA reactions [LH2/LRd].
was represented by the input gas mixture. This mixture was com- CO2 conversion efficiency [%] was calculated based on the CO2
posed of CO2 and H2 according to the stoichiometry of hydrogeno- fraction contained in the gas phase, as described by Eq. (6):
trophic methanogenesis reaction (1:4, equivalent to 15 and 60% of
the total input gas volume), 23% CH4, in order to simulate typical CO2used
gCO2 ¼  100 ð6Þ
biogas composition, and 2% extra H2 as ground gas, which was Q CO2 ;IN
expected to remain unutilized.
where CO2used is the CO2 utilization rate [LCO2/LRd] and was calcu-
Based on these calculations, CH4 production rate derived from
lated as follows (Eq. (7)):
ex situ biogas upgrading was determined as follows (Eq. (1)):
PCH4 ¼ Q CH4 ;OUT  Q CH4 ;IN  Q CH4 ;VFA ð1Þ CO2used ¼ Q CO2 ;IN  Q CO2 ;VFA  Q CO2 ;OUT ð7Þ

where Q CH4 ;IN represents the CH4 loading rate [LCH4/LRd] and Q CH4 ;VFA where Q CO2 ;IN represents the CO2 loading rate [LCO2/LRd].
is the CH4 production rate derived from VFA degradation Finally H2 gas transfer coefficient [day1] was calculated as
[LCH4/LRd]. described by Eq. (8):

Table 1
Reactors’ performances under steady state conditions.

Period GRT [h] Q RC [L/LR.h] Reactor Q IN [L/LR.d] Q OUT;T [L/LR.d] Q OUT [L/LR.d] Output gas composition P CH4 [LCH4/LRd] Y CH4 [LCH4/LH2] gH2 [%] gCO2 [%]
CH4 [%] CO2 [%] H2 [%]
I 15 2.88 R1 1.60 ± 0.01 0.64 ± 0.01 0.543 ± 0.06 90.8 ± 1.4 5.1 ± 0.3 4.0 ± 1.7 0.08 ± 0.05 0.09 ± 0.05 100.0 ± 0.1 89.6 ± 1.4
R2 1.59 ± 0.01 0.64 ± 0.01 0.53 ± 0.05 93.1 ± 2.5 4.4 ± 0.6 2.5 ± 1.9 0.09 ± 0.03 0.09 ± 0.03 100.0 ± 0.1 91.1 ± 1.6
R3 1.60 ± 0.01 0.64 ± 0.01 0.54 ± 0.05 89.5 ± 0.6 6.4 ± 0.4 4.1 ± 0.7 0.09 ± 0.05 0.09 ± 0.05 100.0 ± 0.1 86.2 ± 1.4
R4 1.50 ± 0.01 0.60 ± 0.01 0.59 ± 0.03 92.0 ± 1.3 5.9 ± 1.9 2.1 ± 1.0 0.17 ± 0.03 0.19 ± 0.04 100.0 ± 0.1 85.5 ± 4.5
II 15 5.75 R1 1.58 ± 0.03 0.63 ± 0.01 0.55 ± 0.03 93.7 ± 1.2 4.7 ± 0.6 1.6 ± 0.7 0.11 ± 0.02 0.11 ± 0.02 100.0 ± 0.1 90.7 ± 2.1
R2 1.57 ± 0.02 0.63 ± 0.01 0.54 ± 0.06 94.9 ± 0.8 4.0 ± 0.2 1.2 ± 0.6 0.10 ± 0.05 0.11 ± 0.06 100.0 ± 0.1 92.6 ± 0.5
R3 1.58 ± 0.05 0.63 ± 0.02 0.55 ± 0.05 95.1 ± 0.1 3.2 ± 0.2 1.7 ± 0.2 0.12 ± 0.06 0.13 ± 0.07 100.0 ± 0.1 93.3 ± 0.9
R4 1.42 ± 0.04 0.60 ± 0.01 0.56 ± 0.01 96.3 ± 0.2 2.8 ± 0.4 0.8 ± 0.3 0.21 ± 0.02 0.24 ± 0.03 100.0 ± 0.1 92.9 ± 1.5
III 7 5.75 R1 3.42 ± 0.02 1.37 ± 0.01 1.24 ± 0.03 93.1 ± 1.2 3.0 ± 0.8 3.9 ± 0.9 0.33 ± 0.03 0.16 ± 0.02 100.0 ± 0.1 94.5 ± 1.5
R2 3.43 ± 0.02 1.37 ± 0.01 1.28 ± 0.07 93.9 ± 0.6 2.6 ± 0.5 3.5 ± 0.6 0.38 ± 0.07 0.18 ± 0.04 100.0 ± 0.1 95.4 ± 0.9
R3 3.50 ± 0.03 1.40 ± 0.01 1.28 ± 0.05 93.3 ± 0.7 2.3 ± 0.4 4.5 ± 0.9 0.35 ± 0.05 0.17 ± 0.02 99.9 ± 0.2 96.6 ± 1.9
R4 3.16 ± 0.04 1.26 ± 0.02 1.18 ± 0.05 96.0 ± 0.6 1.4 ± 0.5 2.6 ± 0.4 0.37 ± 0.04 0.19 ± 0.02 100.0 ± 0.1 97.9 ± 0.8
IV 7 10.04 R1 3.43 ± 0.01 1.37 ± 0.01 1.23 ± 0.05 94.8 ± 0.1 2.4 ± 0.4 2.8 ± 0.3 0.35 ± 0.04 0.17 ± 0.02 100.0 ± 0.1 95.5 ± 1.2
R2 3.32 ± 0.02 1.34 ± 0.01 1.25 ± 0.06 94.8 ± 0.9 2.6 ± 0.3 2.6 ± 0.7 0.39 ± 0.07 0.19 ± 0.03 100.0 ± 0.1 95.3 ± 0.9
R3 3.46 ± 0.02 1.39 ± 0.01 1.35 ± 0.08 90.4 ± 1.5 2.7 ± 0.2 6.9 ± 1.4 0.40 ± 0.07 0.19 ± 0.03 98.7 ± 0.7 95.6 ± 0.8
R4 3.17 ± 0.04 1.27 ± 0.02 1187 ± 0.06 95.9 ± 0.5 2.0 ± 0.3 2.2 ± 0.3 0.37 ± 0.06 0.20 ± 0.03 100.0 ± 0.1 97.4 ± 1.4
V 4 10.04 R1 5.93 ± 0.01 2.37 ± 0.01 2.36 ± 0.04 90.4 ± 2.3 2.7 ± 0.4 7.0 ± 2.1 0.72 ± 0.05 0.20 ± 0.01 98.6 ± 1.2 97.7 ± 1.1
R2 5.87 ± 0.01 2.35 ± 0.01 2.28 ± 0.08 91.6 ± 2.4 2.3 ± 0.4 6.1 ± 2.0 0.69 ± 0.08 0.20 ± 0.02 99.0 ± 0.9 98.8 ± 1.0
R3 6.02 ± 0.01 2.41 ± 0.01 2.43 ± 0.08 87.6 ± 0.7 2.7 ± 0.4 9.6 ± 0.6 0.70 ± 0.07 0.19 ± 0.02 96.8 ± 0.5 97.8 ± 1.2
R4 5.54 ± 0.01 2.26 ± 0.01 2.22 ± 0.09 91.6 ± 0.5 2.0 ± 0.3 6.4 ± 0.6 0.73 ± 0.08 0.22 ± 0.03 99.1 ± 0.5 99.3 ± 0.8
VI 4 20.14 R1 5.93 ± 0.03 2.37 ± 0.01 2.37 ± 0.12 92.7 ± 1.9 2.4 ± 0.7 4.9 ± 1.5 0.79 ± 0.11 0.22 ± 0.03 99.6 ± 0.7 97.7 ± 1.7
R2 5.86 ± 0.02 2.34 ± 0.01 2.32 ± 0.16 96.0 ± 0.5 1.4 ± 0.3 2.6 ± 0.5 0.82 ± 0.15 0.23 ± 0.04 100.0 ± 0.1 100.0 ± 1.0
R3 6.04 ± 0.06 2.42 ± 0.02 2.30 ± 0.17 92.5 ± 0.6 2.0 ± 0.5 5.6 ± 0.6 0.70 ± 0.15 0.19 ± 0.04 99.7 ± 0.3 99.3 ± 1.3
R4 5.59 ± 0.01 2.22 ± 0.01 2.16 ± 0.03 94.1 ± 1.2 1.7 ± 0.3 4.2 ± 0.9 0.71 ± 0.02 0.21 ± 0.01 100.0 ± 0.1 99.9 ± 0.6

Q OUT represents the total gas output rate and Q OUT;T the total theoretical gas output rate.
I. Bassani et al. / Bioresource Technology 234 (2017) 310–319 313

rt community, statistical analysis to identify significantly abundance


kL a ¼ ð8Þ
22:4ðH2gTh  H2l Þ differences for each microbe among the samples, was carried out
as previously described by Tsapekos et al. (2016).
where rt is the H2 gas-liquid mass transfer rate, which can be iden-
tified with the H2 utilization rate ðH2used Þ, described by Eq. (5), 22.4
[L/mol] is the gas volume to mole ratio (1 mol gas corresponds to 3. Results and discussion
22.4 L at STP), and H2gTh [M] and H2l [M] are the H2 concentrations
in the gas and in the liquid phase, respectively. At the beginning of Period I, the pH of the reactors was
increased due to the removal of the dissolved CO2 in the liquid
media. The addition of HCl to the daily feeding, in order to decrease
2.4. Analytical methods
the pH values and to maintain them within the optimum range for
methanogenesis (Weiland, 2010), became effective allowing the
The output biogas was recorded in daily basis by an automated
reduction of pH levels during the following experimental periods
gas meter with a 100 mL reversible cycle and registration
(Table 2, and Fig. S2a).
(Angelidaki et al., 1992). Total solids (TS), volatile solids (VS),
Moreover, although the digestate used to inoculate the reactors
Ammonium nitrogen (NH4-N) and Total Kjeldahl Nitrogen (TKN)
had been degassed for one month to avoid any residual biogas pro-
were measured according to the Standard Methods for Examina-
duction, during the whole experimental period; additional biogas
tion of Water and Wastewater (APHA, 2005). Liquid samples from
was produced from degradation of residual organic matter in the
the reactors were collected for pH and VFA analysis every second
digestate, mainly acetate and propionate. Thus, a remarkable VFA
day. The pH was measured immediately after the collection to
reduction was recorded until Period IV, where VFA levels decreased
avoid the CO2 removal from the liquid phase, by digital PHM210
to 0.3 g/L from 1.5 g/L (Period I) (Table 2 and Fig. S2b and S3).
pH meter connected to the Gel pH electrode (pHC3105-8;
Moreover, it was assumed that extra H2 would be utilized to convert
Radiometer analytical). VFA samples were determined as previ-
the CO2 from the biogas produced from acetate and propionate
ously described by (Kougias et al., 2015). The biogas composition
degradation. This assumption was documented by the H2 utilization
was measured every second day by GC-TCD (Mikrolab, Aarhus A/
efficiency ðgH2 Þ which exceeded 100% during the most of the exper-
S, Denmark) as previously described (Kougias et al., 2015). Dis-
solved H2 was measured from liquid samples as described in SI. imental period. Considering both additional CH4 production and CO2
Trace elements analysis was performed on initial digestate and conversion (i.e. from biogas production and upgrade due to VFA
reactors’ liquid samples during steady state operation of period I, degradation), it was calculated that, on average, 0.04 LCH4/LRd were
as described in SI. produced (Table S4).
Finally, the increase in total micronutrients’ content indicates
that the daily provided digestate represented a complete source
2.5. Microbial community composition
of nutrients and its provision rate was in surplus and could ensure
microbial growth and metabolism (Table S2). Therefore, for future
Samples were obtained at the beginning of the experiment and
applications, lower digestate flow rate as nutrient source could be
during reactors’ steady state operation of the last period to ensure
provided.
representative process conditions and microbial community stabil-
ity. Genomic DNA was extracted from reactors’ liquid samples, in
triplicates, with PowerSoilÒ DNA Isolation Kit (MO BIO Laborato- 3.1. Overview of biogas upgrading process performances
ries, Carlsbad, CA). Minor modifications were introduced to the pro-
cedure in order to improve the quality of the genomic DNA The beginning of the experiment mainly represented a start-up
recovered as described by Bassani et al. (2015). Quality and quan- period, during which microbial community adapted to the new
tity of the DNA extracted were determined using NanoDrop (Ther- operation conditions. Nevertheless, microorganisms rapidly
moFisher Scientific, Waltham, MA) and QubitTM fluorimeter (Life started to convert H2 and CO2 to CH4 and thus, in Period I, the
Technologies, Carlsbad, CA). 16S rRNA gene V4 hypervariable region CH4 content increased from 23%, in the input gas, to >91% in the
was amplified with universal primers (Caporaso et al., 2012) and output flow, then stabilizing around 96% in period VI. Approxi-
sequenced using Illumina MiSeq sequencing technology. The mately 2–4% H2 remained unutilized and was recorded in the out-
obtained reads were submitted to the NCBI sequence read archive put gas (Table 1 and Fig. 1). Moreover, despite the fact that the
database (SRA) with accession number SRP091653. Detailed sample experiment started with low CO2 conversion efficiency (gCO2 Þ, it
IDs are listed in Table S3. Analysis of 16S rRNA gene sequences has rapidly increased within the first 10 days stabilizing to 88% at
been performed as described by Kougias et al. (2017). A further the end of Period I. Interestingly, gCO2 continued to increase along
alignment was performed using BLASTN against NCBI 16S rRNA all periods, in all the configurations, achieving >99% in Period VI
database as described by Bassani et al. (2015). A predictive metage- (Table 1 and Fig. 2a).
nomic approach investigating KEGG and COG functional categories As expected, in Periods I and II, in most of the reactor configu-
was applied to the identified genera, as described in SI. Heat maps rations, low CH4 productions rates (P CH4 ) and CH4 yields ðY CH4 )
showing relative abundance and fold change of most relevant oper- were observed (Table 1 and Fig. 2c and d). Their concomitance with
ational taxonomic units (OTUs) were drawn using Multiexperiment high gH2 suggested that the input H2, although efficiently utilized,
Viewer software (MeV) (Howe et al., 2010). was not totally exploited for CH4 production, but rather for other
microbial reactions, such cell maintenance and growth purposes
2.6. Statistical analysis (Kleerebezem and Stams, 2000). This argument, together with
the reduction of micronutrients known to be essential for cells
Descriptive statistics were carried out for all data and mean val- growth, supports the hypothesis of an initial acclimation of the
ues and standard deviations were calculated. The comparison of microbial community (Kayhanian and Rich, 1995) (Table S2). After
quantitative variables (e.g. CH4 content in the output gas) between such adaptation phase, Y CH4 gradually increased reaching values
the different tested configurations to determine significant differ- closer to the stoichiometric maximum of 0.25 LCH4/LH2 (Table 1
ences (p < 0.05) was achieved by conducting one-way analysis of and Fig. 2d). This trend is more visible in Fig. S4, where the stoi-
variance (ANOVA) using Graphpad Prism 5 program (Graphpad chiometric difference between H2 utilization and CH4 production
Software, Inc., San Diego, CA). Moreover, regarding the microbial rates is plotted against time. Values recorded remained higher than
314 I. Bassani et al. / Bioresource Technology 234 (2017) 310–319

Table 2
Reactors’ pH, VFA, dissolved H2 concentration (H2l) and H2 transfer coefficient (kLa) under steady state conditions.

Period Reactor pH TVFA [g/L] Acetate [g/L] Propionate [g/L] H2l [M] kL a [day1]
6
I R1 8.77 ± 0.05 1.77 ± 0.13 1.11 ± 0.10 0.22 ± 0.02 (0.05 ± 0.01)*10 5.76*103
R2 8.75 ± 0.03 1.52 ± 0.11 1.00 ± 0.06 0.15 ± 0.02 (0.06 ± 0.01)*106 10.08*103
R3 8.78 ± 0.07 1.36 ± 0.08 0.88 ± 0.06 0.12 ± 0.01 (0.19 ± 0.07)*106 3.27*103
R4 8.82 ± 0.09 1.62 ± 0.08 1.07 ± 0.08 0.16 ± 0.02 (0.46 ± 0.19)*106 18.34*103
II R1 8.64 ± 0.07 1.46 ± 0.13 0.96 ± 0.08 0.20 ± 0.01 (0.08 ± 0.02)*106 17.03*103
R2 8.66 ± 0.07 1.22 ± 0.13 0.79 ± 0.14 0.14 ± 0.02 (0.31 ± 0.09)*106 18.64*103
R3 8.81 ± 0.04 1.36 ± 0.22 0.90 ± 0.19 0.15 ± 0.03 (0.35 ± 0.12)*106 5.74*103
R4 8.94 ± 0.05 1.65 ± 0.21 1.15 ± 0.17 0.18 ± 0.02 (0.47 ± 0.03)*106 8.62*103
II I R1 8.44 ± 0.06 0.38 ± 0.08 0.14 ± 0.08 0.22 ± 0.01 (0.29 ± 0.10)*106 5.58*103
R2 8.44 ± 0.07 0.45 ± 0.15 0.24 ± 0.13 0.19 ± 0.02 (0.27 ± 0.02)*106 10.07*103
R3 8.57 ± 0.15 1.11 ± 0.24 0.76 ± 0.15 0.19 ± 0.01 (0.18 ± 0.13)*106 2.52*103
R4 8.55 ± 0.10 1.56 ± 0.18 1.06 ± 0.16 0.22 ± 0.01 (0.34 ± 0.13)*106 8.51*103
IV R1 8.33 ± 0.10 0.27 ± 0.04 0.06 ± 0.03 0.17 ± 0.02 (0.36 ± 0.02)*106 2.85*103
R2 8.27 ± 0.11 0.29 ± 0.06 0.12 ± 0.04 0.14 ± 0.02 (0.33 ± 0.07)*106 5.48*103
R3 8.29 ± 0.13 0.24 ± 0.08 0.11 ± 0.04 0.04 ± 0.02 (0.30 ± 0.04)*106 3.27*103
R4 8.30 ± 0.13 0.63 ± 0.12 0.40 ± 0.09 0.13 ± 0.02 (0.27 ± 0.12)*106 13.08*103
V R1 8.20 ± 0.11 0.36 ± 0.07 0.09 ± 0.03 0.14 ± 0.02 (0.24 ± 0.10)*106 7.72*103
R2 8.18 ± 0.06 0.18 ± 0.06 0.09 ± 0.02 0.05 ± 0.02 (0.19 ± 0.05)*106 7.93*103
R3 8.15 ± 0.05 0.34 ± 0.05 0.23 ± 0.02 0.05 ± 0.01 (0.70 ± 0.18)*106 3.79*103
R4 8.12 ± 0.10 0.44 ± 0.06 0.29 ± 0.05 0.07 ± 0.01 (0.16 ± 0.05)*106 5.99*103
VI R1 8.11 ± 0.15 0.17 ± 0.04 0.08 ± 0.02 0.06 ± 0.02 (0.52 ± 0.10)*106 9.32*103
R2 8.03 ± 0.12 0.11 ± 0.07 0.07 ± 0.05 0.02 ± 0.03 (0.23 ± 0.06)*106 15.21*103
R3 8.13 ± 0.17 0.31 ± 0.04 0.22 ± 0.05 0.05 ± 0.01 (0.29 ± 0.03)*106 7.50*103
R4 8.09 ± 0.14 0.18 ± 0.07 0.13 ± 0.05 0.03 ± 0.01 (0.22 ± 0.01)*106 7.47*103

zero during most of the experimental time indicating that the of the proposed concepts under full-scale conditions, it should be
input H2 was partially utilized for microbial maintenance. noted that the characterization of pore sizes as ‘‘large” or ‘‘small”
In summary, the reported results indicate that by applying the refers to a diameter range between 0.4 lm and 2 lm. Thus, it
proposed reactor configurations efficient biogas upgrading was was found that larger pore size resulted in higher output gas qual-
achieved in terms of kinetics of CH4 production, conversion effi- ity compared to the smaller pore diameter. On average, reactors in
ciency and consequent output gas quality, generating biogas with which the gas mixture was distributed through larger pore size dif-
96% CH4 content, which could be employed as transportation fuel fusers (R2 and R4) reached CH4 content 94.2 ± 1.8%, compared to
(Deng and Hägg, 2010). The analyzed biochemical parameters 92.0 ± 2.2% achieved by smaller pore size diffusers (R1 and R3).
demonstrate that the biological conversion of CO2 and H2 to CH4 Accordingly, at the end of the experiment (Period VI), CH4 con-
was efficiently performed as during the whole experimental per- tent stabilized to 94 and 96% in R4 and R2, respectively, compared
iod, no VFA accumulation was recorded (Table 2 and Fig. S2b). This to >92% of R1 and R3 (Table 1 and Fig. 1). Similarly, despite all con-
practically means that the injected CO2 and H2 were exclusively figurations performed at very high gCO2 during the whole experi-
utilized by hydrogenotrophic methanogens to generate methane ment (on average 95%), reactors with larger pore size diffusers
and not by homoacetogenic bacteria that could produce acetate (R2 and R4) showed the best performances (on average 96%). Like-
(Table 2 and Fig. S3a). wise, H2 was efficiently consumed during the whole period, with
A further test was conducted in order to clarify the direct R4 showing the highest gH2 (>99%) (Table 1 and Fig. 2b). This out-
impact of H2 mass transfer rate on upgrading performances. There- come is of interest because smaller pore size devices are expected
fore, based on dissolved and gaseous H2 concentrations, H2 transfer to deliver smaller bubbles, increasing gas-liquid contact and, there-
coefficient ðkL a) has been calculated, according to Eq. (8), and the fore, availability of substrates for microorganisms. A possible
results are shown in Table 2. As expected, dissolved H2 concentra- explanation for the obtained results could be attributed to the
tion (H2l Þ was markedly lower compared to H2 concentration in the higher mixing of reactor’s liquid phase provided by larger pore size
gas phase ðH2gTh Þ (i.e. on average 55 times lower). Nevertheless, devices compared with smaller ones (Merchuk et al., 1998). In fact,
compared to previous studies (Díaz et al., 2015; Luo et al., 2012; proper mixing speed is known to increase gases kL a improving the
Luo and Angelidaki, 2013), remarkable kL a values were achieved contact between gases and reactor’s liquid media (Kramer and
ranging from 2.5⁄103 day1 and 18.6⁄103 day1. The outcomes Bailey, 1991; Luo and Angelidaki, 2012). Specifically, when the
confirm the importance of this parameter to ensure H2 availability bubbles, formed at the pores of the diffusers, circulate in a vigor-
for microorganisms and enlighten the effectiveness of applied ously agitated gas-liquid dispersion, they generate a turbulent
reactor configuration and process parameters for ex situ biogas flow, where shear forces acting on the bubbles break them up into
upgrading process. smaller ones (Villadsen et al., 2011). Therefore, in this experiment,
larger pore size devices presented a positive impact on the gas-
3.2. Effect of diffusion device and pore size on upgrading performances liquid transfer as the intense mixing resulted in a more efficient
gas bubble break, improving the gas-liquid contact. This result
From the analysis of operational parameters, it was demon- was confirmed by measuring the dissolved H2 and consequently
strated that the pore size of the diffusion devices influenced the calculating the kL a coefficient. The outcome of this test proved that
upgrading process, although its contribution was not statistically larger pore size diffusion devices, in particular R2, were able to uti-
significant in all the applied conditions. In order to avoid any con- lize H2 providing the highest kL a (on average 11.2⁄103 day1)
fusion in the interpretation of the results and the implementation (Table 2). Coherently with parameters previously investigated the
I. Bassani et al. / Bioresource Technology 234 (2017) 310–319 315

Fig. 1. Biogas composition of (a) R1, (b) R2, (c) R3 and (d) R4. Continuous lines present gases content at the output of the reactors, while dashed lines illustrate input gas
composition.

highest average PCH4 and Y CH4 were shown by R4, while R2 reported importance for upgrading performances, unexpectedly, diffusion
the highest values at the end of the experiment (0.82 LCH4/LR d device (diffusers and ceramic sponge vs ceramic membrane) did
(p < 0.05) and 0.23 LCH4/LH2, respectively) (Table 1 and not show significant performance distinctions throughout the
Fig. 2c and d). Finally, while diffusers’ pore size resulted of great experiment.
316 I. Bassani et al. / Bioresource Technology 234 (2017) 310–319

Fig. 2. Reactors’ CO2 conversion efficiency (gCO2 ) (a), H2 utilization efficiency (gH2 ) (b), CH4 production rate (P CH4 ) (c) and CH4 yield ðY CH4 ) (d).

3.3. Effect of gas recirculation and input gas flow rate on upgrading affect negatively the process resulting in lower performance effi-
performances ciencies (Kleerebezem and Stams, 2000). In this experiment, a
slight decline of performances, which could have been due to
Notably, in all the reactors increase of gas recirculation rate insufficient H2 solubilisation at this Q IN , has been recorded only
ðQ RC Þ led to a marked improvement of upgrading performances, in Period V, when the highest Q IN was provided. The drop was
resulting in significant higher CH4 content and/or PCH4 in several totally recovered in Period VI, thanks to the application of higher
operating condition and reactor configurations Moreover, in accor- Q RC (Table 1 and Fig. 1 and 2b).
dance with previous studies (Guiot et al., 2011), kL a increased with In conclusion, the obtained results indicate that larger pore size
higher Q RC ; for example, 36% higher kL a was recorded from period devices together with a proper gas recirculation flow managed to
V to period VI (Table 2). These results demonstrate the positive achieve the most efficient biogas upgrading due to the optimal
effect of recirculation on gas-liquid mass transfer rate extending mixing speed and gas retention time. Finally, it was shown that
gas-liquid contact time and therefore enhancing H2 availability the application of a ceramic sponge material having very high sur-
for microorganisms (Guiot et al., 2011). face area (65 m2/reactor) was able to increase the contact between
Nevertheless, change in microbial community living conditions H2 bubbles and liquid and extend the permanence of gases in the
and, more specifically, excessive input gas flow rate (Q IN Þ could liquid media, therefore increasing H2 kL a.
I. Bassani et al. / Bioresource Technology 234 (2017) 310–319 317

3.4. Microbial community composition utilization, pentose and amino acids degradation and peptides
transport, compared to the other genera analyzed (Dataset S2). In
16S rRNA gene sequencing results are summarized in Table S3. addition, its predominance and concomitant increase with Methan-
Rarefaction curves and Principal Component Analysis (PCoA) othermobacter thermautotrophicus, together with the presence of
clearly showed a shift in microbial composition between the initial genes involved in Wood-Ljungdahl pathway, indicates a possible
inoculum used and the reactors content at the end of the experi- role of Anaerobaculum in syntrophic acetate oxidation (SAO) with
ment, with replicate samples clustering together (Fig. S5). More- this methanogen (Dataset S2).
over, rarefaction curves showed an increased species richness in Notably, the second most abundant OTU, accounting for 15%
the reactors after H2 addition. The results reported are focused of the community, was assigned to the newly discovered order
on most abundant members of the microbial community having MBA08 (class Clostridia). In addition, BLASTn search against NCBI
relative abundance >0.5%. database revealed a similarity to the newly identified Hydrogenis-
In accordance with previous studies, Firmicutes was the most pora ethanolica with 90% identity. Nevertheless, being the sequence
represented phylum accounting for the 32% of the microbial com- identity score lower than the threshold for genera classification
munity after H2 addition (Campanaro et al., 2016). Interestingly, (Dataset S1), the taxonomy of this OTU remains uncertain, and
Synergistetes resulted the second most abundant phylum, which defined only at class level as Clostridia sp.1, indicating this microor-
relative abundance doubled reaching >26% at the end of the exper- ganism as a new species. The remarkable abundance of Clostridia
iment. In particular, most abundant OTUs, accounting between 10 sp.1 in the initial inoculum points out a possible important func-
and 15% of the community were assigned to Anaerobaculum mobile tion of this bacterium in AD process. Moreover, its high abundance
(98% identity; increasing < 3-folds; p < 0.05) and family Synergis- at the end of the experiment was in accordance with a recent study
taceae (Dataset S1, Figs. 3 and 4). The occurrence of these microor- on ex situ biogas upgrading, and is indicative of a remarkable resi-
ganisms, together with the increment of other 2 OTUs (19 and 40- lience to the operating conditions suggesting a possible syntrophic
folds, respectively, p < 0.05) assigned to the same phylum, has been relation with hydrogenotrophic methanogens (Kougias et al.,
previously observed in reactors used for biogas upgrading (Bassani 2017).
et al., 2015; Treu et al., 2016b) and is consistent with their role and Further details on microbial community composition are pro-
resilience in AD process (Campanaro et al., 2016; Werner et al., vided in SI. However, among most interesting microorganisms,
2011). Moreover, according to the predictive metagenomic analy- 10 OTUs were assigned to order Thermoanaerobacterales, mostly
sis investigating KEGG and COG functional categories, Anaerobacu- accounting for >1% of the community and increasing at the end
lum is involved in amino acids and carbohydrates metabolism of the experiment (Dataset S1, Figs. 3 and 4). Among them, two
presenting an enriched number of genes for glycolysis, glycerol OTUs were identified as genus Thermacetogenium and species

Fig. 3. Heat maps of relative abundance (>0.5%; left part of the panel) and folds change (log2; right part of the panel) of the most interesting microorganisms populating
initial inoculum and reactors at the end of the experiment. Correspondence between colors and relative abundance or folds change is reported in the scale at the top of each
panel. Folds change is represented in red and green for increased and decreased OTUs, respectively.
318 I. Bassani et al. / Bioresource Technology 234 (2017) 310–319

Fig. 4. Comparison between initial inoculum and samples obtained from each reactor at the end of the experiment (R1–R4 are plotted against the initial inoculum and shown
in panel (a) to (d)). The left part of each panel represents the relative abundance (>0.5%), while the right part shows the folds change of OTUs significantly changing in
abundance.

Tepidanaerobacter syntrophicus (96% identity). As shown by COG between 15 and 120-folds (p < 0.05), with M. thermautotrophicus
and KEGG analysis, genes involved in Wood-Ljungdahl pathway being the most abundant methanogen (2.6% at the end of the exper-
were highly represented confirming the role of these microorgan- iment) (Dataset S1, Figs. 3 and 4). Notably, BLASTn search against
isms in SAO pathway (Treu et al., 2016a; Westerholm et al., 2011) NCBI database revealed 100% similarity of M. thermautotrophicus
(Dataset S2). Moreover, the predictive functional metagenomics with several microbial species, such as Methanobacterium thermag-
suggested the presence in their genomes of genes for glucose and gregans, M. wolfeii, M. thermophilus, M. thermoflexus, and M. defluvii,
pentose metabolism and for amino acids and proteins degradation indicating that the most abundant methanogen populating the
indicating that they would be able to utilize different substrates community was represented by a new species.
(Dataset S2). Additionally, the significant increase in relative abun- The presence of these hydrogenotrophs is in accordance with
dance of 2 OTUs assigned to family Syntrophomonadaceae (30 previous studies showing prevalence and increment of these spe-
folds; p < 0.05) is likely related to their syntrophic relation with cies in biogas upgrading reactors (Bassani et al., 2015; Kougias
hydrogenotrophic methanogens (Dataset S1, Figs. 3 and 4) (Treu et al., 2017; Treu et al., 2016b). Interestingly, several hydrogeno-
et al., 2016a,b). It has been previously reported that members of trophs were maintained in the reactor presumable performing
Syntrophomonas genus encode genes involved in menaquinone the same function, without showing competition. This could have
biosynthesis (Kougias et al., 2016b), and in turn menaquinone resulted in a more robust hydrogenotrophic methanogenic process,
functions as electron carrier during interspecies electron transfer which would be able to tolerate disturbances, as the different
in syntrophic communities (Stams and Plugge, 2009). methanogens might have different tolerance levels. Notably, M.
Regarding Archaea domain, phylum Euryarchaeota accounted for thermautotrophicus was found with the highest relative abundance
<4% of the community. Among most abundant OTUs, 3 OTUs were in R2, the reactor presenting the best Y CH4 and output-gas quality
assigned to Methanothermobacter thermautotrophicus, Methanocul- (96% CH4 content). Additionally, R4, which reached remarkable
leus thermophilus and Methanocorpusculum aggregans (>99% iden- H2 and CO2 conversion efficiencies, showed the highest relative
tity). Interestingly, these microorganisms significantly increased abundance of M. thermophilus (1.34%) suggesting a potential
I. Bassani et al. / Bioresource Technology 234 (2017) 310–319 319

synergistic function between the two methanogens. In conclusion, Götz, M., Lefebvre, J., Mörs, F., Koch, A.M., Graf, F., Bajohr, S., Reimert, R., Kolb, T.,
2016. Renewable power-to-gas: a technological and economic review. Renew.
the proliferation of hydrogenotrophic methanogens and the con-
Energy 85, 1371–1390.
comitant increment of syntrophic bacteria are indicative of the Granovskii, M., Dincer, I., Rosen, M.A., 2006. Economic and environmental
effect of the H2 on the microbial consortium selectively stimulating comparison of conventional, hybrid, electric and hydrogen fuel cell vehicles. J.
hydrogenotrophic methanogenesis. Power Sour. 159, 1186–1193.
Guiot, S.R., Cimpoia, R., Carayon, G., 2011. Potential of wastewater-treating
anaerobic granules for biomethanation of synthesis gas. Environ. Sci. Technol.
4. Conclusions 45, 2006–2012.
Howe, E., Holton, K., Nair, S., Schlauch, D., Sinha, R., Quackenbush, J., 2010. MeV:
multiExperiment viewer. In: Biomedical Informatics for Cancer Research.
The configurations containing larger pore size diffusion devices Spinger, US, pp. 267–277.
showed the best kinetics and output-gas quality due to the Jürgensen, L., Ehimen, E.A., Born, J., Holm-Nielsen, J.B., 2014. Utilization of surplus
electricity from wind power for dynamic biogas upgrading: Northern Germany
achievement of better reactor mixing. The increment of gas recir-
case study. Biomass Bioenergy 66, 126–132.
culation flow rate led to improved output gas quality, generating Kayhanian, M., Rich, D., 1995. Pilot-scale high solids thermophilic anaerobic
biogas with 96% CH4 content that can be used as transportation digestion of municipal solid waste with an emphasis on nutrient
requirements. Biomass Bioenergy 8, 433–444.
fuel. Finally, microbial analysis revealed a significant increase in
Kleerebezem, R., Stams, A.J.M., 2000. Kinetics of syntrophic cultures: a theoretical
relative abundance of hydrogenotrophic archaea, with Methanoth- treatise on butyrate fermentation. Biotechnol. Bioeng. 67, 529–543.
ermobacter thermautotrophicus being the most abundant methano- Kougias, P.G., Boe, K., Einarsdottir, E.S., Angelidaki, I., 2015. Counteracting foaming
gen. The concomitant proliferation of syntrophic bacteria, such as caused by lipids or proteins in biogas reactors using rapeseed oil or oleic acid as
antifoaming agents. Water Res. 79, 119–127.
Anaerobaculum mobile, or members of Thermoanaerobacterales Kougias, P.G., Treu, L., Benavente, D.P., Boe, K., Campanaro, S., Angelidaki, I., 2017.
and Syntrophomonadaceae confirmed the selection-effect of H2 Ex-situ biogas upgrading and enhancement in different reactor systems.
shaping the microbial community and enhancing hydrogeno- Bioresour. Technol. 225, 429–437. http://dx.doi.org/10.1016/j.
biortech.2016.11.124.
trophic pathway. Kougias, P.G., Treu, L., Campanaro, S., Zhu, X., Angelidaki, I., 2016b. Dynamic
functional characterization and phylogenetic changes due to Long Chain Fatty
Acknowledgements Acids pulses in biogas reactors. Sci. Rep. 6. http://dx.doi.org/10.1038/srep28810.
Kramer, H.W., Bailey, J.E., 1991. Mass transfer characterization of an airlift probe for
oxygenating and mixing cell suspensions in an NMR spectrometer. Biotechnol.
We thank Hector Garcia and Hector Diaz for technical assis- Bioeng. 37, 205–209.
tance. Illumina sequencing was performed at the Ramaciotti Centre Lee, J.C., Kim, J.H., Chang, W.S., Pak, D., 2012. Biological conversion of CO2 to CH4
using hydrogenotrophic methanogen in a fixed bed reactor. J. Chem. Technol.
for Genomics (Sydney, Australia). This work was supported by the
Biotechnol. 87, 844–847.
Danish Council for Strategic Research under the project ‘‘SYMBIO  Levene, J.I., Mann, M.K., Margolis, R.M., Milbrandt, A., 2007. An analysis of hydrogen
Integration of biomass and wind power for biogas enhancement production from renewable electricity sources. Sol. Energy 81, 773–780.
Luo, G., Angelidaki, I., 2012. Integrated biogas upgrading and hydrogen utilization in
and upgrading via hydrogen assisted anaerobic digestion”, contract
an anaerobic reactor containing enriched hydrogenotrophic methanogenic
12-132654. culture. Biotechnol. Bioeng. 109, 2729–2736.
Luo, G., Angelidaki, I., 2013. Co-digestion of manure and whey for in situ biogas
upgrading by the addition of H2: process performance and microbial insights.
Appendix A. Supplementary data Appl. Microbiol. Biotechnol. 97, 1373–1381.
Luo, G., Johansson, S., Boe, K., Xie, L., Zhou, Q., Angelidaki, I., 2012. Simultaneous
Supplementary data associated with this article can be found, in hydrogen utilization and in situ biogas upgrading in an anaerobic reactor.
Biotechnol. Bioeng. 109, 1088–1094.
the online version, at http://dx.doi.org/10.1016/j.biortech.2017.03. Merchuk, J.C., Contreras, A., García, F., Molina, E., 1998. Studies of mixing in a
055. concentric tube airlift bioreactor with different spargers. Chem. Eng. Sci. 53,
709–719.
Mikkelsen, M., Jørgensen, M., Krebs, F.C., 2010. The teraton challenge. A review of
References
fixation and transformation of carbon dioxide. Energy Environ. Sci. 3, 43–81.
Sharman, H., 2005. Why wind power works for Denmark. Civ. Eng. 158, 66–72.
Alitalo, A., Niskanen, M., Aura, E., 2015. Biocatalytic methanation of hydrogen and Sovacool, B.K., 2013. Energy policymaking in Denmark: implications for global
carbon dioxide in a fixed bed bioreactor. Bioresour. Technol. 196, 600–605. energy security and sustainability. Energy Policy 61, 829–839.
Angelidaki, I., Ellegaard, L., Ahring, B.K., 1992. Compact automated displacement gas Stams, A.J.M., Plugge, C.M., 2009. Electron transfer in syntrophic communities of
metering system for measurement of low gas rates from laboratory fermentors. anaerobic bacteria and archaea. Nat. Rev. Microbiol. 7, 568–577.
Biotechnol. Bioeng. 39, 351–353. Treu, L., Campanaro, S., Kougias, P.G., Zhu, X., Angelidaki, I., 2016a. Untangling the
APHA, 2005. Standard Methods for the Examination of Water and Wastewater, effect of fatty acid addition at species level revealed different transcriptional
American Water Works Association/American Public Works Association/Water responses of the biogas microbial community members. Environ. Sci. Technol.
Environment Federation. 50, 6079–6090.
Bassani, I., Kougias, P.G., Treu, L., Angelidaki, I., 2015. Biogas upgrading via Treu, L., Kougias, P.G., Campanaro, S., Bassani, I., Angelidaki, I., 2016b. Deeper insight
hydrogenotrophic methanogenesis in two-stage continuous stirred tank into the structure of the anaerobic digestion microbial community: the biogas
reactors at mesophilic and thermophilic conditions. Environ. Sci. Technol. 49, microbiome database is expanded with 157 new genomes. Bioresour. Technol.
12585–12593. 216, 260–266.
Bassani, I., Kougias, P.G., Angelidaki, I., 2016. In-situ biogas upgrading in Tsapekos, P., Kougias, P.G., Treu, L., Campanaro, S., Angelidaki, I., 2016. Process
thermophilic granular UASB reactor: key factors affecting the hydrogen mass performance and comparative metagenomic analysis during co-digestion of
transfer rate. Bioresour. Technol. 221, 485–491. manure and lignocellulosic biomass for biogas production. Appl. Energy 185,
Campanaro, S., Treu, L., Kougias, P.G., De Francisci, D., Valle, G., Angelidaki, I., 2016. 126–135.
Metagenomic analysis and functional characterization of the biogas Villadsen, J., Nielsen, J., Lidén, G., 2011. Gas-liquid mass transfer. In: Bioreaction
microbiome using high throughput shotgun sequencing and a novel binning Engineering Principles. Springer, New York, pp. 459–496.
strategy. Biotechnol. Biofuels 9, 1. Weiland, P., 2010. Biogas production: current state and perspectives. Appl.
Caporaso, J.G., Lauber, C.L., Walters, W.A., Berg-Lyons, D., Huntley, J., Fierer, N., Microbiol. Biotechnol. 85, 849–860.
Owens, S.M., Betley, J., Fraser, L., Bauer, M., Gormley, N., Gilbert, J.A., Smith, G., Werner, J.J., Knights, D., Garcia, M.L., Scalfone, N.B., Smith, S., Yarasheski, K.,
Knight, R., 2012. Ultra-high-throughput microbial community analysis on the Cummings, T.A., Beers, A.R., Knight, R., Angenent, L.T., 2011. Bacterial
Illumina HiSeq and MiSeq platforms. ISME J. 6, 1621–1624. community structures are unique and resilient in full-scale bioenergy
Carton, J.G., Olabi, A.G., 2010. Wind/hydrogen hybrid systems: opportunity for systems. Proc. Natl. Acad. Sci. USA 108, 4158–4163.
Ireland’s wind resource to provide consistent sustainable energy supply. Energy Westerholm, M., Roos, S., Schnürer, A., 2011. Tepidanaerobacter acetatoxydans sp.
35, 4536–4544. nov., an anaerobic, syntrophic acetate-oxidizing bacterium isolated from two
Deng, L., Hägg, M.B., 2010. Techno-economic evaluation of biogas upgrading process ammonium-enriched mesophilic methanogenic processes. Syst. Appl.
using CO2 facilitated transport membrane. Int. J. Greenh. Gas Control 4, 638– Microbiol. 34, 260–266.
646.
Díaz, I., Pérez, C., Alfaro, N., Fdz-Polanco, F., 2015. A feasibility study on the
bioconversion of CO2 and H2 to biomethane by gas sparging through polymeric
membranes. Bioresour. Technol. 185, 246–253.

You might also like