Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

international journal of hydrogen energy 34 (2009) 245–254

Available at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/he

A bench scale study of fermentative hydrogen and methane


production from food waste in integrated two-stage process

Xing Wang*, You-cai Zhao


The State Key Laboratory of Pollution Control and Resource Reuse, Tongji University, Siping road 1239, Shanghai 200092, China

article info abstract

Article history: A two-stage fermentation process combining hydrogen and methane production for the
Received 11 August 2008 treatment of food waste was investigated in this paper. In hydrogen fermentation reactor,
Received in revised form the indigenous mixed microbial cultures contained in food waste were used for hydrogen
27 September 2008 production. No foreign inoculum was used in the hydrogen fermentation stage, the
Accepted 28 September 2008 traditional heat treatment of inoculum was not applied either in this bench scale experi-
Available online 28 November 2008 ment. The effects of the stepwise increased organic loading rate (OLR) and solid retention
time (SRT) on integrated two-stage process were investigated. At steady state, the optimal
Keywords: OLR and SRT for the integrated two-stage process were found to be 22.65 kg VS/m3 d (160 h)
Integrated two-stage process for hydrogen fermentation reactor and 4.61 (26.67 d) for methane fermentation reactor,
Fermentative hydrogen production respectively. Under the optimum conditions, the maximum yields of hydrogen
None-heat treatment of inoculum (0.065 m3 H2/kg VS) and methane (0.546 m3 CH4/kg VS) were achieved with the hydrogen
Indigenous microbial cultures and methane contents ranging from 29.42 to 30.86%, 64.33 to 71.48%, respectively. Biode-
gradability analysis showed that 5.78% of the influent COD was converted to the hydrogen
in H2-SCRD and 82.18% of the influent COD was converted to the methane in CH4-SCSTR
under the optimum conditions.
ª 2008 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights
reserved.

1. Introduction economic feasibility of waste treatment. Moreover, biological


methods produce hydrogen from various renewable resources
Hydrogen has been widely recognized as a clean energy and are less energy-intensive than chemical or electro-
carrier of the future. It can be used as an important indus- chemical ones since they are carried out at ambient temper-
trial raw material in hydrogenation processes [1]. Among ature and pressure.
various hydrogen-producing methods, anaerobic dark Two-phase anaerobic digestion system is often used for
fermentation might be the most promising one since the fermentative hydrogen and methane fermentation. The
rate of hydrogen is higher than that of photo fermentation system refers to the development of unique acid-formers and
and bio-photolysis [2]. methane formers in two separate reactors [3–5]. In such
Dark fermentation is a preferred method for resource a system, only slow-growing acidogens and hydrogen
recovery and energy conversion from food waste. It has production microbes are found in the first phase and involve
several advantages, including volume reduction, waste the production of volatile fatty acids (VFAs) and hydrogen,
stabilization, and biogas recovery. In particular, the two-stage while slow-growing acetogens and methanogens are found in
process including bio-hydrogen and methane production with the second phase, in which VFAs are converted to methane
phase separation has considerable potential to enhance the and carbon dioxide.

* Corresponding author. Tel.: þ86 021 65980609.


E-mail address: wangxing24@126.com (X. Wang).
0360-3199/$ – see front matter ª 2008 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2008.09.100
246 international journal of hydrogen energy 34 (2009) 245–254

Although thermal treatment of the inoculum can, to some


Nomenclature extent, improve the fermentative hydrogen production,
however, it consumes energy, increases the operation cost
TS total solid
and complicates the operation procedure since the inoculum
VS volatile solid
should be first heated at 100  C for more than 15 min before
HRT hydraulic retention time
fermentation. Except for the heat treatment of inoculum,
VFA volatile fatty acid
hydrogen production by fermentative bacteria is also highly
NH4-N ammonia-nitrogen
dependent on the conditions of the process, such as solid
C0 VS concentration of the influent (g/kg)
retention time (SRT), organic loading rate (OLR) and gas partial
r(c) substrate removal rate (kg VS/m3 d)
pressure, which affect the microbial metabolic balance and
q organic loading rate (kg VS/m3 d)
subsequently the fermentation end products [6]. The SRT
m0 feeding rate (kg/d)
determines the substrate uptake efficiency, microbial pop-
mH theoretic maximum H2 yield (m3/kg VS)
ulation, and metabolic pathway. In H2 fermentation, it is
mM theoretic maximum CH4 yield (m3/kg VS)
generally assumed that a high SRT causes the growth of H2
kH first order constant for hydrogen
consumers, including methanogens, and competitors for
fermentation stage (d1)
substrates, such as non-H2-producing acidogens [11].
kM first order constant for methane
However, a low SRT may reduce substrate uptake efficiency,
fermentation stage (d1)
active biomass retention, and therefore, the overall process
H2-SCRD semi-continuous rotating drum
efficiency [12]. Changing the OLR can increase the H2 yield.
CH4-SCSTR semi-continuous stirred tank type
However, there is disagreement in the literature as to whether
reactor
higher H2 yields are achieved with lower or higher OLRs. In
SRT solid retention time
some cases higher OLRs decreased the H2 yield whereas in
COD chemical oxygen demand
others higher OLRs increased the H2 yield. In the latter case, as
t time (d).
OLRs increased the H2 yield usually became constant or
Ct VS concentration in the reactor at time t (g/kg)
eventually began to decrease thereby providing an optimal
Vm volume of the reactor (m3)
OLR (maximum H2 yield) [13,14].
k first order constant (d1)
Based on the above information, we developed a two-stage
m theoretic maximum biogas yield (m3/kg VS)
process combined by a mesophilic hydrogen production
y biogas yield (m3/kg VS)
reactor and a methane production reactor. In hydrogen
fermentation reactor, the heat treatment was not applied in
With regard to two-stage fermentation process, the first stage hydrogen fermentation. The indigenous mixed microbial
(fermentative hydrogen production process) does not signifi- culture contained in food wastes was used for hydrogen
cantly reduce the organic content of the feed. Usually, chemical production. No other foreign seed organisms, such as anaer-
oxygen demand (COD) removal is below 20% during hydrogen obic sludge, waste activated sludge was added into the
production process [6], corresponding to a mean hydrogen hydrogen fermentation reactor.
production of 2.5 mol/mol glucose. The undegraded COD can be The objectives of this study were to investigate the (a)
removed in a subsequent methane fermentation stage with the fermentative production of hydrogen from food waste in
conversion of organic content to CH4. Moreover, the first stage semi-continuous rotating drum (H2-SCRD) at various OLR and
(hydrogen fermentation stage) can also be used as an indepen- SRT using an indigenous mixed microbial culture contained in
dent hydrogen production unit but not as a precursor/pretreat- food waste as the hydrogen producer and (b) subsequent
ment for the methanogenic reactor. The present study focuses anaerobic treatment of the effluent of the H2-SCRD with the
on the exploitation of unsterilized food waste as a source for simultaneous production of methane in a semi-continuous
hydrogen and subsequent methane production. stirred tank type reactor (CH4-SCSTR). The two-stage process
It has been reported that in two-phase anaerobic digestion, was operated at semi-continuous and integrated mode. The
H2 and CO2 were produced in the first stage, the effluent of first effects of the stepwise increased organic loading rate (OLR)
stage was transferred to the second stage to be converted to and solid retention time (SRT) on integrated two-stage process
CH4, few measurements on hydrogen production were were also investigated in this bench scale study. A simple
reported in the literature [7]. Previous study found that only kinetic model for anaerobic digestion based on mass balance
small amount of hydrogen was produced since the hydrogen equation was used to evaluate the performance of the
was utilized by methanogens [8]. hydrogen and methane fermentation phase.
To produce hydrogen from a dark fermentation metabo-
lism, the blocking of the methanogenesis in the anaerobic
pathway is one of the key considerations, due to the conver- 2. Materials and methods
sion of hydrogen to methane in this step. The inhibition of the
methanogenic activity can be achieved by controlling pH, 2.1. Feedstock and inoculum for methane fermentation
temperature (heat treatment) or using additives. Of the
various parameters, heat treatment of the inoculum (seed The indigenous microbial cultures contained in food waste
sludge) was usually applied because high temperature can were used for fermentative hydrogen production. Food waste
inactivate hydrogentrophic methanogens, enrich hydrogen- was collected from the restaurants near the experiment
producing bacteria and select spore-forming anaerobes [9,10]. site, and crushed into small particles (2–4 mm). The chemical
international journal of hydrogen energy 34 (2009) 245–254 247

and physical characteristics of the food waste and inoculum anaerobically at batch mode for about 72 h to activate the
are presented in Table 1. hydrogen producer and subsequently switched to the semi-
An anaerobic granular sludge obtained from a UASB tank continuous mode at designated OLR. The whole process
treating cassava waste was used as inoculum for methano- was conducted with OLR increased in steps and maintained
genic reactor. The inoculum has been sufficiently adapted to for each OLR level of about 30 days. The four levels of OLR
the food waste for over 1 year in another experiment. for H2-SCRD were designed to be 15.10, 22.65, 30.20,
37.75 kg VS/m3 d, respectively. The corresponding solid
2.2. Integrated two-stage fermentation system retention time (SRT) for H2-SCRD was 240, 160, 120, 96 h,
respectively. VS content of effluent from H2-SCRD was
An integrated two-stage fermentation system was designed determined periodically, and the OLR for CH4-SCSTR was
for bench scale study (Fig. 1). The system consisted of calculated to be 2.94, 4.61, 6.28, 8.15 kg VS/m3 d. The corre-
a hydrolysis/acetogenesis rotating drum for hydrogen sponding SRT for methane fermentation step was 40, 26.67,
production (total volume and working volume was 0.45 and 20, 16 d, respectively. In order to better evaluate the
0.2 m3, respectively, stainless steel), a methane fermentation performance of the two-stage anaerobic digestion system,
tank (total volume and working volume was 1.5 and 0.8 m3, the pH in both fermentation stages was not adjusted in this
respectively, stainless steel). The system was operated at the study.
integrated mode. Two buffer tanks were installed in this
system, which were used to temporarily hold the food waste 2.4. Analytic method
and effluent of hydrogen fermentation reactor before they are
pumped into H2-SCRD and CH4-SCSTR. Food waste was first The VFAs were determined with gas chromatography (GC).
added into a buffer tank and then feed into H2-SCRD with The samples were centrifuged at 10000 rpm for 10 min and
a peristaltic pump for fermentative hydrogen production, filtered through a 0.45-mm membrane before the VFA deter-
subsequently the effluent of the H2-SCRD was pumped into mination. The filtrate was collected in a 1.5 ml GC vial, and
the other buffer tank and immediately drained into acidified with 3% H3PO4 to pH 4.0 before assayed on an Agilent
CH4-SCSTR to be converted to methane. In order to save the 6890N GC with flame ionization detector and DB-WAXETR
energy consumption and better reflect the large-scale column (30 m  1.0 mm  0.53 mm). Nitrogen was the carrier
processes, the heat treatment was not applied in the first stage gas and the flux was 25 ml/min. The injection port and the
of hydrogen fermentation. detector were maintained at 220 and 250  C, respectively. The
The rotating speed of H2-SCRD was set at 5 rpm and pro- sample injection volume was 1.0 ml. The oven of GC was pro-
grammed to rotate periodically for 3 min, 5 times per hour by grammed to begin at 110  C and to remain there for 2 min,
Siemens Micromaster Vector 420. CH4-SCSTR was equipped then to increase at a rate of 10  C/min to 220  C, and to hold at
with an impeller fixed in the center of the tank at the speed of 220  C for an additional 2 min.
120 rpm. Due to safety consideration, the biogas was not The concentration of lactate was determined using
combusted to generate heat for the bioreactors. The H2-SCRD a liquid chromatograph (Agilent, 1200) equipped with ultra-
and CH4-SCSTR were constructed with water jacket (8 cm violet detector. ODS Hypersil (4.6 mm  250 mm  5 mm) was
thick) to maintain stable temperature for the fermentation used for the column. Chromatography was performed using
process, in which the water was heated by super-heated binary mobile phase composed of phosphoric acid (0.01 mol/
steam supplied from a nearby steam factory. The temperature l NaH2P04) and methanol, 98:2 (v/v), pH 3.0. UV detection
in water-jackets was set at 40  2  C and controlled by was accomplished at 214 nm. The analysis was carried out
temperature sensors. In addition, the whole system was at a detector temperature of 25  C, flow velocity of 1.0 ml/
equipped with other facilities such as food waste grinder, min, and UV 210 nm. The retention time for lactate was
peristaltic pump, biogas bag, and gas meter. 4.8 min.
TS and VS were analyzed according to the standard
2.3. Semi-continuous experiments for fermentative methods [15]. The biogas produced from H2-SCRD and
hydrogen and methane production CH4-SCSTR was measured using a wet gas meter collected in
gas bags. CH4 and CO2 were determined in a 200 ml sample by
The semi-continuous experiments were carried out with gas chromatography (SHIMADZU GC-14B) with a stainless
four OLRs (feeding rate). For start-up, the H2-SCRD (working steel column packed with Carbosive SII (diameter of 3.2 mm
volume 0.2 m3) was filled with food waste, operated and 2.0 m length) and thermal conductivity detector (TCD).
The temperature of the injection, column and detector was set
at 40, 80 and 90  C, respectively. The carrier was nitrogen and
Table 1 – Characteristics of food waste and inoculum the flow rate used was 30 ml/min. A gas standard consisting of
Parameter Food waste Inoculum 60% (v/v) CH4 and 40% of CO2 was used for gas chromatog-
raphy calibration.
TS (%) 17.60  0.35 5.78  0.42
VS/TS (%) 85.91  1.27 63.66  1.68
pH 4.68  0.14 7.04  0.27
2.5. Kinetic models for two-step anaerobic digestion
Total-N (%) 2.34  0.08 0.42  0.15
TCOD (g O2/kg)a 211.79  11.13 93.114  4.36 The mass balance equation with equal mass flow of input/
output and the derivation methodology for the first order
a In wet weight.
kinetic constant described by Linke [16] was used in this study.
248 international journal of hydrogen energy 34 (2009) 245–254

Methane
hot water tank

CH4-SCSTR
food waste H2-SCRD:
Semi-continuous
rotating drum

CH4-SCSTR:
Semi-continuous
H2-SCRD stirred tank type
buffer tank Residue
reactor

Fig. 1 – Schematic diagram of integrated two-stage fermentation process.

The mass balance equation for each fermentation stage can be (d1); m0 the feeding rate (kg/d); m the theoretic maximum
presented as biogas yield (m3/kg VS); y the biogas yield (m3/kg VS); t the
dc time (d).
V ¼ m0 $c0  m0 $c þ V$rðcÞ (1) Results from bench-scale experiment were used to apply
dt
the model. Parameter values of mH, mM, kH and kM can be esti-
The rate of degradation of the substrates was assumed to
mated using non-linear regression fit to the hydrogen and
follow first-order kinetics, thus the following equations were
methane production data with Eq. (8).
used to describe the H2/CH4 production and substrate degra-
dation in semi-continuous operation
dc 2.6. Determination of biodegradability in H2/CH4 step
 ¼ rðcÞ ¼ k$c (2-1)
dt
Biodegradability can better reflect the conversion rate of the
y ¼ uð1  eÞkt (3) substrate into biogas. In present study, the anaerobic biode-
By integrating Eq. (2-1), we obtained gradabilities of the substrate in hydrogen and methane
fermentation stages were calculated from the maximum
cðtÞ ¼ c0 $ekt (2-2) H2/CH4 yield and the COD of the substrate [17], according to
According to the following combination of equations, we
Hy $0:714
obtained Eqs. (4)–(8) BDH ¼ 100 (a)
WH $CODH
9
dc >
>
¼ 0ðsteady stateÞ >
>
dt >
>
>
>
dc >
> My $2:85
Vm ¼ m0 $c0  m0 $c þ Vm $rðcÞ ð1Þ >
= c0 $k$ct
BDM ¼
WM $CODM
100 (b)
dt q¼ ð5Þ;
dc >
> c0  ct
 ¼ rðcÞ ¼ kc ð2-1Þ >
> where BDH and BDM represent the biodegradability in
dt >
>
>
> hydrogen and methane fermentation reactor at a specific OLR,
c0 >
>
Vm ¼ m0 $ >
;
q respectively. Hy and My represent the hydrogen and methane
q$c0 production at a specific OLR(m3), respectively. CODH repre-
cðtÞ ¼ (6)
q þ c0 $k sents the COD of the food waste (in kg O2/kg VS) and CODM
represents the COD of the effluent of H2-SCRD (in kg O2/kg VS)

cðtÞ ¼ c0 $ekt ð2-2Þ
 c0 m at a specific OLR. WH represents the feed rate of H2-SCRD at
kt ¼ (4)
y ¼ u$ 1  e ð3Þ cðtÞ m  y a specific OLR (in kg VS). Wm represents the feed rate of CH4-
SCSTR at a specific OLR (in kg VS). The constants 0.714 and 2.85
9
c0 m correspond to the COD (kg O2) of 1 m3 H2 and CH4, respectively.
¼ ð4Þ > >
>
>
cðtÞ m  y >
>
>
=
c0 $k$ct m$k$cðtÞ
q¼ ð5Þ y¼ ð7Þ;
c0  ct >
> q
>
>
q$c0 >
> 3. Results and discussion
cðtÞ ¼ ð6Þ >
;
q þ c0 $k
m$k$c0 3.1. Biogas production
y¼ ð8Þ
k$c0 þ q
where c0 is the VS concentration of the influent (g/kg); ct the VS The integrated two-stage process for fermentative hydrogen
concentration in the reactor at time t (g/kg); r(c) the substrate and methane production was operated for about 4 months.
removal rate (kg VS/m3 d); Vm the volume of the reactor (m3); q Using indigenous microbial cultures contained in the food
the organic loading rate (kg VS/m3 d); k the first order constant waste as the hydrogen producer, fermentative hydrogen was
international journal of hydrogen energy 34 (2009) 245–254 249

successfully produced in the first stage without foreign inoc- 0.049 m3/kg VS, which was confirmed to be statistically
ulation and heat treatment. significant (ANOVA, P < 0.05).
After the operation of 34 days, steady state was achieved In this bench scale test, the pH was between 5.2 and 5.8 in
for the integrated two-stage process. Biogas yield ranged from H2-SCRD (Fig. 3(a)) without manual interference, the higher
0.167 to 0.236 m3/kg VS in H2-SCRD, the content of H2 was hydrogen yield (0.071 m3/kg VS) was obtained as OLR was kept
28.03–33.02%, CO2 46.68–51.47%, and CH4 was not detected. In at 15.10 kg VS/m3 d with the pH varying from 5.62 to 5.74, in
CH4-SCSTR, biogas yield reached 0.760–0.874 m3/kg VS at comparison with the hydrogen yield of 0.065, 0.054, 0.049 m3/
steady state, the content of CH4 lied between 58.32% and kg VS at the OLR of 22.65, 30.20, 37.75 kg VS/m3 d, respectively.
71.48%, CO2 between 26.34% and 40.73%. In CH4-SCSTR, the The pH variation in the present study agrees with much of the
hydrogen content was determined to be less than 0.2% work published regarding fermentative hydrogen production
through the entire experiment even when operated at the in which a pH value between 5.0 and 6.0 is considered favor-
maximal OLR (8.15 kg VS/m3 d) where large amount of able [21]. It has been reported that at pH 5.5 a peak of hydrogen
hydrogen producer entered the CH4-SCSTR with the effluent production rate was observed. In addition, VS removal and
of hydrogen fermentation reactor. This result suggested that total VFAs production were higher inside the pH range of
methanogens can quickly dominate the culture although the 5.0–6.0 [22].
H2 producers were present at high OLR [18]. It is obvious that the mixed acid fermentation occurred in
The stepwise increased OLR caused the fluctuation of hydrogen fermentation phase in this study. It is also observed
biogas yields in hydrogen and methane fermentation reactor that the formation of hydrogen was accompanied with that of
as shown in Fig. 2, the contents of hydrogen and methane in the VFAs. According to the compositions of the determined
different reactors were also observed to decrease with the VFAs, it can be concluded that butyrate-producing pathway
increase of OLR. An increase of lactic acid concentration was was the dominant hydrogen fermentation pathway at low
observed (from 2345.62 to 4425.56 mg/l) as OLR increased from OLR, which subsequently shifted to lactic acid-producing
15.10 to 37.75 kg VS/m3 d. It has been reported that lactic acid pathway when OLR increased to high level. In this bench scale
has inhibition effect on hydrogen production [19]. In the test, neither heat treatment of inoculum nor foreign inocula-
present study, the pattern of changes in lactic acid concen- tion was applied in the start-up of H2-SCRD. From a practical
tration seemed to be related with hydrogen production. As the point of view, it is notable that the indigenous food waste
concentration of lactic acid increased from 2345.62 to microbial cultures can be used as the mixed microflora inoc-
4425.56 mg/l, the hydrogen yield dropped from 0.071 to ulum in hydrogen fermentation stage. This can be beneficial
to the operation of full-scale plant since no extra energy would
be consumed for the heat treatment, moreover the operation
of the whole process can be facilitated.

3.2. Effects of OLR and SRT on biogas production

The OLRs for hydrogen fermentation stage and methane


fermentation were both kept at four levels, as shown in
Table 2. Each OLR lasted for 30 day. Negative effects of higher
OLR were exhibited as it stepwise increased, resulting in the
decrease in both hydrogen and the subsequent methane
yields. An increase of OLR from 15.10 to 22.65 kg VS/m3 d
caused an insignificant decrease of hydrogen yield from 0.071
to 0.065 m3/kg VS d (ANOVA, P > 0.05), however, the further
increase of OLR from 22.65 to 30.20 kg VS/m3 caused consid-
erable decline of hydrogen yield from 0.065 to 0.054 kg VS/m3,
which was found to be statistically significant (ANOVA,
P < 0.05).
The methane fermentation, using the effluent from H2-
SCRD as substrate, was performed in CH4-SCSTR at OLR
varying from 2.94 to 8.15 kg VS/m3 d. The methane yields of
0.551  0.001, 0.546  0.014, and 0.525  0.003 m3/kg VS were
found at the lower OLR of 2.94, 4.61, 6.28 kg VS/m3 d, respec-
tively. The statistical difference was found to be insignificant
(ANOVA, P > 0.05). This result indicated that the OLR ranging
from 2.94 to 6.28 kg VS/m3 d did not exhibited significant
inhibition on methanogens in CH4-SCSTR. However, with an
increase of OLR from 6.28 to 8.15 kg VS/m3 d, methane yield
rapidly decreased by 8.57% from 0.525  0.003 to
Fig. 2 – Biogas production in (a) H2-SCRD and 0.480  0.021 m3/kg VS, the statistical difference was found to
(b) CH4-SCSTR. be significant (ANOVA, P < 0.05).
250 international journal of hydrogen energy 34 (2009) 245–254

4 -N (mg/l)

881.65  40.81
1417.82  65.63
2134.68  43.54
2625.45  96.14
4 -N(mg/l)
258.14  24.62
350.87  29.51
422.41  31.95
427.52  30.25

NHþ
NHþ

Lactic acid (mg/l)


Lactic acid (mg/l)

12.46  2.54
8.24  1.27
11.38  3.12
9.78  2.68
2345.62  54.32
3216.78  34.53
3645.38  48.17
4425.56  51.68

Metabolites in the effluent from CH4-SCSTR

Butyric acid (mg/l)


Butyric acid (mg/l)

45.32  12.63
48.16  16.44
67.98  15.98
76.25  17.43
Metabolites in the effluent from H2-SCRD

2641.08  54.65
2307.43  74.01
2411.23  63.22
2235.54  74.83

Propionic acid (mg/l)


Propionic acid (mg/l)

76.73  21.16
87.14  16.76
105.21  19.32
112.80  17.71
1241.13  39.67
1861.35  47.49
1743.84  56.34
1758.65  48.65

Acetic acid (mg/l)

322.66  38.87
338.28  65.43
398.41  54.32
432.09  65.76
Fig. 3 – Profiles of pH, VS concentration and OLR in H2 (a)/
Acetic acid (mg/l)

CH4 and (b) fermentation step of two-stage process,


Table 2 – Operation parameters and profiles of metabolites in H2-SCRD/CH4-SCSTR

4355.41  89.74
4153.46  93.45
3623.16  75.29
3426.17  64.76

respectively.

Different letters indicate a significant difference at 5% level Duncan’s multiple range test.
Ethanol (mg/l)

A decrease of VS removal efficiency was observed as the


OLR was gradually increased to higher level, indicating the
reduction of hydrolysis efficiency in hydrogen fermentation
Ethanol (mg/l)

2320.55  57.43
2102.68  64.39
1748.62  48.27
1412.80  72.63

stage. As shown in Fig. 3(a), VS was destructed from 15.81%


N/A
N/A
N/A
N/A

(influent) to 11.77% at OLR of 15.10 kg VS/m3 d with a VS


removal efficiency of 25.55%. The VS content was further
CH4 yield (m3/kg VS)

increased from 11.77% (OLR 15.10 kg VS/m3 d) to 13.14% at the


OLR of 37.75 kg VS/m3 d with a VS removal efficiency of
b
a
a
a
0.551  0.001
0.546  0.014
0.525  0.003
0.480  0.021

16.89%. The decrease of VS removal efficiency was probably


0.054  0.002 b
0.049  0.001 b
0.071  0.001 a
0.065  0.002 a
(m3/kg VS)

attributed to the increasingly greater OLR, since higher OLR


H2 yield

can lead to shorter solid retention time (SRT) thus cause


insufficient time for hydrolysis of insoluble substrates. Bio-
hydrogen is known to be produced as a by-product in the
hydrolysis of various substrates, e.g., carbohydrate. It should
Feed rate

be noticed that the hydrolysis of substrate is commonly


Feed rate

(kg/d)
(kg/d)

20
30
40
50

known as the rate-limited step, therefore any parameter that


20
30
40
50

can negatively affect the hydrolysis process would decrease


the hydrogen yield, such as the parameter of SRT in the
OLR in CH4-SCSTR

present study. From the perspective of hydrolysis efficiency, it


OLR in H2-SCRD

can be emphasized that the lower OLR (from 15.10 to


22.65 kg VS/m3 d) is more suitable for hydrogen fermentation
(kg VS/m3 d)

(kg VS/m3 d)

stage when the indigenous food waste microbial cultures were


used as hydrogen producer.
15.10
22.65
30.20
37.75

2.94
4.61
6.28
8.15

In methane fermentation stage, the suitable OLR ranged


from 2.94 to 4.61 kg VS/m3 d since higher OLR caused the
international journal of hydrogen energy 34 (2009) 245–254 251

Table 3 – Comparable H2 yields from different substrates


Substrate SRT (h) Reactor pH Temperature ( ) H2 yield (ml/g VS) References
a b
Kitchen waste 96–240 Rotating drum 5.2–5.8 40 0.049–0.071 (0.118 ) Present study
Food waste 126 ASBR 5.3 35 80.9 [12]
Food waste 120 CSTR 5.5 55 125.0 [28]
Household solid waste 48 CSTR 4.8–5.2 37 43.0 [9]
Household solid waste 28.8 CSTR 5.8–6.0 60 46.3 [35]

a m3/kg, VS.
b Theoretical maximum yield.

accumulation of NHþ 4 -N in CH4-SCSTR. It is known that the study showed a significant increase ( p < 0.05) in hydrogen
methane production with good performance is strictly regu- yield (as shown in Table 2) when SRT increased from 120 to
lated by environmental factors, such as pH, temperature, VFA 160 h (corresponding to OLR decreasing from 30.20 to
concentration, and NH4þ-N concentration. High concentra- 22.65 kg VS/m3 d). The further increase of SRT from 160 h (OLR
tion of VFA and NHþ 4 -N can inhibit the methane production. As 22.65 kg VS/m3 d) to 240 h (OLR 15.10 kg VS/m3 d) caused an
shown in Table 3, the total concentration of VFAs in CH4- insignificant ( p > 0.05) increase in hydrogen yield. With regard
SCSTR slightly fluctuated from 457.17 to 630.92 mg/l, the pH to the methane production performance, the statistical
varied in the range of 6.7–7.3. With the low concentration of difference was also confirmed to be insignificant ( p > 0.05)
VFAs and the stable fluctuation in pH, their effects on among the methane yields when SRT was kept at 40 d
methane fermentation can be ruled out from the possibilities (2.94 kg VS/m3 d), 26.67 d (4.61 kg VS/m3 d) and (6.28 kg VS/
that may result in the decrease of methane yield. At this point, m3 d), respectively. Therefore, based on the above observa-
the decrease of methane yield was most likely to be caused by tions on the hydrogen and methane yields and the consider-
the accumulation of NHþ 4 -N. As OLR increased in steps, the ation of fermenter size, the optimal SRT and OLR for the
concentration of NHþ 4 -N considerably increased from 0.551 to integrated two-stage process were considered to be 160 h
2.625 g/l in CH4-SCSTR. High NHþ 4 -N concentration is toxic for (22.65 kg VS/m3 d) and 26.67 d (4.61 kg VS/m3 d), respectively.
anaerobes [23,24]. Methanogens are the least tolerant to
ammonia and most likely to cease growth due to ammonia 3.3. Metabolic products
inhibition [25]. McCarty reported that NHþ 4 -N concentrations
in excess of 3000 mg/l were expected to be toxic at any pH The distribution of metabolites is a crucial signal in the
value [26]. Hobson and Shaw reported that 235 mmol assessment of the efficiency of hydrogen-production course.
(3290 mg/l) of NHþ 4 -N completely prevented the growth of The concentrations of the main metabolic products measured
Methanobacterium formicicum in the pure culture [27]. during bench scale tests are presented in Table 3. In the
In H2 fermentation, it is generally assumed that a high SRT effluent of H2-SCRD, the dominant metabolic products were
causes the growth of H2 consumers, including methanogens ethanol, acetic acid and butyric acid. The sum of which
and non-methanogenic H2-consuming or non-H2-producing accounted for 72.20% (9317.04 mg/l) of the determined
acidogens [11]. However, in the present study, higher SRT metabolites (12903.79 mg/l, NH4-N not included) as OLR was
seemed to be beneficial to H2 fermentation since high SRT can kept at 15.10 kg VS/m3 d, revealing the dominance of ethanol,
provide sufficient time for the hydrolysis of substrate and acetic acid and butyric acid producing bacteria in hydrogen
reduce the influx of lactic acid-producing microorganisms. fermentation reactor. As the OLR further increased to
Besides, using indigenous food waste microbial cultures as the 37.75 kg VS/m3 d, the sum percentage of ethanol, acetic acid
hydrogen producer avoided the existence and prevalence of and butyric acid considerably declined to 7074.51 mg/l,
methanogens, which was commonly thought to be the main amounting up to 53.36% of the total determined metabolites
obstacle in fermentative hydrogen. It was found in this study (13258.72 mg/l). It was found that as the concentrations of
that SRT ranging from 160 to 240 h yielded higher hydrogen acetic acid, butyric acid and ethanol were reduced, the pro-
production, which was statistically greater than that at lower pionic acid and lactic acid concentrations increased. There-
SRT. Previous studies reported the similar observations that fore, it can be inferred that branch reactions for the
a higher SRT was required in bio-hydrogen fermentation due production of propionic and lactic acids occurred.
to the slowly degradable organic compounds [20,28]. The H2 At the OLR of 37.75 kg VS/m3 d, the lactic acid concentra-
yield in this test was comparable to other reported values in tion reached the maximum of 4425.56 mg/l, accounting for
H2 production using organic solid waste, as shown in Table 3. 33.37% of the determined VFAs, which was close to the sum
It should be noticed that under the conditions where none- percentage of ethanol and acetate (36.49%). This result
heat treatment was applied (as in the present study) in demonstrated the bloom of lactic acid-producing microor-
hydrogen fermentation, the SRT was relatively longer than in ganism at the highest OLR of 37.75 kg VS/m3 d since at higher
conventional studies which used heat treatment to select OLR more lactic acid producer may enter the hydrogen reactor
spore-forming bacteria and inactivate the hydrogen- with the influx of food waste.
consumer. As mentioned above, the stepwise increased OLR led to the
SRT is an important parameter in determining the declining rate in hydrolysis of substrate, which was probably
fermentation efficiency and the size of fermenter. Present the main cause for the considerable drop of hydrogen yield.
252 international journal of hydrogen energy 34 (2009) 245–254

Besides that, the increase of lactic acid concentration might be


another cause for the decrease of hydrogen yield. The initial
concentration of lactic acid in food waste was 1435.56 mg/l,
revealing the presence of lactic acid-producer before its
entering into H2-SCRD. In other words, large unexpected
influx of lactic acid producer may enter the H2-SCRD at high
OLR, which converted glucose contained in the food waste
into lactic acid, therefore led to the increase of lactic acid
concentration and the reduction in hydrogen yield. According
to the metabolic pathway of glucose, hydrogen producer, i.e.
Clostridium can hardly produce hydrogen from lactic acid as
a sole carbon source [29]. In the present study, an increase of
lactic acid concentration (from 2345.62 to 4425.56 mg/l) was
observed with the OLR from 15.10 to 37.75 kg VS/m3 d. It was
reported that the conversion efficiency of lactic acid to
hydrogen (0.5%) was much lower than that of glucose or
sucrose [30]. Accordingly, the observed increase of lactic acid
concentration was probably another cause for the decrease of
hydrogen production. In addition, the decrease of hydrogen
yield was also likely to be affected by the increase in propio-
nate concentration, since hydrogen cannot be produced in the
propionate production pathway. As shown in Table 2, the
concentration of propionate was 1241.13 mg/l at the lowest
OLR of 15.10 kg VS/m3 d, which was gradually increased to
1758.65 mg/l with OLR increasing up to 37.75 kg VS/m3 d.
Hence, it was inferred that the propionic acid was an impor-
tant indicator of the variation of bio-hydrogen process. It has
been reported that at pH 6.0  0.1 the propionate producers
were active and competed with the ethanol and butyrate
producers for the substrate in hydrogen fermentation [31].

3.4. H2/CH4 theoretical maximum yield and first order


kinetic constant determination

Fig. 4 – Linear correlation between OLR and H2 (a)/CH4 (b)


The estimated curves with theoretical maximum yield of H2/
yield.
CH4 (mH/mM) plotted against OLR show that the first-order
kinetic model described well the degradation of food waste in
H2/CH4 fermentation, as shown in Fig 4. mH and mM were
obtained by curve fitting from Eq. (8), which resulted in thus enhanced the efficiency of methane fermentation stage
0.118  0.001 m3 H2/kg VS and 0.611  0.015 m3 CH4/kg VS. The and increased the hydrolysis rate since hydrolysis is recog-
theoretical maximum yield of H2/CH4 (mH/mM) was equivalent nized as a rate-limited step where the substrate consists of
to the ultimate anaerobic biodegradability and results when particles [34]. With regard to the first-order constant in
OLR was near zero. hydrogen fermentation, mH (0.173  0.007 d1) was much lower
The range of the first-order constant for putrescible than that of methane fermentation, indicating slower hydro-
components in methane fermentation has been reported to be lysis rate in hydrogen fermentation step under the circum-
between 0.10 and 0.35 d1. The first-order constant kM for stance of no foreign inoculation.
methane fermentation was determined to be
0.264  0.046 d1in the present study. Compared with 3.5. Biodegradability in H2-SCRD and CH4-SCSTR
a previous study, a similar constant for methane fermentation
was obtained for the swine wastes anaerobic digestion The biodegradability of the substrate in H2-SCRD and
(k ¼ 0.259 d1). k-value was also found to be 0.211–0.331 d1 CH4-SCSTR indicated the ratio of substrate decomposed in
during the anaerobic digestion of swine wastes with zeolite each step of two-stage fermentation process. As shown in
added in the range of 0–0.5 g zeolite/g VS [32]. Jokela et al. Fig. 5, the biodegradability in both H2-SCRD and CH4-SCSTR
reported much lower constant for the grey wastes decreased linearly with OLR up to 37.75 and 8.15 kg VS/m3 d,
(0.021–0.058 d1). It shall be noticed that the grey wastes are respectively. Linear fit of biodegradability calculated for
rich in cellulose, which are relatively difficult to be hydrolyzed different SRT in each fermentation stage allowed to obtain
compared with putrescibles [33]. In this case, the high value of regression coefficient R2 of 0.986 and 0.997, respectively,
first-order constant in this study was attributed to the fact which gave a good description of linear correlation between
that the hydrogen fermentation stage pre-hydrolyzed the food OLR and biodegradability. The obtained results, calculated
waste, breaking the molecular structure of organic particles, from Eqs. (a) and (b), demonstrated that at optimum
international journal of hydrogen energy 34 (2009) 245–254 253

stage process with none-heat treatment of inoculum. The


conventional heat treatment was replaced by using indige-
nous food waste microflora as hydrogen producer. The bench
scale test demonstrated that the application of indigenous
food waste microflora was applicable for the H2 and CH4
production in the integrated two-stage fermentation process.
From an industrial and commercial perspective, this can
facilitate the hydrogen fermentation process and save energy
consumption.
The bench scale study focused on the exploitation of
unsterilized food waste as a source for hydrogen and subse-
quent methane production, the indigenous food waste
microflora was used as inoculum. At lower OLR (<22.65 kg VS/
m3 d), acetic acid and butyric acid producing pathway were
the dominant hydrogen fermentation pathway, the hydrogen
yield was not significantly fluctuated. At higher OLR
(>22.65 kg VS/m3 d), a decrease in hydrolysis rate of substrate
and an increase of propionic and lactic acids were observed,
which were considered as the main causes for the decrease in
hydrogen yield when the system was operated at high OLR.
The optimal OLR and SRT for the integrated two-stage
process were considered to be 22.65 kg VS/m3 d (160 h) for
H2-SCRD and 4.61 (26.67 d) for CH4-SCSTR, respectively. Under
the optimum conditions, the maximum yields of H2 and CH4
were 0.065 and 0.546 m3/kg VS d with its composition in biogas
of 29.95% and 67.50%, respectively.
With the application of simple mass balance equation, the
theoretic maximum yield for H2 and CH4 were determined to
be 0.118  0.001 m3 H2/kg VS and 0.611  0.015 m3 CH4/kg VS,
respectively. Under the optimum conditions, biodegradability
analysis showed that 5.78% of the influent COD was converted
to the hydrogen in H2-SCRD and 82.18% of the influent COD
was converted to the methane in CH4-SCSTR, thus a total of
87.96% of substrate was converted to biogas in the integrated
Fig. 5 – Biodegradabilities at different OLR in hydrogen and two-stage process.
methane fermentation stages.

references

conditions (OLRH2: 22.65 kg VS/m3 d, OLRCH4: 4.61 kg VS/m3 d)


5.78% of the influent COD was converted to the hydrogen in [1] Das D, Veziroglu TN. Hydrogen production by biological
H2-SCRD and 82.18% of the influent COD was converted to the processes: a survey of literature. Int J Hydrogen Energy 2001;
methane in CH4-SCSTR, therefore a total of 87.96% of 26:13–28.
[2] Benemann J. Hydrogen biotechnology: progress and
substrate was converted to biogas in the integrated two-stage
prospects. Nat Biotechnol 1996;14:1101–3.
process. [3] Azbar BN, Speece RE. Two-phase, two-stage, and single-stage
Based on the experimental results, the mass balance in the anaerobic process comparison. J Environ Eng 2001;127:240–7.
process is summarized as below. Under the optimum condi- [4] Demirel B, Yenigun O. Two-phase anaerobic digestion
tions (OLRH2: 22.65 kg VS/m3 d, OLRCH4: 4.61 kg VS/m3 d), 50 kg processes: a review. J Chem Technol Biotechnol 2002;77:
of wet food waste (VS 15.12%) could produce hydrogen-rich 743–55.
biogas of 1.64 m3 with hydrogen content of around 29.95% [5] Fox P, Pohland FG. Anaerobic treatment applications and
fundamentals: substrate specificity during phase separation.
from hydrogen production reactor, and methane-rich biogas
Water Environ Res 1994;66:716–24.
of 4.98 m3 with methane content of around 67.50% from [6] Antonopoulou G, Gavala NH, Skiadas IV, Angelopoulos K,
methane production reactor by using the two-stage fermen- Lyberatos G. Biofuels generation from sweet sorghum:
tation process. Fermentative hydrogen production and anaerobic digestion
of the remaining biomass. Bioresour Technol 2008;99:110–9.
[7] Liu DW, Liu DP, Zeng RJ, Angelidaki I. Hydrogen and methane
production from household solid waste in the two-stage
4. Conclusion fermentation process. Water Res 2006;40:2230–6.
[8] Van GS, Sung SW, Lay JJ. Biohydrogen production as
In this bench scale test, the fermentative H2 and CH4 were a function of pH and substrate concentration. Environ Sci
successfully produced from food waste in integrated two- Technol 2001;35:4726–30.
254 international journal of hydrogen energy 34 (2009) 245–254

[9] Mu Y, Wang G, Yu HQ. Response surface methodological [22] Jaime MN, Richard D, Alan G. Hydrogen production from
analysis on biohydrogen production by enriched anaerobic sewage sludge using mixed microflora inoculum: effect of pH
cultures. Enzyme Microb Technol 2006;38:905–13. and enzymatic pretreatment. Bioresour Technol 2008;99:
[10] Chang JJ, Wu JH, Wen FS, Hung KY, Chen YT, Hsiao CL, et al. 6325–31.
Molecular monitoring of microbes in a continuous hydrogen- [23] Sprott GD, Patel GB. Ammonia toxicity in pure cultures of
producing system with different hydraulic retention time. methanogenic bacteria. Syst Appl Microbiol 1986;7:358–63.
Int J Hydrogen Energy 2008;33:1579–85. [24] Hendrinksen HV, Ahring BK. Effects of ammonia on
[11] Hawkes FR, Dinsdale R, Hawkes DL, Hussy I. Sustainable growth and morphology of thermophilic hydrogen-
fermentative hydrogen production: challenges for process oxidizing methanogenic bacteria. FEMS Microbiol Lett
optimisation. Int J Hydrogen Energy 2002;27:1339–47. 1991;85:241–6.
[12] Kim SH, Han SK, Shin HS. Optimization of continuous [25] Kayhanian M. Performance of a high-solids anaerobic
hydrogen fermentation of food waste as a function of solids digestion process under various ammonia concentrations.
retention time independent of hydraulic retention time. J Chem Technol Biotechnol 1994;59:349–52.
Process Biochem 2008;43:213–8. [26] McCarty PL, McKinney RE. Salt toxicity in anaerobic
[13] Wu SY, Hung CH, Lin CN, Chen HW, Lee AS, Chang JS. digestion. J. Water Pollut Control Fed 1961;33:399–415.
Fermentative hydrogen production and bacterial community [27] Hobson PN, Shaw BG. Inhibition of methane production by
structure in high-rate anaerobic bioreactors containing Methanobacterium formicicum. Water Res 1976;10:849–52.
silicone-immobilized and self-flocculated sludge. Biotechnol [28] Shin HS, Youn JH. Conversion of food waste into hydrogen by
Bioeng 2006;93:934–46. thermophilic acidogenesis. Biodegradation 2005;16:33–44.
[14] Yang HJ, Shao P, Lu TM, Shen JQ, Wang DF, Xu ZN, Yuan X. [29] Jo JH, Jeon CO, Lee DS, Park JM. Process stability and
Continuous bio-hydrogen production from citric acid microbial community structure in anaerobic hydrogen-
wastewater via facultative anaerobic bacteria. Int J Hydrogen producing microflora from food waste containing kimchi. J
Energy 2006;31:1306–13. Biotechnol 2007;131:300–8.
[15] APHA. AWWA, WEF: standard methods for the examination [30] Logan BE, Oh SE, Van Ginkel SW. Biological hydrogen
of water and wastewater. Washington, USA:: American production measured in batch anaerobic respirometers.
Public Health Association; 1998. Environ Sci Technol 2002;36:2530–5.
[16] Linke B. Kinetic study of thermophilic anaerobic digestion of [31] Hwang MH, Jang NJ, Hyun SY, Kim IS. Anaerobic bio-
solid wastes from potato processing. Biomass Bioenerg 2006; hydrogen production from ethanol fermentation: the role of
30:892–6. pH. J Biotechnol 2004;111:297–309.
[17] Veeken AHM, Hamelers BVM. Effect of temperature on [32] Montalvo S, Guerrero L, Borja R, Travieso L, Sánchez E,
hydrolysis rates of selected biowaste components. Bioresour Dı́az F. Use of natural zeolite at different doses and dosage
Technol 1999;69:249–54. procedures in batch and continuous anaerobic digestion of
[18] Cooney M, Maynard N, Cannizzaro C, Benemann J. Two- synthetic and swine wastes. Resour Conserv Recy 2006;47:
phase anaerobic digestion for production of hydrogen- 26–41.
methane mixtures. Bioresour Technol 2007;98:2641–51. [33] Jokela JPY, Vavilin VA, Rintala JA. Hydrolysis rates, methane
[19] Noike T, Takabatake H, Mizuno O, Ohba M. Inhibition of production and nitrogen solubilisation of grey waste
hydrogen fermentation of organic wastes by lactic acid components during anaerobic degradation. Bioresour
bacteria. Int J Hydrogen Energy 2002;27:1367–71. Technol 2005;96:501–8.
[20] Lay JJ. Modelling and optimization of anaerobic digested [34] Eastman PA, Rico JL, Polanco FF. Anaerobic treatment of
sludge converting starch to hydrogen. Biotechnol Bioeng cheese whey in a two-phase reactor. Environ Technol 1981;
2000;68:269–78. 12:355–62.
[21] Fang HHP, Liu H. Effect of pH on hydrogen production from [35] Ueno Y, Fukui H, Goto M. Operation of a two-stage
glucose by a mixed culture. Bioresour Technol 2002;82:87–93. fermentation process producing hydrogen and methane
from organic waste. Environ Sci Technol 2007;41:1413–9.

You might also like