You are on page 1of 11

Journal of Cleaner Production 289 (2021) 125775

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

Solid-state anaerobic co-digestion of food waste and cardboard in a


pilot-scale auto-fed continuous stirred tank reactor system
Sameena Begum a, b, Tanmoy Das a, Gangagni Rao Anupoju b, Nicky Eshtiaghi a, *
a
Chemical and Environmental Engineering, School of Engineering, RMIT University, 124 La Trobe St, Melbourne, VIC, 3000, Australia
b
Bioengineering and Environmental Sciences Division, EEFF Department, CSIR-Indian Institute of Chemical Technology, Tarnaka, Hyderabad, 500007, India

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

Article history: Increasing volume of Food waste (FW) and cardboard (CB) which are the two major fractions of
Received 29 January 2020 municipal solid waste increases environmental concerns. This study evaluated the performance of a
Received in revised form pilot-scale auto-fed continuous stirred tank reactor (CSTR) system for the solid-state anaerobic co-
11 September 2020
digestion (ACoD) of acetic acid rich food waste (FW) and cardboard (CB). The objective to co-digest
Accepted 29 December 2020
Available online 2 January 2021
acids rich FW and CB originates from the fact that acetic acid can act as a pre-treatment agent. Can
the frequency of feeding and feeding pattern assists the system to withstand an organic loading rate >4 g
Handling editor: Prof. Jiri Jaromir Klemes volatile solids (VS)/(L.d) without addition of any pH boosting chemicals? Four mixing ratios of 100:0,
80:20, 60:40, 50:50 (v/v %) of FW and CB on biogas yield and process stability at an OLR > 4 g VS/(L.d) at a
Keywords: hydraulic residence time (HRT) of 40 d was investigated. Results revealed that the reactor R1 operated
Food waste with FW:CB in the ratio of 100:0 achieved 76% biodegradability (the ratio of experimental to theoretical
Cardboard methane yield) while it was 39%, 32% and 22% for FW:CB ratios of 80:20, 60:40 and 50:50, respectively.
Operational strategy This study concludes that exposing the CB to acetic acid-rich FW assisted the disintegration and depo-
Auto-fed continuous stirred tank reactor
lymerization of CB without any pre-treatment step. Micro-feeding the reactor on an hourly basis assisted
Anaerobic co-digestion
to overcome the susceptible failure of the system. The outcomes of this study provides new insight into
Biodegradability
the potential of implementing decentralized AD units for the treatment of FW and CB at source, which
not only diverts waste from landfills but creates a revenue model through bioenergy generation and
digestate sales.
© 2021 Elsevier Ltd. All rights reserved.

1. Introduction for instance, waxed CB such as pizza boxes, wet and soiled CB are
not recyclable and they are diverted to the landfills. About 40% of CB
Food waste (FW) is one of the largest components of solid waste is sent to landfills every year which creates a similar environmental
attained increasing attention worldwide due to growing concern hazard like FW (Department of the Environment and Energy, 2019).
about its emission of greenhouse gases and leachate generation. Anaerobic digestion (AD) is one of the most efficient technolo-
One third of the food produced for human consumption is wasted gies for waste treatment and stabilization as it enables the recovery
every year, producing an estimated 1.3 billion tonnes of FW of valuable chemicals (bioethanol, volatile fatty acids (VFA)) and
(Ishangulyyev et al., 2019). In majority of the under developed energy (hydrogen, biogas), promoting a sustainable circular bio-
countries, FW is diverted to open landfills rather than being utilized economy (Kuruti et al., 2015; Begum et al., 2016; Begum et al.,
through alternative and energy-efficient routes of valorization 2018). Increasing interest in AD of FW had resulted in extensive
(Garcia-Garcia et al., 2019). Cardboard (CB) is another major frac- laboratory-scale investigations (as reviewed in Ren et al., 2018;
tion of municipal solid waste. About three tons of wood is required Wang et al., 2018) while mono-digestion of CB is scarcely reported
to produce one ton of CB (General Kinematics, 2019). Recycling CB is because of its poor biodegradability due to lignin despite contain-
the best option to produce new CB which offers dual environmental ing significant organic matter in the form of cellulose and hemi-
and economic benefits. It is also known that not all CB is recyclable cellulose (Li et al., 2019). To the best of our knowledge no full-scale
plant has been installed for the mono-digestion of CB whereas
there are numerous full-scale demonstrations for FW (De Clercq
* Corresponding author. et al., 2019; Yang et al., 2019). The full-scale FW digesters often
E-mail address: nicky.eshtiaghi@rmit.edu.au (N. Eshtiaghi). tend to fail, if operated at higher organic loading rates (OLR) due to

https://doi.org/10.1016/j.jclepro.2020.125775
0959-6526/© 2021 Elsevier Ltd. All rights reserved.
S. Begum, T. Das, G.R. Anupoju et al. Journal of Cleaner Production 289 (2021) 125775

rapid acidification. Most studies to date have focused on the gen- between 24 and 90 d (Kuruti et al., 2017; Irena.org, 2019).
eration of biogas from FW at a maximum OLR of 4 g VS/(L.d) (VS is This study demonstrated the solid-state ACoD of acetic acid rich
volatile solids) using external pH-boosting chemicals to maintain FW and CB in a pilot-scale auto-fed CSTR system. To find the
the pH in the range of 6.8e7.2, as this is a favorable range for the optimal FW:CB mixing ratio, four different mixing ratios of 100:0,
methanogenic phase of AD to occur. It was reported that an OLR 80:20, 60:40 and 50:50 (v/v %) with a corresponding total solids
>4 g VS/(L.d) causes digesters to fail due to overloading (Meng et al., (TS) content of 19%, 21%, 23% and 24% were investigated. This study
2014; Kuruti et al., 2017). attempted to answer two major research questions on reactor
Anaerobic co-digestion (ACoD) of FW and CB is gaining interest operation. Firstly, can an anaerobic digester operated with acetic
to overcome the limitations associated with mono-digestion acid rich FW in mono or co-digestion without addition of any pH
(Guilford et al., 2019). Co-digestion of FW with other co- boosting chemical at an OLR greater than 4 g VS/(L.d) as it is the
substrates such as sewage sludge, cattle manure, yard waste and threshold value for safe operation? Secondly, can the feeding
domestic waste water was proven to be economically feasible to pattern and the frequency of feeding influence the overall stability
overcome the accumulation of VFAs and high solids which leads to of the CSTR system? The study investigated the ACoD of FW and CB
inefficient performance or process failure due to sudden pH drop at a full HRT of 40 d at a corresponding OLR values in the range of
(Panigrahi et al., 2020; Zhang et al., 2013). A recent study reported 4.7e5.7 g VS/(L.d). The CSTR system was initially operated at 90 d
on co-digestion of FW and CB in batch trials by Asato et al. (2019) HRT, followed by 80 d and 60 d at an interval of 10 d with OLR below
investigated the impact of mixing ratios, total solids and food to 4 g VS/(L.d) to allow acclamatization of micro-organisms with the
inoculum (F/I) ratio on biogas or methane yield. They stated that substrate. The CSTR system was operated at a complete HRT of 40 d
greater fractions of FW resulted in higher methane production rates (OLR > 4 g VS/(L.d)) to find the optimal ratio of CB that can be co-
and yields, while greater fractions of CB led to longer lag phases. In digested with FW and overall process stability. The outcome of
another study, Capson-Tojo et al. (2017) studied the impact of F/I this study was used for a techno-economic performance analysis of
ratios of 0.25, 1 and 4 gVS/gVS in dry ACoD of FW and CB and found a full-scale AD plant at one-tonne capacity.
that the reactors produced methane only at F/I of 0.25. An impor-
tant issue for research is that the performance of laboratory scale 2. Materials and Methods
studies may not match the performance at full scale, possibly the
semi-continous and continuous studies would. Therefore, pilot- 2.1. Substrate and inoculum collection
scale demonstrations on ACoD of FW and CB is needed.
Laboratory scale investigations reported on co-digestion of FW 25 L of coarsely ground acetic acid rich FW slurry and inoculum
and CB incorporating physical (size reduction), mechanical (ultra- were collected from a mesophilic anaerobic digester operated with
sound, microwave etc.) and/or chemical (acid/alkali) pre-treatment FW at the Aurora sewage treatment plant (Preston, Australia), in
step demonstrated enhanced process performance but the eco- airtight polypropylene containers. The mesophilic inoculum was
nomic viability is lower (Zeynali et al., 2017). Addition of buffering stored in an incubator at 37  C for two weeks to allow degassing,
chemicals and nutrients in the pre-treatment step would incur and the FW slurry was stored in the refrigerator at 4  C until the
extra operational costs for a full-scale AD plant operated with any start of experiments. The co-substrate used for experimentation
substrate particularly FW and CB. Previous studies have demon- was corrugated CB collected from recycle bins in RMIT University.
strated that AD is very sensitive to disturbances in pH, temperature
and OLR as it is a highly dynamic and multi-step process resulting 2.2. Analytical methods
from a continuous interaction between substrate, microorganisms/
enzymes and operating conditions which creates a delicate equi- Total solids and volatile solids (VS) were measured according to
librium mainly dependent on substrate characteristics (Fisgativa APHA Methods 2540 B and 2540 E respectively (APHA, 1998). The
et al., 2016; Kuruti et al., 2017). The main controls of AD opera- chemical oxygen demand (COD) (total and soluble), ammonia ni-
tion are OLR, hydraulic residence time (HRT), temperature, mixing trogen (NH3eN) and volatile acids (VA) were determined by
and the feeding pattern. When translating the results from batch to colorimetric techniques using a HACH (Model: DR/5000 U) spec-
continuous, the frequency of feeding and the feeding pattern is trophotometer according to APHA Methods 8000, 10031 and 8196,
often overlooked (Svensson et al., 2018). For instance, one time respectively (Zahan et al., 2018). The samples were centrifuged
feeding in done in batch reactors, one time a day in continuously (Eppendorf 5702, Germany) at 10,000 rpm for 5 min and then
stirred tank reactors (CSTR) while the full-scale AD plants are fed filtered through 0.45 mm filter paper (mixed cellulose-ester mem-
continuously throughout the day at regular intervals. Feeding brane filter, Advantec, Japan) to measure the soluble organic con-
patterns and its frequency are the two significant parameters that stituents. The density of the CB and FW slurry was calculated from
influences the overall performance of the digesters (Ziels et al., its mass and volume. The slurry’s pH was measured using a cali-
2017). A case study investigated by Svenson et al. (2018) proved brated pH meter (Thermo Orion, Model 550 A), and its elemental
that frequently fed digesters produced 20% more methane from FW composition was determined using a LECO (TruMac) analyzer.
and had lower effluent concentrations of long chain fatty acids
compared to one time fed reactor. Therefore, to assess the techno- 2.3. Preparation of feed mixture (FW and CB) and characterization
economic performance analysis of a full-scale AD plant based on
laboratory scale outcomes, similar feedstock, feeding pattern and The corrugated CB sheets were cut into pieces of 3e4 cm in
frequency of feeding to the reactors should be taken into consid- length using scissors. The density of the FW was measured and
eration. A new operational approach to treat FW and CB containing found to be 1060 kg/m3 while the density of CB was 72 kg/m3. The
high solids >20% without any buffering chemicals and/or pre- feed mixture of FW and CB was prepared in ratios of 100:0, 80:20,
treatment step, is required to fulfil essential conditions such as 60:40 and 50:50 on the basis of volume. For example, to prepare a
robustness and high adaptive capacity of the AD system. Stable mixture of FW:CB 50:50, the equivalent weight of 50 mL CB
operation at high OLR, low HRT and good economic performance (0.0036 g) was added to 50 mL of FW slurry and soaked for 1 h. The
are necessary for a process to be implemented at full scale soaked feed mixture was ground into a paste using a heavy-duty
(Cavinato et al., 2013). The optimum HRT required for AD of sub- blender. Characteristics of the raw FW slurry, feed mixture sam-
strates such as animal manures, FW and agricultural waste falls ples, CB and inoculum are presented in Table 1. The pH of the acetic
2
S. Begum, T. Das, G.R. Anupoju et al. Journal of Cleaner Production 289 (2021) 125775

Table 1
Physico-chemical characteristics of inoculum, CB and feed mixtures.

Parameter Units Inoculum CB FW:CB (100:0) FW:CB (80:20) FW:CB (60:40) FW:CB (50:50)

pH () 7.4 ± 0.2 ND 3.6 ± 0.1 3.7 ± 0.1 3.7 ± 0.1 3.8 ± 0.1
TS % 4.2 ± 0.1 93 ± 0.2 19 ± 0.2 21 ± 0.2 23 ± 0.2 24 ± 0.2
VS % 2.4 ± 0.1 85 ± 0.1 18 ± 0.1 20 ± 0.1 21 ± 0.1 22 ± 0.1
tCOD g/L 37.6 ± 0.2 ND 624 ± 0.2 550 ± 0.2 518 ± 0.2 520 ± 0.2
sCOD g/L 11.6 ± 0.1 ND 68 ± 0.1 70 ± 0.1 71 ± 0.1 72 ± 0.1
NH3eN g/L 2.5 ± 0.1 ND 1.3 ± 0.1 0.4 ± 0.1 0.8 ± 0.1 0.7 ± 0.1
VA g/L 0.3 ± 0.02 ND 42 ± 0.1 43 ± 0.1 43.5 ± 0.1 43 ± 0.1
Carbon (C) % ND 49 51 ND ND ND
Hydrogen (H) % ND 5.4 10 ND ND ND
Nitrogen (N) % ND 0.3 0.7 ND ND ND
Sulfur (S) % ND 0.3 0.2 ND ND ND
Oxygen (O) % ND 0.2 0.4 ND ND ND

ND-Not determined.

acid rich FW slurry was lower i.e. < 4.0 and the carbon to nitrogen includes feed, reactor and digestate valves. The temperature of each
(C/N) ratio of FW and CB was high i.e. 73 and 163 respectively, reactor is controlled by an electric heating jacket wrapped around
which are not favorable for the AD process, but the operational the external reactor wall. A touchscreen programmable logic
strategy significantly reduced the impact of these limitations. controller (PLC) is used to program the feeder mechanism and for
data logging and monitoring. Gas generation is measured using a
2.4. Equipment description liquid displacement gas flow meter module, which is a single
perspex block with six perspex tumbling buckets of active volume
The equipment used in the present study was the patented 22e24 mL each. Each reactor is also equipped with 1 L bottles for
auto-fed CSTR system (Lobster-max-I model) with four reactors the collection of digestate.
namely R1, R2, R3 and R4 as shown in Fig. 1. The reactors or
anaerobic digesters are cylindrical vessels made of stainless steel 2.5. Startup and operational strategy
with a reactor volume of 5.2 L and an effective working volume of
5 L and feed syringe volume of 1.5 L each. Each reactor is equipped The feeders of each of the reactors were initially filled using the
with its own beam to automatically feed the digester at the set feed top-up feed syringe, keeping the reactor and feed valve open to
rate and time. The feed from the feeder syringe is delivered to the prevent a gas lock. The FW:CB ratios in the feed mixture were
reactor from a feeder port at the bottom using a unique patented 100:0, 80:20, 60:40 and 50:50 for R1eR4 respectively. The feeder
feeding system that moves vertically upwards. A non-return syringes were topped up with fresh feed periodically using the top-
membrane value is fitted between the digester and the feeder sy- up syringe, which pushes the piston downwards automatically.
ringes to prevent the movement of digestate into the feeder section. Subsequently, the reactors were inoculated with 5 L of inoculum
Mixing in the reactors is performed by a gas-tight mixer shaft until it started to flow into the digestate bottles by keeping the
connected to a 24 V DC motor rotating at 45 rpm. The equipment reactor-feed isolation valve closed and the digestate valve open.

Fig. 1. The Lobster-max-I auto-fed CSTR digester system.

3
S. Begum, T. Das, G.R. Anupoju et al. Journal of Cleaner Production 289 (2021) 125775

The tubing from the reactors was connected to the gas flow meter gVSadded) was observed over days 11e20 (HRT of 80 d) in R1, while
module to estimate the biogas generation potential of the FW and in R2 it decreased by 10% (from 0.30 to 0.27 L/gVSadded) and in R3
CB mixture in each reactor. and R4 it increased by 79% (0.58e1.04 L/gVSadded) and 18.4%
After filling the feeder syringes and inoculating the reactors, the (0.38e0.45 L/gVSadded) respectively. Fig. 2A also shows that during
input parameters (dynamic parameters) were set in the PLC days 21e30 (HRT of 60 d) the biogas yield increased by 28.6%
controller as summarized in Table 2. The temperature in the re- (1.15e1.48 L/gVSadded) and 185% (0.27e0.77 L/gVSadded) in R1 and
actors was maintained at 37 ± 0.1  C and the experiments were run R2 respectively, while in R3 and R4 it decreased by 41.3%
over 70 d. The reactors were operated at a stepwise decreasing HRT (1.04e0.61 L/gVSadded) and 28.8% (0.45e0.32 L/gVSadded) corre-
of 90 d, 80 d and 60 d (for a duration of 10 d each to allow sufficient spondingly. At the final and full HRT of 40 d, average biogas yields of
time for the inoculum to acclimatize with feed mixture) and a full 0.47, 0.31, 0.07 and 0.18 L/gVSadded were observed from R1, R2, R3
HRT of 40 d at an OLR between 4.7 and 5.7 g VS/(L.d). The beam and R4 respectively, with the system attaining stability. These ob-
speed of the reactors was calibrated early in the experiment and servations indicated that the biogas yield decreased as the HRT
maintained until the end (Table 2). The number of feed cycles per decreased in reactors R1 and R2, and in contrast, in reactors R3 and
day, feed cycle period and the feed delivery rate were constant R4 the biogas yield increased as the HRT decreased. This implies
throughout the experiment. The reactors were fed every hour i.e. 24 that an increase in the CB:FW ratio improved the buffering capacity
feed cycles per day. As the HRT decreased from 90 d to 40 d, OLR, and enhanced the synergistic methanogenic mechanism.
feed volume per day, feed volume and feed time per cycle increased Fig. 2B reveals that the cumulative biogas yield in R1 was 0.74 L/
accordingly, as presented in Table 2. In spite of the unfavourable pH gVSadded which is significantly higher than the yield achieved from
of the feed (<4.0), the solid-state ACoD of FW and CB mixture was R2, R3 and R4 (0.32, 0.30 and 0.21 L/gVSadded respectively). It can be
feasible with process stability because of frequent micro-feeding clearly stated that the biodegradability of the FW:CB in the ratio of
mechanism on an hourly basis. 100:0 in R1 was higher than the biodegradability of FW:CB in the
ratio of 80:20, 60:40 and 50:50 in R2, R3 and R4 correspondingly. It
3. Results and discussion is known that VFA is the precursor for methanogens for the pro-
duction of biogas through methanogenesis (Begum et al., 2020). In
3.1. Performance analysis of solid-state ACoD of FW and CB in an reactor R1 the feed contained only FW without any CB, therefore
auto-fed CSTR system the VFA available in FW was utilized by methanogens to produce
more biogas, while in reactors R2, R3 and R4 the VFA was being
3.1.1. Impact of HRT on daily and cumulative biogas yield absorbed by CB for the degradation of cellulose. The average cu-
The impact of HRT on daily and cumulative biogas yield over the mulative biogas yield in R1 and R3 sharply increased as HRT
digestion time during mesophilic ACoD of FW and CB in different decreased from 90 d to 60 d, while in R2 and R4 the difference in
mixing ratios in reactors R1, R2, R3 and R4 is shown graphically in average cumulative biogas yield was smaller (Fig. 2B). At the
Fig. 2. Fig. 2A shows that the daily biogas yield increased sharply beginning of 40 d HRT, the average cumulative biogas yield
from 0.1 to 2.0 L/gVSadded during the first 10 d of start-up (HRT of decreased in reactors R1, R3 and R4, while in R2 it increased
90 d) in R1, whereas in R2, R3 and R4 the increase in biogas yield slightly. It is interesting to note from Fig. 2B that the reactor R2 with
was 0.05e0.6 L/gVSadded. This could be attributed to the volume of FW:CB 80:20 attained stability with stable biogas yield in the range
inoculum (abundance of methanogens) available initially compared of 0.3e0.35 L/gVSadded within the first 10 d of operation (HRT of
to the volume of the input substrate in all the reactors. Subse- 90 d) while in reactor R4 the system attained stability in the first 15
quently, a sharp decrease in biogas yield of 42.5% (from 2.0 to 1.15 L/ days but the biogas yield relatively decreased by 40% (0. 35 to

Table 2
The fixed and dynamic paramaters showing the operating conditions of auto-fed CSTR system.

Reactor/Operational parameters R1 (FW:CB, 100:0) R2 (FW:CB, 80:20) R3 (FW:CB, 60:40) R4 (FW:CB, 50:50)

Fixed parameters
Beam speed (mm/h) 108 111 111 111
Feed cycle period (min) 60 60 60 60
Feed cycles per day 24 24 24 24
Feed delivery rate (mL/min) 12.75 12.75 12.75 12.75
Dynamic parameters
HRT (d) 90 90 90 90
OLR (g VS/(L.d) 2 2.3 2.4 2.5
Feed volume per day (mL) 57.7 57.7 57.7 57.7
Feed volume per cycle (mL) 2.4 2.4 2.4 2.4
Feed time per cycle (s) 12.2 12.2 12.2 12.2
HRT (d) 80 80 80 80
OLR (g VS/(L.d) 2.3 2.6 2.7 2.9
Feed volume per day (mL) 65 65 65 65
Feed volume per cycle (mL) 2.7 2.7 2.7 2.7
Feed time per cycle (s) 12.73 12.73 12.73 12.73
HRT (d) 60 60 60 60
OLR (g VS/L/d) 3.1 3.5 3.6 3.8
Feed volume per day (mL) 86.6 86.6 86.6 86.6
Feed volume per cycle (mL) 3.6 3.6 3.6 3.6
Feed time per cycle (s) 16.98 16.98 16.98 16.98
HRT (d) 40 40 40 40
OLR (g VS/(L.d) 4.7 5.2 5.5 5.7
Feed volume per day (mL) 130 130 130 130
Feed volume per cycle (mL) 5.4 5.4 5.4 5.4
Feed time per cycle (s) 25.47 25.47 25.47 25.47

4
S. Begum, T. Das, G.R. Anupoju et al. Journal of Cleaner Production 289 (2021) 125775

Fig. 2. Impact of HRT on biogas generation profile of anaerobic co-digestion of high-solids FW and CB at four mixing ratios (A) Daily biogas yield; (B) Cumulative biogas yield.

0.21 L/gVSadded). In reactors R1 and R3, the biogas yield sharply respectively. This could be attributed to the frequent micro-feeding
increased till 30 days of operation (i.e. from HRT of 90 de60 d) strategy of the substrate (which was at low pH (<4.0)) in the re-
followed by a decrease at a complete HRT of 40 d. This signifies that actors on an hourly basis which facilitated the complete utilization
the reactors R1 and R3 took longer time to attain stability. of the organic matter by the methanogens, in addition to the OLR
remaining below the threshold of 4 g VS/(L.d). This could be
credited to the feeding pattern, that is, the feed volume per cycle
3.1.2. Impact of OLR on cumulative biogas yield
(mL) at HRTs of 90 d, 80 d and 60 d, which was 2.37 mL, 2.7 mL, and
The increase in OLR corresponding to the decrease in HRT from
3.6 mL respectively per hour in all the reactors (Table 2). No buff-
90 d to 40 d and its impact on cumulative biogas yield is presented
ering chemicals were added to the substrate (which was at low
in Fig. 3. In reactors R1 and R3, the cumulative biogas yield
pH < 4.0) to increase the pH to neutral before feeding, because the
increased sharply from 0.1 to 1.1 L/gVSadded and 0.07e0.57 L/
ratio of volume of inoculum in the reactor to volume of feed fed to
gVSadded, respectively, as the OLR increased from 2 to 3.1 gVS/(L.d)
the reactors per hour was high, which balanced the pH in the
in R1 and 2.3 to 3.5 gVS/(L.d) in R3 respectively on the basis of HRT.
reactor by creating self-buffering in the system. This suggests that
In reactors R2 and R4, gradual increase in cumulative biogas yield of
the buffering capacity in the reactors could be retained with
0.05e0.38 L/gVSadded and 0.05e0.35 L/gVSadded were observed as
smaller amounts of feed being fed to the reactors frequently.
the OLR increased from 2.4 to 3.6 gVS/(L.d) and 2.5 to 3.8 gVS/(L.d),
5
S. Begum, T. Das, G.R. Anupoju et al. Journal of Cleaner Production 289 (2021) 125775

Fig. 3. Impact of increasing OLR as a result of decreasing HRT on cumulative biogas yield during solid-state anaerobic co-digestion of FW and CB.

Decrease in cumulative biogas yield from 1.1 to 0.74 L/gVSadded, chemicals in comparison to other mixing ratios of FW:CB. Never-
0.57 to 0.29 L/gVSadded and 0.35 to 0.29 L/gVSadded were observed in theless, the biogas yield from monodigestion of FW (i.e. FW:CB
reactors R1, R3 and R4 after the beginning of 40 d HRT; i.e. increase 100:0) was superior among all the ratios.
in OLRs from 3.1 to 4.7 gVS/(L.d) in R1; 3.6 to 5.5 gVS/(L.d) in R3;
and 3.8 to 5.7 gVS/(L.d) in R4 correspondingly. In reactor R2, the 3.1.3. Impact of volatile acids (VA) present in FW on CB degradation
decrease in biogas yield from 0.38 to 0.32 L/gVSadded as the OLR at OLRs of 2e5.7 g VS/(L.d)
increased from 3.5 to 5.2 gVS/(L.d) was lower compared to those in It is well documented that the ACoD process performance of
other reactors (Fig. 3). This could be due to the impact of the FW:CB lignocellulosic biomass, such as wheat straw, rice straw and CB,
mixing ratio (80:20) that was adopted by the methanogens, even improves after pre-treatments such as acid, alkali, ultrasonication
though the feed volume per cycle increased to 5.4 mL as the OLR and/or combinations of these (Zeynali et al., 2017). This is because
increased. This could be the reason reactor R2 attained stability in the pre-treatment disintegrates the structure of the biomass and
biogas production with in the first 10 days of operation. As the OLR cell lysis occurs, which releases the intracellular organic matter
increased in reactors R1, R3 and R4, the feed volume per cycle (mL) (monomers such as sugars, amino acids and glycerol) and makes
also increased from 3.6 mL to 5.4 mL per hour (Table 2), which was these more bioavailable thereby improving the reaction rates and
reflected in the increased amount of CB fed to the reactors in R3 and efficiency of the digestion process. In particular, acid (H2SO4) and/or
R4 that resulted in the decrease in biogas yield. However, in reactor alkaline (NaOH) pre-treatment of lignocellulosic biomass improves
R1 the decrease might have been due to organic shock loading. the biogas yield compared to biomass without any pre-treatment
Although a decrease in biogas yield from 1.1 to 0.74 L/gVSadded was (Mosier, 2005). In the present study, the VA in the form of acetic
observed in R1, it was still significant as the reactor was operated at acid (CH3COOH) which is a weak acid favored the efficient solubi-
an OLR >4 g VS/(L.d) which is a threshold value for safe operation lization of CB by disintegrating its structure during 90 d HRT,
with FW, however an HRT of 40 d might have led to attain stability although it is a weak acid in FW. Acetic acid is also a major
even at higher OLR. In reactors R3 and R4, lower biogas yield was degradation product of newsprint paper, among other low molec-
observed at the full HRT of 40 d because the CB:FW ratios were ular weight carboxylic acids such as formic, lactic, malonic, malic
40:60 and 50:50 respectively, which resulted in lower biodegrad- and succinic acids (Gibson and Watt, 2010). C3 and C4 alkyl
ability of CB at reduced HRT. The overall performance of the CSTR substituted cyclohexanols and C4 to C12 aliphatic alcohols, among
system showed that the reactor with FW:CB of 80:20 is better in numerous volatile organic compounds, have also been identified as
terms of attaining quick process stability and stable biogas yield main degradation products in cotton and chemical wood pulp
during ACoD without the need of any pre-treatment or buffering books. Dupont and Tetreault (2000) observed immediate
6
S. Begum, T. Das, G.R. Anupoju et al. Journal of Cleaner Production 289 (2021) 125775

significant depolymerization of cellulosic material after it was in the range of 45e65% in all the four reactors. This decline in VS
exposed to acetic acid in the concentration range of 20e200 mg/m3 removal efficiencies is mainly because of the high solids present in
(277e2,777 mg/kg) but not after exposure to 3 mg/m3 (41 mg/kg). the substrate (i.e. > 19% TS, which is regarded as solid-state or dry
Lignocellulose has a recalcitrant and crystalline structure that re- AD). It is interesting to note that the digestate pH in all the reactors
sists bacterial attack and biodegradation. Subjecting CB to acetic was between 6.8 and 7.5 throughout the operation, which suggests
acid-rich FW increased the biodegradability of lignocelluloses by that efficient digestion had occurred (Fig. 4). The current study
decreasing its crystallinity and eliminating its network structure. investigated the performance of FW and CB ACoD at an HRT of 40 d
Therefore, a major finding of this study is that acetic acid-rich FW to understand its techno-economic feasibility at full-scale.
can be utilized to degrade CB during ACoD of FW and CB which Although the co-substrate CB is not as easily biodegradable as
eliminated the use of external pre-treament chemicals. FW, the reactor start-up approach and micro-feeding strategy
The concentration of acetic acid in FW that interacted with the improved the overall performance of FW and CB co-digestions
amount of CB fed to the reactors R2, R3 and R4 and its impact on substantially and the system attained steady state without the
biogas yield against OLR and HRT, is summarized in Table 3. To need for external buffering. As noted earlier, depolymerization of
make a comparison of biogas yield obtained in CSTRs as a function cellulose by exposure to acetic acid is more likely to occur over the
of OLR irrespective of the FW:CB mixing ratios, biogas yield of long term. Overall, to enhance the performance of FW and CB
reactor R1 is also presented in Table 3. Dupont and Tetreault (2000) anaerobic co-digestion, a pre-treatment step is necessary.
found that the degradation of cellulose by exposure to acetic acid is
dependent on the concentration of the acid and also on the aging or 3.3. Mass balance
exposure time. Table 3 shows that the increase in the CB ratio in FW
from 20% in R2 to 40% in R3 and 50% in R4 decreased the concen- The CSTR system was continuously operated for 70 d, with
tration of acetic acid from 3.1 to 1.5 to 1.2 gVA/gCB, respectively. decreasing HRT from 90 d to 80 d followed by 60 d and 40 d. The TS
Moreover, in reactor R2, the average cumulative biogas yield in the total solid of substrate which is the sum of VS and fixed solids
increased from 0.24 to 0.35 (L/gVSadded) as the HRT reduced from (FS). The mass balance in terms of input TS, VS and FS fed to the
90 d to 40 d (Table 3). In reactors R3 and R4 the ratio of CB in FW reactors and the output in terms of biogas generation, FS and un-
was higher, resulting in lower interaction of VA with CB and digested VS is shown in Fig. 5. During the digestion time of 70 d, the
reduced degradation, leading to lower biogas yield than 20% CB in reactors R1eR4 were fed with TS of 1386 kg, 1530 kg, 1679 kg and
FW. Dupont and Tetreault (2000) also reported that the depoly- 1750 kg, of which the total VS was 1310 kg, 1460 kg, 1535 kg and
merization of cellulose by exposure to acetic acid is more likely to 1600 kg respectively. FS is the non-biodegradable inert component
occur over the long term, and the same was observed in R3 and R4. of the digestate. Biogas generation of 974 L, 551 L, 452 L and 333 L
From HRT of 90 de60 d, the average cumulative biogas yield over 70 d of digestion time was observed from the total VS fed to
increased from 0.31 to 0.56 (L/gVSadded) and 0.27 to 0.35 (L/gVSad- R1eR4 respectively. The undigested fraction of VS and FS consti-
ded) in reactors R3 and R4, while at 40 d, the biogas yield decreased. tutes the TS in the digestate. The biogas generation and VS removal
This implies that co-digestion of 40% and 50% CB in FW needs in R1 were higher than in R2, followed by R3 and R4. The major
longer HRT. In reactor R1, as the OLR increased from 2 to 4.7 gVS/ reason was the composition of the substrate, because the CB de-
(L.d), the biogas yield decreased from 0.92 to 0.74 L/gVSadded. The grades more slowly than FW.
decrease in biogas could be attributed to organic shock loading at
an OLR 4.7 gVS/(L.d) but the system still attained stability at higher 3.4. Theoretical v/s experimental methane yield of solid-state ACoD
OLR with notable biogas yield. of FW and CB

3.2. Impact of increasing CB content on VS removal efficiencies and The theoretical methane yield (TMY) from FW, CB, and the
digestate pH during ACoD of FW and CB mixture of FW and CB in different ratios was estimated from the
elemental composition of FW and CB (Table 1) using Eq. (1)
The digestate samples collected from the digestate bottles of (Achinas and Euverink, 2016).
reactors R1, R2, R3 and R4 after every five days were analyzed for VS
and pH. The VS removal efficiencies and pH values of the digestate 22:4*½2a þ 8b  4c  3d e
8  4
TMY ¼
are shown in Fig. 4. The VS removal efficiencies were between 75 12:017*a þ 1:0079*b þ 15:999*c þ 14:0067*d þ 32:065*e
and 90% during the first 20 d (i.e., HRT of 90 and 80 d). At an HRT of (1)
60 d, the VS removal efficiencies declined by 10% i.e. 65e80%. This
reduction in VS removal efficiency could be due to the decrease in Where a, b, c, d and e are the masses of the elements C, H, O, N and S
HRT and an increase in OLR. This observation elucidates the exis- respectively divided by their molar masses in the numerator. The
tence of an inverse correlation between organic matter removal TMY calculated using Eq. (1) does not determine the biodegrad-
and OLR. The VS removal efficiencies further decreased at an HRT of ability and non-biodegradability of the substrate, therefore, a
40 d, but the system reached steady state with an average removal comparison of the calculated TMY of the substrate and the

Table 3
Impact of VA concentration in FW on the hydrolysis of CB (g VA/g CB) and biogas yield in the reactors R2 (FW:CB 80:20), R3 (FW:CB 60:40) and R4 (FW:CB 50:50) at OLRs of
2.3e5.7 g VS/L/d compared to rector R1 (FW:CB 100:0) at OLR of 2e4.7 g VS/L/d.

R1 (FW:CB 100:0) R2 (FW:CB 80:20) R3 (FW:CB 60:40) R4 (FW:CB 50:50)

HRT OLR (gVS/L/ Biogas yield (L/ OLR (gVS/L/ gVA/ Biogas yield (L/ OLR (gVS/L/ gVA/ Biogas yield (L/ OLR (gVS/L/ gVA/ Biogas yield (L/
(d) d) gVS) d) gCB gVS) d) gCB gVS) d) gCB gVS)

90 2 0.92 2.3 3.0 0.24 2.4 1.5 0.31 2.5 1.2 0.26
80 2.3 1.0 2.6 3.0 0.32 2.7 1.5 0.49 2.9 1.2 0.30
60 3.1 1.1 3.5 3.0 0.34 3.6 1.5 0.56 3.8 1.2 0.35
40 4.7 0.74 5.2 3.0 0.35 5.5 1.5 0.39 5.7 1.2 0.21

7
S. Begum, T. Das, G.R. Anupoju et al. Journal of Cleaner Production 289 (2021) 125775

Fig. 4. Impact of FW:CB mixing on volatile solids removal efficiency and digestate pH of ACoD of CB and FW over 70 d digestion time.

experimental methane yield (EMY) was essential to find the actual 3.5. A techno-economic analysis of solid-state ACoD of FW and CB
biodegradability of the feedstock. The methane (CH4) content in the at full-scale
present study was assumed to be 65%, which is a standard for FW
(Irena.org, 2019; Isola et al., 2018; Kuo and Dow, 2017; Metcalf and The profitable operation of an AD unit relies on low capital and
Eddy, 1991). The TMY of FW:CB in the ratio of 100:0 and 0:100 was operational costs. Most of the studies in literature investigated the
estimated considering the elemental composition of the substrates biomethane potential of FW and CB as lab-scale batch experiments
individually, while the TMY for mixture was estimated based on the that do not represent the output that could be achieved from a full-
number of g of VSadded (i.e. gVSadded of FW þ gVSadded of CB). A scale AD unit. In the present study, a pilot-scale CSTR system that
prediction of TMY was made based on the cumulative gVSadded could represent a full-scale AD unit was operated with FW and
from the mixture. The biodegradability (%) was estimated by the mixture of FW and CB in different mixing ratios for 70 d. Table 5
ratio of EMY to TMY, as shown in Eq. (2). presents the estimates of techno-economic analysis of a full-scale
AD unit based on the performance of the pilot-scale CSTR system.
EMY It is experimentally proven in the present study that the AD of
Biodegradability ð%Þ ¼ *100% (2)
TMY FW alone results in better output than the co-digestion of FW and
CB. The performance output from ACoD of FW and CB illustrated
The TMY, EMY and the biodegradability from the reactors are
that R2 with FW:CB of 80:20 outperformed compared to R3 (FW:CB
summarized in Table 4, which shows that the EMY in each reactor is
60:40) and R4 (FW:CB 50:50), and the same is expected at full-scale
lower than the TMY. An important finding of this study is that the
(Table 5). The HRT and power required for the operation of the AD
biodegradability of FW in reactor R1 was 76%, while in reactors R2,
unit was kept constant in all cases. The water and power require-
R3 and R4 it was 39%, 32%, and 22% respectively. This could be
ment per day were considered from the literature (Begum et al.,
attributed to the fact that lignocellulosic substrates like CB degrade
2017a, b; Kuruti et al., 2017). The operational cost of the AD unit,
slowly due to their recalcitrant characteristic and the presence of
which includes manpower, water and power, was estimated be-
lignin compared to sugar-containing substrates like FW. In addi-
tween 10,080 $/y in all the cases as the water, power and manpower
tion, the TS concentration of all the feed mixtures was higher than
required is assumed to be constant in all the cases (Table 5). It can
18%, so poor accessibility of microbes to organic matter due to
be illustrated from Table 5 that the biogas generation from R1 (i.e.
reduced mobility inside the digester could explain the lower
1000 kg/d of FW) is 133 m3/day whereas for R2, R3 and R4, it is
biodegradability of CB. Lower fluidity due to higher rheological
52.5 m3/d, 40 m3/d and 27 m3/d, respectively. The poor perfor-
properties of digester media in 40% CB (R3) and 50% CB (R4) in FW
mance of R3 and R4 interms of biogas generation can be attributed
resulted in lower biodegradability than in R1 and R2 (Miryahyaei
to the increase of CB ratio. Previous research on 1000 kg FW AD
et al., 2019). Therefore, the biodegradability of FW and CB is
plants demonstrated a maximum biogas generation between 80
lower than that of FW alone, because there will always be complex
and 150 m3/day (Kuruti et al., 2017) which corroborates with the
biochemical reactions produced by the co-digestion of the different
present study i.e. 133 m3/day (Table 5). No previous researchers
substrates.
8
S. Begum, T. Das, G.R. Anupoju et al. Journal of Cleaner Production 289 (2021) 125775

waste disposal rather than earning profits, the predicted techno-


economics are highly promising. The complete economic feasi-
bility of these plants can be assessed when the capital cost of these
plants is taken into consideration which is not accounted in this
study. This study suggests that the installation of decentralized AD
units of this kind at the source of waste generation would eliminate
the costs of transportation of waste to centralized landfill and
produce indirect economic benefits through reduction of environ-
mental contamination and associated harm to public health.

3.6. Further research and recommendations

This study concludes that there remains further scope for


research on solid-state ACoD of acetic acid rich FW and CB partic-
ularly with regard to pre-treatment. The present study found that
the biogas yield decreased at an HRT of 40 d with an OLR >4 g VS/
(L.d), even after the system attained stability. The biodegradability
of the ACoD of FW and CB mixture ranged from 22 to 39%, which
could be further enhanced by the introduction of physical, chemical
and/or biological pre-treatment steps, depending on the composi-
tion of waste and the operating conditions. In addition to the pre-
treatment, further studies on rheological properties of FW and CB
with high solids is necessary because the accessibility of organic
matter to the microbes can be influenced by the rheology of the
reactor contents. The cost of pre-treatment and the value of the
extra methane yield must be taken into account in the overall
economic feasibility analysis of an AD unit particularly for FW:CB
ratios of 60:40 and 50:50. The investment costs for pre-treatment
of recalcitrant substrates like CB are high due to the necessary
process engineering steps. Longer HRTs may improve plant effi-
ciency but requires more capital investment due to large reactor
volumes because the land located near waste generation sources is
inevitably expensive. Research into the detailed techno-economic
feasibility analysis of a full-scale AD plant with a pre-treatment
step and rheological properties has not been undertaken to date
Fig. 5. Overall mass balance in reactor inlets and outlets (TS: Total solid, VS: Volatile which is highly recommended.
solid, FS: fixed solids).
4. Conclusion
Table 4
Comparison of the theoretical and experimental methane yield of FW and CB. This study evaluated the performance of a solid-state ACoD of
acetic acid rich FW and CB in a pilot-scale auto-fed CSTR system at a
Reactor name TMY (L CH4/g EMY (L CH4/g Biodegradability
complete HRT of 40 d. ACoD of FW and CB in the ratios of 100:0,
VSadded) VSadded) (%)
80:20, 60:40 and 50:50 with TS concentration between 19% and
FW:CB (0:100) 0.45 e e 24%.
R1 (FW:CB 0.63 0.48 76%
100:0)
This study concluded that the solid-state ACoD of FW and CB
R2 (FW:CB 0.61 0.24 39% without any pre-treatment or addition of buffering chemicals is
80:20) feasible if the micro-feeding strategy and long HRT (for the long-
R3 (FW:CB 0.6 0.19 32% term exposure of CB to acetic acid-rich FW) are implemented.
60:40)
Micro-feeding strategy and frequent feeding pattern on an hourly
R4 (FW:CB 0.6 0.13 22%
50:50) basis assisted in overcoming the acidification phenomena of FW and
attaining stability even at OLR >4 gVS/(L.d). The lower biogas yield
for reactors with CB content more than 40% is due to poor biode-
gradability of CB due to lower interaction of acetic acid (CH3COOH)
have reported results for full-scale co-digestion of FW and CB, while
in FW with CB. The organic matter removal efficiencies (for 50% CB
this study has made an attempt to predict the performance at full-
addition to no CB in food waste) declined from 75 e 90% to 45e65%
scale.
as the HRT was shortened from 90 d to 40 d. The techno-economic
As can be seen from Table 5 that the difference a positive profits
feasibility analysis indicated the system is profitable even without
that can be achieved from R1 and R2, while for R3 and R4 there is no
any pre-treatment for the reactors with foodwaste only or food-
profit. The lower revenue and no net profit in the case of R3 and R4
waste with 20% CB. To make the digestion of more than 40% CB in
can be attributed to lower degradation of CB. Notwithstanding the
food waste profitable, a pre-treatment method need be envisaged. If
fact that the lower revenues can be attributed to the usage of biogas
the safe waste disposal is the primary concern rather than earning
for electricity generation, rather than its usage for heat applications
profits, the installation of full-scale plants of FW and CB can be an
such as cooking fuel. This is because the cost of power is relatively
ideal solution. Installation of AD units at sources of waste generation
lower than the cooking fuel. Nevertheless, if the installation of this
would also eliminate the cost that would otherwise be incurred for
kind of full-scale plants are considered as a solution for the safe
transportation of waste to a centralized landfill location.
9
S. Begum, T. Das, G.R. Anupoju et al. Journal of Cleaner Production 289 (2021) 125775

Table 5
A techno-economic analysis of ACoD of FW and CB at different mixing ratios in a micro-feeding CSTR.

R1 (FW:CB 100:0) R2 (FW:CB 80:20) R3 (FW:CB 60:40) R4 (FW:CB 50:50)

Technical details of the AD unit


Quantity of substrate (kg/d) 1000 800 kg FW and 14.4 kg CB 600 kg FW and 29 kg CB 500 kg FW and 36 kg CB
Organic matter per ton of substrate (kg VS/d) 180 164 139 126
Volume of digester (m3) 38.3 38.5 38.2 38.6
HRT (d) 40 40 40 40
Water requirement (L/d) 250 250 250 250
Power required for operation (kWh/d) 15 15 15 15
Biogas generation (m3/d) 133 52.5 40 27
Electricity generation from biogas (kWh/d) 266 105 80 54
Digestate generation (L/d) 892 732 570 495
Operational cost
Manpower required for (1 person for 1 h)/y @ $25/h ($/d) 25 25 25 25
Cost of water ($/d) 1.0 1.0 1.0 1.0
Cost of power required ($/d) 4.2 4.2 4.2 4.2
Total operating cost ($/d) 30 30 30 30
No. of working days 336 336 336 336
Total operating cost ($/y) 10,080 10,080 10,080 10,080
Revenue from the plant
Revenue from electricity generation ($/d) 75 29.4 22.4 15
Revenue from digestate sales ($/d) 9 7.3 6 5
Total revenue from the plant ($/d) 84 37 28.4 20
Total revenue from the AD unit ($/y) 28,224 12,432 9,542 6,720
Profit (revenue minus operating cost) ($/y) 18,144 2,352 () 538 () 3,360

1 kWh ¼ 28 cents (Canstarblue.com.au, 2020); Water ¼ $0.0036/L (Team Poly Water Tanks | Water Solutions for Li et al., 2019); 1 m3 of biogas ¼ 2 kWh (Aqper.com, 2020);
Digestate ¼ $0.01/L (Kuruti et al., 2017); $ ¼ AUD (Australian Dollar).

CRediT authorship contribution statement Begum, S., Golluri, K., Anupoju, G.R., Ahuja, S., Gandu, B., Kuruti, K., Maddala, R.K.,
Yerramsetti Venkata, S., 2016. Cooked and uncooked food waste: A viable
feedstock for generation of value added products through biorefinery approach.
Sameena Begum: Methodology, Formal analysis, Investigation, Chem. Eng. Res. Des. 107, 43e51. https://doi.org/10.1016/j.cherd.2015.10.0322.
Writing - original draft, Visualization. Tanmoy Das: Investigation. Begum, S., Anupoju, G.R., Sridhar, S., Bhargava, S.K., Jegatheesan, V., Eshtiaghi, N.,
Gangagni Rao Anupoju: Supervision. Nicky Eshtiaghi: Supervi- 2018. Evaluation of single and two stage anaerobic digestion of landfill leachate:
Effect of pH and initial organic loading rate on volatile fatty acid (VFA) and
sion, Writing - review & editing, Conceptualization, Project biogas production. Bioresour. Technol. 251, 364e373. https://doi.org/10.1016/j.
administration. biortech.2017.12.069.
Begum, S., Juntupally, S., Anupoju, G.R., Eshtiaghi, N., 2020. Comparison of meso-
philic and thermophilic methane production potential of acids rich and high-
Declaration of competing interest strength landfill leachate at different initial organic loadings and food to
inoculum ratios. Sci. Total Environ. 715, 136658.
Capson-Tojo, G., Trably, E., Rouez, M., Crest, M., Steyer, J.P., Delgenes, J.P., Escudie
, R.,
The authors declare that they have no known competing 2017. Dry anaerobic digestion of food waste and cardboard at different substrate
financial interests or personal relationships that could have loads, solid contents and co-digestion proportions. Bioresour. Technol. 233,
appeared to influence the work reported in this paper. 166e175. https://doi.org/10.1016/j.biortech.2017.02.126.
Cavinato, C., Bolzonella, D., Pavan, P., Fatone, F., Cecchi, F., 2013. Mesophilic and
thermophilic anaerobic co-digestion of waste activated sludge and source sor-
Acknowledgement ted biowaste in pilot- and full-scale reactors. Renew. Energy 55, 260e265.
https://doi.org/10.1016/j.renene.2012.12.044.
De Clercq, D., Wen, Z., Song, Q., 2019. Innovation hotspots in food waste treatment,
The authors acknowledge RMIT University, Melbourne, Australia biogas, and anaerobic digestion technology: a natural language processing
and the Council of Scientific and Industrial Research e Indian approach. Sci. Total Environ. 673, 402e413. https://doi.org/10.1016/
Institute of Chemical Technology for financial support and a j.scitotenv.2019.04.051.
Department of the Environment and Energy, 2019, 12.6.19. http://www.
scholarship for Sameena Begum. The authors also acknowledge the environment.gov.au/protection/waste-resource-recovery/national-waste-
Aurora sewage treatment plant (Preston, Australia) for providing reports/national-waste-report-2018.
Dupont, A.L., Te treault, J., 2000. Cellulose degradation in an acetic acid environ-
food waste and inoculum, and Mr Shoutian Li for his help in reactor
ment. Stud. Conserv. 45, 201e210. https://doi.org/10.1179/sic.2000.45.3.201.
operation. Fisgativa, H., Tremier, A., Dabert, P., 2016. Characterizing the variability of food
waste quality: a need for efficient valorisation through anaerobic digestion.
Waste Manag. 50, 264e274. https://doi.org/10.1016/j.wasman.2016.01.041.
References Garcia-Garcia, G., Stone, J., Rahimifard, S., 2019. Opportunities for waste valorisation
in the food industry e a case study with four UK food manufacturers. J. Clean.
Achinas, S., Euverink, G.J.W., 2016. Theoretical analysis of biogas potential predic- Prod. 211, 1339e1356. https://doi.org/10.1016/j.jclepro.2018.11.269.
tion from agricultural waste. Resour. Technol. 2, 143e147. https://doi.org/ General Kinematics, 2019, 12.6.19. https://www.generalkinematics.com/blog/the-
10.1016/j.reffit.2016.08.001. life-cycle-of-cardboard/.
Asato, C.M., Gonzalez-Estrella, J., Skillings, D.S., Castano ~, A.V., Stone, J.J., Gibson, L.T., Watt, C.M., 2010. Acetic and formic acids emitted from wood samples
Gilcrease, P.C., 2019. Anaerobic digestion of synthetic food waste-cardboard and their effect on selected materials in museum environments. Corrosion Sci.
mixtures in a semi-continuous two-stage system. Sustain. Energy Fuels 3, 52, 172e178. https://doi.org/10.1016/j.corsci.2009.08.054.
3582e3593. https://doi.org/10.1039/c9se00667b. Guilford, N.G.H., Lee, H.P., Kanger, K., Meyer, T., Edwards, E.A., 2019. Solid-state
Begum, S., Ahuja, S., Anupoju, G.R., Kuruti, K., Juntupally, S., Gandu, B., Ahuja, D.K., anaerobic digestion of mixed organic waste: the synergistic effect of food waste
2017a. Process intensification with inline pre and post processing mechanism addition on the destruction of paper and cardboard. Environ. Sci. Technol. 53,
for valorization of poultry litter through high rate biomethanation technology: 12677e12687. https://doi.org/10.1021/acs.est.9b04644.
a full scale experience. Renew. Energy 114, 428e436. https://doi.org/10.1016/ Ishangulyyev, R., Kim, S., Lee, S.H., 2019. Understanding food loss and waste-why
j.renene.2017.07.049. are we losing and wasting food? Foods 8, 297. https://doi.org/10.3390/
Begum, S., Ahuja, S., Rao, A., Kiran, G., Gandu, B., Kuruti, K., Swamy, Y., Ahuja, D., foods8080297.
2017b. Significance of decentralized biomethanation systems in the framework Isola, C., Sieverding, H.L., Asato, C.M., Gonzalez-Estrella, J., Litzen, D., Gilcrease, P.C.,
of municipal solid waste treatment in India. Curr. Biochem. Eng. 4, 2e8. https:// Stone, J.J., 2018. Life cycle assessment of portable two-stage anaerobic digestion
doi.org/10.2174/2212711902999151001134836.

10
S. Begum, T. Das, G.R. Anupoju et al. Journal of Cleaner Production 289 (2021) 125775

of mixed food waste and cardboard. Resour. Conserv. Recycl. 139, 114e121. j.biortech.2004.06.025.
Kuo, J., Dow, J., 2017. Biogas production from anaerobic digestion of food waste and Panigrahi, S., Sharma, H.B., Dubey, B.K., 2020. Anaerobic co-digestion of food waste
relevant air quality implications. J. Air Waste Manag. Assoc. 67, 1000e1011. with pretreated yard waste: a comparative study of methane production, ki-
Kuruti, K., Begum, S., Ahuja, S., Anupoju, G.R., Juntupally, S., Gandu, B., Ahuja, D.K., netic modeling and energy balance. J. Clean. Prod. 243, 118480.
2017. Exploitation of rapid acidification phenomena of food waste in reducing Ren, Y., Yu, M., Wu, C., Wang, Q., Gao, M., Huang, Q., Liu, Y., 2018. A comprehensive
the hydraulic retention time (HRT) of high rate anaerobic digester without review on food waste anaerobic digestion: research updates and tendencies.
conceding on biogas yield. Bioresour. Technol. 226, 65e72. https://doi.org/ Bioresour. Technol. 247, 1069e1076. https://doi.org/10.1016/
10.1016/j.biortech.2016.12.005. j.biortech.2017.09.109.
Kuruti, K., Gangagni Rao, A., Gandu, B., Kiran, G., Mohammad, S., Sailaja, S., Svensson, K., Paruch, L., Gaby, J.C., Linjordet, R., 2018. Feeding frequency influences
Swamy, Y.V., 2015. Generation of bioethanol and VFA through anaerobic process performance and microbial community composition in anaerobic di-
acidogenic fermentation route with press mud obtained from sugar mill as a gesters treating steam exploded food waste. Bioresour. Technol. 269, 276e284.
feedstock. Bioresour. Technol. 192, 646e653. https://doi.org/10.1016/ Wang, P., Wang, H., Qiu, Y., Ren, L., Jiang, B., 2018. Microbial characteristics in
j.biortech.2015.05.104. anaerobic digestion process of food waste for methane productioneA review.
Li, D., Song, L., Fang, H., Teng, Y., Cui, H., Li, Y.-Y., Liu, R., Niu, Q., 2019. Optimization of Bioresour. Technol. 248, 29e36. https://doi.org/10.1016/j.biortech.2017.06.152.
biomethane production in mono-cardboard digestion: key parameters influ- WPCF, A., 1998. APHA. Stand. Methods Exam. Water Wastewater, twentieth ed. Am.
ence, batch test kinetic evaluation, and DOM indicators variation. Energy Fuel. Public Heal. Assoc. (APHA), Washingt. DC.
33, 4340e4351. Yang, Y., Bao, W., Xie, G.H., 2019. Estimate of restaurant food waste and its biogas
Measuring small-scale biogas capacity and production, 2016. Int. Renew. Energy production potential in China. J. Clean. Prod. 211, 309e320. https://doi.org/
Agency, 12.6.19. https://www.irena.org/-/media/Files/IRENA/Agency/ 10.1016/j.jclepro.2018.11.160.
Publication/2016/IRENA_Statistics_Measuring_small-scale_biogas_2016.pdf. Zahan, Z., Othman, M.Z., Muster, T.H., 2018. Anaerobic digestion/co-digestion kinetic
Meng, Y., Shen, F., Yuan, H., Zou, D., Liu, Y., Zhu, B., Chufo, A., Jaffar, M., Li, X., 2014. potentials of different agro-industrial wastes: a comparative batch study for C/
Start-up and operation strategies on the liquefied food waste anaerobic N optimisation. Waste Manag. 71, 663e674. https://doi.org/10.1016/
digestion and a full-scale case application. Bioproc. Biosyst. Eng. 37, 2333e2341. j.wasman.2017.08.014.
https://doi.org/10.1007/s00449-014-1211-8. Zeynali, R., Khojastehpour, M., Ebrahimi-Nik, M., 2017. Effect of ultrasonic pre-
Metacalf, E., Eddy, H., 1991. Wastewater Engineering-Treatment. Reuse, Disposal. treatment on biogas yield and specific energy in anaerobic digestion of fruit
Miryahyaei, S., Olinga, K., Abdul Muthalib, F.A., Das, T., Ab Aziz, M.S., Othman, M., and vegetable wholesale market wastes. Sustain. Environ. Res. 27, 259e264.
Baudez, J.C., Batstone, D., Eshtiaghi, N., 2019. Impact of rheological properties of https://doi.org/10.1016/j.serj.2017.07.001.
substrate on anaerobic digestion and digestate dewaterability: new insights Zhang, C., Xiao, G., Peng, L., Su, H., Tan, T., 2013. The anaerobic co-digestion of food
through rheological and physico-chemical interaction. Water Res. 150, 56e67. waste and cattle manure. Bioresour. Technol. 129, 170e176. https://doi.org/
https://doi.org/10.1016/j.watres.2018.11.049. 10.1016/j.biortech.2012.10.138.
Mosier, N., Wyman, C., Dale, B., Elander, R., Lee, Y.Y., Holtzapple, M., Ladisch, M., Ziels, R.M., Beck, D.A.C., Stensel, H.D., 2017. Long-chain fatty acid feeding frequency
2005. Features of promising technologies for pretreatment of lignocellulosic in anaerobic codigestion impacts syntrophic community structure and bio-
biomass. Bioresour. Technol. 96, 673e686. https://doi.org/10.1016/ kinetics. Water Res. 117, 218e229.

11

You might also like