J of Chemical Tech Biotech - 2022 - Moreno Andrade - Biohydrogen Production From Food Waste and Waste Activated Sludge in

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Research Article

Received: 23 June 2022 Revised: 3 August 2022 Accepted article published: 12 September 2022 Published online in Wiley Online Library: 30 September 2022

(wileyonlinelibrary.com) DOI 10.1002/jctb.7238

Biohydrogen production from food waste and


waste activated sludge in codigestion:
influence of organic loading rate and changes
in microbial community
Iván Moreno-Andrade,* María José Berrocal-Bravo and
Idania Valdez-Vazquez

Abstract
BACKGROUND: Food waste (FW) and waste activated sludge (WAS) are complementary substrates that improve H2 production
by dark fermentation. However, there is little experience in the process performance in the long-term operation of reactors
co-processing FW–WAS. This study aims to determine the optimal FW–WAS ratio in batch H2 production and then feed this
FW–WAS ratio to an anaerobic sequencing batch reactor (SBR) to evaluate the influence of three organic loading rates (OLR)
– 15, 22 and 45 g volatile solids (VS) L−1 d−1 – on H2 performance and microbial composition.
RESULTS: Batch tests showed that the FW–WAS ratio of 90–10 increased 22% of the H2 production compared to the individual
FW. In the SBR operation, the OLR significantly influenced H2 performance, reaching the highest productivity of 733 ± 282
mL H2 L−1 d−1 at an OLR of 22 g VS L−1 d−1. The cooperative lactic acid cross-feeding between Olsenella and Megasphaera
resulted in the highest H2 productivity at an OLR of 22 g VS L−1 d−1. OLR promoted the prevalence of different taxa of hydro-
lytic bacteria, where Enterococcus and Veillonella were positively correlated, with the highest substrate hydrolysis of 50% at an
OLR of 15 g VS L−1 d−1. Prevotella, a genus with a wide spectrum of hydrolytic enzymes, was predominant in the SBR operation
regardless of the OLR.
CONCLUSION: The SBR operation at OLR of 22 g VS L−1 d−1 improved H2 production by applying a 90–10 FW–WAS ratio. The
change in OLR promoted different taxa of hydrolytic species in the SBR; however, Prevotella spp. prevailed at the three OLRs.
© 2022 The Authors. Journal of Chemical Technology and Biotechnology published by John Wiley & Sons Ltd on behalf of Society
of Chemical Industry (SCI).

Keywords: hydrogen; Megasphaera; organic loading rate; Prevotella; sequencing batch reactor

INTRODUCTION wastes helps decrease the carbohydrate/protein ratio, provides


Globally, there is a crisis related to the generation and disposal of alkalinity and micronutrients, and dilutes toxic substances, result-
food waste (FW).1 In 2018, FW production reached about 1.3 bil- ing in the improvement of the H2 production.6-8 Other advan-
lion metric tons worldwide, of which 95% ended at landfill sites tages of the codigestion of wastes are the nutritional balance of
generating greenhouse gases and contributing to global warm- the substrates to obtain an optimal C/N ratio, the pH adjustment
ing. However, FW generation constitutes one opportunity to pro- to those values required for the bioprocess selected and the dilu-
duce hydrogen (H2), a net-zero-emission biofuel. H2 production tion of toxic molecules minimizing the growth inhibition; all these
by biological methods such as dark fermentation is more sustain- factors improve the yield of biogas or fermentation products com-
able than conventional routes due to its lower requirements of pared with mono-digestion.9,10
energy and chemicals.2 Because the composition of FW varies
depending on its origin (e.g., carbohydrate content, C/N ratio
* Correspondence to: Iván Moreno-Andrade, Laboratory for Research on
and others), the H2 yield ranges from 50 to 150 L H2 kg−1 of vola- Advanced Processes for Water Treatment, Unidad Academica Juriquilla,
tile solids (VS).3-5 One inherent challenge for producing H2 from Instituto de Ingeniería, Universidad Nacional Autónoma de México, Blvd.
FW is its high carbohydrate/protein ratio, which results first in Juriquilla 3001, 76230 Querétaro, Mexico. E-mail: imorenoa@iingen.
the accumulation of volatile fatty acids (VFA) that decrease the unam.mx
pH and finally inhibit the fermentative activity of H2-producing Laboratory for Research on Advanced Processes for Water Treatment, Unidad
bacteria.6 To solve this operational failure without consuming Academica Juriquilla, Instituto de Ingeniería, Universidad Nacional Autónoma
large amounts of alkalis, the co-fermentation of FW with other de México, Juriquilla, Mexico
230

© 2022 The Authors. Journal of Chemical Technology and Biotechnology published by John Wiley & Sons Ltd on behalf of Society of Chemical
Industry (SCI).
This is an open access article under the terms of the Creative Commons Attribution-NonCommercial License, which permits use, distribution and
reproduction in any medium, provided the original work is properly cited and is not used for commercial purposes.
10974660, 2023, 1, Downloaded from https://analyticalsciencejournals.onlinelibrary.wiley.com/doi/10.1002/jctb.7238 by Birla Institute of Technology & Science, Wiley Online Library on [09/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Biohydrogen from codigestion of food waste and waste activated sludge www.soci.org

Wastewater treatment plants generate large amounts of waste The inoculum used was an anaerobic sludge that was provided
activated sludge (WAS), which also has the potential to produce by a brewery. This sludge was pretreated by drying at 105 °C for
H2 by dark fermentation. However, due to its low carbohydrate/ 24 h, ground and sifted to obtain a particle size <65 mm, select-
protein ratio (<0.1) and high content of non-biodegradable mate- ing the H2-producing microorganisms according to several
rials, the resulting H2 yields are low: below 30 L kg−1 VS.6-8,11 authors.16-19 The characterization of the pretreated inoculum
Despite this, WAS possesses interesting properties for its co- showed the next values in g/ginoculum:total solids (TS)
processing with FW, such as high alkalinity and trace metals act- 0.98 ± 0.01, VS 0.69 ± 0.02, COD 2.13 ± 0.3, carbohydrates 0.07
ing as micronutrients.12 The co-processing of FW with WAS has ± 0.01 and proteins 0.14 ± 0.03.
been demonstrated to improve the H2 production rate compared
to the processing of the individual wastes.11-14 Codigestion of Hydrogen potential production test
WAS with other organic wastes has demonstrated to improve Different combinations of FW–WAS (% VS) were evaluated to test
sludge dewaterability, increase buffer capacity, accelerate initial their effects on hydrogen potential production (HPP) at
biogas production time and balance C/N ratio.10 10 g VS L−1: 90–10, 80–20, 70–30, 60–40, 50–50, 30–70 (ratios
Kim et al.14 determined an optimal FW–WAS ratio of 87–13%, are expressed on a VS basis). Treatments with individual wastes
which increased the H2 production compared to the individual of FW and WAS (100–0 and 0–100) were used for benchmarking
processing of FW. Later, another study confirmed that a slight later. The HPP test was carried out by using an Automatic Meth-
addition of WAS (10:1 on a chemical oxygen demand basis) signif- ane Potential Test System equipment (AMPTS II, Bioprocess Con-
icantly enhanced the H2 fermentation performance.11 On the trol, Sweden). The standardized protocol for determining
other hand, Liu et al.13 reported that the co-processing of FW with biohydrogen potential reported by Carrillo-Reyes et al.19 was used
WAS did not improve the H2 yield. These contradictory results as a guideline. Batch reactors with a total volume of 600 mL and a
may be related to the feedstocks’ initial characteristics such as working volume of 460 mL with 140 mL of headspace were used.
chemical oxygen demand (COD), volatile solids (VS), carbohydrate Nutrients and alkalinity buffer were added according to Mizuno
content and C/N ratio, which vary depending on their origin, as et al.,20 and deoxygenated water was added to fill the total vol-
aforementioned. The C/N ratio is an important parameter since, ume. The initial pH was adjusted to 7.0 using a solution of
generally, FW has a higher C/N ratio than WAS, and co- 0.1 mol L−1 NaOH according to Ramos et al.,18 and each batch test
fermentation can be adjusted to an optimal C/N value (15–30); was inoculated with 0.7 g of the pretreated inoculum. Intermittent
the production of soluble proteins, carbohydrates and short-chain mixing at 120 rpm (1 min for every 3 min during the experiment)
fatty acid is enhanced consequently, and the dewaterability of was applied. Incubation was at a constant temperature of 37°C. All
anaerobic digestate was improved.10 treatments were run in triplicate. Endogenous H2 production by
Previous studies have demonstrated the convenience of the the inoculum alone was subtracted from all treatments in the
FW–WAS co-processing for improving H2 production; however, HPP calculation.
all the studies were limited to batch tests.11-14 In continuous
reactors, there could be an inoculum adaptation, substrate Reactor experimental set-up
depletion or accumulation of putative toxic compounds, among The FW–WAS ratio for which the highest HPP was reached was
other phenomena present in batch fermentations. Therefore, applied to H2 production in a sequencing batch reactor (SBR).
the potential of H2 production previously reported during the The SBR was constructed as an acrylic cylinder with a reaction vol-
co-processing of FW–WAS could be improved in continuous ume of 2 L, a working volume of 1.5 L and headspace of 0.5 L, and
reactors. The organic loading rate (OLR) is a key parameter for an exchange volume of 50%. The reactor was inoculated with 20 g
H2 production from FW in continuous bioreactors.15 OLR influ- anaerobic sludge pretreated at 105 °C for 24 h, inhibiting metha-
ences the hydrolysis of particulate substrate, removal, and con- nogenic microorganisms. The inoculum was added once at the
version of carbohydrates, consequently leading to the efficient initial operation of the reactor. Initial activation of the inoculum
production of H2 and VFA. As a result, this study aims to deter- was obtained in ten operation cycles (reaction time 12 h) using
mine the optimal FW–WAS ratio for the improvement of H2 pro- 15 g L−1 glucose and mineral medium: 41.6 g NH4Cl, 4 g
duction in batch tests and then to determine the effects of the K2HPO4, 2 g MgCl2.6H2O, 1.6 g FeSO4·.7H2O, 40 mg CoCl2.6H2O,
OLR on H2 productivity and microbial composition in a discon- 40 mg MnCl2.4H2O, 40 mg KI, 8 mg NiCl2.6H2O, 8 mg ZnCl2.
tinuous reactor operated for 110 days with the optimal After the inoculum activation, the experiments applying a fixed
FW–WAS ratio. feedstock at 15 g VS L−1 (FW–WAS ratio of 90–10) were carried
out. Reaction time was 12, 8 and 4 h, related to the OLR tested
(15, 22 and 45 g VS L d−1, respectively). The fill, settle and draw
EXPERIMENTAL phases were 5, 60 and 5 min, respectively. The reactor had deflec-
Feedstocks and inoculum tors in the walls assuring a complete mixing. A temperature con-
FW was obtained from a university cafeteria daily for 2 weeks; troller based on a recycling water pump was used to maintain a
only the fermentable matter was preserved at −20°C (bones constant temperature of 37 ± 1°C inside the reactor (Thermo Sci-
and inert materials such as plastic or paper were discarded). The entific, Waltham, MA, USA). pH was maintained at 5.5 by a control-
daily samples were mixed, homogenized and mashed with a ler (pH 140 series, EUTECH Instruments, USA) using a combined
blender (1 HP grinder JR MJ22, Torrey Mexico), obtaining a parti- electrode (BNC, Sensorex), supplying NaOH 1 mol L−1. The stirrer
cle size <0.5 mm. This mixture was stored for further experiments and pumps for filling and drawing were timed and controlled by
at −20°C to avoid its auto-fermentation. WAS was obtained from a software programmed in LabVIEW (National Instruments, USA)
conventional municipal wastewater treatment plant after dewa- using a USB6008 data acquisition card (National Instruments,
tering with a press and stored at 4°C; no further pretreatment USA) connected to a personal computer.
was applied. The characteristics of FW and WAS are listed in The biogas flow rate was measured online using a flowmeter
231

Table 1. (ADM 1000, Agilent, Santa Clara, USA). The accumulated biogas

J Chem Technol Biotechnol 2023; 98: 230–237 © 2022 The Authors. wileyonlinelibrary.com/jctb
Journal of Chemical Technology and Biotechnology published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry (SCI).
10974660, 2023, 1, Downloaded from https://analyticalsciencejournals.onlinelibrary.wiley.com/doi/10.1002/jctb.7238 by Birla Institute of Technology & Science, Wiley Online Library on [09/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.soci.org I Moreno-Andrade, MJ Berrocal-Bravo, I Valdez-Vazquez

volume was obtained with respect to time by numerical integra- RESULTS AND DISCUSSION
tion implemented in LabVIEW. Biogas samples were taken daily Feedstocks (FW and WAS) characterization
to evaluate the percentage of H2. Table 1 shows the characterization of FW and WAS. The moisture
The cumulative H2 production (H) of the experimental data was content of both residues was higher than 60%, making them suit-
fitted to a modified Gompertz equation (Eqn 1): able for the wet anaerobic digestion process. FW was composed
   of 96% easy-to-degrade material with low alkalinity. These two
R max characteristics most likely result in the quick accumulation of
Hðt Þ=H max exp −exp ð⊗−tÞ+1 ð1Þ
H max VFA, acidification and, finally, inhibition of dark fermentation. On
the other hand, WAS had alkalinity threefold higher than
where H(t) (mL H2 L−1) is the total amount of hydrogen produced FW. The co-processing of FW with WAS indeed increases the sys-
during the experiment (h), Hmax (mL) is the maximum amount of tem's buffer capacity and reduces the process inhibition result of
hydrogen produced, Rmax (mL H2 h−1) is the maximum hydrogen the VFA accumulation. Also, WAS may be dosed during FW fer-
rate and ⊗ (h) is the lag time before the exponential hydrogen mentation for pH control between 5.5 and 7.0, optimizing H2
production. production,18-19 which would avoid the requirements for costly
alkalis. Trace metals were significantly higher in WAS. Some of
Analytical analysis these metals are necessary for metabolism, cellular growth, and
Biogas composition (CH4, CO2 and H2) was analyzed using an SRI activation and function of enzymes and coenzymes in dark fer-
8610C gas chromatograph (SRI Instruments, USA) equipped with mentation for hydrogen production.25 In particular, Fe and Ni
a thermal conductivity detector and a 30 m long (0.53 mm ID) are essential for the biosynthesis of hydrogenases, which are
Carboxen-1010 PLOT column. The operating conditions were set responsible for the reduction of protons, resulting in the forma-
as follows: the carrier gas was nitrogen at a flow rate of tion of H2.26 Also, WAS has a higher concentration of Ca and Al,
4.5 mL min−1; the temperature of the injector was 200 °C; the col- which are metals that help growth and cell retention, so they
umn temperature was 100 °C; and the temperature of the detec- are necessary to avoid the washout of cells from continuous
tor was fixed at 230 °C. bioreactors.27
Contents of VFA (acetic, butyric, propionic and valeric acids) and
solvents (ethanol and butanol) were quantified using a gas chro-
Hydrogen potential production test
matograph (Varian model 3300, CA, USA) coupled to a flame ion-
Figure 1 shows (A) lag time (⊗), (B) Rmax (maximum hydrogen pro-
ization detector and equipped with a 15 m long (0.53 mm id)
duction rate) and (C) Hmax (total amount of hydrogen produced)
Zebron ZB-FFAP column. The injector and detector temperatures
for the individual wastes and the FW–WAS ratios under evalua-
were maintained at 190 and 210 °C, respectively. Column temper-
tion. In general terms, the co-processing of FW–WAS improved
ature was maintained at 45 °C for 1.5 min, after which it was
the fermentation performance. The first beneficial effect of WAS
increased to 135 °C at a rate of 8 °C min−1. The carrier gas was
was the immediate reduction of the lag time (⊗) from 17 h in
nitrogen at 9.5 mL min−1. The lactate content was determined
the individual FW to 5 h for the FW–WAS ratios between 70–30
by ion chromatography using a Dionex ICS-1500 system with an
REIC IonPac AS23 250 × 4 mm column. The eluent, 0.8 mmol L−1
NaHCO3, was mixed with 4.5 mmol L−1 Na2CO3 using an isocratic
flow of 3 mL min−1 at 30 °C.
The COD was measured spectrophotometrically using Hach Table 1. Physicochemical characterization of the feedstocks
methods according to the manufacturer's instructions (HACH,
Parameter FW WAS
CO, USA). TS and VS were determined according to standard
methods.21 Total carbohydrates were measured by the phenol- TS (g/g) 0.37 0.16
sulfuric acid method.22 All the results were analyzed by an ANOVA VS (g/g) 0.35 0.09
with a post-hoc Tukey test with Rstudio (Desktop 2022.07.2+576) VS/TS 95.6 56.2
to compare them among treatments. Alkalinity (mg CaCO3 L−1) 160 535
CODtotal (mg g−1 TS) 1212 1200
Microbial community analysis CODsoluble (mg g−1 TS) 356 168
During the SBR operation, cell samples were collected to char- Carbohydratestotal (mg g−1 TS) 198 42
acterize the microbial composition and preserved at −4°C Carbohydratessoluble (mg g−1 TS) 152 5
pending analysis. Genomic DNA was extracted from cell sam- Proteinstotal (mg g−1 TS) 236 119
ples using the PowerSoil DNA isolation kit (MO BIO Laborato- Proteinssoluble (mg g−1 TS) 57.3 —
ries, Carlsbad, CA, USA) according to the manufacturer's Density (kg L−1) 1.04 0.91
instructions. The DNA concentration was quantified by spec- pH 3.9 8.4
trophotometry using a NANODrop 2000c (Thermo Scientific, Calcium (mg L−1) 7.3 197
USA). The extracted DNA was submitted to RTL Genomics Aluminum (mg L−1) 1.1 31
(Lubbock, TX, USA) for Illumina MiSeq sequencing using the Iron (mg L−1) 0.5 56
universal primers 515F (5´GTGCCAGCMGCCGCGGTAA) and Zinc (mg L−1) 1.4 1.7
806R (5GGACTACHVGGGTWTCTAAT) of the 16S rDNA gene. Cupper (mg L−1) 0.01 1.0
The pipeline procedure for the sequence analysis and further Manganese (mg L−1) 0.2 0.7
analysis of the identified operational taxonomic units have Nickel (mg L−1) 0.01 0.5
been explained previously.23 Pearson's correlations between Potassium (mg L−1) 27 14
the bacterial populations and process parameters were carried Sodium(mg L−1) 127 78
232

out with PAST software.24

wileyonlinelibrary.com/jctb © 2022 The Authors. J Chem Technol Biotechnol 2023; 98: 230–237
Journal of Chemical Technology and Biotechnology published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry (SCI).
10974660, 2023, 1, Downloaded from https://analyticalsciencejournals.onlinelibrary.wiley.com/doi/10.1002/jctb.7238 by Birla Institute of Technology & Science, Wiley Online Library on [09/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Biohydrogen from codigestion of food waste and waste activated sludge www.soci.org

Figure 1. Kinetic parameters and H2 production performance for each of the FW–WAS ratios in batch test: (A) lag time; (B) Rmax; (C) Hmax; (D) metabolite
production at different FW–WAS ratios. The dotted line represents the adjustment to a fourth-order polynomial equation.

and 30–70 (P < 0.001). Rmax had contrasting values for the individ- Hydrogen production from FW–WAS codigestion in
ual wastes, almost ten times higher for FW than WAS, with a the SBR
decreasing trend as WAS increased in the FW–WAS ratios. Hmax The inoculum was activated with glucose according to Castillo-
showed a similar decrease as WAS increased in the FW–WAS Hernández et al.,17 and the FW–WAS ratio of 90–10 was then fed
ratios; nevertheless, the 90–10 ratio showed a higher Hmax than to the SBR. The reactor was operated for 110 cycles at different
the individual wastes and the other FW–WAS ratios (P < 0.01). OLRs: 15 g VS L−1 d−1 (cycles 10–40), 22 g VS L−1 d−1 (cycles
The individual WAS had a very low Hmax, a result of its low avail- 50–80) and 45 g VS L−1 d−1 (cycles 80–110). Figure 2 shows the
ability of carbohydrates compared to FW. VFA production was variations in compositions of H2, CH4 and CO2 during the SBR
notably higher in the 90–10 ratio compared to the other treat- operation. The H2 percentage averaged 46 ± 15% during the
ments. It is important to highlight that WAS addition did not glucose activation, decreasing with the substrate change to
change the composition of the main produced VFA (Fig. 1D), FW–WAS. The change in OLR influenced the H2 percentages; the
which were acetate and butyrate, both related to H2 production.26 OLR of 22 g VS L−1 d−1 resulted in the highest H2 percentage of
Nonetheless, WAS addition reduced the amount of ethanol, which 33 ± 8% (Fig. 2), while the increase in OLR resulted in the lowest
possibly was produced by the activation of solventogenesis in the H2 percentage of 14 ± 8% (P > 0.05).
H2-producing bacteria, resulting from VFA accumulation in ratios
with higher content of FW.
These results agreed with those obtained by Kim et al.,14
who proposed an optimal codigestion FW–sewage sludge Table 2. Volatile solids and carbohydrates removal in H2 potential
ratio of 87–13% for improvement of the H2 production in production test
comparison with the individual FW. A small dose of WAS
Mixture ratio VS Carbohydrate
improved the H2 production, most likely due to its contribu-
FW–WAS (%) removal (%) removal (%)
tion of alkalinity, which stabilized the pH of the system, and
micronutrients like Ni and Fe, which are important for the 100–0 53 ± 2.1 93 ± 0.5
biosynthesis of enzymes such as hydrogenases, responsible 90–10 51 ± 1.6 95 ± 1.2
for H2 production.14,28,29 80–20 49 ± 2.2 91 ± 2.0
Table 2 shows the removal of VS and carbohydrates. The effi- 70–30 42 ± 4.3 88 ± 4.0
ciency of VS removal decreased as WAS increased in the ratios. 60–40 44 ± 3.8 95 ± 3.7
VS removal was in the same order as that previously reported 50–50 36 ± 5.1 89 ± 1.5
for similar FW–WAS ratios.11-13 Carbohydrate removal averaged 30–70 23 ± 7.2 90 ± 3.5
91.6% in all the treatments, independent of the FW–WAS 0–100 12 ± 5.3 92 ± 2.3
233

ratio (P < 0.05).

J Chem Technol Biotechnol 2023; 98: 230–237 © 2022 The Authors. wileyonlinelibrary.com/jctb
Journal of Chemical Technology and Biotechnology published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry (SCI).
10974660, 2023, 1, Downloaded from https://analyticalsciencejournals.onlinelibrary.wiley.com/doi/10.1002/jctb.7238 by Birla Institute of Technology & Science, Wiley Online Library on [09/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.soci.org I Moreno-Andrade, MJ Berrocal-Bravo, I Valdez-Vazquez

Figure 2. Biogas composition during SBR operation at different OLRs.

Figure 3(A) shows the cumulative H2 production fitting the


modified Gompertz model (dotted line) for each SBR cycle, while
Table 3 shows the kinetic parameters and removal obtained at dif-
ferent OLRs. The results showed that the maximal value of Hmax
and Rmax (235 ± 27 mL H2 Lreactor−1 and 62 ± 15
mL H2 Lreactor−1 h−1, respectively) were obtained when an OLR
of 22 g VS L−1 d−1 was applied. The lag time did not show differ-
ences independently of the OLR applied. These results contrast
with results obtained in a batch test of FW–WAS codigestion,
where the lag time ranged from 3 to 18.7 h.11,13 The reduction
of lag time in the long-term operation of discontinuous reactors
is related to the microbial acclimatization of the microorganisms
and the presence of remnant substrate from previous cycles – in
our case, an exchange volume of 50%. The lag time results agree
with those obtained in SBR systems producing H2 from FW as
mono-substrate.30,31
The H2 productivities mirror the H2 percentages where the OLR
of 22 g VS L−1 d−1 resulted in the highest productivity of 733
± 282 mL H2 Lreactor−1 d−1, and the increase in OLR resulted in a
decrease in productivity to 383 mL H2 Lreactor−1 d−1 (Table 3).
From cycle 29, CH4 began to be detected in the biogas (5 ±
2.6%). Methane production is related to the presence of hydroge-
notrophic methanogens that consume H2 and CO229 and are
entered into the SBR with the non-sterile feedstocks.17,32
The maximum hydrolysis was obtained with an OLR of
15 g VS L−1 d−1 (50%) (Table 3). This agrees with previous results
using only FW,15 inferring that WAS does not affect the system's
hydrolysis percentage. Hydrolysis can be increased if the HRT
and solids retention time (SRT) are increased. However, the
increase in HRT and SRT increases the abundance of methano-
gens (growth rates are between 0.0167 and 0.02 h−1), reducing
H2 production.15,20 Efficient hydrolysis contributes to VS removal
and the transformation of particulate material into soluble
(e.g., COD), increasing its availability to microorganisms. Total car-
bohydrates removal showed no significant differences comparing
the three OLR tested in our study (>80%).
The OLR influences Hmax, Rmax and YH2 (Fig. 3B), demonstrating
the process's best performance at an OLR of 22 g VS L−1 d−1,
where the H2 production and YH2 were higher than reported in
a fermentative continuously stirred tank reactors (CSTR) using a
FW–WAS rate of 1:5 with a hydraulic retention time (HRT) of
3 days, reporting 8.6 ± 4.8 mL H2 g−1 VS.33 The tendency (adjust-
Figure 3. (A) Kinetics of the H2 production at different OLRs. The dotted line
represents the Gompertz model adjustment. (B) Effect of OLR on Rmax, Hmax
ment of the results to a polynomial model) of Fig. 3(B) shows that
and YH2. The dotted line represents the adjustment to a second-order polyno- Hmax, Rmax and YH2 can be optimized, possibly at an OLR in a range
234

mial equation. (C) Metabolite production at different OLRs in the SBR system. of 25–35 g VS L−1 d−1; however, optimization experiments need

wileyonlinelibrary.com/jctb © 2022 The Authors. J Chem Technol Biotechnol 2023; 98: 230–237
Journal of Chemical Technology and Biotechnology published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry (SCI).
10974660, 2023, 1, Downloaded from https://analyticalsciencejournals.onlinelibrary.wiley.com/doi/10.1002/jctb.7238 by Birla Institute of Technology & Science, Wiley Online Library on [09/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Biohydrogen from codigestion of food waste and waste activated sludge www.soci.org

Table 3. Kinetic parameters and removal at different OLR

OLR (g VS L−1 d−1) 15 22 45

Hmax (mL Lreactor−1) 140 ± 15 235 ± 27 43 ± 4


Rmax (mL Lreactor−1 h−1) 44 ± 8 62 ± 15 13 ± 4
⊗ (h) 1.6 ± 0.2 1.2 ± 0.1 1.2 ± 0.3
YH2 (mL H2 g−1 VS) 9.3 ± 1.0 15.6 ± 1.8 2.8 ± 0.2
Hydrolysis (%) 50 ± 5 37 ± 7 27 ± 5
Carbohydrate removal (%) 80.4 ± 13.0 80.8 ± 9.2 86.8 ± 15.2
VS removal (%) 61.3 ± 5.3 66.1 ± 4.5 45.2 ± 7.0
CODtotal removal (%) 34 ± 5 15 ± 7 12 ± 4
Productivity 364 ± 120 733 ± 282 383 ± 41
(mL H2 Lreactor−1 d−1)

Figure 5. Pearson's correlation between the dominant bacterial popula-


tions and process parameters.

generation of propionate (838 ± 87 mg L−1), a metabolite pro-


duced from the consumption of substrate and hydrogen.35
By increasing the OLR from 15 to 22 g VS L−1 d−1, the metabolic
pathways were dominated by the acetic and butyric acid fermen-
tations. This OLR increases butyrate production (604
± 90 mg L−1) and maintains the acetic fermentation (producing
1178 mg L−1). In this case, acetic consumption could be low since
propionate generation decreased to 235 ± 30 mg L−1. This fact
was reflected in higher H2 production (Hmax and Rmax). The OLR
of 45 g VS L−1 d−1 showed a decrease in VFA generation, increas-
Figure 4. Relative abundance of the community genera in the SBR oper-
ing lactic acid production. In this case, there was no predominant
ated at different OLRs.
fermentation pathway since each metabolite (acetate, butyrate,
propionate and lactate) was produced at a similar concentration
to be proposed to confirm it. The difference in the possible opti- range. The presence of lactate and ethanol promoted the con-
mal OLR, compared with others in literature, can be attributed sumption of substrate not associated with the H2 generation
to the FW–WAS ratio, operation regime and OLR applied. Optimi- and has been reported in the fermentation of FW.14,15,17,30
zation of these parameters will contribute to the increase in H2 The microbial characterization showed that Prevotella was the
production and energy recovery in continuous and discontinuous dominant genus independently of the OLR (Fig. 4). Prevotella
reactors. spp. is a non-spore-forming, obligate anaerobic bacteria with
Metabolite production (Fig. 3C) shows that the main VFA in all hemicellulase activity whose fermentation end products consist
the HRT was acetate, followed by butyrate (except in the OLR of of acetic and succinic acids and smaller quantities of lactic acid
15 g VS L−1 d−1, where the propionate was the second main and H2.36
VFA). The final H2 yield depends on the primary metabolite path- When the OLR increased from 15 to 22 g VS L−1 d−1, Mega-
way, the acetate pathway provides the maximum yield of sphaera became the second dominant species. Megasphaera has
4 mol H2 molglucose−1, while the butyrate and ethanol pathways been well recognized as an H2 producer from lactic and acetic
are limited to 2 mol H2 molglucose−1.34 Isobutyrate and isovalerate acids.37 Both genera have been previously reported as the domi-
were detected in low concentrations (<30 mg L−1) independently nant genera in bioreactors that produce H2 from FW,15,17,31,34-39
of the OLR. Caproate was only detected in the OLR of indicating that FW–WAS co-processing did not change the most
22 g VS L−1 d−1 at a low concentration (<20 mg L−1). The highest abundant members in the microbial community. The increase in
acetate production (1510 ± 155 mg L−1) was obtained with an OLR from 22 to 45 g VS L−1 d−1 stimulated the abundance of Mit-
OLR of 15 g VS L−1 d−1. However, H2 production was not the suokella spp. to 21%. Mitsuokella spp. are non-spore-forming,
highest under this condition. The accumulation of acetate in the strictly anaerobic, fermentative bacteria previously reported in
medium does not necessarily imply a higher production of H2 H2-producing reactors.40 Mitsuokella is able to hydrolyze phytate,
since several microbial species can convert H2 and CO2 into ace- the principal storage form of phosphorus in foods.41 It is possible
tate. These species include homoacetogenic microorganisms that to infer that the higher availability of phytate in the SBR operated
reduce CO2 to acetate through the acetyl-CoA pathway. Addition- at an OLR of 45 g VS L−1 d−1 stimulated the growth of Mitsuokella,
235

ally, H2 production under this condition was affected by the a fact that needs further research.

J Chem Technol Biotechnol 2023; 98: 230–237 © 2022 The Authors. wileyonlinelibrary.com/jctb
Journal of Chemical Technology and Biotechnology published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry (SCI).
10974660, 2023, 1, Downloaded from https://analyticalsciencejournals.onlinelibrary.wiley.com/doi/10.1002/jctb.7238 by Birla Institute of Technology & Science, Wiley Online Library on [09/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.soci.org I Moreno-Andrade, MJ Berrocal-Bravo, I Valdez-Vazquez

Interestingly, Megasphaera and Olsenella coexisted in the SBR, REFERENCES


being positively correlated with H2 production (Fig. 5). Olsenella 1 Melikoglu M, Lin C and Webb C, Analyzing global food waste problem:
spp. are anaerobic bacteria that produce lactic acid as the major pinpointing the facts and estimating the energy content. Open Eng
end product with no hydrolytic activity.42 Therefore, Olsenella 3:157–164 (2013).
2 Sanchez A, Ayala OR, Hernandez-Sanchez P, Valdez-Vazquez I and de
could be providing lactic acid to Megasphaera, resulting in the León-Rodríguez A, An environment-economic analysis of hydrogen
highest H2 production at an OLR of 22 g VS L−1 d−1. Two lactic production using advanced biorefineries and its comparison with
acid bacteria – Lactobacillus and Lactococcus – also coexisted with conventional technologies. Int J Hydrogen Energ 45:7994–28006
Megasphaera, most likely providing them substrate (lactic acid) for (2020).
H2 production. On the other hand, Enterococcus and Veillonella 3 Kim SH and Shin HS, Effects of base-pretreatment on continuous
enriched culture for hydrogen production from food waste. In J
were positively correlated, with the substrate hydrolysis being Hydrogen Energ 33:5266–5274 (2008).
more abundant at an OLR of 15 g VS L−1 d−1. Enterococcus 4 Pan J, Zhang R, El-Masahd HM, Sun H and Ying Y, Effect of food to
degrades complex polysaccharides like xylan,43 while Veillonella microorganism ratio on biohydrogen production from food waste
produces acetic and propionic acids as the major end products.44 via anaerobic fermentation. Int J Hydrogen Energ 33:6968–6975
(2008).
In a previous study, a synthetic consortium formed by Enterococ- 5 Kim DH, Kim SH and Shin HS, Hydrogen fermentation of food waste
cus (non-amylolytic), Veillonella (non-amylolytic) and Eubacterium without inoculum addition. Enzyme Microb Technol 45:181–187
(amylolytic) succeed in hydrolyzing and fermenting 97% of amy- (2009).
lomaize starch granules that contained a resistant starch frac- 6 Zhang P, Liu C, Zheng Y, Zhao Y and Zhen G, Statistical key factor opti-
mization of conditions for biohydrogen production from sewage
tion.45 In this synthetic consortium, Veillonella hydrolyzed
sludge and food waste by anaerobic codigestion. Energy Fuel 33:
maltose and maltodextrins and utilized lactic acid excreted by 11163–11172 (2019).
the other members. In the SBR, Veillonella may be involved in 7 Siddiqui Z, Horan NJ and Salter M, Energy optimisation from co-
the solubilization of the most recalcitrant starch in FW at an OLR digested waste using a two-phase process to generate hydrogen
of 15 g VS L−1 d−1, contributing to the highest hydrolysis rate of and methane. Int J Hydrogen Energ 36:4792–4799 (2011).
8 Xu R, Zhang K, Liu P, Khan A, Xiong J, Tian F et al., A critical review on
50%. An interesting fact was that the abundance of Enterococcus the interaction of substrate nutrient balance and microbial commu-
and Veillonella increased with Prevotella abundance. Previous nity structure and function in anaerobic co-digestion. Bioresour
studies found that Veillonella along with Prevotella form part of Technol 247:1119–1127 (2018).
the oral microbiota forming biofilms.45 This background, along 9 Chen S, Tao Z, Yao F, Wu B, He L, Hou K et al., Enhanced anaerobic co-
with our results, indicated that these three species were associ- digestion of waste activated sludge and food waste by sulfidated
microscale zerovalent iron: insights in direct interspecies electron
ated possibly with biofilms hydrolyzing the different fractions of transfer mechanism. Bioresour Technol 316:123901 (2020).
carbohydrates present in FW. Finally, Prevotella, despite its well- 10 Yang Q, Wu B, Yao F, He L, Chen F, Ma Y et al., Biogas production from
known hydrolytic activity, was not correlated with the substrate anaerobic co-digestion of waste activated sludge: co-substrates and
hydrolysis or carbohydrate removal. This phenomenon may be influencing parameters. Rev Environ Sci Biotechnol 18:771–793
(2019).
related to its broad spectrum of hydrolytic enzymes, including 11 Kim DH, Kim SH, Shin HS, Kim MS and Shin HS, Sewage sludge addition
xylanases,43 glycosidases, lipases and proteinases.46 Therefore, to food waste synergistically enhances hydrogen fermentation per-
the convenient activity of Prevotella was persistent with the OLR formance. Bioresour Technol 102:8501–8506 (2011).
changes, emerging as a potential keystone species that calls for 12 Mamimin C, Probst M, Gómez-Brandón M, Podmirseg ME, Insam H,
further research. Reungsang A et al., Trace metals supplementation enhanced micro-
biota and biohythane production by two-stage thermophilic fer-
mentation. Int J Hydrogen Energ 44:3325–3338 (2019).
13 Liu X, Li R, Ji M and Han L, Hydrogen and methane production by co-
digestion of waste activated sludge and food waste in the two-stage
CONCLUSIONS fermentation process: substrate conversion and energy yield. Biore-
In batch tests, a FW–WAS ratio of 90–10 increased the H2 produc- sour Technol 146:317–323 (2013).
tion (Hmax) by 22% compared to the individual FW. The SBR oper- 14 Kim SH, Han SK and Shin HS, Feasibility of biohydrogen production by
anaerobic co-digestion of food waste and sewage sludge. Int J
ation showed that OLR influenced Hmax and Rmax, obtaining the Hydrogen Energ 29:1607–1616 (2004).
best performance at an OLR of 22 g VS L−1 d−1. The cooperative 15 Santiago SG, Morgan-Sagastume JM, Monroy O and Moreno-
interaction between Olsenella and Megasphaera resulted in the Andrade I, Biohydrogen production from organic solid waste in a
highest H2 production from FW–WAS. The OLR changes pro- sequencing batch reactor: an optimization of the hydraulic and
moted different taxa of hydrolytic bacteria, while Enterococcus solids retention time. Int J Hydrogen Energ 45:25681–25688 (2020).
16 Li Z, Chen Z, Ye H, Wang Y, Luo W, Chang JS et al., Anaerobic co-
and Veillonella were stimulated at low OLR, and Mitsuokella was digestion of sewage sludge and food waste for hydrogen and VFA
stimulated at an OLR as high as 45 g VS L−1 d−1. The wide spec- production with microbial community analysis. Waste Manag 78:
trum of hydrolytic activities of Prevotella possibly allowed its prev- 789–799 (2018).
alence in the SBR. 17 Castillo-Hernández A, Mar-Alvarez I and Moreno-Andrade I, Start-up
and operation of continuous stirred-tank reactor for biohydrogen
production from restaurant organic solid waste. Int J Hydrogen Energ
40:17239–17245 (2015).
18 Ramos C, Buitrón G, Moreno-Andrade I and Chamy R, Effect of the ini-
ACKNOWLEDGEMENTS tial total solids concentration and initial pH on the biohydrogen pro-
The financial support by DGAPA-UNAM through project PAPIIT duction from cafeteria food waste. Int J Hydrogen Energ 37:13288–
IN102722 is acknowledged. Gloria Moreno, Angel Hernández 13295 (2012).
and Jaime Perez are acknowledged for their technical assistance. 19 Carrillo-Reyes J, Tapia-Rodríguez A, Buitrón G, Moreno-Andrade I,
Palomo-Briones R, Razo-Flores E et al., A standardized biohydrogen
potential protocol: an international round robin test approach. Int
J Hydrogen Energ 44:26237–26247 (2019).
20 Mizuno O, Dinsdale R, Hawkes FR, Hawkes DL and Noike T, Enhance-
CONFLICT OF INTEREST ment of hydrogen production from glucose by nitrogen gas sparg-
236

The authors declare no conflicts of interest. ing. Bioresour Technol 73:59–65 (2000).

wileyonlinelibrary.com/jctb © 2022 The Authors. J Chem Technol Biotechnol 2023; 98: 230–237
Journal of Chemical Technology and Biotechnology published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry (SCI).
10974660, 2023, 1, Downloaded from https://analyticalsciencejournals.onlinelibrary.wiley.com/doi/10.1002/jctb.7238 by Birla Institute of Technology & Science, Wiley Online Library on [09/04/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Biohydrogen from codigestion of food waste and waste activated sludge www.soci.org

21 APHA. Standard methods for the examination of water and wastewa- 34 Cabrol L, Morone A, Tapia-Venegas E, Steyer J-P and Ruiz-Filippi TE,
ter, 21st ed. Am Water Work Assoc Public Work Assoc Environ Fed. Microbial ecology of fermentative hydrogen producing biopro-
Washington, DC, USA. (2005). cesses: useful insights for driving the ecosystem function. FEMS
22 Dubois M, Gilles KA, Hamilton JK, Rebers PA and Smith F, Colorimetric Microbiol Rev 41:158–181 (2017).
method for determination of sugars and related substances. Anal 35 Motte JC, Trably E, Escudié R, Hamelin J, Steyer JP, Bernet N et al., Total
Chem 28:350–356 (1956). solids content: a key parameter of metabolic pathways in dry anaer-
23 Barragán-Trinidad M, Carrillo-Reyes J and Buitrón G, Hydrolysis of obic digestion. Biotechnol Biofuels 6:164 (2013).
microalgal biomass using ruminal microorganisms as a pretreat- 36 Wu CC, Johnson JL, Moore WEC and Moore LVH, Emended descriptions
ment to increase methane recovery. Bioresour Technol 244:100– of Prevotella denticola, Prevotella loescheii, Prevotella veroralis and
107 (2017). Prevotella melaninogenica. Int J Syst Bacteriol 42:536–541 (1992).
24 Hammer Ø, Harper DAT and Paul DR, Past: paleontological statistics 37 Ohnishi A, Abe S, Bando Y, Fujimoto N and Suzuki M, Rapid detection
software package for education and data analysis. Palaeontol Elec- and quantification methodology for genus Megasphaera as a hydro-
tron 4:art 4 (2001). gen producer in a hydrogen fermentation system. Int J Hydrogen
25 Balachandar G, Khanna N and Das D, Biohydrogen production from Energ 37:2239–2247 (2012).
organic wastes by dark fermentation, in Biohydrogen, ed. by 38 Feng K, Li H and Zheng C, Shifting product spectrum by pH adjustment
Pandey A, Chang J-S, Hallenbeck PC and Larroche C. Elsevier, during long-term continuous anaerobic fermentation of food waste.
Burlington, MA, USA, pp. 103–144 (2013). Bioresour Technol 270:180–188 (2018).
26 Hallenbeck P, Fundamentals on Biohydrogen, in Biohydrogen, Vol. 588, 39 Mariakakis I, Bischoff P, Krampe J, Meyera C and Steinmetz H, Effect of
1st edn, ed. by Pandey A, Chang JS, Hallenbeck P and Larroche C. organic loading rate and solids retention time on microbial popula-
Elsevier, Netherlands (2013). tion during bio-hydrogen production by dark fermentation in large
27 Yuan Z, Yang H, Zhi X and Shen J, Increased performance of continu- lab-scale. Int J Hydrogen Energ 36:10690–10700 (2011).
ous stirred tank reactor with calcium supplementation. Int J Hydro- 40 Muñoz-Paez K, Alvarado-Michi EL, Moreno-Andrade I, Buitrón G and
gen Energ 7:2622–2626 (2010). Valdez-Vazquez I, Comparison of suspended and granular cell
28 Yang H and Shen J, Effect of ferrous iron concentration on anaerobic anaerobic bioreactors for hydrogen production from acid agave
bio- hydrogen production from soluble starch. Int J Hydrogen Energ bagasse hydrolyzates. Int J Hydrogen Energ 45:275–285 (2020).
31:2137–2146 (2006). 41 Lan GQ, Ho YW and Abdullah N, Mitsuokella jalaludinii sp. nov., from the
29 Zabranska J and Pokorna D, Bioconversion of carbon dioxide to meth- rumens of cattle in Malaysia. Int J Syst Evol Microbiol 52:713–718
ane using hydrogen and hydrogenotrophic methanogens. Biotech- (2002).
nol Adv 36:707–720 (2018). 42 Han KI, Lee KC, Eom MK, Kim JS, Suh MK, Park SH et al., Olsenella faeca-
30 Kim SH and Shin HS, Effects of base-pretreatment on continuous lis sp. nov., an anaerobic actinobacterium isolated from human fae-
enriched culture for hydrogen production from food waste. Int J ces. Int J Syst Evol Microbiol 69:2323–2328 (2019).
Hydrogen Energ 19:5266–5274 (2008). 43 Valdez-Vazquez I, Pérez-Rangel M, Tapia A, Buitrón G, Molina C,
31 Baldi F, Pecorini I and Iannelli R, Comparison of single-stage and two-stage Hernández G et al., Hydrogen and butanol production from native
anaerobic co-digestion of food waste and activated sludge for hydro- wheat straw by synthetic microbial consortia integrated by species
gen and methane production. Renew Energy 143:1755–1765 (2019). of enterococcus and clostridium. Fuel 159:214–222 (2015).
32 Santiago SG, Trably E, Latrille E, Buitrón G and Moreno-Andrade I, The 44 Jumas-Bilak E, Carlier JP, Jean-Pierre H, Teyssier C, Gay B, Campos J
hydraulic retention time influences the abundance of Enterobacter, et al., Veillonella montpellierensis sp. nov., a novel, anaerobic, Gram-
clostridium, and lactobacillus during the hydrogen production from negative coccus isolated from human clinical samples. Int J Syst Evol
food waste. Lett Appl Microbiol 69:138–147 (2019). Microbiol 54:1311–1316 (2004).
33 Moreno-Andrade I, Carrillo-Reyes J, Santiago SG and Bujanos- 45 Kolenbrander P, The genus Veillonella. PRO 4:1022–1040 (2006).
Adame MC, Biohydrogen from food waste in a discontinuous pro- 46 Gharbia SE and Shah HN, Hydrolytic enzymes liberated gram-
cess: effect of HRT and microbial community analysis. Int J Hydrogen negative anaerobes. FEMS Immunol Med Microbiol 6:139–146
Ener 48:17246–17252 (2015). (1993).

237

J Chem Technol Biotechnol 2023; 98: 230–237 © 2022 The Authors. wileyonlinelibrary.com/jctb
Journal of Chemical Technology and Biotechnology published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry (SCI).

You might also like