Nativa, Sinop, v. 11, n. 1, p. 108-114, 2023.
Pesquisas Agrárias e Ambientais
DOI: https://doi.org/10.31413/nativa.v11i1.14583
ISSN: 2318-7670
Generation of biogas and thermal energy at the Bolo das Oliveiras
Agroindustry, Pombal, Paraíba, Brazil
José Joaquim de SOUZA NETO1, Bruno Fonsêca FEITOSA2* ,
Roberlucia Araujo CANDEIA3, Mônica Tejo CAVALCANTI3,4 , Adriana Silva LIMA3
1Federal University of Paraíba, João Pessoa, PB, Brazil.
2State University of Campinas, Campinas, SP, Brazil.
3Federal University of Campina Grande, Pombal, PB, Brazil.
4National Institute of the Semiarid Region, Campina Grande, PB, Brazil.
*E-mail: brunofonsecafeitosa@live.com
Submission: 10/31/2022; Accepted on 04/10/2023; Published on 04/13/2023.
ABSTRACT: This study aimed to assess the potential for generation of thermal energy from biogas produced
by a rural biodigester in the Bolo das Oliveiras Agroindustry, Pombal/PB, Brazil. The biodigester was fed every
two days with 0.30 m3 of biomass (mixture of water and bovine manure), retention time of 45 days. Affluent
and effluent samples were collected every 15 days for 75 days. The affluent had a higher (p < 0.05) solids
contents than the effluent. The highest dissolved oxygen concentration (6.67 mg L−1) was observed in the
affluent. The effluent had lower (p < 0.05) total alkalinity than the affluent at all sampling times. CH4 values
were higher than CO2 values throughout the experiment. Biogas also contained trace proportions of H2S and
NH3 (2/3 and 1/3 ppMV, respectively). CH4 emissions were estimated at 10.58 m3 day−1. CH4 was the major
constituent of biogas, as indicated by flame combustion behavior. Generation of biogas and thermal energy at
the Bolo das Oliveiras Agroindustry may be economically feasible, providing a minimum monthly savings of
R$ 1,582.00.
Keywords: biodigester; effluent; methane.
Geração de biogás e energia térmica na agroindústria Bolo das Oliveiras,
Pombal, Paraíba, Brasil
RESUMO: Este estudo teve como objetivo avaliar o potencial de geração de energia térmica a partir do biogás
produzido por um biodigestor rural na Agroindústria Bolo das Oliveiras, Pombal/PB, Brasil. O biodigestor foi
alimentado a cada dois dias com 0,30 m3 de biomassa (mistura de água e esterco bovino), tempo de retenção
hidráulica de 45 dias. Amostras de afluentes e efluentes foram coletadas a cada 15 dias durante 75 dias. O
afluente apresentou teores de sólidos maiores (p < 0,05) do que o efluente. A maior concentração de oxigênio
dissolvido (6,67 mg L−1) foi observada no afluente. O efluente apresentou alcalinidade total menor (p < 0,05)
do que o afluente em todos os tempos de amostragem. Os valores de CH4 foram superiores aos valores de CO2
durante todo o experimento. O biogás também continha traços de H2S e NH3 (2/3 e 1/3 ppMV,
respectivamente). As emissões de CH4 foram estimadas em 10,58 m3 dia−1. O CH4 foi o principal constituinte
do biogás, conforme indicado pelo comportamento da combustão da chama. A geração de biogás e energia
térmica na Agroindústria Bolo das Oliveiras pode ser economicamente viável, proporcionando uma economia
mensal mínima de R$ 1.582,00.
Palavras-chave: biodigestor; efluente; metano.
1. INTRODUCTION
In Brazil, cattle farming is one of the most economically
important agricultural activities. The cattle population of
Paraíba State increased by 1.6% in 2017, reaching about 1.25
million head (BRASIL, 2019). However, Brazilian agricultural
enterprises generate high amounts of organic matter and have
significant consequences on the environment (SAADY;
MASSÉ, 2015).
In recent decades, the scientific community has shown
great interest in the development of sustainable alternatives
to minimize waste generation, promoting changes in the
management of cattle waste (RIOS; KALTSCHMITT, 2016).
As animal manure is mainly composed of carbon, hydrogen,
nitrogen, and oxygen, it holds potential in thermal and
electric energy production (PIÑAS et al., 2018). There are,
however, some drawbacks in the use of animal manure, such
as low biodegradability resulting from its high content of
lignocellulosic fibers
(ANDRIAMANOHIARISOAMANANA et al., 2017).
Organic matter decomposition can be exploited to
produce biogas and electric power, contributing to the
reduction of unit costs in agricultural production. Some
researchers investigated the feasibility of using biogas to meet
daily cooking energy needs (SANTOS et al., 2018). An
example of such a system has been implemented in the Bolo
das Oliveiras Agroindustry, Pombal, Paraíba, Brazil, which
had a monthly demand of 14 liquefied petroleum gas (LPG)
cylinders (13 Kg each). To partially replace the use of LPG,
Souza Neto et al.
Nativa, Sinop, v. 11, n. 1, p. 108-114, 2023.
109
we developed a rural, sertanejo-type biodigester and a biogas
purification system that is fed with cattle waste supplied by
residents of the Várzea Comprida dos Oliveiras community.
The purpose of this study was to assess the potential for
generation of biogas and thermal energy of the biodigester
developed for the Bolo das Oliveiras Agroindustry. We
investigated the physical and chemical characteristics of the
influent and effluent at different sampling times, evaluated
biogas quality in terms of its composition, flame color,
combustion characteristics, and CH4 emission, and estimated
the potential savings obtained from thermal energy
generation.
2. MATERIAL AND METHODS
2.1. Location of the experiment
The research project was developed in the Bolo das
Oliveiras Agroindustry (Figures 1A), located in the rural
community of Várzea Comprida dos Oliveiras, Pombal,
Paraíba, Brazil. According to the 2019 census, Pombal has
32,801 inhabitants and an area of 889 km2 (IBGE, 2019). The
rural community is 9.81 km away from the Center for
Agrofood Sciences and Technology (CCTA) of the Federal
University of Campina Grande (UFCG), Pombal campus,
Paraíba, Brazil, where the laboratory analyses were
performed. Geographical coordinates for the site were
determined using a global positioning system (GPS) device
(Garmin®, model 010-01199-10) and are given in UTM
(Universal Transverse Mercator) format based on the
rectangular coordinate system: 625653 m E and 9252998 m
S.
A biodigester (Figure 1B) and biogas purification system
was installed at the Bolo das Oliveiras Agroindustry.
According to Ribeiro Filho et al. (2017), the region has hot
semi-arid climate.
2.2. Biodigester features
The biodigester has a total volume of 14.8 m3. The system
consists of (I) a 0.25 m3 feed tank, (II) a fermenting chamber
with a total capacity of 14.8 m3, (III) a 5 m3 gas holder, (IV)
a pressure adapter, (V) a primary water filter, and (VI) a 0.38
m3 overflow tank (Figure 2).
2.3. Purification system features
The biogas purification system comprises three treatment
columns (Table 1). The system also includes a compressor
that delivers the gas to an adapted oven through a 38 m long
line, operating at a pressure of about 110 lbf in−2.
Figure 2. Bolo das Oliveiras Agroindustry in Pombal, PB, Brazil (A) and area of implantation of the biodigester in the countryside (B).
Figura 2. Agroindústria Bolo das Oliveiras em Pombal, PB, Brasil (A) e área de implantação do biodigestor no interior (B).
Figure 2. Longitudinal section of the sertanejo-type biodigester and purification system. A, chemical solution; B, hydrogen sulfide (H2S) gas
filter; C, water column; D, gas compressor.
Figura 2. Corte longitudinal do biodigestor tipo sertanejo e sistema de purificação. A, solução química; B, filtro de gás sulfureto de hidrogénio
(H2S); C, coluna de água; D, compressor de gás.
A
B
Generation of biogas and thermal energy at the Bolo das Oliveiras Agroindustry, Pombal, Paraíba, Brazil
Nativa, Sinop, v. 11, n. 1, p. 108-114, 2023.
110
Table 1. Description of treatment columns of the biogas
purification system.
Tabela 1. Descrição das colunas de tratamento do sistema de
purificação de biogás.
Column
Description
A
Contains sodium hydroxide (NaOH) and removes
carbon dioxide (CO
2
).
B
Lined internally with iron filings for removal of
hydrogen sulfide (H
2
S) from the biogas mixture.
C
Contains artisanal well water and removes CO
2
, H
2
S,
ammonia (NH3), and other gases. A safety seal
prevents the reflux of gas to the biodigester.
2.4. Methods
During the study period, the digestion chamber was fed
every two days with 0.30 m3 of biomass (a mixture of 200 L
of water and 100 Kg of bovine manure). The water used was
from an artesian well and a simple effluent treatment system
installed at the agroindustry. Bovine manure was collected
from a herd of about 40 animals raised in a semi-feedlot
system, most of which were lactating cows. The amount of
manure generated by the cattle is equivalent to that produced
by five feedlot animals (10 Kg of manure per day), according
to data from Sganzerla (1983) adapted by Colatto; Langer
(2012). A hydraulic retention time of 45 days was adopted.
Influent and effluent samples were collected every 15 days
for 75 days.
2.5. Physical and chemical analyses
Physical and chemical analyses were performed in five
replications, according to the recommendations of Baird et
al. (2023). Fixed solids (FS, combustion in a muffle furnace
at 550 °C for 1 h), total solids (TS, combustion in a muffle
furnace at 550 °C for 1 h, followed by oven-drying at 105 °C
for 24 h), and volatile solids (VS) were calculated according
to Eqs. (1), (2), and (3), respectively. Electrical conductivity
(direct electrode reading), dissolved oxygen (measured using
a portable oxygen meter under controlled aerobic conditions
at 20 °C for 5 days), pH (measured using a pH meter
calibrated with pH 4.0, 7.0, and 10.0 buffer solutions), and
total alkalinity (TA, volumetric determination using a
standardized solution of sulfuric acid and calculated using
Eq. 4) were also determined.
FS (mg L) = ()×
(01)
TS (mg L) = ()×
(02)
VS (mg L) = ()×
(03)
TA (mg HCO L ) =  × × 
(04)
where: W0 is the initial weight; W1 the dry weight; W2 the final
weight; Vs the sample volume; V the volume of base used for
titration; M the base molarity; and 61000 the mass of HCO3
expressed in milligrams.
2.6. Biogas qualification and economy estimate
The proportion and composition of gases (CH4, CO2,
H2S, and NH3) in biogas was determined on-site at the time
of sampling using a kit developed by Kunz; Sulzbach (2007).
Flame color and combustion behavior were analyzed in a
biogas stove and oven. CH4 production, expressed in m3, was
calculated using Eqs. (5) and (6) and applied to estimate the
economic savings of thermal energy generation.
CH4 (m3 day1)=××××
(05)
d =
(06)
where: Mt is the total amount of manure fed to each biodigester unit
(Kg day−1), t is the sampling time (days); Na is the number of animals
producing waste, Pb is the biogas production (Kg biogas Kg−1
manure); C is the CH4 content of biogas; Ve is the specific volume
of CH4 (0.67 Kg m−3); d is the density of CH4 (0.72 Kg m−3;, m is
the mass of CH4, and V is the volume of CH4.
Biogas production was determined using the energy
conversion value for cattle recommended by the National
Biomass Reference Center (CENBIO), as adapted by
Colatto; Langer (2011).
2.7. Statistical analysis
The Shapiro-Wilk Test was applied to assess the
normality of the data. Physical and chemical data were
subjected to analysis of variance in a completely randomized
design using Assistat software version 7.7 beta (SILVA;
AZEVEDO, 2016). Means were compared by Tukey’s test at
the 5% significance level (p < 0.05).
3. RESULTS
3.1. Physical analyses
Table 2 describes the physical characteristics of influent
and effluent from the biodigester after 45, 60, and 75 days of
digestion.
The TS (3407.23–3446.67 mg L−1) and VS (3399.25–
3434.58 mg L−1) of effluents at 45 and 60 days of digestion
did not differ (p > 0.05) from each other, nor did the FS
content (6.96–7.98 mg L−1) of effluents at 60 and 75 days.
Effluent EC remained significantly higher (p < 0.05) than
influent, with no differences (p > 0.05) between effluents
collected at different times.
Table 2. Physical characteristics of the affluent and effluent of a sertanejo-type biodigester after 45, 60, and 75 days of digestion.
Tabela 2. Características físicas do afluente e efluente de um biodigestor do tipo sertanejo após 45, 60 e 75 dias de digestão.
Sample Time (days)
Parameter
TS (mg L
−1
)
FS (mg L
−1
)
VS (mg L
−1
)
EC (mS cm
−1
)
Affluent
-
3672.17 ± 1.02
a
14.52 ± 0.28
a
3657.64 ± 0.74
a
3.76 ± 0.04
b
Effluent
45
3446.67 ± 0.92
b
12.09 ± 0.81
b
3434.58 ± 0.82
b
4.01 ± 0.04
a
60
3407.23 ± 0.87
b
7.98 ± 0.50
c
3399.25 ± 0.51
b
3.95 ± 0.10
a
75
3190.85 ± 0.80
c
6.96 ± 0.05
c
3183.89 ± 0.81
c
3.96 ± 0.05
a
CV (%)
1.70
4.82
1.72
1.60
p
-
value
<0.0001
<0.0001
<0.0001
0.0059
Values are presented as the mean ± standard deviation. Means in a column followed by the same letter are not significantly different at p < 0.05 by Tukey’s
test. -, relative to all sampling periods; TS, total solids; FS, fixed solids; VS, volatile solids; EC, electrical conductivity; CV, coefficient of variation.
Souza Neto et al.
Nativa, Sinop, v. 11, n. 1, p. 108-114, 2023.
111
3.2. Chemical analyses
The chemical characteristics of influent and effluents at
45, 60, and 75 days are presented in Table 3. DO values
differed significantly from each other (p < 0.05). As expected,
the influent had the highest DO content (6.67 mg L−1).
Nevertheless, digester effluents had high DO concentrations
at all sampling times.
A relationship was observed between effluent pH values,
which did not differ (p > 0.05) from those of the influent only
on days 45 and 60. Effluents had lower TA than the influent
at all sampling times, with significant differences between
effluent samples (p < 0.05).
Table 3. Chemical characteristics of the affluent and effluent of a sertanejo-type biodigester after 45, 60, and 75 days of digestion.
Tabela 3. Características químicas do afluente e efluente de um biodigestor do tipo sertanejo após 45, 60 e 75 dias de digestão.
Sample Time (days)
Parameter
DO (mg L
−1
)
pH
TA (mg HCO
3
L
−1
)
Affluent
-
6.67 ± 0.06
a
6.67 ± 0.12
b
1259.67 ± 0.58
a
Effluent
45
4.83 ± 0.06
b
6.89 ± 0.06
ab
1258.67 ± 0.58
a
60
4.00 ± 0.10
c
6.92 ± 0.06
ab
1250.67 ± 0.58
b
75
3.60 ± 0.10
d
7.12 ± 0.17
a
1233.67 ± 0.58
c
CV (%)
1.71
1.58
0.05
p
-
value
<0.0001
0.007
<0.0001
Values are presented as the mean ± standard deviation. Means in a column followed by the same letter are not significantly different at p < 0.05 by Tukey’s
test. -, relative to all sampling times; DO, dissolved oxygen; TA, total alkalinity; HCO3, hydrogen carbonate; CV, coefficient of variation.
3.3. Biogas qualification and economy estimate
Figure 3A and B shows the results of biogas quality
testing, indicating the concentrations of CH4 and CO2, as well
as the proportions of H2S and NH3.
CH4 levels were higher than those of CO2 throughout the
experiment (Figure 3A). In the current study, the highest CH4
concentration in biogas (78%) was observed at 60 days after
digestion, not differing from those observed at 45 and 75
days (76%). The proportions of H2S and NH3 remained
stable throughout the experimental period (2/3 H2S and 1/3
NH3 ppMV). Figure 4 show photograph of the biogas flame.
The predominantly light blue color of the biogas flame
indicated a high concentration of CH4. We estimated that
CH4 was produced at a rate of 10.58 m3 day−1 or 317.26 m3
month−1 (30 days) under the current experimental
conditions. A CH4 volume of 317.26 m3 is equivalent, in
terms of energy, to 17–18 LPG cylinders of 13 Kg.
Figure 3. Concentrations of methane (CH4) and carbon dioxide (CO2) (A), and proportions of hydrogen sulfide (H2S) and ammonia (NH3)
(B) in biodigester effluents collected after 45, 60, and 75 days of digestion.
Figura 3. Concentrações de metano (CH4) e dióxido de carbono (CO2) (A), e proporções de sulfeto de hidrogênio (H2S) e amônia (NH3)
(B) em efluentes de biodigestores coletados após 45, 60 e 75 dias de digestão.
Figure 4. Photograph of the biogas flame on a stove top.
Figura 4. Fotografia da chama de biogás em cima de um fogão.
4. DISCUSSION
Solids determination is important, as it provides a
characterization of biodegradable matter, which directly
influences the efficiency of anaerobic digestion. Organic
matter is consumed and transformed into biogas through the
action of microorganisms; thus, the higher the amount of
biodegradable matter, the higher the biogas production
potential (HASSANEEN et al., 2020). The values of TS, FS,
and VS throughout the experimental period can be associated
with substrate degradation (Table 2). The solids contents of
the influent were significantly higher (p < 0.05) than those of
the effluents, as is expected for biogas generation, according
to Xiao et al. (2018).
The biodegradation rate of organic matter and,
consequently, the rate of biogas production depend on
A
B
Generation of biogas and thermal energy at the Bolo das Oliveiras Agroindustry, Pombal, Paraíba, Brazil
Nativa, Sinop, v. 11, n. 1, p. 108-114, 2023.
112
substrate composition and nutrient content, biomass/water
ratio, inoculum source, and process conditions, such as pH,
temperature, and hydraulic retention time (MONLAU et al.,
2015). As discussed by Simm et al. (2016), biodegradation
rate can also be influenced by digester configuration, such as
by the use of mechanical stirring and post-fermentation.
It is likely that EC values (Table 2) were associated with
the physicochemical characteristics of water from artesian
wells, as previously observed by Farhat et al. (2018). Artesian
well water was used to dilute bovine manure before feeding
the biodigester, possibly increasing the salt content of the
organic material. No study has investigated the correlation
between substrate salt concentration and biogas production.
McVoitte et al. (2019), however, identified that substrate
pretreatment may influence biogas yield and quality.
Furthermore, Arelli et al. (2018) reported that biofertilizers
with high salt concentrations obtained by digestion may
damage the soil and water bodies if used without
pretreatment.
The DO results (Table 3) suggest the occurrence of
microbial multiplication, indicating the need for post-
treatment to avoid pollution or contamination of water
bodies, soil, and air when using the effluent as an organic
fertilizer (ORRICO et al. 2016). TA, volatile fatty acids, and
pH are crucial to assess the level of substrate stability in the
digester and prevent system souring (GUIMARÃES et al.,
2018; JANKE et al., 2018). All pH values were within the
optimal range suggested for anaerobic digestion (pH 6.0–8.0)
by previous studies (GARDONI; AZEVEDO 2019),
obviating the need for pH correction.
According to Rosli et al. (2016), the TA below the affluent
(Table 3) is indicative of an expressive buffering capacity,
considered positive for anaerobic digestion. TA levels greater
than 1000 mg HCO3 L−1 are recommended to maintain a
neutral pH. Normally, TA ranges from 1000 to 5000 mg
HCO3 L−1 in anaerobic processes (PANYAPING;
MOONTEE, 2017).
Similarly, Simm et al. (2016) found high CH4
concentrations in biogas obtained by anaerobic digestion of
crude glycerin. Campos et al. (2013) reported CH4
concentrations of 48.60 to 68.14% in unpurified biogas
during 86 days of anaerobic digestion of coffee wastewater.
According to Piñas et al. (2018), biogas composition may be
influenced by substrate type and animal diet. Leite et al.
(2015), in studying dense sludge samples, observed that the
proportions of H2S and NH3 vary according to the type of
organic matter used for anaerobic digestion. The light of the
biogas flame (Figure 4) in agreement with the findings of
Mario et al. (2015). Calza et al. (2015) highlighted that the
calorific value of biogas varies according to the amount of
CH4 in the gaseous compound.
Considering that a conventional LPG cylinder of 13 Kg
costs R$ 113.00, we estimated that biogas generation
afforded a monthly savings of R$ 1,582.00 for the Bolo das
Oliveiras Agroindustry. As the agroindustry consumes only
14 cylinders per month, biogas generation provides a surplus
of four cylinders (R$ 452.00). Assuming the installation costs
of the biodigester to be R$ 9500.00, it is estimated that the
time for a return on investment is 6 months. This shows that
the sertanejo-type biodigester can be an economically feasible
investment with a short payback period.
5. CONCLUSIONS
The results of this experiment demonstrate that the solids
contents of biodigester effluents are influenced by organic
matter biodegradation rate. Substrate pretreatment was
recommended for higher biogas production and quality.
Chemical analysis underscored the importance of effluent
post-treatment to minimize environmental impacts. The
biogas had a high CH4 concentration, as evidenced by the
predominantly blue flame. The use of biogas for generation
of thermal energy in the Bolo das Oliveiras Agroindustry
seems to be economically feasible, capable of affording
minimum monthly savings of R$ 1,582.00. Biogas plants are
an efficient solution to improve manure management,
mitigate environmental impacts, and stimulate investment in
renewable energy. However, further studies are needed to
improve the efficiency, monitoring, and savings of biogas
generation.
6. REFERENCES
ANDRIAMANOHIARISOAMANANA, F. J.; SAIKAWA,
A.; TARUKAWA, K.; QI, G.; PAN, Z.; YAMASHIRO,
T.; IWASAKI, M.; IHARA, I.; NISHIDA, T.; UMETSU,
K. Anaerobic codigestion of dairy manure, meat and
bone meal, and crude glycerol under mesophilic
conditions: synergistic effect and kinetic studies. Energy
for Sustainable Development, v. 40, p. 11-18, 2017.
https://doi.org/10.1016/j.esd.2017.05.008
ARELLI, V.; BEGUM, S.; ANUPOJU, G. R.; KURUTI, K.;
SHAILAJA, S. Dry anaerobic co-digestion of food waste
and cattle manure: Impact of total solids, substrate ratio
and thermal pre treatment on methane yield and quality
of biomanure. Bioresource Technology, v. 253, p. 273-
280, 2018.
https://doi.org/10.1016/j.biortech.2018.01.050
BRAZIL. Ministry of Agriculture. Agricultural and Livestock
Plan. Secretary of Agricultural Policy. Agricultural and
Livestock Plan 2018-2019, 2019. Available at:
<http://www.agricultura.gov.br/assuntos/politica-
agricola/plano-agricola-e-pecuario>
CALZA, L. F.; LIMA, C. B.; NOGUEIRA, C. E. C.;
SIQUEIRA, J. A. C.; SANTOS, R. F. Cost assessment of
biodigester implementation and biogas-produced energy.
Journal of the Brazilian Association of Agricultural
Engineering, v. 35, n. 6, p. 990-997, 2015.
http://dx.doi.org/10.1590/1809-4430-
Eng.Agric.v35n6p990-997/2015
CAMPOS, C. M. M.; PRADO, M. A. C.; PEREIRA, E. L.
Anaerobic digestion of wastewater from coffee and
chemical analysis of biogas produced using gas
chromatography: quantification of methane, and
potential energy gas exchanger. Bioscience Journal, v.
29, n. 3, p. 570-581, 2013.
COLATTO, L.; LANGER M. Biodigestor - solid livestock
waste for energy production. Unoesc & Ciência
ACET, v. 2, n. 2, p. 119-128, 2011.
FARHAT, A.; MILADI, B.; HAMDI, M.; BOUALLAGUI,
H. Fermentative hydrogen and methane co-production
from anaerobic co-digestion of organic wastes at high
loading rate coupling continuously and sequencing batch
digesters. Environmental Science and Pollution
Research, v. 25, n. 28, p. 27945-27958, 2018.
https://doi.org/10.1007/s11356-018-2796-2
Souza Neto et al.
Nativa, Sinop, v. 11, n. 1, p. 108-114, 2023.
113
GARDONI, R. A. P.; AZEVEDO, M. A. Study of the
biodegradation of poultry carcasses through the
composting process in closed discontinuous biodigesters.
Revista Engenharia Sanitária, v. 24, n. 3, p. 425-429,
2019. https://doi.org/10.1590/s1413-41522019118916
GUIMARÃES, C. S.; MAIA, D. R.; SERRA, E. G.
Construction of biodigesters to optimize the production
of biogas from anaerobic co-digestion of food waste and
sewage. Energies, v. 11, n. 4, p. 1-10, 2018.
https://doi.org/10.3390/en11040870
HASSANEEN, F. Y.; ABDALLAH, M. S.; AHMED, N.;
TAHA, M. M.; ELAZIZ, S. M. M. A.; EL-MOKHTAR,
M. A.; BADARY, M. S.; ALLAM, N. K. Innovative
nanocomposite formulations for enhancing biogas and
biofertilizers production from anaerobic digestion of
organic waste. Bioresource Technology, v. 309,
e123350, 2020.
https://doi.org/10.1016/j.biortech.2020.123350
IBGE. Instituto Brasileiro de Geografia e Estatística.
Population, 2019. Available
at:<https://cidades.ibge.gov.br/brasil/pb/pombal/pan
orama>.
JANKE, L.; WEINRICH, S.; LEITE, A. F.; STRAUBER,
H.; RADETSKI, C. M.; NIKOLAUSZ, M.; NELLES,
M.; STINNER, W. Year-round biogas production in
sugarcane biorefineries: Process stability, optimization
and performance of a two-stage reactor system. Energy
Conversion and Management, v. 168, p. 188-199,
2018. https://doi.org/10.1016/j.enconman.2018.04.101
KUNZ, A.; SULZBACH, A. Portable Biogas Kit: For
analyzing the concentration of methane gas, carbon
dioxide, ammonia and hydrogen sulphide in biogas.
KUNZ, A. [et al.]. Brazil. Patent 012070001117. 09 Oct.
2007. Available at:
<https://gru.inpi.gov.br/pePI/jsp/patentes/PatenteSea
rchBasico.jsp>
LEITE, W.; MAFFAZZIOLI, E.; GUIMARÃES, L.;
MAGO, A. D.; BELLI FILHO, P. Comparison of
organic loading rate and hydraulic retention time effects
on the mesophilic anaerobic digestion of thickened waste
activated sludge. Engenharia Sanitária e Ambiental, v.
20, n. 4, p. 581-588, 2015.
http://dx.doi.org/10.1590/S1413-
41522015020040105625
MARIO, J. S.; COELHO, M. A. A.; SCHAEFFER, L.;
ROSSINI, E. G. Preliminary study for compression of
biogas in cylinders for domestic consumption. Revista
Espacios, v. 36, n. 6, p. 1-11, 2015.
MCVOITTE, W. P. A.; CLARK, O. G. The effects of
temperature and duration of thermal pretreatment on the
solid-state anaerobic digestion of dairy cow manure.
Heliyon, v. 5, e02140, 2019.
https://doi.org/10.1016/j.heliyon.2019.e02140
MONLAU, F.; SAMBUSITI, C.; FICARA, E.;
ABOULKAS, A.; BARAKAT, A.; CARRERE, H. New
opportunities for agricultural digestate valorization:
current situation and perspectives. Energy &
Environmental Science, v. 9, p. 2600-2621, 2015.
https://doi.org/10.1039/C5EE01633A
ORRICO, A. C. A.; LOPES, W. R. T.; MANARELLI, D. M.;
ORRICO JUNIOR, M. A. P.; SUNADA, N. S.
Anaerobic co-digestion of dairy cattle manure and waste
oil. Journal of the Brazilian Association of
Agricultural Engineering, v. 36, n. 3, p. 537-545, 2016.
PANYAPING, K.; MOONTEE, P. Potential of biogas
production from mixed leaf and food waste in anaerobic
reactors. Journal of Material Cycles and Waste
Management, v. 20, p. 723-737, 2017.
http://dx.doi.org/10.1007/s10163-017-0629-x
PIÑAS, J. A. V.; VENTURINI, O. J.; LORA, E. E. S.;
ROALCABA, O. D. C. Technical assessment of mono-
digestion and co-digestion systems for the production of
biogas from anaerobic digestion in Brazil. Renewable
Energy, v. 117, p. 447-458, 2018.
https://doi.org/10.1016/j.renene.2017.10.085
BAIRD, R.; EATON, A.; RICE, E.; BRIDGERWATER, L.
Standard methods for the examination of water and
wastewater. 24 ed. New York: American Public Health
Association, 2023. 1624p.
RIOS, M.; KALTSCHMITT, M. Electricity generation
potential from biogas produced from organic waste in
Mexico. Renewable and Sustainable Energy Reviews,
v. 54, p. 384-395, 2016.
https://doi.org/10.1016/j.rser.2015.10.033
ROSLI, N. S.; IDRUS, S.; DAUD, N.; AHSAN, A.
Assessment of potential biogas production from rice
straw leachate in upflow anaerobic sludge blanket reactor.
International Journal of Smart Grid and Clean
Energy, v. 5, n. 3, p. 135-143, 2016.
http://dx.doi.org/10.12720/sgce.5.3.135-143
RIBEIRO FILHO, J. C.; PALÁCIO, H. A. Q.; ANDRADE,
E. M.; SANTOS, J. C. N.; BRASIL, J. B. Rainfall
characterization and sedimentological responses of
watersheds with different land uses to precipitation in the
semiarid region of Brazil. Revista Caatinga, v. 30, n. 2,
p. 468-478, 2017. https://doi.org/10.1590/1983-
21252017v30n222rc
SAADY, N. M. C.; MASSÉ, D. I. High rate psychrophilic
anaerobic digestion of high solids (35%) dairy manure in
sequence batch reactor. Bioresource Technology, v.
186, p. 74-80, 2015.
https://doi.org/10.1016/j.biortech.2015.03.038
SANTOS, I. F. S.; VIEIRA, N. D. B.; NÓBREGA, L. G. B.;
BARROS, R. M.; TIAGO FILHO, G. L. Assessment of
potential biogas production from multiple organic wastes
in Brazil: Impact on energy generation, use, and
emissions abatement. Resources, Conservation and
Recycling, v. 131, p. 54-63, 2018.
https://doi.org/10.1016/j.resconrec.2017.12.012
SGANZERLA, E. Biodigestor: a solution. Porto Alegre:
Agriculture, 1983. 88p.
SILVA, F. A. Z.; AZEVEDO, C. A.V. The assistat software
version 7.7 and its use in the analysis of experimental
data. African Journal of Agricultural Research, v. 11,
n. 39, p. 3733-3740, 2016.
https://doi.org/10.5897/AJAR2016.11522
SIMM, S.; ORRICO, A. C. A.; ORRICO JUNIOR, M. A. P.;
SUNADA, N. S.; SCHWINGEL, A. W.; COSTA, M. S.
S. M. Crude glycerin in anaerobic co-digestion of dairy
cattle manure increases methane production. Scientia
Agricola, v. 74, n. 3, p. 175-179, 2016.
https://doi.org/10.1590/1678-992x-2016-0057
XIAO, B.; ZHANG, W.; WU, J.; QIANG, H.; LIU, J.; LI, Y.
Temperature-phased anaerobic digestion of food waste:
A comparison with single-stage digestions based on
performance and energy balance. Bioresource
Generation of biogas and thermal energy at the Bolo das Oliveiras Agroindustry, Pombal, Paraíba, Brazil
Nativa, Sinop, v. 11, n. 1, p. 108-114, 2023.
114
Technology, v. 249, p. 826-834, 2018.
http://dx.doi.org/10.1016/j.biortech.2017.10.084
Author Contributions:
All authors of this work contributed equally in all
functions in the article, from its conception to the writing. All
authors read and agreed to the published version of the
manuscript.
Funding: No funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement:
Study data can be obtained by request to the corresponding
author or the second author, via e-mail. It is not available on the
website as the research project is still under development.
Conflicts of Interest:
The authors declares no conflict of interest. Supporting entities
had no role in the design of the study; in the collection, analyses, or
interpretation of data; in the writing of the manuscript, or in the
decision to publish the results.