ArticlePDF Available

Abstract and Figures

New technologies based on the anaerobic digestion process make it possible to manage problematic waste. Methane efficiency depends largely on the level of the hydration of the substrates used for biogas production and their ability to decompose easily. The aim of this study was to present the current state of knowledge and practices in substrate hydration characteristics, focusing on pretreatment methods as the preferred method for improving efficiency. The paper discusses issues related to the degree of hydration of substrates in the context of their use in biogas plants. Reference was also made to topics related to the transportation and logistics of raw material supply regarding environmental impact. Biogas plant projects should be expanded to include an element related to assessing the impact of raw material deliveries on the immediate environment. Previous papers have not sufficiently analyzed the aspect related to the hydration of substrates used in anaerobic digestion processes. The presented and discussed research results can be implemented to optimize biogas plant water management processes. By replacing standard feedstock transportation methods with a pipeline, the environmental impact can be reduced by nearly ten times.
Content may be subject to copyright.
Citation: Czekała, W.; Nowak, M.;
Bojarski, W. Characteristics of
Substrates Used for Biogas
Production in Terms of Water
Content. Fermentation 2023,9, 449.
https://doi.org/10.3390/
fermentation9050449
Academic Editor: Alessia Tropea
Received: 11 April 2023
Revised: 25 April 2023
Accepted: 4 May 2023
Published: 9 May 2023
Copyright: © 2023 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
fermentation
Review
Characteristics of Substrates Used for Biogas Production in
Terms of Water Content
Wojciech Czekała * , Mateusz Nowak and Wiktor Bojarski
Department of Biosystems Engineering, Pozna´n University of Life Sciences, Wojska Polskiego 50,
60-637 Pozna´n, Poland; mateusz.nowak@up.poznan.pl (M.N.); wiktor.bojarski@up.poznan.pl (W.B.)
*Correspondence: wojciech.czekala@up.poznan.pl; Tel.: +48-749-265-938
Abstract:
New technologies based on the anaerobic digestion process make it possible to manage
problematic waste. Methane efficiency depends largely on the level of the hydration of the substrates
used for biogas production and their ability to decompose easily. The aim of this study was to present
the current state of knowledge and practices in substrate hydration characteristics, focusing on
pretreatment methods as the preferred method for improving efficiency. The paper discusses issues
related to the degree of hydration of substrates in the context of their use in biogas plants. Reference
was also made to topics related to the transportation and logistics of raw material supply regarding
environmental impact. Biogas plant projects should be expanded to include an element related to
assessing the impact of raw material deliveries on the immediate environment. Previous papers have
not sufficiently analyzed the aspect related to the hydration of substrates used in anaerobic digestion
processes. The presented and discussed research results can be implemented to optimize biogas
plant water management processes. By replacing standard feedstock transportation methods with a
pipeline, the environmental impact can be reduced by nearly ten times.
Keywords: anaerobic digestion; fermentation; water content; pretreatment; biomass; biogas
1. Introduction
1.1. Climate Problems
Climate change resulting from relentless economic development poses a significant
threat to future generations. It is necessary to make reasonable and practical changes,
primarily in economic sectors related to food production and energy [
1
]. One of the
most noticeable climate changes around the world is the phenomenon of global warming.
Emissions of greenhouse gases such as methane, carbon dioxide, and nitrous oxide, for
example, contribute to this situation [
2
]. Replacing fossil fuels with biomass-derived fuels
is one way to reduce greenhouse gas emissions into the environment. Biogas produced
by anaerobic digestion is a solution for the environmental and energy crisis [
3
]. The use
and production of this type of biofuel are possible both in developing countries whose
economy is based on agriculture and in developed ones that produce significant amounts of
waste with high energy potential. Despite the possibilities, the development level of biogas
technologies in many countries currently varies [
4
]. This sector ’s problems on the road to
expansion are often due to inadequate legal conditions in this area and a lack of capital.
However, it should be noted that countries in Europe and Asia building biogas plants are
solving an important aspect: the possibility of processing various types of waste [5].
1.2. Biomass Characteristics and Utilization
The quantity and quality of research on processing and ways to use hard-to-find
biomass have recently increased significantly [
6
]. Biomass is a valuable energy material.
Biomass includes products and wastes of plant and animal origins. Sawdust is also an
example of biomass, the waste generated after felling trees [
7
]. Some other agricultural
Fermentation 2023,9, 449. https://doi.org/10.3390/fermentation9050449 https://www.mdpi.com/journal/fermentation
Fermentation 2023,9, 449 2 of 15
wastes, e.g., olive pruning and tea leaf residues and wetland vegetations, are also biomass-
rich lignocellulosic materials that can be subsequently reused for biogas production [
8
,
9
].
Ongoing research indicates that it is possible to use available biomass sources, such as
composting or anaerobic digestion [
10
,
11
]. By processing and removing excess moisture
from such renewable substrates, the possibilities for their use can be expanded. Biomass,
briquettes or pellets are an alternative that reduces fossil fuel extraction [
12
]. In addition,
biomass is also used to produce biochemical compounds or biocatalysts that enable chemi-
cal transformations [
13
]. In a breakdown of the world’s total energy demand, biomass, and
waste provide 55.7 EJ, ranking fourth. For years, the top three have consistently been coal,
oil, and natural gas [
14
]. Biomass is the primary energy resource in developing countries
with limited access to fossil fuels [15].
The use of waste biomass is currently becoming a popular solution worldwide [
16
].
One can distinguish between sources of traditional biomass that have been used since
the beginning of time on earth, such as wood, and biomass, which has been divided
into three distinct generations [
17
]. The first generation of biofuels are products derived
from crops that compete with human food production. This includes liquid biofuels
produced from feedstocks with high starch and sucrose content. There are concerns about
the competition between food and energy production from such sources, hence, second-
generation biofuels have been introduced [
18
,
19
]. A good example is biofuels from inedible
plants or agricultural and forestry residues. The second generation is thus a solution to
produce green energy while processing waste [
20
]. The last type is third generation biofuels,
which use algae for energy. These are microorganisms characterized by their easy culture
and simple cellular structure. Major advantages of algae are: no competition with food
crops for arable land, high growth rates, and low fractions of lignin, which reduces the need
for energy-intensive pretreatment. However, some disadvantages, such as the presence
of high water content, seasonal chemical composition, and the occurrence of inhibitory
phenomena during anaerobic digestion, make algal biofuels not yet economically feasible,
even though they are more environmentally friendly than fossil fuels. They also thrive in
polluted water, which they filter during growth [21].
There are various ways and technologies to convert biomass to energy. The current
climate policy of EU countries seeks to reduce greenhouse gas emissions by 40% by 2030
from 1990 levels [
22
]. Biomethane facilities play an essential role in the energy transition,
enabling the production of biomethane to replace the need for natural gas. As is well
known, using and replacing hard coal with natural gas or biomethane reduces smog
and high particulate matter emissions in many countries, including Poland [
23
,
24
]. The
increase in demand for gaseous fuel contributes to developing and constructing new
biogas plants. This potential is significant in countries producing significant amounts of
waste with appropriate energy parameters [
25
]. The versatility of modern biogas plants
makes it possible to use a variety of feedstock substrates. The waste processing through
anaerobic digestion and the generation of environmentally friendly renewable energy fits
perfectly with the goal of climate neutrality [
26
]. The resulting by-product of the process,
the digestate, is a valuable organic fertilizer used in the agricultural sector [
27
,
28
]. Research
conducted by Kumar et al. [
27
] suggest the sustainable up-cycling of spent mushroom
substrates inspired by a circular economy approach through a synergistic production of
bioenergy and secondary fruit crops, which could potentially contribute to minimize the
carbon footprints of the mushroom production sector.
1.3. Water Requirements for Biomass Production
In addition to the fertilizer aspect, energy crops require certain amounts of water
for growth and development. Despite the availability of good fertilizer in the form
of digestate, the current problem for agriculture is drought. Research conducted by
Mathioudakis et al.
[
29
] presents results on substrate conversion technology in terms of
water use. According to the paper’s authors, biogas is the energy source with the smallest
water footprint, among other biomass conversion technologies, such as pyrolysis and
Fermentation 2023,9, 449 3 of 15
gasification [
28
,
29
]. A closed-loop operation of the plant has a beneficial effect on water
management. It is possible to use digestate pulp for fertilizer or, after careful separation, also
for crop irrigation [
30
,
31
]. Comparing total freshwater demand globally, the agricultural
sector uses about 92% of the resource [
32
]. Authors Mekonnen and Gerbens-Leenes [
33
]
point to a high water footprint for certain biofuels, particularly the first hydropower gen-
eration. For biomass from crops directed for use as biogas feedstock, the total water
footprint for this type of agricultural production always remains the same. However, some
directed crops have a relatively high water footprint due to limited opportunities to harvest
residues [16,33].
Research on the carbon footprint is increasingly being analyzed, but recent years have
also directed another essential aspect—water footprint analysis. The study aim was to dis-
cuss the topic of water content in substrates used for biogas production. The main chapters
analyze our own and the literature research considering the importance of the proper water
management of modern biogas plants. The paper also presents the latest methods used
during substrate pretreatment that increase the efficiency of anaerobic digestion processes.
Activities and topics related to the work are necessary to increase the efficiency of water
use and conservation, which can contribute to achieving climate neutrality.
2. Characteristics of Substrates Used for Anaerobic Digestion
2.1. Types and Potential of Raw Materials Used for Biogas Production
By analyzing the literature on the use of raw materials and waste for biogas production,
it is predicted that it is possible to reach 108 EJ by 2030. Using the potential of biomass, it is
possible to replace 20% of primary energy in this way [
34
]. Regardless of the projections,
reaching this level is challenging for the biofuel sector due to the limited availability of
raw materials and the difficulty of their conversion. The residues used present significant
differences in chemical composition and dry matter content. In addition, some chemical
compounds in the feedstock can severely limit the biogas production potential, negatively
affecting the anaerobic digestion [
35
]. Using biogas facilities is, therefore, a rational alter-
native to reduce the consumption of fossil fuels. In the anaerobic digestion process, it is
possible to use a variety of substrates, including highly hydrated waste from the agro-food
industry. After anaerobic digestion, the waste is transformed into an environmentally
safe substance as digestate [
36
,
37
]. In addition, biogas facilities are part of conducting a
closed-loop economy. In the case of biogas combustion in cogeneration engines, the emitted
carbon dioxide is reabsorbed in the photosynthesis process of plants that are substrates in
the anaerobic digestion process, such as corn. The seasonality of the production of some of
the substrates is also an important aspect. In most installations, it is necessary to build a
storage area to ensure continuous production [38,39].
Dividing the substrates used in anaerobic digestion into appropriate categories is
possible. According to Yuan and Gerbens-Leenes [
32
], the first category is biomass from
agriculture. Substrates can come from crops directly directed to produce high-energy
crops, e.g., corn, and from crop residues and residues such as manure. Forestry-related
residues are included in the following category. This sector is characterized by logging
waste or waste from the paper industry. All other municipal and industrial organic wastes
were included in the third category of anaerobic digestion substrates [
32
]. The primary
criterion for selecting a suitable feedstock mix is biogas efficiency, feedstock availability,
and economic and environmental aspects. The introduction of an additional substrate can
have a beneficial effect on increasing the energy efficiency of the process. Carbohydrates
and proteins are characterized by a faster rate of decomposition compared to a substance
made up of fats. However, according to Table 1, it is fats that have the highest methane
efficiency [
40
]. An essential parameter for the substrates used in the anaerobic digestion
process is also the C/N ratio, which should be at an equilibrated level to ensure the stability
of biogas production [41].
The potential for using selected feedstocks for biogas production is determined by
methane efficiency. In addition to the differences in methane efficiency, an important aspect
Fermentation 2023,9, 449 4 of 15
is the availability of substrates near the biogas plant [
42
]. Table 1shows the theoretical
potential for the methane efficiency of selected compounds contained in substrates used by
biogas plants.
Table 1. Maximum theoretical biogas efficiency (own elaboration based on [4345]).
Component Methane Efficiency
(m3kg1VS) Reference
Carbohydrates 0.42 [43]
Proteins 0.50 [43]
Fats 1.01 [4]
Table 1shows that the most significant potential in the amount of methane produced is
in the substrates containing significant amounts of fat. Types of fat-rich substrates include
oil leachates. A side effect of using fat-rich substrates is the formation of long-chain fatty
acids in the decomposition process, which become toxic to bacteria at excessive concen-
trations [
46
]. Another group of substrates exhibiting high methane efficiency are products
containing significant amounts of proteins. During the decomposition of proteins, NH
4+
compounds are formed, which are converted into ammonia compounds at a later stage of
the process. Too high a concentration of ammonia can lead to the inhibition of anaerobic
digestion. However, it is possible for microorganisms to adapt to an environment with
higher concentrations of ammonia [
47
,
48
]. In addition to protein, lipid, and carbohydrate
compounds, it is also possible to use biomass with high levels of lignocellulose. Lignocellu-
losic biomass makes it more difficult to decompose the substance, which translates into
lower methane yields compared to the other groups [49].
The key findings and the preferred direction of biogas plants is to process and use
problematic substrates to manage and treat waste, which is also costly to dispose of [
50
]. An
essential aspect of the efficient operation of biogas plants is to ensure the correct proportions
of the feedstock, which allows for the optimal anaerobic digestion process and the growth
of new structures of methane bacteria. Meeting the difficulty of achieving a high level
of biogas efficiency is not possible if the anaerobic digestion process is carried out using
only one substrate. New trends indicate using two or more substrates, most often in the
co-digestion process, thus increasing plant profitability [51,52].
2.2. Water Content of Substrates vs. the Environmental Effect of Their Transport
Using various feedstocks for biogas production often necessitates transporting and
delivering them to facilitie several or tens of kilometers away. Ongoing research indicates
that the logistics of supplying feedstock substrates and the export and management of
digestate are components of biogas plant operations that adversely affect the environment.
The distance involved in transportation has the most significant impact on environmental
pollution and affects global warming in particular [
53
,
54
]. Another critical issue is how raw
materials are stored on the biogas plant site before they are used for anaerobic digestion.
Adequate storage has a positive effect on preserving the freshness of the products and
keeping their methane efficiency constant. Water in the form of moisture bound in many
substrates creates problems during improper storage, as it can lead to the appearance of
bacteria and mold [55].
Based on the problems associated with the operation of a biogas plant, Tucki et al. [
56
]
analyzed factors affecting operational efficiency. The main aspect impacting the efficiency
of biogas plants is the transportation distance of substrates. In addition, it is also necessary
to prepare appropriate technological infrastructure to facilitate the process of feeding
digesters. The authors also point out an essential and often overlooked aspect: the need to
manage the digestate. The use of the by-product in question on one’s fields or the provision
of guarantees in receiving the digestate determines the efficiency of the plant’s operation.
However, the analyzed article does not specify a vital aspect: the minimum distances that
Fermentation 2023,9, 449 5 of 15
guarantee the economic justification of the transportation of pulp and substrates. It is
proposed to research transportation’s economic and environmental aspects [56].
The carbon footprint associated with substrate supply is also significantly affected
by the parameter of dry matter and dry organic matter. Running the anaerobic digestion
process with a highly hydrated substrate decreases methane efficiency per ton of fresh
matter. Therefore, it is a good practice to use and collect the raw material from a fixed
location to ensure commensurate parameters for a specific type of substrate. In addition, if
the owner of a biogas plant contracts a significant amount of substrate, it is necessary to
establish acceptable deviations from the assumed parameters indicative of its quality. An
important aspect is also the possibility of transporting acceptable and, at the same time,
maximum quantities of transported substrates by vehicles with a large capacity [57,58].
Research conducted by Muradin and Foltynowicz [
59
] describes the environmental im-
pact aspect of feedstock transports for an example of biogas plants. The authors conducted
their experiments with four different biogas plants. This impact was expressed in [Pt]
points based on ISO 14040-44 and LCIA Impact 2002+ [
58
,
59
]. The results on the logistics
of transporting raw materials to biogas plants are presented in the analyzed article. The
standard transportation of slurry using a farm tractor with a barrel to a modern pipeline
was compared. The results also show that transportation has a negative impact on the
environmental effect. Using a pipeline, it is possible to reduce the negative environmental
impact by up to 10 times compared to the standard transportation method [59].
As mentioned earlier, the key findings is the dry matter level for liquid and solid
substrates. The feedstock dry matter parameter mainly determines the methane efficiency.
In general, the higher the feedstock dry matter value, the higher the methane efficiency per
ton of fresh substrate mass [60].
3. Feedstock Hydration and Anaerobic Digestion Technology
Processing biomass to produce biogas can be carried out in many different ways. Clas-
sification of the anaerobic digestion process requires the determination of an appropriate
criterion. The most popular division of anaerobic digestion technologies is based on the
temperature range at which the process occurs. Determining the optimal temperature level
is related to the type of bacteria involved in the anaerobic decomposition process. The
optimum temperature value allows for the proper decomposition of organic matter and
provides a suitable environment for a particular group of bacteria. Using this variable, the
different types of anaerobic digestion technology have been characterized as psychrophilic,
mesophilic, and thermophilic [
61
]. Another possibility is the division resulting from the
number of process steps, the method of substrate dosing, or the degrees of separation
of the different phases of anaerobic digestion. Depending on the type of substrate pro-
cessed, the critical factor determining the anaerobic digestion technology is the degree of
hydration. The dry matter content of the chamber determines the type of technology used,
distinguishing between wet and dry anaerobic digestions [62].
In biogas plants, the average daily feedstock requirement ranges from a few to tens
of Mg of feedstock. A biogas plant with a capacity of 0.5 MW
e
requires a daily supply
of 30 Mg of feedstock to the digester [
63
]. A biogas facility’s size and electrical capacity
also depend on factors related to the amount and type of substrates used, which are found
in the immediate vicinity. Regardless of the biogas plant size, each consists of the same
essential components, including those related to feedstock storage. In the case of solid
feedstock such as corn silage, for example, it is essential to ensure proper conditions during
preparation and storage [
55
]. A portion of waste is characterized by odor nuisance and
requires storage in storage halls with biological or activated carbon filters. In the case
of liquid substrates, problems are encountered with the need to install mixers or heating
systems at the storage site. Therefore, the choice of the appropriate anaerobic digestion
technology can be determined, in addition to the availability of raw material and the
convenience of its storage [
55
,
64
]. In addition, it should be noted that the selected anaerobic
Fermentation 2023,9, 449 6 of 15
digestion method is not determined by the dry weight compactness of the individual
feedstock but by the hydration of the feedstock mixture entering the digester [64].
3.1. Wet Anaerobic Digestion
For years, the wet anaerobic digestion process has continued to be the most common
and dominant way of treating various types of substrates processed in biogas plants [
62
].
This type of anaerobic digestion is popular due to the nature of the process, in which
the bacteria for each of the different phases require an aqueous environment in their
surroundings. If the reactor mixture is not hydrated enough, the bacteria will have difficult
access to the substrate, which can cause only partial decomposition and a decreased
biogas production. Moisture is also a factor that provides a substrate for bacteria to grow
and multiply, ensuring the high efficiency of the ongoing process. In addition, for wet
technology, natural process inhibitors are distributed throughout the mixture, ensuring
their low concentration and not interfering with the process. Low hydraulic retention
time (HRT) for highly hydrated substrates such as slurry can result in the leaching of
microorganisms. When selecting substrates for biogas plants, it is necessary to take into
account the dry matter input and dry organic matter input load of the digester [65,66].
Dry matter values in the reactor chamber are conventional and are defined differently
depending on the available literature sources. The most commonly accepted values for the
wet process are <10–15% dry matter [
67
69
]. The low solids content and liquid consistency
create the possibility of pumping the mixture between chambers and the subsequent
management of digestion residues [
30
]. In wet technology, operating biogas plants are
characterized by a full flow of substances through the entire volume of the reactor. Before
refilling the reactor with a new portion of the substrate, a conditioning process is used,
i.e., adding process water or the liquid fraction of the digestate to maintain the assumed
level of the dry mass of the feedstock. Ensuring a suitable mixing ratio is possible through
vertical and horizontal mechanical mixers or a hydraulic mixing process (liquid recycling).
The advantages of building this type of plant are the ease of construction of the tank and
the possibility of carrying out maintenance work without emptying the reactor [
69
]. Based
on research conducted by Luning et al. [
70
], a mass flow model was developed for the wet
anaerobic digestion process for a full-scale plant, as shown in Figure 1.
Fermentation 2023, 9, x FOR PEER REVIEW 6 of 15
liquid substrates, problems are encountered with the need to install mixers or heating
systems at the storage site. Therefore, the choice of the appropriate anaerobic digestion
technology can be determined, in addition to the availability of raw material and the con-
venience of its storage [55,64]. In addition, it should be noted that the selected anaerobic
digestion method is not determined by the dry weight compactness of the individual feed-
stock but by the hydration of the feedstock mixture entering the digester [64].
3.1. Wet Anaerobic Digestion
For years, the wet anaerobic digestion process has continued to be the most common
and dominant way of treating various types of substrates processed in biogas plants [62].
This type of anaerobic digestion is popular due to the nature of the process, in which the
bacteria for each of the dierent phases require an aqueous environment in their sur-
roundings. If the reactor mixture is not hydrated enough, the bacteria will have dicult
access to the substrate, which can cause only partial decomposition and a decreased bio-
gas production. Moisture is also a factor that provides a substrate for bacteria to grow and
multiply, ensuring the high eciency of the ongoing process. In addition, for wet technol-
ogy, natural process inhibitors are distributed throughout the mixture, ensuring their low
concentration and not interfering with the process. Low hydraulic retention time (HRT)
for highly hydrated substrates such as slurry can result in the leaching of microorganisms.
When selecting substrates for biogas plants, it is necessary to take into account the dry
maer input and dry organic maer input load of the digester [65,66].
Dry maer values in the reactor chamber are conventional and are dened dierently
depending on the available literature sources. The most commonly accepted values for the
wet process are <1015% dry maer [6769]. The low solids content and liquid consistency
create the possibility of pumping the mixture between chambers and the subsequent man-
agement of digestion residues [30]. In wet technology, operating biogas plants are charac-
terized by a full ow of substances through the entire volume of the reactor. Before rell-
ing the reactor with a new portion of the substrate, a conditioning process is used, i.e.,
adding process water or the liquid fraction of the digestate to maintain the assumed level
of the dry mass of the feedstock. Ensuring a suitable mixing ratio is possible through ver-
tical and horizontal mechanical mixers or a hydraulic mixing process (liquid recycling).
The advantages of building this type of plant are the ease of construction of the tank and
the possibility of carrying out maintenance work without emptying the reactor [69]. Based
on research conducted by Luning et al. [70], a mass ow model was developed for the wet
anaerobic digestion process for a full-scale plant, as shown in Figure 1.
Figure 1. Mass balance in the wet anaerobic digestion process (own elaborated based on [70]).
One biogas plant analyzed by Luning et al. [70] used a wet substrate separation pro-
cess to eliminate sand and other materials that adversely aect the process. The process
uses the phenomenon of sedimentation. The substrate is then mixed and directed to the
digester. The nal stage is dewatering the digestion residue using belt presses. It should
be noted that dewatering the digestate from the wet anaerobic digestion process results in
Figure 1. Mass balance in the wet anaerobic digestion process (own elaborated based on [70]).
One biogas plant analyzed by Luning et al. [
70
] used a wet substrate separation process
to eliminate sand and other materials that adversely affect the process. The process uses
the phenomenon of sedimentation. The substrate is then mixed and directed to the digester.
The final stage is dewatering the digestion residue using belt presses. It should be noted
that dewatering the digestate from the wet anaerobic digestion process results in about
40% wastewater and 42% solid fraction. This waste should be managed appropriately. The
costs associated with the disposal of solid fraction waste are significantly higher than the
disposal of process wastewater. This aspect has a significant impact on the profitability of
the biogas plant and the resulting water footprint. In the case of dry anaerobic digestion,
the final mass balance is in the ratio of 10% wastewater and 77% solid fraction digestate.
Fermentation 2023,9, 449 7 of 15
This ratio facilitates the introduction of technology for drying the solid fraction and using
it as an energy fuel [69].
3.2. Dry Anaerobic Digestion
The dry anaerobic digestion method allows for the processing of various types of
the feedstock of varying compositions, such as manure or food waste [
11
]. Using such
substrates without prior dilution and preparation is impossible in wet technology, as they
exceed standard equipment’s ‘pumpability’ limit. The finished batch substrate mixture
in the reactor contains a high amount of dry matter at levels ranging between 20 and
40% [
69
]. Large-scale biogas production using dry anaerobic digestion technology is
currently considered a fledgling method. This may be due to the lack of prevalence among
most industrial biogas plants [
11
,
70
]. However, differences in process parameters and the
feedstock used and the bacteria for dry and wet conditions are similar [71].
A system based on dry anaerobic digestion is devoid of the drawbacks of a wet
digestion process. It is possible to conduct anaerobic digestion in a discontinuous process,
and no additional energy inputs are required for mixing and grinding the feedstock [
72
].
The dry process prevents the formation of foam and crusts on the surface of the digester
and also counteracts the phenomenon of sedimentation [
73
]. The benefit of dry technology
is the processing of waste in its original form. In addition, a process of this type does not
require additional diluent liquid and, according to Jha et al. [
74
], is capable of producing
a higher volume of methane per m
3
of bioreactor volume [
74
]. A 540–750 NL kg VS
biogas efficiency was obtained for the substrate studied using the wet anaerobic digestion
process [
75
]. Comparing methane efficiency for the dry process with the wet process in the
literature has been analyzed at the laboratory scale [
74
] and at the large scale of industrial
biogas plants [
70
]. The authors of the paper [
75
] established, based on their research, the
higher methane efficiency of the dry-fermentation process of grass compared to the dry-wet
method. A 420–540 NL kg VS biogas efficiency was obtained for the substrate studied
using the dry anaerobic digestion process [
75
]. A study by Angelonidi and Smith [
66
]
compared nine plants using dry and wet technologies for anaerobic digestion processes.
For the dry method, a greater flexibility related to the type of feedstock used, reduced water
consumption, shorter retention times, and an easier management of the end products of
the process were evaluated. In contrast, the wet technology was characterized by a more
favorable final energy balance [69].
In recent years, Europe has seen progress related to dry anaerobic digestion efficiency
by up to 50%, although methane anaerobic digestion is responsible for managing only 35%
of all waste [
76
]. Additionally, on other continents, e.g., in China, interest in dry anaerobic
digestion has increased, which is an ideal solution for managing the
0.9 billion tons
of
straw generated annually in China. Straw is a problematic feedstock for anaerobic di-
gestion, as there are difficulties with its breakdown by bacteria, so pretreatment may be
necessary [
77
]. The use of dry anaerobic digestion technology allows for treating wastes
containing high values of organic matter in their composition, such as municipal solid
waste and agricultural waste [
76
]. By analyzing the water footprint and substrate hydration
for dry digestion technology, it was estimated that it requires four to ten times less water
than wet processes [
78
]. This fact also has the benefit of reducing the volume of the digester
and the savings associated with substrate pretreatment processes, as it only requires the re-
moval of very large particles—larger than 5 cm [
79
]. Challenges posed by the dry anaerobic
digestion process relate to generating a homogeneous mixture and transportation. The use
of equipment for transporting solid substrates is more expensive than for liquid substrates.
It is possible to design and build pumps, belts, or screw conveyors transporting substances
with a high dry matter value [
80
]. The authors of the paper [
81
] show the problem associ-
ated with the inhibition of the dry fermentation process when the concentration of total
solids is above 40%. At such a high level of total solids, methane bacteria have difficulty
decomposing substrates. An excessively high amount of total solids can also cause plant
heating problems [78,81].
Fermentation 2023,9, 449 8 of 15
4. Substrate Pretreatment as Key to Increasing Anaerobic Digestion Process Efficiency
Biogas production technology is constantly optimized for biological processes and
chemical reactions to increase productivity and reduce operating costs [
82
]. It is theoretically
possible to carry out the methane anaerobic digestion process for the organic fraction of
any substrate. Depending on availability, it is possible to process wood, crop residue,
lignocellulosic waste, food waste, or cotton stalks [
19
]. Some feedstocks are not naturally
adapted to anaerobic digestion, as they are characterized by complicated or very long
decomposition by bacteria. In addition, some of the raw materials may also contain natural
inhibitors of the process. Therefore, it is becoming necessary to develop technology related
to the pretreatment of substrates allowing the management of difficult-to-degrade wastes,
which can contribute to increasing the potential of biogas facilities worldwide [83,84].
Most raw materials inherently characterized by susceptibility to fermentation, such as
slurry and sewage sludge, are used. This type of waste, too, often requires pretreatment
processes due to excessively high levels of hydration, which reduces methane efficiency.
Using appropriate pretreatment technology, it is possible to increase the availability of valu-
able biogas substrates while helping to reduce environmental pollution [
83
,
85
]. Substrates
containing large amounts of cellulose (40–60%) and hemicellulose (20–40%) after pretreat-
ment have high biogas potential [
86
]. Waste containing lignocellulosic compounds is one
of the most considerable organic resources in the world, providing an average of about
200 billion tons
per year [
87
]. Organic wastes, such as bird feathers, contain more than 90%
creatine in their content [
88
]. The use of pretreatment allows creatine to be broken down
into oligomers, which are easily submitted to anaerobic fermentation [
89
]. Waste from fruit
and food processing are also substrates with high methane potential after pretreatment [
83
].
In recent years, research has been conducted to manage waste polyethylene terephthalate
(PET). Yoshida et al. [
89
] found that a new type of bacteria Ideonella sakaiensis 201-F6 can
degrade PET compounds, but further research is needed to expand the scale and possibili-
ties of this type of technology [
89
]. An innovative option for converting problematic waste
is gasification technology before anaerobic digestion. Gasification converts the substrate
into synthesis gas, which is directed into the environment with anaerobic bacteria, thus
producing biogas [90]. Combining the two types of technology brings some benefits [83].
Substrate pretreatment processes are designed to improve availability and facilitate
anaerobic microorganisms’ utilization of organic matter. Several pretreatment types are
classified as chemical, biological, mechanical, and thermal, respectively [
91
]. In addition,
substrates are often prepared by removing solid contaminants such as sand and metals,
which can be found in municipal waste. In addition, pretreatment allows for grinding
the feedstock or even removing process inhibitors such as oils. An important aspect also
concerns the compaction of substrates before they are delivered to the biogas plant site.
This procedure is an opportunity that allows for reductions related to greenhouse gas
emissions. The reason for the unfavorable water and carbon footprint of biogas plants is
the necessity to transport substrate characterized by low biogas efficiency and then manage
the highly hydrated digestate pulp [
83
,
92
,
93
]. Similarly, in the case of chemical methods,
various types of chemical compounds, such as strong acids and alkalis, are used. The acidic
chemical treatment process allows lignocellulosic compounds.
Analyzing the aspect of the application of storage solution leads to significant cost
reduction at the storage and handling stage of the biomass supply chain, which results in
considerable cost savings for the whole biomass logistics function. This reduction exceeds
by far the extra cost imposed by biomass material losses. However, side-effects of applying
cheaper storage solutions without biomass drying, such as a heating value reduction in the
biomass, health and fire risks, etc., should be further investigated [94,95].
4.1. Chemical and Biological Pretreatment
The use of the pretreatment of substrates by biological and chemical methods is rela-
tively the same. In the case of chemical methods, various types of chemical compounds,
such as strong acids and alkalis, are used. The acidic chemical treatment process allows lig-
Fermentation 2023,9, 449 9 of 15
nocellulosic compounds to be broken down into simpler monosaccharides [
96
]. In addition,
the acidic reaction of this type of treatment can be controlled by the presence of hydrolytic
bacteria [
97
]. High costs resulting from the pretreatment of substrates by acidic methods
limit the development of this technology on the industrial scale of anaerobic digestion
processes [
98
]. Another type of chemical method is alkaline compounds such as hydroxides,
which allow them to be used at ambient temperatures [
99
]. The anaerobic digestion process
may require the addition of alkaline reagents to balance the pH, and alkaline pretreatment
is a preferred method over acidic pretreatment [
100
]. Research by Liew et al. [
101
] showed
that the pretreatment of leaves using a 3.5% NaOH solution allows for a 20% increase in
methane efficiency within three days. This type of treatment contributes to an increased
anaerobic digestion efficiency for substrates containing lignocellulose compounds. Con-
tinuous anaerobic digestion with substrates after alkaline pretreatment can lead to salt
accumulation and pH increase, thus contributing to the inhibition of the methanation
process [101,102]. Chemical treatment processes are also not widely used on an industrial
scale due to the lack of economic viability currently resulting from the high cost of alkali.
However, this type of treatment may be the only option for highly acidic or lignin-rich
substrates to ensure a stable anaerobic digestion process [103].
Biological pretreatment of substrates can be carried out at a low temperature and
without additional chemicals. It includes processes under aerobic and anaerobic conditions,
but this type of substrate preparation is not used for municipal waste [
96
]. The treatment
conditions of anaerobic technology are often used as a pre-acidification or dark anaerobic di-
gestion. This technology makes it possible to separate the methane generation process from
hydrolysis and acid production. The pH value for the pre-acidification stage is between
4 and 6, which inhibits the methane generation process and causes volatile fatty acids to
accumulate. In a study by Liu et al. [
104
], a 21% higher methane efficiency was achieved by
introducing a pre-acidification stage for household waste. In addition, it resulted in the ben-
efit of a higher methane concentration in biogas [
103
,
104
]. Compared to other pretreatment
methods, biological methods require a more extended period of substrate preparation while
having low efficiency [
105
]. Research conducted by
Mshandete et al.
[
106
] also presents
pretreatment methods involving aeration. Their tests showed an increased methane pro-
duction efficiency of 26% for a 9 h treatment stage. Prolongation of the assumed aeration
stage leads to the aerobic breakdown of the substrate and reduces the amount of methane
produced [106].
4.2. Mechanical and Thermal Pretreatment
Pretreatment of substrates using mechanical and thermal methods is categorized as a
physical process. The rationale for using mechanical pretreatment is to reduce the particle
size of substrates while increasing the surface area of bacterial activity [
18
]. In addition,
grinding the substrates facilitates the mixing process in the digesters and reduces the
problem associated with dross formation. Most industrial biogas plants use mechanical
processing of substrates, allowing for uniformity of the feedstock mixture [
101
]. Nah
et al. [
107
] conducted research related to substrate pretreatment using high-pressure particle
breaking. This process reduced the hydraulic retention time of sludge from 13 days to
6 days
while maintaining the same process efficiency. This research was conducted only on
a laboratory scale, and the final stage did not test the methane content of the biogas [
107
].
Fine grinding of waste containing lignocellulosic compounds increases the anaerobic
digestion process’s efficiency, but too fine particles can cause acidification because they
dissolve quickly [
96
]. In addition, the disadvantages of mechanical treatment processes
include the possibility of damage from inert materials such as stones [
96
]. A number of
researchers have studied the effect of knife milling on biogas production. Menind and
Normak [
108
] found that about a 10% higher gas yield was obtained when hay was ground
to 0.5 mm compared to 20–30 mm. A different investigation found that grinding sisal fibers
from 100 mm to 2 mm produced about 20–25% higher gas yields [109].
Fermentation 2023,9, 449 10 of 15
The thermal pretreatment process is based on processing substrates under high tem-
peratures (typically 125 to 190
C) and pressure while maintaining the set conditions for
about an hour [
96
]. Biogas plants using substrates after thermal pretreatment can use an
increased reactor load factor and conduct a more stable anaerobic digestion process [
108
].
A crucial factor during thermal treatment is selecting an appropriate temperature value.
If the value is too high, feedstock substrates can be degraded. In addition, conducting
the thermal process allows the destruction of pathogenic pathogens contained in some
wastes [
96
,
110
]. An example of the thermal pretreatment of feedstock on an industrial
scale is the thermo-druck-hydrolyse (TDH) process. The technology has been used for
diluting to about 10–15% dry weight of kitchen waste and plastics. The TDH reactor is
pressurized for about 20 min at 20 to 30 bar and 170–200
C. It was found that this type of
technology increased biogas production by 20 to 30% for the substrates analyzed. Thermal
pretreatment plays a unique role in locations with a free supply of waste heat from plants
such as factories or power plants [
96
]. Table 2shows the advantages effect of pretreatment
on the fermentation process from selected methods.
Table 2. Effect of pretreatment on the fermentation process (own elaboration based on [94107]).
Pretreatment Process Advantages Reference
Chemical decomposition of lignocellulose
only solution for highly acidic substrate [98,101,102]
Biological
Mechanical
used as pre-acidification
low temperature
increasing the surface area for bacterial activity
[96,104]
[94,97,105]
Thermal more stable anaerobic digestion process [94,95,107]
5. Conclusions and Directions for Further Research
In addition to managing waste, biogas plants also process significant amounts of ballast
water in anaerobic digestion. The agricultural sector and related agricultural production
are the largest consumers of water resources. The development and exploitation of biogas
plants that use agricultural residues optimize aspects of the water footprint. In addition,
waste biomass is generally characterized by high hydration, so pretreatment methods are
required. Removing excess moisture even before transporting the substrates to the biogas
plant site is good practice.
Recently, the situation related to protecting the world’s water resources is also not
insignificant. Creating assumptions in the context of water limits in the future can help
avoid the problem of unsustainable water consumption. Numerous studies are being
conducted to determine the water footprint for products and industrial plants, including
biogas plants and their substrates. It is becoming necessary to look for solutions related
to the low efficiency of the anaerobic digestion process when over-hydrated substrates
are used. Due to the multitude of available technologies related to the pretreatment of
substrates and for the direct execution of the anaerobic digestion process, further research
needs to optimize the discussed solutions. It should also be emphasized that pretreatment
processes are often an expensive and energy-intensive technology. The most common
method used to pretreat substrates is a mechanical one. Reducing the particle size of
substrates results in a greater surface accessibility for microorganisms.
Transportation distance is essential in the production and delivery of substrate for a
biogas plant. Depending on the conditions of the selected location for a newly built biogas
plant, the impact of harmful greenhouse gas emissions associated with transportation
should be considered. It is proposed to initiate new research on transportation elements and
biogas substrate management. The use of modern technologies related to the continuous
analysis of selected substrate parameters and control of delivery processes can allow for a
breakthrough optimization of the biogas sector in Poland and the world.
Fermentation 2023,9, 449 11 of 15
Author Contributions:
Conceptualization, M.N. and W.C.; methodology, M.N. and W.C.; software,
M.N.; validation, M.N. and W.C.; formal analysis, M.N. and W.C.; investigation, M.N. and W.C.;
resources, M.N. and W.C.; data curation, M.N. and W.C.; writing—original draft preparation, M.N.,
W.B. and W.C.; writing—review and editing, M.N. and W.C.; visualization, M.N.; supervision, W.C.;
funding acquisition, M.N. and W.C. All authors have read and agreed to the published version of
the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest:
The authors declare no conflict of interest. The funders 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.
References
1.
Sanson, A.V.; Van Hoorn, J.; Burke, S.E.L. Responding to the Impacts of the Climate Crisis on Children and Youth. Child Dev.
Perspect. 2019,13, 201–207. [CrossRef]
2.
Lamb, W.F.; Wiedmann, T.; Pongratz, J.; Andrew, R.; Crippa, M.; Olivier, J.G.J.; Wiedenhofer, D.; Mattioli, G.; Khourdajie, A.A.;
House, J.; et al. A review of trends and drivers of greenhouse gas emissions by sector from 1990 to 2018. Environ. Res. Lett.
2021
,
16, 073005. [CrossRef]
3.
Ahmad, S.; Razzaq, S.; Rehan, M.A.; Hashmi, M.A.; Ali, S.; Amjad, M.S.; Mehmood, U. Experimental Investigation of Novel
Fixed Dome Type Biogas Plant using Gas Recovery Chamber in Rural Areas of Pakistan. Int. J. Renew. Energy Res. IJRER
2019
,9,
1537–1547.
4.
Patinvoh, R.J.; Taherzadeh, M.J. Challenges of biogas implementation in developing countries. Curr. Opin. Environ. Sci. Health
2019,12, 30–37. [CrossRef]
5.
Jarrar, L.; Ayadi, O.; Al Asfar, J. Techno-economic Aspects of Electricity Generation from a Farm Based Biogas Plant. J. Sustain.
Dev. Energy Water Environ. Syst. 2020,8, 476–492. [CrossRef]
6.
Adeleke, A.; Odusote, J.; Lasode, O.; Ikubanni, P.; Madhurai, M.; Paswan, D. Evaluation of thermal decomposition characteristics
and kinetic parameters of melina wood. Biofuels 2022,13, 117–123. [CrossRef]
7.
Ibikunle, R.A.; Titiladunayo, I.F.; Lukman, A.F.; Dahunsi, S.O.; Akeju, E.A. Municipal solid waste sampling, quantification and
seasonal characterization for power evaluation: Energy potential and statistical modelling. Fuel 2020,277, 118122. [CrossRef]
8.
Werghemmi, W.; Fayssal, S.A.; Mazouz, H.; Hajjaj, H.; Hajji, L. Olive and green tea leaves extract in Pleurotus ostreatus var. florida
culture media: Effect on mycelial linear growth rate, diameter and growth induction index. IOP Conf. Ser. Earth Environ. Sci.
2022
,
1090, 012020. [CrossRef]
9.
Elbagory, M.; El-Nahrawy, S.; Omara, A.E.-D.; Eid, E.M.; Bachheti, A.; Kumar, P.; Abou Fayssal, S.; Adelodun, B.; Bachheti, R.K.;
Kumar, P.; et al. Sustainable Bioconversion of Wetland Plant Biomass for Pleurotus ostreatus var. florida Cultivation: Studies on
Proximate and Biochemical Characterization. Agriculture 2022,12, 2095. [CrossRef]
10.
Czekała, W.; Janczak, D.; Pochwatka, P.; Nowak, M.; Dach, J. Gases Emissions during Composting Process of Agri-Food Industry
Waste. Appl. Sci. 2022,12, 9245. [CrossRef]
11.
Jansson, A.T.; Patinvoh, R.J.; Sárvári Horváth, I.; Taherzadeh, M.J. Dry Anaerobic Digestion of Food and Paper Industry Wastes at
Different Solid Contents. Fermentation 2019,5, 40. [CrossRef]
12.
Adeleke, A.A.; Odusote, J.K.; Ikubanni, P.P.; Orhadahwe, T.A.; Lasode, O.A.; Ammasi, A.; Kumar, K. Ash analyses of bio-coal
briquettes produced using blended binder. Sci. Rep. 2021,11, 547. [CrossRef] [PubMed]
13.
Liu, H.; Kumar, V.; Jia, L.; Sarsaiya, S.; Kumar, D.; Juneja, A.; Zhang, Z.; Sindhu, R.; Binod, P.; Bhatia, S.K.; et al. Biopolymer
poly-hydroxyalkanoates (PHA) production from apple industrial waste residues: A review. Chemosphere
2021
,284, 131427.
[CrossRef] [PubMed]
14.
World Energy Outlook 2020—Analysis. Available online: https://www.iea.org/reports/world-energy-outlook-2020
(accessed on 1 February 2023).
15.
Aberilla, J.M.; Gallego-Schmid, A.; Azapagic, A. Environmental sustainability of small-scale biomass power technologies for
agricultural communities in developing countries. Renew. Energy 2019,141, 493–506. [CrossRef]
16.
Anal, A.K. Quality Ingredients and Safety Concerns for Traditional Fermented Foods and Beverages from Asia: A Review.
Fermentation 2019,5, 8. [CrossRef]
17.
Gerbens-Leenes, W.; Vaca-Jiménez, S.; Mekonnen, M. Burning Water, Overview of the Contribution of Arjen Hoekstra to the
Water Energy Nexus. Water 2020,12, 2844. [CrossRef]
18.
Irfan, M.; Zhao, Z.-Y.; Panjwani, M.K.; Mangi, F.H.; Li, H.; Jan, A.; Ahmad, M.; Rehman, A. Assessing the energy dynamics of
Pakistan: Prospects of biomass energy. Energy Rep. 2020,6, 80–93. [CrossRef]
Fermentation 2023,9, 449 12 of 15
19.
Branco, R.H.R.; Serafim, L.S.; Xavier, A.M.R.B. Second Generation Bioethanol Production: On the Use of Pulp and Paper Industry
Wastes as Feedstock. Fermentation 2019,5, 4. [CrossRef]
20.
Robak, K.; Balcerek, M. Review of Second Generation Bioethanol Production from Residual Biomass. Food Technol. Biotechnol.
2018,56, 174–187. [CrossRef]
21.
Chowdhury, H.; Loganathan, B. Third-generation biofuels from microalgae: A review. Curr. Opin. Green Sustain. Chem.
2019
,20,
39–44. [CrossRef]
22. Nugent, N.; Rhinard, M. The ‘political’ roles of the European Commission. J. Eur. Integr. 2019,41, 203–220. [CrossRef]
23.
Khan, M.U.; Lee, J.T.E.; Bashir, M.A.; Dissanayake, P.D.; Ok, Y.S.; Tong, Y.W.; Shariati, M.A.; Wu, S.; Ahring, B.K. Current status of
biogas upgrading for direct biomethane use: A review. Renew. Sustain. Energy Rev. 2021,149, 111343. [CrossRef]
24. Wielgosinski, G.; Czerwi´nska, J. Smog Episodes in Poland. Atmosphere 2020,11, 277. [CrossRef]
25.
Kozłowski, K.; Pietrzykowski, M.; Czekała, W.; Dach, J.; Kowalczyk-Ju´sko, A.; Jó´zwiakowski, K.; Brzoski, M. Energetic and
economic analysis of biogas plant with using the dairy industry waste. Energy 2019,183, 1023–1031. [CrossRef]
26.
Theuerl, S.; Herrmann, C.; Heiermann, M.; Grundmann, P.; Landwehr, N.; Kreidenweis, U.; Prochnow, A. The Future Agricultural
Biogas Plant in Germany: A Vision. Energies 2019,12, 396. [CrossRef]
27.
Kumar, P.; Eid, E.M.; Taher, M.A.; El-Morsy, M.H.E.; Osman, H.E.M.; Al-Bakre, D.A.; Adelodun, B.; Abou Fayssal, S.; Goala, M.;
Mioˇc, B.; et al. Biotransforming the Spent Substrate of Shiitake Mushroom (Lentinula edodes Berk.): A Synergistic Approach to
Biogas Production and Tomato (Solanum lycopersicum L.) Fertilization. Horticulturae 2022,8, 479. [CrossRef]
28. Czekała, W. Digestate as a Source of Nutrients: Nitrogen and Its Fractions. Water 2022,14, 4067. [CrossRef]
29.
Mathioudakis, V.; Gerbens-Leenes, P.W.; Van der Meer, T.H.; Hoekstra, A.Y. The water footprint of second-generation bioenergy:
A comparison of biomass feedstocks and conversion techniques. J. Clean. Prod. 2017,148, 571–582. [CrossRef]
30.
Czekała, W.; Lewicki, A.; Pochwatka, P.; Czekała, A.; Wojcieszak, D.; Jó´zwiakowski, K.; Waliszewska, H. Digestate management
in Polish farms as an element of the nutrient cycle. J. Clean. Prod. 2020,242, 118454. [CrossRef]
31.
Mudryk, K.; Fr ˛aczek, J.; Jewiarz, M.; Wróbel, M.; Dziedzic, K. Analysis of Mechanical Dewatering of Digestate. Agric. Eng.
2016
,
20, 157–166. [CrossRef]
32.
Yuan, Z.L.; Gerbens-Leenes, P.W. Biogas feedstock potentials and related water footprints from residues in China and the
European Union. Sci. Total Environ. 2021,793, 148340. [CrossRef] [PubMed]
33. Mekonnen, M.M.; Gerbens-Leenes, W. The Water Footprint of Global Food Production. Water 2020,12, 2696. [CrossRef]
34.
IRENA. Renewable Energy Prospects: China, REmap 2030 Analysis; IRENA: Abu Dhabi, UAE, 2014; Available online: http:
//www.irena.org/remap/IRENA_REmap_China_report_2014.pdf (accessed on 10 March 2023).
35.
Langeveld, H.; Kalf, R.; Elbersen, H.W. Bioenergy production chain development in the Netherlands: Key factors for success.
Biofuels Bioprod. Biorefining 2010,4, 484–493. [CrossRef]
36.
Li, Y.; Chen, Y.; Wu, J. Enhancement of methane production in anaerobic digestion process: A review. Appl. Energy
2019
,240,
120–137. [CrossRef]
37.
Mikhaylov, A.; Moiseev, N.; Aleshin, K.; Burkhardt, T. Global climate change and greenhouse effect. Entrep. Sustain. Issues
2020
,7,
2897–2913. [CrossRef]
38.
Abbas, I.; Liu, J.; Noor, R.S.; Faheem, M.; Farhan, M.; Ameen, M.; Shaikh, S.A. Development and performance evaluation of small
size household portable biogas plant for domestic use. Biomass Convers. Biorefinery 2022,12, 3107–3119. [CrossRef]
39.
Ellacuriaga, M.; García-Cascallana, J.; Gómez, X. Biogas Production from Organic Wastes: Integrating Concepts of Circular
Economy. Fuels 2021,2, 144–167. [CrossRef]
40.
Agabo-García, C.; Pérez, M.; Rodríguez-Morgado, B.; Parrado, J.; Solera, R. Biomethane production improvement by enzymatic
pre-treatments and enhancers of sewage sludge anaerobic digestion. Fuel 2019,255, 115713. [CrossRef]
41.
Mao, C.; Wang, Y.; Wang, X.; Ren, G.; Yuan, L.; Feng, Y. Correlations between microbial community and C:N:P stoichiometry
during the anaerobic digestion process. Energy 2019,174, 687–695. [CrossRef]
42.
Kushkevych, I.; Víezová, M.; Vít ˇez, T.; Kováˇc, J.; Kaucká, P.; Jesionek, W.; Bartoš, M.; Barton, L. A new combination of substrates:
Biogas production and diversity of the methanogenic microorganisms. Open Life Sci. 2018,13, 119–128. [CrossRef]
43.
Schnürer, A. Biogas Production: Microbiology and Technology. Adv. Biochem. Eng. Biotechnol.
2016
,156, 195–234. [CrossRef]
[PubMed]
44.
McGenity, T.J.; Timmis, K.N.; Fernández, B.N. Hydrocarbon and Lipid Microbiology Protocols; Springer: Berlin/Heidelberg, Germany, 2016.
45.
Nwokolo, N.; Mukumba, P.; Obileke, K.; Enebe, M. Waste to Energy: A Focus on the Impact of Substrate Type in Biogas Production.
Processes 2020,8, 1224. [CrossRef]
46.
Rasit, N.; Idris, A.; Harun, R.; Wan Ab Karim Ghani, W.A. Effects of lipid inhibition on biogas production of anaerobic digestion
from oily effluents and sludges: An overview. Renew. Sustain. Energy Rev. 2015,45, 351–358. [CrossRef]
47.
Rajagopal, R.; Massé, D.I.; Singh, G. A critical review on inhibition of anaerobic digestion process by excess ammonia. Bioresour.
Technol. 2013,143, 632–641. [CrossRef]
48.
Westerholm, M.; Moestedt, J.; Schnürer, A. Biogas production through syntrophic acetate oxidation and deliberate operating
strategies for improved digester performance. Appl. Energy 2016,179, 124–135. [CrossRef]
49.
Solowski, G.; Konkol, I.; Cenian, A. Production of hydrogen and methane from lignocellulose waste by fermentation. A review of
chemical pre-treatment for enhancing the efficiency of the digestion process. J. Clean. Prod. 2020,267, 121721. [CrossRef]
Fermentation 2023,9, 449 13 of 15
50.
Ullah Khan, I.; Hafiz Dzarfan Othman, M.; Hashim, H.; Matsuura, T.; Ismail, A.F.; Rezaei-DashtArzhandi, M.; Wan Azelee, I.
Biogas as a renewable energy fuel—A review of biogas upgrading, utilization and storage. Energy Convers. Manag.
2017
,150,
277–294. [CrossRef]
51.
Perez-Esteban, N.; Vinardell, S.; Vidal-Antich, C.; Peña-Picola, S.; Chimenos, J.M.; Peces, M.; Dosta, J.; Astals, S. Potential of
anaerobic co-fermentation in wastewater treatment plants: A review. Sci. Total Environ. 2022,813, 152498. [CrossRef]
52.
Abubakar, A.M.; Yunus, M.U. Reporting Biogas Data from Various Feedstock. Int. J. Form. Sci. Curr. Future Res. Trends
2021
,11,
23–36.
53.
Bacenetti, J.; Lovarelli, D.; Ingrao, C.; Tricase, C.; Negri, M.; Fiala, M. Assessment of the influence of energy density and feedstock
transport distance on the environmental performance of methane from maize silages. Bioresour. Technol.
2015
,193, 256–265.
[CrossRef]
54.
Huopana, T.; Song, H.; Kolehmainen, M.; Niska, H. A regional model for sustainable biogas electricity production: A case study
from a Finnish province. Appl. Energy 2013,102, 676–686. [CrossRef]
55.
Bochmann, G.; Montgomery, L.F.R. 4—Storage and pre-treatment of substrates for biogas production. In The Biogas Handbook;
Wellinger, A., Murphy, J., Baxter, D., Eds.; Woodhead Publishing Series in Energy; Woodhead Publishing: Cambridge, UK, 2013;
pp. 85–103. ISBN 978-0-85709-498-8. [CrossRef]
56.
Tucki, K.; Piatkowski, P.; Wojcik, G. Selected aspects from the analysis of the agricultural biogas plants sector in Poland. Energy
Mark. 2016,122, 54–58.
57. May, G.; Piekarski, W.; Kowalczyk-Ju´sko, A. Logistics of supplying raw material to an agricultural biogas plant. Logistics 2014.
58.
Jolliet, O.; Margni, M.; Charles, R.; Humbert, S.; Payet, J.; Rebitzer, G.; Rosenbaum, R. IMPACT 2002+: A New Life Cycle Impact
Assessment Methodology. Int. J. Life Cycle Assess. 2003,8, 324–330. [CrossRef]
59. Muradin, M.; Foltynowicz, Z. Logistic aspects of the ecological impact indicators of an agricultural biogas plant. Logforum 2018,
14, 535–547. [CrossRef]
60.
Dach, J.; Pulka, J.; Janczak, D.; Lewicki, A.; Pochwatka, P.; Oniszczuk, T. Energetic Assessment of Biogas Plant Projects Based on
Biowaste and Maize Silage Usage. In IOP Conference Series: Earth and Environmental Science; IOP Publishing: Bristol, UK, 2020;
Volume 505, p. 012029.
61.
Borek, K.; Romaniuk, W. Biogas Installations for Harvesting Energy and Utilization of Natural Fertilisers. Agric. Eng.
2020
,24,
1–14. [CrossRef]
62.
Stanbury, P.F.; Whitaker, A.; Hall, S.J. Principles of Fermentation Technology; Elsevier: Amsterdam, The Netherlands, 2013;
ISBN 978-1-4832-9291-5.
63. Stürmer, B. Feedstock change at biogas plants–Impact on production costs. Biomass Bioenergy 2017,98, 228–235. [CrossRef]
64.
Kory´s, K.A.; Latawiec, A.E.; Grotkiewicz, K.; Kubo ´n, M. The Review of Biomass Potential for Agricultural Biogas Production in
Poland. Sustainability 2019,11, 6515. [CrossRef]
65.
Nordberg, A.; Jarvis, A.; Stenberg, B.; Mathisen, B.; Svensson, B.H. Anaerobic digestion of alfalfa silage with recirculation of
process liquid. Bioresour. Technol. 2007,98, 104–111. [CrossRef]
66.
Adekunle, K.F.; Okolie, J.A. A Review of Biochemical Process of Anaerobic Digestion. Adv. Biosci. Biotechnol.
2015
,06, 205.
[CrossRef]
67.
Shahriari, H.; Warith, M.; Hamoda, M.; Kennedy, K.J. Effect of leachate recirculation on mesophilic anaerobic digestion of food
waste. Waste Manag. 2012,32, 400–403. [CrossRef] [PubMed]
68.
Weiland, P. Biogas production: Current state and perspectives. Appl. Microbiol. Biotechnol.
2010
,85, 849–860. [CrossRef] [PubMed]
69.
Angelonidi, E.; Smith, S. A comparison of wet and dry anaerobic digestion processes for the treatment of municipal solid waste
and food waste. Water Environ. J. 2015,29, 549–557. [CrossRef]
70.
Luning, L.; van Zundert, E.H.M.; Brinkmann, A.J.F. Comparison of dry and wet digestion for solid waste. Water Sci. Technol.
2003
,
48, 15–20. [CrossRef] [PubMed]
71.
Stolze, Y.; Zakrzewski, M.; Maus, I.; Eikmeyer, F.; Jaenicke, S.; Rottmann, N.; Siebner, C.; Pühler, A.; Schlüter, A. Comparative
metagenomics of biogas-producing microbial communities from production-scale biogas plants operating under wet or dry
fermentation conditions. Biotechnol. Biofuels 2015,8, 14. [CrossRef] [PubMed]
72.
Czubaszek, R.; Wysocka-Czubaszek, A.; Banaszuk, P. GHG Emissions and Efficiency of Energy Generation through Anaerobic
Fermentation of Wetland Biomass. Energies 2020,13, 6497. [CrossRef]
73.
Abdelsalam, E.M.; Samer, M.; Amer, M.A.; Amer, B.M.A. Biogas production using dry fermentation technology through
co-digestion of manure and agricultural wastes. Environ. Dev. Sustain. 2021,23, 8746–8757. [CrossRef]
74.
Jha, A.K.; Li, J.; Zhang, L.; Ban, Q.; Jin, Y. Comparison between Dry and Wet Anaerobic Digestions of Cow Dung under mesophilic
and thermophilic conditions. Adv. Water Resour. Prot 2013,1, 28–38.
75.
Vogel, T.; Ahlhaus, M.; Barz, M. Optimization of the biogas production from grass by dry-wet fermentation. In Proceedings of the
Eighth International Scientific Conference—Engineering for Rural Development, Jelgava, Latvia, 28–29 May 2009; LLU: Loma
Linda, CA, USA, 2009.
76.
Rocamora, I.; Wagland, S.T.; Villa, R.; Simpson, E.W.; Fernández, O.; Bajón-Fernández, Y. Dry anaerobic digestion of organic waste:
A review of operational parameters and their impact on process performance. Bioresour. Technol. 2020,299, 122681. [CrossRef]
77.
Yu, Q.; Liu, R.; Li, K.; Ma, R. A review of crop straw pre-treatment methods for biogas production by anaerobic digestion in
China. Renew. Sustain. Energy Rev. 2019,107, 51–58. [CrossRef]
Fermentation 2023,9, 449 14 of 15
78.
Shapovalov, Y.; Zhadan, S.; Bochmann, G.; Salyuk, A.; Nykyforov, V. Dry Anaerobic Digestion of Chicken Manure: A Review.
Appl. Sci. 2020,10, 7825. [CrossRef]
79.
Rapport, J.L.; Zhang, R.; Williams, R.B.; Jenkins, B.M. Anaerobic Digestion technologies for the treatment of Municipal Solid
Waste. Int. J. Environ. Waste Manag. 2012,9, 100–122. [CrossRef]
80. Mata-Alvarez, J. Biomethanization of the Organic Fraction of Municipal Solid Wastes; IWA Publishing: Barcelona, Spain, 2003.
81.
Kothari, R.; Pandey, A.K.; Kumar, S.; Tyagi, V.V.; Tyagi, S.K. Different aspects of dry anaerobic digestion for bio-energy: An
overview. Renew. Sustain. Energy Rev. 2014,39, 174–195. [CrossRef]
82.
Anukam, A.; Mohammadi, A.; Naqvi, M.; Granström, K. A Review of the Chemistry of Anaerobic Digestion: Methods of
Accelerating and Optimizing Process Efficiency. Processes 2019,7, 504. [CrossRef]
83.
Patinvoh, R.J.; Osadolor, O.A.; Chandolias, K.; Sárvári Horváth, I.; Taherzadeh, M.J. Innovative pre-treatment strategies for biogas
production. Bioresour. Technol. 2017,224, 13–24. [CrossRef]
84.
Maneein, S.; Milledge, J.J.; Nielsen, B.V.; Harvey, P.J. A Review of Seaweed pre-Treatment Methods for Enhanced Biofuel
Production by Anaerobic Digestion or Fermentation. Fermentation 2018,4, 100. [CrossRef]
85.
Cao, B.; Zhang, T.; Zhang, W.; Wang, D. Enhanced technology based for sewage sludge deep dewatering: A critical review. Water
Res. 2021,189, 116650. [CrossRef] [PubMed]
86.
Kang, Q.; Appels, L.; Tan, T.; Dewil, R. Bioethanol from Lignocellulosic Biomass: Current Findings Determine Research Priorities.
Sci. World J. 2014,2014, e298153. [CrossRef]
87.
Zhang, Y.-H.P. Reviving the carbohydrate economy via multi-product lignocellulose biorefineries. J. Ind. Microbiol. Biotechnol.
2008,35, 367–375. [CrossRef]
88.
Patinvoh, R.J.; Feuk-Lagerstedt, E.; Lundin, M.; Sárvári Horváth, I.; Taherzadeh, M.J. Biological pre-treatment of Chicken Feather
and Biogas Production from Total Broth. Appl. Biochem. Biotechnol. 2016,180, 1401–1415. [CrossRef]
89.
Yoshida, S.; Hiraga, K.; Takehana, T.; Taniguchi, I.; Yamaji, H.; Maeda, Y.; Toyohara, K.; Miyamoto, K.; Kimura, Y.; Oda, K. A
bacterium that degrades and assimilates poly(ethylene terephthalate). Science 2016,351, 1196–1199. [CrossRef] [PubMed]
90.
Phillips, J.R.; Huhnke, R.L.; Atiyeh, H.K. Syngas Fermentation: A Microbial Conversion Process of Gaseous Substrates to Various
Products. Fermentation 2017,3, 28. [CrossRef]
91.
Taherzadeh, M.J.; Karimi, K. Pretreatment of Lignocellulosic Wastes to Improve Ethanol and Biogas Production: A Review. Int. J.
Mol. Sci. 2008,9, 1621–1651. [CrossRef] [PubMed]
92.
Brentrup, F.; Hoxha, A.; Christensen, B. Carbon Footprint Analysis of Mineral Fertilizer Production in Europe and Other World Regions.
In Proceedings of the 10th International Conference on Life Cycle Assessment of Food, Dublin, Ireland, 19–21 October 2016.
93.
Bharathiraja, D.B.; Sudharsana, T.; Jayamuthunagai, J.; Ramanujam, P.K.; Sivasankaran, C.; Iyyappan, J. Biogas production—A
review on composition, fuel properties, feed stock and principles of anaerobic digestion. Renew. Sustain. Energy Rev.
2018
,90,
570–582. [CrossRef]
94.
Liu, Z.-H.; Qin, L.; Jin, M.-J.; Pang, F.; Li, B.-Z.; Kang, Y.; Dale, B.E.; Yuan, Y.-J. Evaluation of storage methods for the conversion
of corn stover biomass to sugars based on steam explosion pretreatment. Bioresour. Technol.
2013
,132, 5–15. [CrossRef] [PubMed]
95.
Rentizelas, A.; Tolis, A.; Tatsiopoulos, I. Logistics issues of biomass: The storage problem and the multi-biomass supply chain.
Renew. Sustain. Energy Rev. 2009,13, 887–894. [CrossRef]
96.
Ariunbaatar, J.; Panico, A.; Esposito, G.; Pirozzi, F.; Lens, P.N.L. Pre-treatment methods to enhance anaerobic digestion of organic
solid waste. Appl. Energy 2014,123, 143–156. [CrossRef]
97. Wan, C.; Li, Y. Fungal pre-treatment of lignocellulosic biomass. Biotechnol. Adv. 2012,30, 1447–1457. [CrossRef]
98.
Kumar, D.; Murthy, G.S. Impact of pre-treatment and downstream processing technologies on economics and energy in cellulosic
ethanol production. Biotechnol. Biofuels 2011,4, 27. [CrossRef]
99.
Mao, C.; Feng, Y.; Wang, X.; Ren, G. Review on research achievements of biogas from anaerobic digestion. Renew. Sustain. Energy
Rev. 2015,45, 540–555. [CrossRef]
100.
Li, H.; Li, C.; Liu, W.; Zou, S. Optimized alkaline pre-treatment of sludge before anaerobic digestion. Bioresour. Technol.
2012
,123,
189–194. [CrossRef] [PubMed]
101.
Liew, L.N.; Shi, J.; Li, Y. Enhancing the solid-state anaerobic digestion of fallen leaves through simultaneous alkaline treatment.
Bioresour. Technol. 2011,102, 8828–8834. [CrossRef] [PubMed]
102.
Chen, Y.; Cheng, J.J.; Creamer, K.S. Inhibition of anaerobic digestion process: A review. Bioresour. Technol.
2008
,99, 4044–4064.
[CrossRef] [PubMed]
103.
Montgomery, L.; Bochmann, G. Pretreatment of Feedstock for Enhanced Biogas Production; IEA Bioenergy: Paris, France, 2014;
ISBN 978-1-910154-05-2.
104.
Liu, D.; Liu, D.; Zeng, R.J.; Angelidaki, I. Hydrogen and methane production from household solid waste in the two-stage
fermentation process. Water Res. 2006,40, 2230–2236. [CrossRef] [PubMed]
105.
Zheng, Y.; Zhao, J.; Xu, F.; Li, Y. Pretreatment of lignocellulosic biomass for enhanced biogas production. Prog. Energy Combust.
Sci. 2014,42, 35–53. [CrossRef]
106.
Mshandete, A.; Björnsson, L.; Kivaisi, A.K.; Rubindamayugi, S.T.; Mattiasson, B. Enhancement of anaerobic batch digestion of
sisal pulp waste by mesophilic aerobic pre-treatment. Water Res. 2005,39, 1569–1575. [CrossRef]
107.
Nah, I.W.; Kang, Y.W.; Hwang, K.-Y.; Song, W.-K. Mechanical pre-treatment of waste activated sludge for anaerobic digestion
process. Water Res. 2000,34, 2362–2368. [CrossRef]
Fermentation 2023,9, 449 15 of 15
108. Barber, W.P.F. Thermal hydrolysis for sewage treatment: A critical review. Water Res. 2016,104, 53–71. [CrossRef]
109.
Menind, A.; Normak, A. Study on grinding biomass as pre-treatment for biogasification. In Agronomy Research; Estonian
Agricultural University: Tartu, Estonia, 2003; p. 155.
110.
Skiadas, I.V.; Gavala, H.N.; Lu, J.; Ahring, B.K. Thermal pre-treatment of primary and secondary sludge at 70 degrees C prior to
anaerobic digestion. Water Sci. Technol. 2005,52, 161–166. [CrossRef]
Disclaimer/Publisher’s Note:
The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.
... Several factors, such as retention time, influence the composition of biogas. Retention time strongly correlates with the substrate's methane content and carbon oxidation state (Czekała et al., 2023). The reactor configuration also influences biogas output. ...
Article
Full-text available
The rising demand for renewable energy sources has fueled interest in converting biomass and organic waste into sustainable bioenergy. This study employs a bibliometric analysis (2013-2023) of publications to assess trends, advancements, and future prospects in this field. The analysis explores seven key research indicators, including publication trends, leading contributors, keyword analysis, and highly cited papers. We begin with a comprehensive overview of biomass as a renewable energy source and various waste-to-energy technologies. Employing Scopus and Web of Science databases alongside Biblioshiny and VOSviewer for analysis, the study investigates publication patterns, citation networks, and keyword usage. This systematic approach unveils significant trends in research focus and identifies prominent research actors (countries and institutions). Our findings reveal a significant increase in yearly publications, reflecting the growing global focus on biomass and organic waste conversion. Leading contributors include China, the United States, India, and Germany. Analysis of keywords identifies commonly used terms like "biofuels," "pyrolysis," and "lignocellulosic biomass." The study concludes by proposing future research directions, emphasizing advanced conversion technologies, integration of renewable energy sources, and innovative modelling techniques.
... Biogas composition including methane, carbon dioxide, and hydrogen sulphide are demonstrated in Figure 2a-c, respectively. Methane and carbon dioxide have a major portion in biogas mixture, and the other components are present in small amounts depending on substrate characteristics, as Czekała et al. [58] reported substrates with higher portions of carbohydrate and protein have a high hydrolysis rate, but substrates with fat have a comparatively high methane content in biogas. Methane contents of biogas produced from 100:0, 90:10, 70:30, 50:50, 30:70, 10:90, and 0:100 test groups were 52.78%, 51.80%, 51.68%, 51.76%, 51.25%, 51.15%, and 51.61%, respectively (Figure 2c). ...
Article
Full-text available
Anaerobic co-digestion (ACoD) of cow dung (CD) and maize cob (MC) may be envisaged as the best way to enhance biomethane formation and production of nutrient-enriched fertilizer for the implementation of a circular bio-economic system. The study aimed to find out the optimum ratio for the highest biogas production to produce heat and energy and also the generation of nutrient enriched organic fertilizer to use in crop land. A batch study was carried out for 99 days in an incubator maintaining 35 ± 1 °C temperature for seven different test groups of CD and MC (100:0, 90:10, 70:30, 50:50, 30:70, 10:90, and 0:100). The highest biogas production (356.6 ± 21.2 mL/gVS) was at 50:50 ratio with 138.05% and 32.02% increments compared to the digestion of CD and MC alone, respectively. Kinetic modeling showed the best fit using a Logistic model to evaluate ACoD of CD and MC mathematically. ACoD of available CD and MC in Bangladesh could produce 716.63 GWh/yr electricity for consumption and a large volume of nitrogen-enriched fertilizer to use in nitrogen deficit soil. There was no significant difference in nutrient enrichment among different test groups. Awareness about ACoD technology and proper use of digestate might bring this technology to field-level utilization and thus help to implement the circular bio-economic concept through zero waste generation.
... Comparatively, these moisture percentages were lower than 21.31% previously reported for PP by Ologunde et al. (2022). High moisture content, while essential for microbial activity, can be counterproductive during the stages of transportation and storage, affecting the substrate's quality prior to the AD process (Czekała et al. 2023). Therefore, it is imperative to determine the moisture content appropriate for a particular substrate to achieve optimal biogas production. ...
Article
The increasing accumulation of agricultural residues like rice husk (RH) and plantain peels (PP) poses environmental challenges, necessitating efficient waste management strategies. The study explores the potential of anaerobic co-digestion (AcoD) as a sustainable solution, specifically investigating its capacity for biogas generation from these agricultural residues. The primary objective is to determine the optimal substrate mixing ratios (SMRs) to maximize biogas yield. In-depth examination revealed that the highest biogas production reached a significant 2880 mL with a blend of 60% RH and 40% PP (RH60PP40). Additionally, the 80% RH and 20% PP composite (RH80PP20) demonstrated a substantial yield of 1996 mL. However, when plantain peels were used as the major substrate, biogas outputs decreased to 1250 mL and 173 mL for RH40PP60 and RH20PP80, respectively. Synergistic indexes (SI), measuring compatibility, reported values of 1.36 and 1.96 for the most promising samples, underscoring their optimal blending for biogas enhancement. From the perspective of digestate quality, plantain peel-based digestate (PP100D) stood out as a leading biofertilizer candidate due to its enriched nutrient profile. For the kinetic analysis, the logistic model was identified as the most predictive, outperforming the exponential and modified Gompertz models in mapping biogas production dynamics. Conclusively, the study accentuates that strategically optimized anaerobic co-digestion (AcoD) of RH and PP not only amplifies biogas outputs but also presents a viable, sustainable avenue for managing the environmental concerns associated with unchecked agricultural residue accumulation.
... In addition, it was shown in the work that already as a result of significant losses, primarily from nitrogen compounds, the composition of the digestate may be richer in these nutrients [136][137][138]. ...
Article
Full-text available
The use of methane fermentation in mesophilic conditions for the energy use of cow manure and additional co-substrates from the farm can bring a small dairy farm (140 dairy cows) financial benefits of up to EUR 114,159 per year. Taking into account the need to pay for emissions calculated as carbon dioxide equivalent, this profit could be reduced to EUR 81,323 per year. With the traditional direct use of manure, this profit would drop by as much as 60% to the level of EUR 33,944 per year. Therefore, the introduction of fees for emissions may significantly burden current dairy farms. As has already been shown, just compacting and covering the manure (which costs approx. EUR 2000 per year for 140 cows) would give almost twice as much profit—EUR 64,509 per year. Although an investment in a small biogas plant with a cogeneration unit on a family dairy farm may have a payback period of less than 6.5 years and a return of capital employed of 16%, most small farms in the world will not be able to afford its construction without external subsidies. At the same time, it would make it possible to reduce emissions by almost 270 times—from 41,460 to 154 tons of CO2eq per year—and the possibility of preserving valuable nutrients and minerals and supporting soil properties in the digestate. Therefore, it seems necessary for Europe to introduce a support system for small- and medium-sized farms with this type of investment in the near future in a much larger form than it has been so far.
... Another solution that brings measurable profits is the construction of a local biogas plant with a capacity of 250-500 kW and use of other waste streams [76,[120][121][122][123]. In the analyzed case, it would be an installation intended for a dozen local farmers (up to 10 km from the installation) and waste streams other than manure. ...
Article
Full-text available
The aim of the publication was to analyze investments in biogas plants with a cogeneration unit for an average size dairy farm. The basis for the calculation was the use of cow manure as the only substrate in methane fermentation. The economic balance also includes ecological and service aspects. The study also shows how much energy and quality potential is lost due to improper manure management and what impact a single farm with dairy cows has on the emission of carbon dioxide equivalent. It has been estimated that as a result of improper storage of manure, even 2/3 of its fertilizing, energy and economic value can be lost, while causing damage to the environment. It has been estimated that for a single farm with 100 cows, without government mechanisms subsidizing investments in RES, the payback period exceeds 15 years, and the Return of Capital Employed is slightly more than 6%.
Article
Full-text available
The management of biodegradable waste from various sectors of economy is an essential element in terms of environmental protection. The paper discusses issues related to the possibility of bio-waste treatment using anaerobic digestion technologies and composting processes, highlighting the conditions for the processes and their advantages and disadvantages. The challenges of overproduction of bio-waste faced by highly developed countries around the world are also presented. Research showed that the anaerobic digestion of this waste combines both biofuel production and a circular economy. The popularity of this method is linked, among others to a low cost of raw materials and wide range of possible uses for biogas (i.e. electricity, heat, or biomethane). In addition, an alternative bio-waste management option, compost production, was discussed. The study aimed to compare anaerobic and aerobic bio-waste management processes.
Article
Full-text available
Anaerobic digestion is traditionally used for treating organic materials. This allows the valorization of biogas and recycling of nutrients thanks to the land application of digestates. However, although this technology offers a multitude of advantages, it is still far from playing a relevant role in the energy market and from having significant participation in decarbonizing the economy. Biogas can be submitted to upgrading processes to reach methane content close to that of natural gas and therefore be compatible with many of its industrial applications. However, the high installation and operating costs of these treatment plants are the main constraints for the application of this technology in many countries. There is an urgent need of increasing reactor productivity, biogas yields, and operating at greater throughput without compromising digestion stability. Working at organic solid contents greater than 20% and enhancing hydrolysis and biogas yields to allow retention times to be around 15 days would lead to a significant decrease in reactor volume and therefore in initial capital investments. Anaerobic digestion should be considered as one of the key components in a new economy model characterized by an increase in the degree of circularity. The present manuscript reviews the digestion process analyzing the main parameters associated with digestion performance. The novelty of this manuscript is based on the link established between operating reactor conditions, optimizing treatment capacity, and reducing operating costs that would lead to unlocking the potential of biogas to promote bioenergy production, sustainable agronomic practices, and the integration of this technology into the energy grid.
Article
Full-text available
Due to fossil-fuel-limitation constraints, new energy sources are being sought. On the other hand, organic fertilizers that can be used in agriculture are increasingly being sought. One of the renewable energy sources is biogas produced from substrates large in organic matter. Apart from biogas, the product of the anaerobic digestion process is digestate. Due to the high content of nutrients, mainly nitrogen, this product can be successfully used as a fertilizer. This study aims to determine the content of total nitrogen (Ntot) and its selected fractions in the raw and processed digestate from agricultural biogas plants. The nitrogen fractions included N-NH4, N-NO3, and Norganic. The total nitrogen content (Ntot) and its fraction in raw digestate were determined. Samples used for the research came from five agricultural biogas plants. Separation into liquid and solid fractions is one of the methods for digestate management. The nitrogen content in selected samples obtained after separation of digestate in a biogas plant and on a laboratory scale was also checked. The obtained results show that digestate from agricultural biogas plants is a nitrogen-rich fertilizer. The content of Ntot in the tested samples ranged from 1.63 g∙kg−1 to 13.22 g∙kg−1 FM. The N-NH4 content in the analyzed material ranged from 0.75 to 4.75 g∙kg−1 FM. The determined physical and chemical properties confirm that the raw and processed digestate is characterized by appropriate fertilization properties, with particular emphasis on the content of Ntot and the share of its mineral forms. Based on the chemical composition, digestate from agricultural biogas plants can be considered a multi-component fertilizer.
Article
Full-text available
The abundant biomass growth of aquatic macrophytes in wetlands is one of the major concerns affecting their residing biota. Moreover, the biomass degenerates within the wetlands, thereby causing a remixing of nutrients and emission of greenhouse gases. Therefore, it is crucial to find sustainable methods to utilize the biomass of aquatic macrophytes devoid of environmental concerns. The present study investigates the utilization of the biomass of three aquatic macrophytes, including the lake sedge (CL: Carex lacustris Willd.), water hyacinth (EC: Eichhornia crassipes Mart. Solms), and sacred lotus (NL: Nelumbo nucifera Gaertn.) to produce oyster (Pleurotus ostreatus var. florida) mushrooms. For this purpose, different combinations of wheat straw (WS: as control) and macrophyte’s biomass (WH) such as control (100% WH), CL50 (50% WH + 50% CL), CL100 (100% CL), EC50 (50% WH + 50% EC), EC100 (100% EC), NL50 (50% WH + 50% NL), and NL100 (100% NL) were used for P. florida cultivation under controlled laboratory conditions. The results showed that all selected combinations of wheat straw and macrophyte biomass supported the spawning and growth of P. florida. In particular, the maximum significant (p < 0.05) growth, yield, bioefficiency, proximate, and biochemical parameters were reported using the WH substrate followed by CL, NL, and EC biomass, which corresponds to the reduction efficiency of the substrate parameters. Therefore, the findings of this study reveal that the biomass of selected aquatic macrophytes can be effectively utilized for sustainable mushroom cultivation while minimizing the risk associated with their self-degeneration.
Article
Full-text available
The exponential disposal of agro-industrial wastes onto the environment has endangered all forms of life. The implementation of these wastes in mushroom production is an eco-friendly and promising solution. The effect of olive and tea leaves extracts represented in culture media treatments: potato dextrose agar (PDA) 80% + tea extract (TE) 20% (T2), PDA 80% + olive extract (OE) 20% (T3), PDA 80% + OE 10% + TE 10% (T4), PDA 70% + OE 20% + TE 10% (T5) on mycelial growth diameter (MGD) and linear growth rate (MLGR) of Pleurotus ostreatus var. florida was compared to the one of PDA used as control (T1) at 22, 25 and 28°C inoculation temperatures. Optimum MGD was observed at 28°C in T1 (day 2), while it was significantly increased by 1.1–1.4 folds (days 4, 6, 8) in culture media containing plant extracts nevertheless the inoculation temperature. MLGR was improved by 102%–145% in olive/tea culture media compared to PDA nevertheless the inoculation temperature. Optimum growth induction index (GII) was observed in T3 (22.2%) at 22°C, T5 (21.9%) at 25°C and T4 (18.2%) at 28°C. These findings suggest the combination of olive and tea leaves extracts in the production of Pleurotus ostreatus .
Article
Full-text available
The vegetable production is an important part of agriculture sector in every country. In Poland, vegetables and fruits production covering the area of no more than 3% of agricultural land, is more than 36% of plant production and 14–15% of the whole agricultural production. The study aim was to determine the management possibilities of the selected waste from vegetable production in composting process. Laboratory tests were carried out using the bioreactor set-up with capacity of 165 dm3, respectively, for each chamber. The composting process has been tested for the following mixtures: K1—cabbage leaves, tomato dry leaves + manure and slurry additive; K2—cabbage leaves, solid fraction from biogas plant + manure and straw additive; K3—cabbage leaves, onion husk + straw additive. In all three composts the thermophilic phase occurred which indicates that the process ran correctly. In each chamber, the temperature exceeded 70 °C and its maximum value during the experiment was 77.5 °C for K2 compost. The article discusses changes in O2, CO2, NH3 and H2S emissions during composting. The carbon dioxide concentration in the exhausted gas from analyzed composts and the ratio with oxygen they testify to the decomposition of raw materials in the composting process. The results showed that the agri-food waste can be a proper substrate for composting production. Due to legal regulations and the increase in prices of mineral fertilizers, the development of the compost market should be expected.
Article
Full-text available
The knowledge of nutrient composition of specific substrate(s) for anaerobic digestion for the production of biogas can provide first-hand information on the possible outcome of digesting such feedstock. It will also help in planning the construction of large-scale biogas plants based on the awareness of the substrates output quantity of biodegradation products. This paper aims to present feedstock information, yield of the bioprocess and bioenergy capacity of products from anaerobic digestion for comparison, studies and analysis.
Article
Full-text available
Fermentation (not anaerobic digestion) is an emerging biotechnology to transform waste into easily assimilable organic compounds such as volatile fatty acids, lactic acid and alcohols. Co-fermentation, the simultaneous fermentation of two or more waste, is an opportunity for wastewater treatment plants (WWTP) to increase the yields of sludge mono-fermentation. Most publications have studied waste activated sludge co-fermentation with food waste or agri-industrial waste. Mixing ratio, pH and temperature are the most studied variables. The highest fermentation yields have been generally achieved in mixtures dominated by the most biodegradable substrate at circumneutral pH and mesophilic conditions. Nonetheless, most experiments have been performed in batch assays which results are driven by the capabilities of the starting microbial community and do not allow evaluating the microbial acclimation that occurs under continuous conditions. Temperature, pH, hydraulic retention time and organic load are variables that can be controlled to optimise the performance of continuous co-fermenters (i.e., favour waste hydrolysis and fermentation and limit the proliferation of methanogens). This review also discusses the integration of co-fermentation with other biotechnologies in WWTPs. Overall, this review presents a comprehensive and critical review of the achievements on co-fermentation research and lays the foundation for future research.
Article
Full-text available
Biomass and other carbonaceous materials can be gasified to produce syngas with high concentrations of CO and H2. Feedstock materials include wood, dedicated energy crops, grain wastes, manufacturing or municipal wastes, natural gas, petroleum and chemical wastes, lignin, coal and tires. Syngas fermentation converts CO and H2 to alcohols and organic acids and uses concepts applicable in fermentation of gas phase substrates. The growth of chemoautotrophic microbes produces a wide range of chemicals from the enzyme platform of native organisms. In this review paper, the Wood–Ljungdahl biochemical pathway used by chemoautotrophs is described including balanced reactions, reaction sites physically located within the cell and cell mechanisms for energy conservation that govern production. Important concepts discussed include gas solubility, mass transfer, thermodynamics of enzyme-catalyzed reactions, electrochemistry and cellular electron carriers and fermentation kinetics. Potential applications of these concepts include acid and alcohol production, hydrogen generation and conversion of methane to liquids or hydrogen.
Article
Anaerobic digestion produces biogas, a mixture of CH4 and CO2, where CH4 is a low cost, environmentally friendly, and renewable energy source. The application of biogas production is increasing rapidly as a means of reducing the pollution impact of organic biomasses. However, biogas contains unwanted elements such as hydrogen sulfide, carbon monoxide, siloxanes, and carbon dioxide. To remove these elements, several biogas upgrading technologies like water scrubbing, amine scrubbing, pressure swing adsorption, and membrane separation have been developed and are being used at various commercial scales. Problems with these methods are high energy consumption, the use of expensive chemicals, and high operating cost. Therefore, a major effort is currently underway to improve the design of existing methods as well as developing innovative new upgrading technologies such as cryogenic separation and biological upgrading. This review intends to provide a comprehensive overview of the limitations with the existing upgrading technologies along with recent advances in physical, chemical, and biological biogas upgrading technologies (e.g., pressure swing adsorption, membrane separation, biochar adsorption and CO2 conversion by biological organisms) and further into possible future solutions, such as hybrid systems. Comparative studies of process complexities and associated economic concerns are also provided, and future perspectives that may facilitate research into sustainable biogas upgrading technologies are discussed, focusing in particular on cryogenic separation, novel biological techniques, biochar based upgrading and hybrid technologies incorporating two or more different methods seamlessly integrated.
Article
Apple pomace, the residue which is left out after processing of apple serves as a potential carbon source for the production of biopolymer, PHA (poly-hydroxyalkanoates). It is rich in carbohydrates, fibers and polyphenols. Utilization of these waste resources has dual societal benefit- waste management and conversion of waste to an eco-friendly biopolymer. This will lower the overall economics of the process. A major limitation for the commercialization of biopolymer in comparison with petroleum derived polymer is the high cost. This article gives an overview of valorization of apple pomace for the production of biopolymer, various strategies adopted, limitations as well as future perspectives.