You are on page 1of 145

MINISTÉRIO DA EDUCAÇÃO

Universidade Federal de Ouro Preto


Graduate Program in Environmental Engineering – PROAMB

UNIVERSIDADE FEDERAL DE OURO PRETO


GRADUATE PROGRAM IN ENVIRONMENTAL
ENGINEERING

BIOGAS PRODUCTION BY ANAEROBIC CO-


DIGESTION USING BYPRODUCTS OF FIRST-
AND SECOND-GENERATION BIOETHANOL
PRODUCTION PROCESS

Oscar Fernando Herrera Adarme

Ouro Preto
2020
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

BIOGAS PRODUCTION BY ANAEROBIC CO-DIGESTION USING


BYPRODUCTS OF FIRST- AND SECOND-GENERATION BIOETHANOL
PRODUCTION PROCESS

Oscar Fernando Herrera Adarme

Thesis submitted to the Graduate Program in


Environmental Engineering, Universidade Federal
de Ouro Preto, as partial requirement to obtain the
title of PhD in Environmental Engineering

Research area: Environmental Technology


Research line: Bioenergy
Advisor: PhD. Sérgio Francisco de Aquino
Co-Advisor: PhD. Bruno Eduardo Lobo Baêta

Ouro Preto
2020
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

ABSTRACT

In this thesis anaerobic co-digestion (AcD) of sugarcane biorefinery byproducts,


i.e. hemicelluloses hydrolysate (HH) (obtained by hydrothermal pretreatment of
sugarcane bagasse), vinasse (generated during 1G ethanol production process), yeast
extract (YE) (which can be obtained from spent/excess yeast) and sugarcane bagasse fly
ashes (SBFA), was optimized by means of biochemical methane potential (BMP)
experiments. The best experimental conditions of AcD (25-75% HH-to-vinasse mixture
ratio; 1.0 g L-1 YE; 15 g L-1 SBFA and 100-0% HH-to-vinasse; 1.5 g L-1 YE; 45 g L-1
SBFA) led to the production of 0.279 and 0.267 Nm3 of CH4 per kg of chemical oxygen
demand (COD) removed with an energy surplus of 0.43 and 0.34 MJ kg SB-1,
respectively. Adsorption experiments using SBFA were carried out and showed this
residue could adsorb up to 61.71 and 58.21 mg g-1 of 5-Hydroxymethyl-2-furfuraldehyde
and 2-furfuraldehyde, thereby reducing toxicity and improving biogas production during
AcD of the aforementioned byproducts. Subsequently, two mesophilic systems fed
continuously with the mixture of HH and vinasse and inoculated with SBFA and YE were
used in order to validate batch data. A single stage hybrid anaerobic reactor (HAR) was
compared with a two-stage acidogenic-methanogenic system formed by an acidogenic
structured-bed reactor (ASTBR) followed by an UASB methanogenic reactor. The
organic loading rate (OLR) (from 0.9 to 10.8 g COD L-1 d-1) applied to these systems was
attained by fixing the hydraulic retention time (HRT) in the reactors (17.5 h in HAR; 6 h
in the ASTBR and 19.9 h in the UASB) and changing the influent COD concentration by
means of dilution. The results showed the feasibility of applying the two-stage system
(ASTBR/UASB) to treat a HH-vinasse mixture, leading to a global COD removal higher
than 80% and methane yield of 0.245 NL CH4 g CODr-1. In its turn, the single stage
(HAR) system led to 65% of COD removal and 0.205 NL CH4 g CODr-1 of methane yield.
Microbial community analyses of sludge collected from the ASTBR at different
operational conditions (OLR = 3.75 g COD L-1 d-1 (phase I) and OLR = 10.81 g COD L-
1
d-1 (phase III)) revealed the main species were Clostridium beijerinckii (61.9%) and
Desulfovibrio desulfuricans (19.1%) for phase I; and Lactobacillus casei (66.3 %) and
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

Lactococcus lactis (22.4 %) for phase III. Regarding the UASB reactor, the main
identified species were Longilinea arvoryzae (15.5%), Methanosaeta concilii (19.9 %)
and Syntrophus aciditrophicus (15.4%) at phase I; and Longilinea arvoryzae (31.7%),
Methanosaeta concilii (10.6 %) and Pleomorphomonas oryzae (10.1%) at phase III. In its
turn, the main species identified in the HAR sludge were Bacteroides graminisolvens
(11.3%), Desulfovibrio desulfuricans (12.6 %) and Methanosaeta concilii (16.4%.) Both
systems (single HAR and ASTBR-UASB) exhibited a stable long-term operation (240
days) with low VFA accumulation (average of 550 mg/L at HAR and 625 mg/L at
UASB). Finally, economic indicators showed it is not viable to integrate the 1G-2G
sugarcane mill including the AcD unit when considering the current technologies for 2G
ethanol production in view of their higher capital and operational costs. However, when
fixing the experimental methane yield as 0.245 Nm3 kg COD-1r , which was obtained for
the two-stage system, and using 50% of bagasse surplus, it was possible to achieve
internal rate of return (IRR), return on investment (ROI), and payback period values of
21.8%, 59.50% and 10.55 years respectively. When experimental methane yield was
increased in 10% and keeping the same 50% of surplus bagasse use, it was possible to
achieve internal rate of return (IRR), return on investment (ROI), and payback period
values of 26%,89.05% and 5.36 years, respectively. Thus, the results evidenced the
importance of investment in R&D especially in 2G ethanol production (available bagasse)
but also in the anaerobic co-digestion of byproducts to reach the viability of this agro-
industrial business.
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

TABLE OF CONTENTS

1 INTRODUCTION 14

2 LITERATURE REVIEW 17

SUGARCANE BIOREFINERY 17

2.1.1 First-generation Bioethanol Production (1G) 19

2.1.2 Second generation ethanol production process (2G) 23

ANAEROBIC DIGESTION 25

2.2.1 Anaerobic Co-Digestion 27

2.2.2 Anaerobic co-digestion of bioethanol production wastes 30

CONCLUSIONS FROM LITERATURE REVIEW - THESIS PREMISES 36

3 HYPOTHESES 38

CENTRAL 38

COMPLEMENTARY HYPOTHESES 38

4 OBJETIVES 39

GENERAL 39

SPECIFIC 39

5 USE OF ANAEROBIC CO-DIGESTION AS AN ALTERNATIVE TO ADD VALUE TO SUGARCANE


BIOREFINERY WASTES 40

ABSTRACT 40

INTRODUCTION 40

MATERIAL AND METHODS 42

5.3.1 Chemicals 42

5.3.2 Sugarcane Biorefinery Wastes 43

5.3.3 Hydrothermal Pretreatment of SB 43

5.3.4 AcD of Hemicelluloses Hydrolysate and Vinasse 43

5
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

5.3.5 Design of Experiments and Statistical Analysis 45

5.3.6 Effect of Yeast Extract Addition on AcD 46

5.3.7 Effect of Sugarcane Bagasse fly ash addition on AcD 46

5.3.8 Modeling the Methane Production in AcD Experiments 46

5.3.9 Evaluation of Toxicants Adsorption onto SBFA 47

5.3.10 Analytical Methods 48

5.3.10.1 Characterization of gas fractions 48

5.3.10.2 Characterization of liquid fractions 48

5.3.10.3 Characterization of solid fractions 49

5.3.11 Energy Balance 50

RESULTS AND DISCUSSION 52

5.4.1 Physicochemical Characterization of Liquid and Solid Fractions 52

5.4.2 Anaerobic co-digestion 58

5.4.2.1 Effect of vinasse and HH mixture on methane production 58

5.4.2.2 Effect of YE addition on the methane production 65

5.4.2.3 Effect of SBFA addition on the methane production 69

5.4.3 Mechanism of SBFA Adsorption during AcD. 72

5.4.4 Energy Analysis 76

6 CONTINUOUS BIOGAS PRODUCTION BY ANAEROBIC CO-DIGESTION OF SUGARCANE


BYPRODUCTS: ANALYSIS OF REACTOR CONFIGURATIONS 79

ABSTRACT 79

INTRODUCTION 80

MATERIAL AND METHODS 81

6.3.1 Feedstocks for anaerobic co-digestion (substrates and inoculum) 81

6.3.2 Experimental setup and operating conditions 82

6.3.3 Performance evaluation: analytical methods 85


6
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

6.3.3.1 Characterization of gas fractions 85

6.3.3.2 Characterization of liquid fractions 85

6.3.3.3 Characterization of solid products 86

6.3.3.4 Metagenomic characterization of reactors sludge. 87

6.3.3.5 Kinetic analysis 87

6.3.4 Hydrodynamic analysis 88

RESULTS AND DISCUSSION 88

6.4.1 Organic matter removal at different organic loading rates 88

6.4.1.1 Single stage system 88

6.4.1.2 Two-stage system 91

6.4.2 Biogas production and methane yield 93

6.4.2.1 Single stage system 93

6.4.2.2 Two-stage system 94

6.4.3 Volatile fatty acid analysis 96

6.4.3.1 VFA profile in the acidogenic reactor (ASTBR) 96

6.4.3.2 VFA profile in methanogenic reactors of single (HAR) and two-stage (UASB)
systems 97

6.4.4 Microbial community analysis at different OLR 99

6.4.4.1 Single stage system 99

6.4.4.2 Two-stage system 101

6.4.5 Hydrodynamics and kinetic analysis 104

7 TECHNO-ECONOMIC ANALYSIS OF ANAEROBIC CO-DIGESTION OF SUGARCANE


BIOREFINERY WASTES 108

ABSTRACT 108

INTRODUCTION 109

METHODOLOGY 110

7
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

7.3.1 Basic inputs and scenario description 110

7.3.2 Estimates and calculations 113

7.3.3 Economic assessment methodology 113

RESULTS AND DISCUSSION 114

7.4.1 Energy Assessment 114

7.4.2 Economic Assessment 117

8 CONCLUSIONS 121

9 FUTURE WORKS 123

10 REFERENCES 124

8
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

LIST OF FIGURES

Figure 2.1 Relationship between bioeconomy and sustainable chemistry for the use of
sugarcane. (adapted from (Vaz, 2017)). ......................................................................... 18
Figure 2.2 Flowchart for first- and second-generation ethanol production processes using
sugarcane. Adapted from (Hoarau, J et al., 2018) .......................................................... 20
Figure 2.3 Conceptual flow diagram of anaerobic co-digestion processes (Adapted from
(Siddique e Wahid, 2018)) ............................................................................................. 28
Figure 5.1 Typical FTIR spectra of the yeast extract sample used in this study. ........... 55
Figure 5.2 FTIR spectra for sugarcane bagasse fly ash .................................................. 57
Figure 5.3 Pareto’s chart of the standardized effects for anaerobic co-digestion of HH and
vinasse ............................................................................................................................ 60
Figure 5.4 Biochemical methane potential for different mixtures of HH-to-vinasse and
F/M ratio ......................................................................................................................... 61
Figure 5.5 Fitting (line (two-phase model) and dashed line(Gompertz model)) of the
experimental data (o) for the best condition of BMP ..................................................... 62
Figure 5.6 Response surface plot for BMP experiments according to the Doehlert
experimental design ........................................................................................................ 64
Figure 5.7 Methane production (open bar) and VFA(filled bar) (sum of formic acid, acetic
acid, propionic acid, butyric acid, isobutyric acid, valeric acid and isovaleric acid
concentrations) accumulation (line and dot) for anaerobic co-digestion in the presence of
yeast extract at three different conditions. ...................................................................... 67
Figure 5.8 Accumulated methane production (open bar) and VFA (filled bar) (sum of
formic acid, acetic acid, propionic acid, butyric acid, isobutyric acid, valeric acid and
isovaleric acid concentrations) accumulation for two different anaerobic co-digestion
conditions (25-75% HH-to-Vinasse, 1.0 g L-1 YE, 15 g L-1 SBFA and 100-0% HH-to-
Vinasse, 1.5 g L-1 YE, 45 g L-1 SBFA). ....................................................................... 70
Figure 5.9 a)BET isotherm fitted (full line) to experimental data (circles) of COD
adsorption using a mixture of 25-75 % of HH-Vinasse and 1.0 gL-1 of Yeast extract and

9
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

SBFA as adsorbent at concentrations of 1.5 to 60 g L-1; b) BET adsorption isotherm for


experimental data obtained using a mixture of 100-0 % of HH-Vinasse and 1.5 gL-1 of
Yeast extract ................................................................................................................... 75
Figure 5.10 Energy balance for the process proposed using AcD of HH and vinasse with
addition of YE (1.0 and 1.5 g L-1) and SBFA (15, 45 and 75 g L-1) for two proportions
of substrates. ................................................................................................................... 77
Figure 5.11 Flowchart of the process proposed for managing sugarcane and its wastes at
an 1G/2G ethanol production process in a biorefinery concept. .................................... 78
Figure 6.1 Schematic diagram of two stage anaerobic system consisting in acidogenic
structured bed reactor (ASTBR) and upflow anaerobic sludge blanket (UASB).......... 84
Figure 6.2 Schematic diagram of single stage anaerobic system using a hybrid reactor
(HAR) ............................................................................................................................. 84
Figure 6.3 COD behavior in HAR along the three operational phases with varying OLR.
........................................................................................................................................ 90
Figure 6.4 COD profile at influent and effluent of ASTBR (a) and UASB (b) reactors
operated at different OLRs ............................................................................................. 92
Figure 6.5 Biogas production and methane yield for three different OLR in HAR. ...... 94
Figure 6.6 Granules sampled from HAR at two different OLR corresponding with phase
I and phase III ................................................................................................................. 94
Figure 6.7 Biogas production and methane yield for three different OLR in (a) ASTBR
system and (b) two stage anaerobic system. ................................................................... 96
Figure 6.8 Total volatile fatty acids (VFA) profile in mg/L (line and square) and is
mgCOD/L (dashed line) accumulated in the ASTBR along the three operational phases.
........................................................................................................................................ 97
Figure 6.9 Total volatile fatty acids accumulated (line and square) as function of
volumetric methane production rate (dashed line) in HAR (a) and UASB (b) reactors. 99
Figure 6.10 Relative abundance of microbial communities at species level for HAR in
two different OLR ........................................................................................................ 100
Figure 6.11 Venn diagram of species identified in HAR submitted to two different OLR
...................................................................................................................................... 101

10
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

Figure 6.12 a)The Relative abundance of microbial communities at species level for
ASTBR anaerobic system in two different OLR b)The Relative abundance of microbial
communities at species level for S-UASB anaerobic system in two different OLR ... 103
Figure 6.13 Venn diagram of species identified at the single (HAR) and two-stage
(ASTBR+UASB) anaerobic systems ........................................................................... 104
Figure 7.1 Flowchart of 2G+AcD scenario .................................................................. 111
Figure 7.2 Energy production for three scenarios of surplus bagasse usage for 2G ethanol
production and biogas production at single- or two-stage anaerobic systems. E BB –
energy obtained by bagasse burning; E HT – energy spent on hydrothermal pretreatment;
E 1Stage – energy obtained as biogas in the HAR reactor; E 2Stage – energy obtained as
biogas in the ASTBR-UASB system. ........................................................................... 115
Figure 7.3 Energy production for three scenarios of surplus bagasse usage for 2G ethanol
production and biogas production at single- or two-stage anaerobic systems considering
an increase of 10% in methane yield. E BB – energy obtained by bagasse burning; E HT
– energy spent on hydrothermal pretreatment; E 1Stage – energy obtained as biogas in
the HAR reactor; E 2Stage – energy obtained as biogas in the ASTBR-UASB system.
...................................................................................................................................... 116
Figure 7.4 Incremental cash flow analysis for the scenarios of 2G ethanol production
integrated to biogas production from vinasse and HH AcD according to bagasse use: a)
20% of bagasse for 2G ethanol), b) 50% of bagasse for 2G ethanol c) 80% of bagasse for
2G ethanol .................................................................................................................... 119
Figure 7.5 Incremental cash flow analysis for the scenarios of 2G ethanol production
integrated to biogas production from vinasse and HH AcD according to bagasse use and
considering a methane yield increase of 10%: a) 20% of bagasse for 2G ethanol), b) 50%
of bagasse for 2G ethanol c) 80% of bagasse for 2G ethanol ...................................... 120

11
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

LIST OF TABLES
Table 2.1 Main residues formed in the sugar and ethanol production process. ............. 22
Table 2.2 Physicochemical characteristics of vinasse obtained from annexed and
autonomous ethanol production in sugar-alcohol biorefinery. (Marafon et al., 2020;
Santos et al., 2019) ......................................................................................................... 23
Table 2.3 Anaerobic co-digestion studies of vinasse with different substrates (to be
continued). ...................................................................................................................... 32
Table 5.1 Doehlert experimental design for anaerobic co-digestion of vinasse and
hemicellulose hydrolysates (HH) ................................................................................... 45
Table 5.2 Physicochemical characterization of vinasse, hemicelluloses hydrolysate, yeast
extract and sugarcane bagasse fly ash ............................................................................ 53
Table 5.3 Physicochemical characterization of sugarcane bagasse fly ash by XRF ...... 54
Table 5.4 Analysis by ICP OES for sugarcane bagasse ash digested............................. 58
Table 5.5 Kinetic constants for methane production for best anaerobic co-digestion
process using substrates from sugarcane biorefinery. .................................................... 63
Table 5.6 Experimental design for anaerobic co-digestion of vinasse, hemicellulose
hydrolysates (HH), yeast extract and sugarcane bagasse ashes...................................... 66
Table 5.7 Isotherm parameters estimated by the models for experiments 13 (25-75% HH-
to-vinasse, 1.0 g L-1 YE) and 14 (pure HH, 1.5 g L-1 YE) using SBFA as adsorbent at
concentrations of 1.5 to 60 g L-1. .................................................................................... 74
Table 6.1 Operational conditions for single and two-stage anaerobic reactors .............. 83
Table 6.2 Hydrodynamic parameters obtained for single and two-stage anaerobic systems
...................................................................................................................................... 105
Table 6.3 - Kinetic parameters for COD removal in single and two-stage anaerobic
reactors at different operational phases. ....................................................................... 106
Table 7.1 - Data of 1G2G sugarcane ethanol production used in the simulations,
considering annexed distilleries.................................................................................... 112
Table 7.2 – Anaerobic Co-digestion system parameters for simulation....................... 112

12
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

Table 7.3 – Economical analysis for each scenario of surplus bagasse used at an integrated
1G-2G plant and AcD with a fixed methane yield of 0.245 Nm3 kg CODr -1 ............. 118
Table 7.4 – Economical analysis for each scenario of surplus bagasse use at an integrated
1G-2G plant and AcD with a fixed methane yield of 0.301 Nm3 kg CODr -1 ............. 119

13
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

1 INTRODUCTION

The constant and growing interest in cleaner, safer, lower-cost energy systems has
increased attention on world biofuel demand over the past 30 years. Bioethanol, one of
the leading sustainable biofuels, reduces carbon monoxide and carbon dioxide levels in
relation to fossil fuel (Silva et al., 2020). Such compound has been classified by the US
Environmental Agency (US EPA) as a biofuel that can reduce greenhouse gas (GHG)
emissions by up to 61% when compared to gasoline (Delgado et al., 2017). A wide range
of renewable raw materials can be used for ethanol production, among these fermentable
sugars (sugarcane, beet, sorghum); starches (corn, potato, rice, wheat, agave); and
cellulosic material (straw, grasses, ears of corn, wood, sugarcane bagasse). However,
despite the potential of various raw materials, due to regional availability and
technological challenges, around 75% of ethanol traded worldwide in 2016 was produced
from corn and sugarcane (Bergmann et al., 2018). Therefore, studies aimed at researching
and developing appropriate methodologies to achieve complete environmental, technical
and economic management of raw materials and products should be encouraged.
In Brazil, sugarcane is the main source of biomass for first generation (1G) ethanol
production by using extracted sugars, and several by-products associated with this
production process are generated. To put it in context, according to recent data from the
National Supply Company (CONAB, 2019) for the 2019/2020 harvest, sugarcane
production was 642.7 million tons, and in the same production chain 30.1 million tons of
sugar were obtained along with 33.8 billion liters of ethanol. In this sense, a by-product
of interest that could be better used is sugarcane bagasse (SB), which is composed mainly
of polymeric units in the form of cellulose (30 to 45%), hemicelluloses (25 to 30%) and
lignin (25 to 30 %) and whose yield is estimated at 280 kg per ton of sugarcane (with
50% humidity). Currently, much of the SB generated in the sugar-alcohol process is
burned in steam-generating boilers (Bergmann et al., 2018).
As previously presented, the SB has the potential to be better used since it has
high contents of cellulose and hemicelluloses. Currently a large number of studies are
being carried out to enable the production of second generation (2G) ethanol from the SB
generated in the 1G ethanol production process (Dias et al., 2013). However, to enable
14
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

the use of this material for such purpose, a pretreatment step capable of removing lignin
and hemicellulose (C5 sugars) is required. Currently one of the most promising and
feasible pretreatment techniques for industrial implementation is the hydrothermal
process. This process employs water at high temperature (it is also called liquid hot water
pretreatment) whose main objective is to extract hemicellulose (C5 sugars) and partially
degrade lignin from lignocellulosic biomass (Wertz et al., 2018).
It is now known that the use of hemicelluloses hydrolysates (rich in pentoses or
C5 sugars) for the production of 2G ethanol still has technical drawbacks since
conventional yeast use hexoses (C6 sugars) as substrate for ethanol production.
Therefore, the production of ethanol from C5 sugars normally requires the use of
genetically modified microorganisms, which implies in higher costs (Wertz et al., 2018).
An alternative for using such pentose-rich hydrolysate is to use anaerobic biotechnology
to produce other biofuels and / or value-added products (Safari, Karimi e Shafiei, 2016;
Santos et al., 2018; Saripan e Reungsang, 2013). Indeed, anaerobic digestion (AD) is a
promising process for renewable energy production, as it allows the conversion of organic
matter, into combustible gases (methane and hydrogen) and valuable by-products (e.g.
organic acids).
Commonly used substrates for biogas production include sewage and industrial
effluent / waste, either raw or in combination with each other. However, there are several
studies showing that pentose liquor can be converted to methane and hydrogen in
anaerobic processes (Baêta, Lima, Adarme, et al., 2016; Baêta, Lima, Filho, et al.,
2016;Ribeiro et al., 2017; Rabelo et al., 2011; Kaparaju et al., 2009). On the other hand,
although it is a feasible technological route, there is still a lack of information and
uncertainty in relation to the most common AD process drawbacks such as: lack of
nutrients, alkalinity and the presence of inhibitory compounds.
In addition to SB, the sugar and alcohol sector is also responsible for the
generation of vinasse, which is a liquid effluent characterized by high levels of organic
matter, potassium and sulfur. Currently much of this vinasse is used in fertigation,
although many studies have shown the environmental impacts of this practice in relation
with GHG emissions and soil salinization by the potassium (Haandel, Van e Lier, Van,

15
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

2015). On the other hand, different research groups indicate the possibility of using this
effluent for biogas production via anaerobic digestion (Marafon et al., 2020 ;Colturato,
2015 ;Fuess et al., 2018; Fuess, Garcia e Zaiat, 2018). However, although the literature
contains a lot of information on the use of anaerobic treatment schemes, the results are
highly variable and dependent to the final products desired (H2, CH4, volatile fatty acids
(VFA)), as well as, to the process parameters including pH, temperature, alkalinity and
hydraulic retention time.
Another kind of byproducts, like ashes from SB burning and excess of yeast from
fermentation step, are formed in the 1G ethanol production process. This products have
specialty characteristics because of, for example, surplus yeast can be used as nutritional
supplement due to its highly contents of proteins and vitamin B (Bergmann et al., 2018).
Moreover, sugarcane bagasse ashes have some applications as alkali agent, in order to
correct the pH of soil, fertilizer and adsorbent material (Cacuro, T A e Waldman, 2015).
Some constituents of vinasse (e.g. nutrients such as potassium), disposable yeast
(e.g. vitamins and minerals) or ashes (e.g. trace metals, alkalinity) can enhance the
anaerobic degradation of hemicelluloses hydrolysates. In addition, dilution of vinasse
with hemicelluloses hydrolysates may reduce the toxicity caused by sulfur compounds,
thereby increasing the molar fraction of methane in biogas. Is important to highlighted
that nowadays, this sugarcane biorefinery ‘residues’ can be applied in the same life cycle
of sugarcane; however, its huge production does not allow its complete insertion.
Based on the above, the evaluation of hydrogen and methane production from
anaerobic co-digestion of vinasse, hemicellulosic hydrolysate and other residual streams
from sugarcane biorefinery, becomes of great relevance since the energy vectors
generated by this process can economically make possible the integration of 1G and 2G
ethanol production.

16
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

2 LITERATURE REVIEW

SUGARCANE BIOREFINERY
The economic development of different biomass-based production chains is part
of the research, development and innovation (R & D & I) agenda in countries such as
Germany, France, USA, Brazil, China and South Korea, among others. This mobilizes
large amounts of public and private efforts and resources, aiming at the optimal use of
biomass for the recovery of added value products and reduction of environmental impacts.
In this context, Brazil is one of the leading producers of biomass for food and biofuels,
hence these efforts are of great importance for maintaining a positive economic scenario
with fewer negative impacts (Vaz, 2017). According to data from the National Supply
Company, Brazil is the world's largest sugarcane producer, and a production of 642,7
million tons has been forecast for the 2019/20 harvest (Companhia Nacional de
Abastecimento (CONAB), 2019). This encourages the use of sugarcane bagasse as
platform for the recovering of more sustainable products and services.

Concepts such as biorefinery and green chemistry focus on the use of biomass,
similar to the use of oil, in order to generate added value through the production of fuels,
heat, energy, chemicals and materials. These concepts consider sustainable integrated
systems (raw materials-processes-products-wastes) according to technical parameters
that take into account, among other things, mass and energy balances and life cycle
analysis. Thus, a sugarcane-based biorefinery is a clear example of how to define a
relationship between bioeconomic and renewable chemistry concepts and assumptions
(Figure 2.1), thereby allowing the implementation of innovative proposals, both
technically and economically (Joelsson et al., 2016; Vaz, 2017).

Biorefinery classification can be made based on the type of raw material, type of
technology and associated methods (1st or 2nd generation) employed. Currently, first
generation (1G) biorefineries produce their main products from sugars and starches
available in crops. However, this raw material is the basis for food production and
represents about 1% of the available plant biomass. In the case of 2nd generation (2G)

17
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

biorefineries, the sources of raw materials are lignocellulosic biomass (e.g. sugarcane
bagasse - SB), generated in agricultural and industrial activities. Such biomasses are
basically formed by compounds with polymeric characteristics such as cellulose (30 to
45%), hemicelluloses (25 to 30%) and lignin (25 to 30%). Cell wall carbohydrates present
in the form of cellulose (glucose) and hemicellulose (arabinose, galactose, glucose, xylose
and mannose) can then be exploited for their conversion into different types of fuels
(ethanol, methane, hydrogen, among others). However, the complexity of interactions of
the three main constituents in the cell wall is the main challenge for recovering
carbohydrates from lignocellulosic biomass (Baruah et al., 2018).

Figure 2.1 Relationship between bioeconomy and sustainable chemistry for the use of
sugarcane. (adapted from (Vaz, 2017)).

In Brazil, the use of sugarcane as raw material for biorefineries is desirable


because of its potential to provide the following compounds and materials (Vaz, 2017):

✓ Sugars: sucrose (broth), glucose (derived from cellulose and sucrose inversion),
and xylose (derived from hemicellulose);
✓ Bioethanol: from sucrose fermentation

18
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

✓ Natural polymers or macromolecules: cellulose, hemicellulose and lignin, which


are present in bagasse and straw (lignocellulosic biomass) and can be converted
into sources of hexose, pentose and phenol, respectively;
✓ Vinasse or wastewater rich in organic matter;
✓ Other compounds, such as higher alcohols and carbon dioxide (CO2) at a
compatible purity content for food and chemical use.

2.1.1 First-generation Bioethanol Production (1G)

The flowchart presented in Figure 2.2 illustrates the concept of biorefinery using
sugarcane. Initially the sugarcane is harvested in the field and goes through a cleaning
step to be later crushed so that the juice is extracted. The extracted juice (broth) is sent to
the 1G sugar and ethanol production processes while some of the bagasse is used in a
CHP (Combined Heat and Power) cogeneration system operating on a steam cycle. The
juice sent for sugar and ethanol production is treated in different processes for the removal
of soluble and insoluble impurities. The clarified broth goes through cooking
(concentration), crystallization and drying steps in order to obtain sugar. Mixing the
clarified broth with molasses forms the wort that is subjected to fermentation processes.
The ‘wine’ obtained after fermentation is then sent to distillation to obtain hydrated
ethanol, which goes through a later dehydration step to obtain anhydrous ethanol using
different technologies such as extractive distillation or adsorption onto molecular sieves
(Albarelli, 2013).

19
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

Figure 2.2 Flowchart for first- and second-generation ethanol production processes using sugarcane. Adapted from (Hoarau, J et al., 2018)

20
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

The ‘wastes’ generated in the sugar and ethanol production processes are normally
applied in the same life cycle of sugarcane, hence they are normally referred to as
‘byproducts’. The main byproducts formed, their usual destination and some
technological routes used for processing them are presented in Table 2.1.

Among the residues formed in the process, the vinasse is the most studied due to
its pollutant characteristics. This effluent presents high acidity and high concentrations of
organic matter, nitrogen and sulfur (Table 2.2). The characteristics cited and associated
with the large volume generated during ethanol production process give to this effluent
especial attention in relation to its final disposal. In Brazil, it is estimated that for each
liter of alcohol produced in the 1G ethanol production process, it is generated
approximately 10 to 15 liters of vinasse (Moraes, Zaiat e Bonomi, 2015a).

The technological advances and research development during the 1980s and
1990s, allowed the 1G vinasse to be used in bio-fertigation processes, which apparently
solved the problem of disposing this residue by reducing the amount of chemical
compounds (nutrients) used by the sugar industry for soil fertilization. However, the use
of vinasse for bio-fertigation is still a controversial topic regarding the benefits for
sugarcane cultivation and the consequences for the environment. Actually, it is not
possible to state that this final destination, although permitted by federal laws, is free from
environmental problems (Moraes et al., 2014).
On the other hand, the use of SB as solid fuel generate a by-product called ashes
whose qualitative composition and applications, as presented in Table 2.1, vary according
to the material and parameters of incineration (e.g. temperature, incineration time and
humidity of the incinerated material) used, confirming that such residue is heterogeneous
in composition and morphology (Cacuro, Thiago A. e Waldman, 2015). The average
amount of ashes generated by the SB combustion is 25 kg for each ton of dry incinerated
bagasse (Sales e Lima, 2010) so that for a medium-sized plant (between 1 and 2 million
of sugarcane processed per season) about 1800 tons of ashes per season are generated.
This confirms that proper management of ashes is indispensable to minimize the
associated environmental impacts
21
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Table 2.1 Main residues formed in the sugar and ethanol production process.
Wastes Main composition Usual destination Technological approach used References

Filter cake High concentrations of Fertilizer for - Biogas production (Leite, Janke, Lv, et al., 2015); (Janke, L, Leite, A.,
calcium, nitrogen and sugarcane growth et al., 2016);(Moore, Nogueira e Kulay, 2017);
potassium with varying - Metal adsorption (Barros et al., 2017); (Santos et al., 2016); (López
compositions depending on González, Pereda Reyes e Romero Romero, 2017)
sugarcane variety

Excess Yeast Vitamin B, Proteins and Sold as animal - Nutritional supplements (Bergmann et al., 2018); (Victral et al., 2016);
Extract Aminoacids feed - Redox mediators for dye (Bahia et al., 2018); (Bagudo et al., 2010)
removal

Vinasse Nitrogen, Phosphorus, - Fertilizer for - Biofuels: biogas (methane and (Hoarau et al., 2018)
Potassium, Sulfates, among sugarcane growth hydrogen) and biodiesel;
others - Fertigation - Enzymes;
- Treatment for - Organic acids and alcohol with
organic matter high value added;
removal - Humus (composting)

Ashes High concentration of silica, Discarded in - Sand and silica substitute in (Chingono et al., 2018);
oxides of Ca, Mg, Mn, K, Al landfills or used as the preparation of concretes and (Ferreira et al., 2015); (Cacuro, Thiago A. e
as well as P, Na and S. fertilizer in ceramics Waldman, 2015); (Sudibyo et al., 2018)
Trace metals and unburned sugarcane - Adsorption (Teixeira et al., 2014); (Sales e Lima, 2010)
carbon. plantation - Catalytic processes

22
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Table 2.2 Physicochemical characteristics of vinasse obtained from annexed and


autonomous ethanol production in sugar-alcohol biorefinery. (Marafon et al., 2020;
Santos et al., 2019)

Sugarcane vinasse
Parameters From annexed ethanol From autonomous
production ethanol production
pH 4.63 4.50
CODsoluble (g/L) 35.36 14.51
BOD (g/L) 23.18 8.79
N (g/L) 0.48 - 0.71 0.45-1.6
SO42- (g/L) 2.0 1.22
P (g/L) 0.09-0.2 0.2-3.03
K (g/L) 3.34-4.0 3.74-7.83

The residue called filter cake, generated in the filtration and clarification of the
extracted sugarcane juice, is normally applied as fertilizer in the fields to grow sugarcane.
Table 2.1 illustrates some other technological routes (e.g. production of adsorbents and
biogas) that can be used depending on the compositional characteristics of the cake. The
application of filter cake as a substitute for traditional potassium inputs during sugarcane
grow provides good results for agriculture (Moore, Nogueira e Kulay, 2017). However,
the application and storage process must be controlled since this material also has a high
organic load and is a potential source of pollution (Albarelli, 2013).

2.1.2 Second generation ethanol production process (2G)

It is estimated that by burning about half of the sugarcane bagasse generated by


the sugar-alcohol industry is enough to provide the heat and energy to run the production
process (Silva et al., 2020). Therefore, the remaining (excess) fraction of sugarcane
bagasse that is not burnt and used in the CHP cycle of the 1G process can be used for 2G
ethanol production. As shown in Figure 2.2, some unit operations and processes must be
added in order to process lignocellulosic biomass (LB) and obtain more biofuel.

23
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Consequently, to make feasible the use of this material for such purpose, a pretreatment
stage of LB, capable of separating lignin and hemicellulose (C5 sugars) from cellulose,
is necessary (Figure 2.2).

The pretreatments used for biomass deconstruction can be classified as physical,


chemical or biological (Zheng et al., 2014). They allow a better use of LB by fractioning
its main components in order to obtain high levels of fermentable compounds. Thus,
pretreatment processes represent a significant cost in the use of lignocellulosic waste as
raw material for fuel production processes (ethanol, biogas) or bioproducts (Santos et al.,
2020a). In view of this, the search for a pretreatment that offers low cost and guarantees
high performance in bioprocesses is a very important issue.

Among the pretreatment techniques available, the so called liquid hot water
(LHW) or hydrothermal (HT) or autohydrolysis pretreatment is highlighted. Such process
allows the solubilization of hemicelluloses, and consists in the use of water that acts as a
solvent and nucleophile in the hydrolysis reactions of hemicelluloses at high temperatures
(160 to 230 ° C) and pressures ( > 5 MPa) (Ruiz, Thomsen e Trajano, 2017). The main
advantages of this technique are the non-addition of chemical/aggressive compounds such
as acids, thereby dispensing the use of reactors built with corrosion resistant materials.
This makes the technique more economically and environmentally attractive for large-
scale applications (Baruah et al., 2018).

The fermentable sugars generated in the HT process, mainly pentoses, can be


directly used for the production of 2G ethanol as long as genetically modified
microorganisms or wild lineages isolates with co-fermentation capacities are employed
in the fermentation stage, which makes the process less feasible or attractive (Wertz et
al., 2018)(Cadete et al., 2017; Valinhas et al., 2018). On the other hand, an alternative for
extracting energy from this fraction is the production of biogas using anaerobic systems.
In this sense, Baêta et al., (2016) showed that high biogas production efficiencies can be
obtained if macro and micronutrients are added to hemicelluloses hydrolysate.

24
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

An important drawback for the use of pentoses liquor, generated by HT of SB, is


the formation of toxic or inhibitory compounds such as 2-furfuraldehyde (FF) and 5-
hydroxymethyl-2-furfuraldehyde (HMF). The presence of these compounds is usually a
limiting factor for bioconversion processes that use pentose liquor as a substrate (Barakat
et al., 2012).Thus, techniques that allow the removal of toxic compounds may be
necessary during the production process aiming its best application.

Other inhibitory compounds generated during HT of SB are lignin degradation


products. A variety of aromatic, polyaromatic, phenolic and aldehyde compounds can be
produced from this fraction of SB. Phenolic compounds are highly toxic for
microorganisms and cause changes in cell division and loss of integrity in biological
membranes, thereby affecting their ability to act as selective barriers and enzyme
matrices. As a consequence, cell growth and membrane assimilation of key compounds
are affected (Baêta, 2016).

Finally, in the case of vinasse resulting from sugarcane bagasse processing to


obtain ethanol (process 2G), it is possible to find some different characteristics although
the fermentation process it is carried out in the same vat. This is because of undigested
residue obtained after enzymatic hydrolysis if often composed mainly of lignin, while the
composition can vary depending on the method of pretreatment (Mathew et al., 2018).In
this sense, one of the main differences is the high levels of COD found in this vinasse,
which can be 4 to 5 times higher when compared to 1G vinasse. In addition, other factors
that are highlighted for this residue are the presence of toxic compounds such as phenolic
derivatives present in the raw material; the low or almost zero nutrient content (nitrogen,
phosphorus and potassium), which prevents the 2G vinasse be used for fertigation in
sugarcane crops (Cammarota et al., 2012; Moraes, Zaiat e Bonomi, 2015b).

ANAEROBIC DIGESTION
Anaerobic digestion (AD) is a biological process in which anaerobic
microorganisms convert organic matter into biogas. AD can be performed in one or more
25
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

reactors, known as digesters. Anaerobic degradation of organic matter is a biochemically


complex process that is subject to the formation of several compounds and intermediate
reactions, catalyzed by different microorganisms and enzymes.
The anaerobic digestion process can be divided into four stages: i) hydrolysis, ii)
acidogenesis, iii) acetogenesis and iv) methanogenesis. The phases are performed by
different types of microorganisms (bacteria and archaea) that work under conditions of
syntrophic relationship with each other (Aquino et al., 2019). Normally, single stage
digestion (1 reactor) is prone to be subjected to operational (e.g. organic or hydraulic
overload) and environmental (e.g. low temperature, sudden presence of toxic compounds
in feed) conditions that may trigger and result in lower methane generation due to the
accumulation of volatile fatty acids. VFA accumulation leads to lower pH and worsens
the growth conditions for methanogenic and acetogenic microorganisms due to
thermodynamic inhibition (Aquino e Chernicharo, 2005), which may result in complete
reactor failure. To solve this problem, two-stage reaction systems have been used, that is,
the metabolic processes of acidogenic and methanogenic microorganisms are spatially
separated. Generally, digestion of the initial stage (acidogenic) is operated at lower pH (5
to 6) and shorter hydraulic detention times (0.2 to 0.4 days) while the second phase
(methanogenic) is performed at pH close to neutrality (6.8 to 7.2) and higher hydraulic
retention time (0.5 to 2 days) (Siddique e Wahid, 2018).
At the end of the last century, several commercial AD processes were developed
in order to be used for the treatment of 1G vinasse. However, they were rejected by the
sugar and alcohol industry because the generation of energy from biogas was not
economically viable (Manochio, 2015). Nevertheless, the academic sectors continued to
develop research to improve process conditions to make it feasible the production of
energy from biogas generated by the anaerobic digestion of vinasse. In this sense, the
most significant contributions in this area are reported by España-Gamboa et. al (2011),
Cabrera-Díaz et al., (2017), Fuess, L T et al., (2017), Fuess et al., (2018), and in Hoarau,
Julien et al.,(2018).

26
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

2.2.1 Anaerobic Co-Digestion

Anaerobic co-digestion (AcD) is an alternative that can be used to enhance energy


and value-added byproducts recovery from AD, minimizing environmental impacts in the
1G and 2G ethanol production chain thereby contributing to the sustainability of the
sector (Moraes, Zaiat e Bonomi, 2015a), (Navaneethan et al., 2011). This process consists
in mixing two or more different substrates (e.g. hemicelluloses hydrolysate + vinasse +
excess yeast extract + sugarcane bagasse boiler ashes) into the anaerobic reactor with the
aim of increasing its stability and biogas production and enhancing its energy balance by
strengthening the system buffer capacity and C / N ratio.
The success of this alternative is based on the fact that mixing different substrates
can guarantee an ideal condition for the growth of anaerobic microorganisms. This is
because co-digestion of different substrates often supplements the medium with macro
and micronutrients that are present in the components of the mixture. In addition, co-
digestion of different substrates allows compounds considered individually toxics for the
anaerobic microorganisms, such as furfurals, phenols and soluble lignin fragments
present in the hemicelluloses hydrolysate, to be diluted to concentrations lower than those
that cause toxicity or inhibition to the microbial consortium (Maison, Nava e Peláez,
2009).
A schematic diagram for anaerobic co-digestion process is shown in Figure 2.3.
Some important properties must be considered in order to take full advantage of co-
digestion, including the co-substrates characteristics, the inhibitions induced by them and
their applied organic loads (Shah et al., 2015). In addition, for enhancing system synergy
and optimizing methane production, it is essential to select and blend appropriate
proportions of co-substrates.

27
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 2.3 Conceptual flow diagram of anaerobic co-digestion processes (Adapted from
(Siddique e Wahid, 2018))
The most important factors for the optimization of AcD are the chemical
properties of the substrates used, the initial C/N ratio, the use of additives as well as the
values of process parameters such as temperature, pH, hydraulic retention time (HRT),
and applied organic loading rate (OLR). Initially, the complete knowledge of the
physicochemical properties (carbohydrates, proteins and lipids) of the substrates allows
one to understand their biodegradability, bioaccessibility and bioavailability (Hagos et
al., 2017). For instance, substrates rich in sugars and other carbohydrates may lead to
volatile fatty acids (VFAs) accumulation, thereby inhibiting acetogenesis and limiting
subsequent methanogenesis.
The relationship between the amount of carbon and nitrogen present in the raw
materials (C/N ratio) is extremely important for AcD since insufficient amounts of carbon
or nitrogen can limit process performance. In other words, a high C/N ratio provides
insufficient nitrogen to maintain active cell biomass, resulting in lower biogas production,
28
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

and a low C/N ratio may increase the risk of ammonia accumulation, which is toxic for
methanogenic archaea (Tabatabaei e Ghanavati, 2018). On the other hand, it is difficult
to define precisely which ratio is optimal as the best C/N value can be affected by different
factors such as substrate type, trace element composition, chemical composition and
biodegradability (Shah et al., 2015).
Temperature is one of the main factors for the survival of microorganisms during
AcD processes. There are three operating ranges that can be used in anaerobic reactors:
psychrophilic (25 °C), mesophilic (around 35 °C) and thermophilic (around 55 °C).
Normally, microorganisms can grow better in the mesophilic and thermophilic
temperature ranges so that an increase in temperature generally has a positive effect on
their metabolic rate which accelerates the digestion process. On the other hand, the
thermophilic process is more difficult to control and requires more energy to keep the
reactor temperature constant. It is known that significant temperature changes can
adversely affect microbial growth and reduce biogas production (Chernicharo, 2007).
pH is another parameter that significantly affects the anaerobic process since
enzymatic reactions of microorganisms depend on this parameter. For biogas production,
different microorganisms require different optimum pH values for growth, although at
neutral pH the overall better condition is attained. Indeed, maintaining optimal pH values
for each group of anaerobic microorganisms (acidogenic phase at pH around 5 to 6;
methanogenic phase at pH around 7), was the main reason for developing two-stage
reactors (Siddique e Wahid, 2018). Alternatively, to avoid problems with pH variation
during anaerobic digestion, some authors recommend the use of alkaline chemicals such
as CaCO3, NaHCO3 or NaOH, resulting in a significant increase in operating costs.
According to some authors, alkalinity supplementation can become a bottleneck for
implementation of AcD at the industrial scale (Boncz et al., 2012; Fuess, L.T., Araujo JR,
et al., 2017; Janke, L. et al., 2016).
Several aspects of reactor design employed for AcD processes are also critical. It
is important for a large-scale process to operate the reactor at high OLR and low HRT,
thereby ensuring optimal biogas yields and lower costs. In this sense, different reactor
configurations, such as the one or two stages configurations; continuous or batch feeding
29
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

arrangements; as well as low and high rate systems have been investigated, as described
by Shah et al., (2015). In addition, reactors that use biomass immobilization are very
interesting because they have longer cell retention time, which improves both reactor
performance and stability, even though a lower HRT is used (Haandel, Van et al., 2006).
Finally, it is well known that AcD and AD are universal technologies with
significant advantages for energy generation from waste; and the need to ensure the
economic viability of treatment facilities has been drawing attention to process
improvement techniques (Romero-Güiza et al., 2016). This can be done, for instance, by
dosing additives to improve microbial activity and/or reduce concentrations of inhibitory
compounds. Among the additives of AD one can highlight the supplementation of trace
metals (Fe, Ni, Mo, Co, W and Se) that are indispensable cofactors in enzymatic reactions
involved in methane formation (Janke, L. et al., 2016). In addition, the use of
nanoparticles and additives that promote the immobilization of biomass and/or the
adsorption of inhibitory compounds (activated carbon, polymers and zeolites) have been
investigated in recent years. Such strategies are viable alternatives to improve the
performance of bioreactors whose costs and effects are still a matter of debate amongst
specialists (Romero-Güiza et al., 2016).
Some studies show, for example, that powdered activated carbon (PAC) can help
in the adaptation of microorganisms, improving reactor stability by adsorbing inhibiting
compounds and enhancing organic matter degradation (Hu E Stuckey, 2007; Akram E
Stuckey, 2008; Baêta et al., 2013; Santos, 2017). In addition, due to its low selectivity
and high adsorption capacity, PAC adsorbs recalcitrant soluble compounds, increasing
their retention time in the reactor; which leads to better COD removal and higher methane
production (Baêta, 2012; Martinez et al., 2017).

2.2.2 Anaerobic co-digestion of bioethanol production wastes

Considering the technological possibilities applicable to treat residues from first-


and second-generation ethanol production processes, Table 2.3 shows there are a wide
variety of alternatives. It is seen that vinasse can be co-digested with substrates such as
straw, manure, rumen, coffee wastes, glycerol, whey, filter cake, among others.
30
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

According to the literature review, it is possible to highlight the use of different reactor
types and configurations such as completely stirred tank reactor (CSTR), Up-flow
anaerobic sludge blanket reactor (UASB), anaerobic sequential batch biofilm reactor
(AnSSBR), Up-flow fixed film reactor (UAFF) operating in batch, semi-batch and
continuous feed regimes. Moreover, it is seen there is no definite trend regarding the
addition of nutrients and additives as well as proportion of the waste mixtures used,
thereby indicating that co-digestion is still a poorly mature technology in this sector.
Consequently, a large number of questions must be answered to achieve technical,
economic and environmental feasibility (Cesaro e Belgiorno, 2015; Moraes et al., 2014;
Moraes, Zaiat e Bonomi, 2015b).
Among the most significant works, it is important to highlight the results and
operating conditions developed by Lovato et al., (2018), López González, Pereda Reyes
Romero Romero, (2017) and Albanez et al., (2018). These authors studied residues from
the sugar-alcohol industry and investigated changes in the proportion of substrate mixture
(0-100%) as well as the addition of nutrients and support media on reactor efficiency
aiming the production of methane and hydrogen. On the other hand, among the reviewed
works, only that proposed by Uellendahl e Ahring, (2010) used effluents from the 2G
ethanol production process, mainly the effluent from yeast fermentation whose substrate
was wheat straw pretreated, in order to obtain biogas using a thermophilic UASB reactor
inoculated with granular sludge from reactors treating effluent from a brewery.
Finally, no studies were found that integrate - via AcD - pentose liquor (generated
by hydrothermal SB pretreatment) with 1G vinasse. Moreover, the use of excess yeast
extract as source of nutrients and the addition of sugarcane bagasse ashes as an alkalinity
agent, trace metal source and support medium in anaerobic systems has not yet been
reported in the literature. Thus, research using the anaerobic co-digestion of vinasse,
hemicelluloses hydrolysate and other residues from sugarcane industry, with the aim of
producing biogas (CH4), can be regarded as novel and strategic. This technological route
could allow the environmental adaptation of these effluents and increase the energy
production of the sugar and alcohol sector, hence contributing to ensure technical,
environmental and economic viability of the integration of 1G-2G ethanol production chains.
31
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

Table 2.3 Anaerobic co-digestion studies of vinasse with different substrates (to be continued).
Co-substrates Reactor type and operational conditions System characteristics Target final Main results References
product
Effluent from UASB, T: 53°C and 38°C, OLR: 3.5 kg VS m -3d-1, V: COD/VS: 2.2 CH4 YM: 270 to 340 L CH4 kg VS-1 (Uellendahl e Ahring,
ethanol 770 L, 310 days of operation, NaOH for alkalinity 2010)
fermentation
(wheat straw and
granular sludge)
Vinasse (beet)- CSTR, T: 37°C, V: 20L, HRT: 30 to 36 d, limestone CODV:421.6 to 541.9 CH4 YM: 235.7 to 267.4 L CH4kg VS-1 (Moraes et al., 2015)
manure-straw for alkalinity, addition of nutrient solution, OLR: 2- gL-1, CODeffluent: 39 gL-
3 g SV L-1 d-1 1
, VSTS-1 straw: 90.63
%

Vinasse(wine)- UAFF continuous, T: 55°C, V: 3.45 L, OLR: 14.9 to CODV: 10 gL-1 CH4 Removal efficiency: COD: 87% (Perez et al., 2006)
cutting oil 22.3 kg CODm-3d-1, HRT: 0.15 to 0.80 d, NaOH for CODCOW: 2.5 gL-1 YM: 0.034 to 0.003 m3 CH4g
wastewater alkalinity, the support medium is sintered glass Mixture composition: COD-1 removed
(COW) beads 57 to 43% and 33 to
67%
Vinasse (V)- Batch, T: 37°C, V: 1L CODV: 22 gL-1 CH4 Removal efficiency: COD: 60% (Rodríguez et al., 2007)
Waste active CODWAS:33 gL-1 Methane rate generation
sludge (WAS) Mixture: V-WAS from 14.34 g CH4 g COD-1 day-1
0-100% to 100-0%
Vinasse Batch (30 days), T: 25°C, V: 0.6 L CODV:299 gL-1 CH4 YM: 6.49 NmL CH4g VS-1 (Iqbal Syaichurrozi and
(Ethanol)-Rumen- 1 g of urea as N source, Sumardiono, 2014)
Urea 10% Rumen as
inoculum

32
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

Table 2-3 Anaerobic co-digestion studies of vinasse with different substrates (continuation).
Co-substrates Reactor type and operational conditions System characteristics Target final Main results References
product
Vinasse--Sugar Enzymatic pretreatment, batch (30 days), T: 37°C, CODV: 978.23 g kg TS-1 CH4 YM: 769 to 499.7 mL CH4g VS-1 (Ziemiński e Kowalska-
beet pulp silage V: 1 L, Na2CO3 for alkalinity, agitation: 4 rpm CODSBPS: 1020.5 g kg
Wentel, 2015)
(SBPS) TS-1
Mixtures: 1:3.1:1,3:1
of CODV : CODSBPS
Vinasse (rice CSTR, T: 35°C, V: 9L, HRT: 20 d, agitation: 60 rpm, Mixture: 1:1, VSTS-1 CH4 Removal efficiency VS: 55% (Dai et al., 2015)
wine) – dead Vinasse: 87.5%, VSTS-1 YM: 0.4 L CH4kg VS-1
dead animals (thermal pretreated), OLR: 6.8 kg
animals (pork) meat pork: 95.5%
VSm-3 d-1
Ethanol AnSBBR semi-batch and batch, V: 2L, T: 30°C, T of CODV: 25 gL-1 CH4 Removal efficiency COD: 79- (Lovato et al., 2018)
vinasse(V)-Whey cycle: 8 h, polyurethane as support material, OLR: CODW: 1 gL-1 87%
(W) 6.4 kg COD m-3 day-1, Mixture V-W from 0- YM: 289.4 to 204.16 NmL CH4g
NaHCO3 for alkalinity 100% and from 100- COD-1
0%
Vinasse -Glycerol- Batch, T: 37 °C, V: 1 L Mixtures of 1.73 kg CH4 YM:208-305 L CH4 kg VS-1 add (Gil et al., 2017)
manure-sludge- COD kg VS-1
vegetable waste
– leachate
Vinasse- Pretreatment with ultrasound, semi-batch, T: CODV: 51.88 gL-1 H2-CH4 60% H2-40% CH4 and 20% H2 (Martinez-Jimenez et
sugarcane straw 55°C, V: 4.3 L, agitation: 500 rpm, NaOH for VS straw: 94.56 % combined 80% CH4 (v/v)
al., 2017)
alkalinity
Vinasse–Coffee CSTR, T: 55 °C, HRT: 55 d, OLR: 0.16 to 0.37 kg VS CODV: 51.88 gL-1 CH4 YM: 0.140 to 0.210 L CH4 g VS-1 (Pinto et al., 2018)
wastes (CW) m-3 d-1 CODCW: 426 gL-1
Mixture: 70%:30%
(V:CW)

33
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

Table 2-3 Anaerobic co-digestion studies of vinasse with different substrates (continuation)
Co-substrates Reactor type and operational conditions System characteristics Target Main results References
final
Product
Vinasse-Food Semi-batch (77 days), T: 37°C, V: 4.3 L, 2 stages, agitation: 50 rpm CODV: 21.8 gL-1 CH4 YM: 320 NmL (Náthia-Neves
waste (FW) CODRR: 61 gL-1 CH4 g VS-1 et al., 2018)
Mixture: 80-20% (V:FW) YH: 300 NmL
CH4 g VS-1
Vinasse-filter Batch (30 days), Semi-batch (112 days), T: 37°C, V: 0.5 L (batch)- COOV: 62.3 gL-1 CH4 YM: 0.231 to (López
cake (press mud) 50 L (semi-batch), OLR: 0.7-3 g COD L-1 d-1 COOPM: 284.5 gL-1 Mixture: from 0- 0.248 NL CH4 g González,
100% and from100-0% V-PM COD-1added Pereda Reyes e
YM: 172.9 to Romero
368.2 NL CH4 Romero, 2017)
kg VS-1 (semi-
batch)
Vinasse-Manure Batch (22 days), V: 0.6L, NaOH for alkalinity CODV: 38.8 gL-1 CH4 YM: 63.9 mL (Silva e Abud,
VS Manure: 80% CH4 g COD-1 2017)
Manure 0.5;3;5.5 %

Vinasse Batch (6 days), T: 35°C, V: 4L, nutrient solution (trace metals) CODV: 59.7 gL-1 H2 YH: 0.227- (García-
(Tequila)- CODN: 25.1 gL-1 0.2397 NL H2 g Depraect et al.,
Nejayote Mixture from 0-100% and from COD-1removed 2017)
(wastewater 100-0% V-N
from Mexican
corn tortilla
processes)

34
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

Table 2-3 Anaerobic co-digestion studies of vinasse with different substrates (continuation).
Co-substrates Reactor type and operational conditions System characteristics Target final Main results References
product
Vinasse- Brewery Batch (18 days), T: 50°C, V: 1L Mixture 1:1 (V- H2 YH: 4 mL H2 g TS-1 (Xin et al., 2017)
sludge Sludge), 10 % TS

Vinasse-Molasses AnSBBR, T: 30°C, cycle: 3 to 4 h, agitation: 200 CODV: 25 gL-1 H2 Removal efficiency COD: 20% (Albanez et al., 2018)
rpm, polyurethane as support material, nutrient CODMolasses: 1100 gL-1 YH: 29.1 NmL H2 g COD-1
solution, OLR: 5.6 g COD/L d Mixture of V-
Molasses: 0 to 100%
and 100 to 0%
Vinasse Batch, CSTR, Semi-batch (24 h), V: 0.3 L, T: 60°C, CODV: 2048 mg g VS-1 H2 and CH4 YH: 57.8 to 74 NmL H2 g VS-1 (Wang et al., 2011,
(cassava)-waste Agitation: 120 rpm, 2 stages. CODCassava sludge: 1945 YM: 326.5 to 361 NmL CH4 g VS- 2012, 2013)
sludge (cassava)- mgg VS-1 1

bovine-piggery CODbovine manure: 1312


manure- Waste mgg VS-1
active sludge CODPiggery manure: 1115
(WAS) mgg VS-1
CODWAS: 1750 mgg VS-1
Mixture: 7:1 (Vinasse-
Co-substrate)
Vinasse Semi-batch (640 days), T: 37°C, V: 5L, HRT: 45 d, Mixture: 85-15 % (V- CH4 YM: 0.311 NL CH4 g VS-1 (Westerholm, Hansson
(Fermented OLR: 2.8 g VS L-1d-1, nutrient solution (trace Manure) based on VS e Schnürer, 2012)
cerals) -Manure minerals)
UAFF= Up-flow fixed film reactor, YM= Methane yield, YH= Hydrogen yield, VS= Volatile solids, TS= Total solids, OLR=organic load rate, AnSBRR=
anaerobic sequential batch biofilm reactor, UASB= Up-flow anaerobic sludge blanket reactor

35
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

CONCLUSIONS FROM LITERATURE REVIEW - THESIS PREMISES


Annually large amounts of lignocellulosic residues are discharged to the
environment as a result of agricultural or agroforest production. Such destination
represents a loss in potential of energy production, besides generating, in some cases,
environmental impacts. In this context, Brazil takes a prominent position, due to the fact
it is the third largest agricultural exporter (OMC , 2010) and generates a large amount of
bio-waste from the agricultural sector. To put it in context, the National Supply Company
(CONAB) reported that for the 2019/2020 biennium the total production of sugarcane
destined for the sugar-alcohol industry was 642.7 million tons (Companhia Nacional de
Abastecimento (CONAB, 2019).
During the 1G ethanol production process, a large amount of liquid effluent
(called vinasse) is obtained from the distillation of the fermented wine. This effluent can
become an environmental hassle depending on its final destination due to the high organic
load and volume generated. According to data from CONAB, the ethanol production
estimated for the 2019/20 harvest is 33.8 billion liters, representing an increase of 4.6%
when compared with the previous harvest (CONAB, 2019). Consequently, the amount of
vinasse generated in this season could reach the mark of 405.6 billion liters (Moraes, Zaiat
e Bonomi, 2015b).
Regarding the possibility of expanding ethanol production, the production of 2G
ethanol appears as an interesting alternative, since it would be derived from sugarcane
bagasse generated in the 1G ethanol production process. In this sense, the volume of
ethanol produced could be increased without the need to expand the area for sugarcane
growth. However, in order to attain the full sustainability of such production process, the
management of residues from sugarcane bagasse pretreatment, particularly the
hemicellulose hydrolysates, must be well understood to minimize and avoid future
environmental problems (Eggert e Greaker, 2014; Macrelli, Mogensen e Zacchi, 2012a;
Moraes, Zaiat e Bonomi, 2015b).
Nowadays, fertigation is method used by the sugar-alcohol industry for the final
disposal of vinasse. Although this management has positive impacts on crop growth, it
also has many negative environmental impacts, adversely affecting soil, groundwater and
36
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

atmosphere. Anaerobic digestion of vinasse for biogas production is a relatively well


understood process in which operational improvement can be made through the adoption
of different bioreactor configurations. However, this process presents economic and
technological bottlenecks for full scale implementation due to the need of using chemicals
that help in the nutritional balance, to buffer the system (alkalinity) and to minimize the
problems related to sulfur in the production streams.
On their turn, other wastes from the bioethanol industry have interesting
characteristics for anaerobic digestion applications. As examples, excess or discharged
yeasts from fermentation vats have high levels of nitrogen, vitamins and minerals; and
the ashes from bagasse burning boilers have metallic oxides, mainly alkaline, silica and
carbon (inert, amorphous) that could improve anaerobic digestion when used as additives
to increase alkalinity and trace metals.
According to the aforementioned, it is evident that the energetic, economic and
environmental sustainability of the integrated production of 1G-2G ethanol must consider
the better use and destination for residues (e.g. sugarcane bagasse, vinasse,
hemicelluloses hydrolysate, excess/spent yeast and sugarcane bagasse boiler ashes)
generated in both processes. A friendly alternative for the correct use and disposal of these
residues could involve the use of anaerobic co-digestion for the production of energy as
biogas (CH4 and / or H2). This thesis aims to investigate this possibility in order to
guarantee processes efficiency and cost minimization for the final disposal of sugar-
alcohol industry residues.

37
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

3 HYPOTHESES

CENTRAL
Mesophilic anaerobic co-digestion of residues produced during first
(1G) and second generation (2G) ethanol processes is technically and
economically sustainable and contributes to the integration (1G + 2G) of
sugarcane biorefinery.

COMPLEMENTARY HYPOTHESES
CH.1 - Addition of sugarcane bagasse boiler ashes and excess yeast
(as extract) in the anaerobic digestion of residual streams of 1G (vinasse) and
2G (pentoses liquor) ethanol production, allows the improvement of C / N
ratios, nutrient (macro-micro) and alkalinity supplementation, thereby
improving biogas yields;
CH.2 - Separation of acidogenic and methanogenic stages for
anaerobic co-digestion of hemicelluloses hydrolysates, vinasse, ashes and
excess yeast extract helps to increase the organic biodegradability and biogas
production in continuously fed reactors;
CH.3 - Physicochemical characteristics of sugarcane bagasse boiler
ashes enhance anaerobic digestion by the adsorption of inhibitors compounds
present in vinasse and hemicelluloses hydrolysate, hence improving the
production of biogas in the acidogenic (hydrogen) and methanogenic
(methane) reactors.

38
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

4 OBJETIVES

GENERAL
Evaluate the potential of biogas production by anaerobic co-digestion of
wastes (vinasse, excess yeast extract, sugarcane bagasse boiler ashes and
hemicelluloses hydrolysate) from first- and second-generation ethanol production
process.

SPECIFIC
1. Determine, on a batch bench-scale, the influence of hemicelluloses
hydrolysate and vinasse composition on CH4 production during mesophilic
anaerobic co-digestion without nutrient supplementation;

2. Verify the effect yeast extract and boiler ashes addition, as sources of
nutrients and alkalinity, on CH4 production during mesophilic anaerobic co-
digestion of vinasse and hemicelluloses hydrolysates;

3. Investigate the kinetics of anaerobic co-digestion of vinasse,


hemicelluloses hydrolysate, excess yeast extract and sugarcane bagasse boiler
ashes from first- and second-generation sugarcane ethanol production process;

4. Evaluate the stability and performance of biogas production, in single


and two stage anaerobic reactors continuously fed with a mixture of vinasse-
hemicelluloses hydrolysate and inoculated with excess yeast extract and
sugarcane bagasse boiler ashes, using optimal predetermined batch conditions;

5. Perform technical and economical analyses of the anaerobic co-


digestion process of hemicelluloses hydrolysate and vinasse according to the
concept of integrated sugarcane biorefinery.

39
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

5 USE OF ANAEROBIC CO-DIGESTION AS AN ALTERNATIVE TO ADD


VALUE TO SUGARCANE BIOREFINERY WASTES

ABSTRACT

In this chapter we present data that aimed to address the objectives 1,2 and 3
(section 4.2) and test the hypotheses CH1 and CH3 (section 3.2) of this thesis. In this
study the anaerobic co-digestion (AcD) of sugarcane biorefinery byproducts, i.e.
hemicelluloses hydrolysate (HH) (obtained by hydrothermal pretreatment of sugarcane
bagasse), vinasse (generated during 1G ethanol production process), yeast extract (YE)
(which can be obtained from spent/excess yeast) and sugarcane bagasse fly ashes (SBFA),
was optimized by means of biochemical methane potential (BMP) experiments. The best
experimental conditions of AcD (25-75% HH-to-vinasse mixture ratio; 1.0 g L-1 YE;
15 g L-1 SBFA and 100-0% HH-to-vinasse; 1.5 g L-1 YE; 45 g L-1 SBFA) led to the
production of 0.279 and 0.267 Nm3 of CH4 per kg of chemical oxygen demand (COD)
removed with an energy surplus of 0.43 and 0.34 MJ kg SB-1, respectively. Adsorption
experiments using SBFA were carried out and showed this residue could adsorb up to
61.71 and 58.21 mg g-1 of 5-Hydroxymethyl-2-furfuraldehyde and 2-furfuraldehyde,
thereby reducing toxicity and improving biogas production during AcD of the
aforementioned byproducts.

INTRODUCTION

Increasing global demand for energy and the environmental impacts associated to
the use of fossil fuels have increased the attention for cleaner, safer and cheaper energy
sources. Energy sources such as ethanol, biodiesel and biogas which can be obtained from
lignocellulose biomass (LB) are some of the options to supply part of the global energy
demand.
In the Brazilian context, sugarcane is the main biomass source for the production
of first-generation (1G) ethanol from sugarcane juice. According to the Brazilian National
Supply Company (CONAB), the 2017/2018 sugarcane harvest season was about 633.2
40
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

million tons (CONAB, 2018). The sugarcane processing generated 37.87 million tons of
sugar, 27.76 billion liters of ethanol and sugarcane bagasse (SB) at a rate of 250-280 kg
SB (with 50% moisture)/ton sugar cane milled. Currently, most (65-75%) of the SB
generated in the mills is burned into boilers for steam and power generation (Bergmann
et al., 2018). Therefore, the remaining SB could be used to produce second-generation
(2G) ethanol.
In Brazil, the main SB fractionating route currently used by biorefineries for
second generation ethanol (2G) production consists of: (1) pretreatment of SB to separate
hemicelluloses from cellulignin; (2) hydrolysis of cellulignin to obtain fermentable
sugars; and (3) fermentation of C6 sugars into ethanol, followed by distillation (Rocha-
Meneses et al., 2017). Such steps in a 2G plant generate a liquid stream called
hemicellulose hydrolysate or pentoses-rich liquor, which is obtained after SB
pretreatment, in addition to the known liquid and solid residues of 1G plant, i.e. vinasse,
spent yeasts, filter cake and ashes (Rebelato et al., 2019). These by-products could be
used as raw materials for the production of value-added chemicals and biofuels (O’Hara
e Mundree, 2016). One option to add value to the solid and liquid streams from 1G and/or
2G plant is to use anaerobic co-digestion (AcD) to recover energy as biogas (methane and
hydrogen).
Vinasse is most studied because of its polluting potential and the large volume
generated per liter of ethanol produced (≈10-15 liters) (Hoarau, J. et al., 2018). However,
biogas produced directly from vinasse fermentation in anaerobic reactors has high sulfur
content (mainly H2S from sulfate reducing bacteria), which is corrosive for engines and
contributes to sulfur emissions to atmosphere, thereby requiring specific biogas treatment
for its removal (Colturato, 2015). On the other hand, the hemicelluloses hydrolysate (HH)
produced in the thermal pretreatment of SB has low sulfur content and can be used for
biogas production since it is not straightforward to obtain ethanol from C5 sugar
fermentation (Baêta, Lima, Adarme, et al., 2016).

41
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Another residue from ethanol/sugar production is the sugarcane bagasse fly ash
(SBFA) whose production is estimated at about 6 kg of ash per ton of sugarcane
processed. The most common use for this solid residue is as plant fertilizer and soil
amender, even though with significant drawbacks (Rodríguez-díaz et al., 2015). On the
other hand, recent researches have suggested that SBFA has good adsorptive properties
(Chingono et al., 2018) for various compounds and could be used as an inorganic additive
in the AD (Sudibyo et al., 2018).
At the moment, most of the studies available in the literature reported individual
uses for each residue generated by the sugar-alcohol industry and to the best of our
knowledge there is no published paper that evaluated the influence of all these residues
combined on the anaerobic digestion. Therefore, the aim of this study was to evaluate the
methane production from the AcD of both liquid (HH and vinasse) and solid residues
(yeast extract (from spent yeasts) and SBFA) generated by an integrated 1G and 2G
ethanol plant. The influence of substrate composition (vinasse-to-HH volumetric mixture
ratio) and C/N ratio on methane production, as well as the role of SBFA on the adsorption
of compounds toxic to the AcD were also investigated. Energy balances of the best
scenarios were estimated to show that the AcD of the aforementioned wastes can
contribute to close loops and improve the sustainability of the sugar-alcohol industry.

MATERIAL AND METHODS

5.3.1 Chemicals

Cyclohexane, ethanol (99.5%), hydrochloric acid (36-37 wt.% in H2O), sulfuric


acid (95-98%), sodium carbonate, and sodium hydroxide were purchased from Synth
(Brazil). Chromatography grade sulfuric acid (99.999%), nitric acid (0.034 mol L-1) /
dipicolinic acid (0.014 mol L-1) solution, sodium bicarbonate (0.1 mol L-1) solution and
chromatography standards (cellobiose, D-glucose, D-xylose, L-arabinose, acetic acid,
formic acid, propionic acid, isobutyric acid, butyric acid, valeric acid, isovaleric acid, 5-
42
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

hydroxymethyl-2-furfuraldehyde (HMF) and 2-furfuraldehyde (FF)) were purchased


from Sigma-Aldrich. Yeast extract (YE, powder) was purchased from HiMedia
Laboratories (Brazil). Licowax C Micropowder PM (Amide wax) was purchased from
Clariant (Brazil). Nitrogen (99.9999%) was purchased from White Martins Praxair
Technology Inc. (Brazil).

5.3.2 Sugarcane Biorefinery Wastes

Sugarcane bagasse (SB) was provided by Jatiboca Sugar and Ethanol Plant
(Urucânia, MG, Brazil), and it was sampled in the 2016/2017 harvest season. SB was air-
dried under sunlight to approximately 12% humidity and stored in a closed container.
Vinasse was sampled from a first-generation (1G) ethanol distillation system at Agropéu
Sugar and Ethanol Plant (Pompéu, MG, Brazil). SB fly ash (SBFA) was sampled after
SB combustion in a high-pressure boiler and was provided by the Minas Gerais Sugarcane
Industries Association (SIAMIG, MG, Brazil).

5.3.3 Hydrothermal Pretreatment of SB

Hydrothermal pretreatment of SB was carried out in a 2 L reactor vessel (Parr


Instruments, model 4848) using the optimized experimental conditions reported by Baêta
et al., (2016a), i.e. temperature of 183°C, reaction time of 41 min and liquid-to-solid ratio
of 3.94 mL/g SB. The reactor vessel was loaded with 120 g of SB, on dry-weight basis,
at each run. After the reaction time, the reactor vessel was quenched in a water-ice bath
and opened. The pretreated SB was dewatered in a hydraulic press (SOLAB, model SL-
10) at a pressure of 9 tons and the hemicelluloses hydrolysate (HH) was collected and
stored at -20°C.

5.3.4 AcD of Hemicelluloses Hydrolysate and Vinasse

Batch anaerobic co-digestion (AcD) experiments were performed in triplicate to


assess the biochemical methane potential (BMP) of the main substrates, i.e. vinasse and
hemicelluloses hydrolysate (HH). All experiments were performed at mesophilic
43
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

conditions in glass bottles of 120 mL. The glass bottles were placed in a thermostatic
orbital shaker incubator at 180 rpm and 35 ± 1°C. The liquid volume of each glass bottle
was 60 mL, leaving a headspace of equal volume. A substrate volume of 6 mL was used
in each experiment. The inoculum amount depended on the chemical oxygen demand
(COD) fed to each glass bottle in order to keep a desired food-to-microorganism ratio
(F/M). The effects of the substrate composition, i.e. mixture of vinasse and HH, and
inoculum amount were evaluated using an experimental design (Section 5.3.5). The initial
pH was adjusted to the 6.9-7.2 range using 2 mol L-1 aqueous HCl solution or aqueous
Na2CO3 solution. After adjusting the suspension pH, all glass bottles were sealed and then
purged with N2 for 3 min with a flowrate of 109.3 mL N2 min-1 in order to reduce the O2
concentration in the reaction flasks. All experiments were performed without the addition
of macro- and micro-nutrients as well as alkalinity. A control test, i.e. bottles incubated
without substrate, was also included in the experiments to evaluate the endogenous
activity of the inoculum.

The inoculum was a mixture of fresh bovine manure (FBM) - sampled from local
cow farms - and sewage sludge sampled from a pilot-scale mesophilic upflow anaerobic
sludge blanket (UASB) reactor fed with sewage at the Centre for Research and Training
on Sanitation (CePTS), UFMG/Copasa (Arrudas Wastewater Treatment Plant, Belo
Horizonte, MG, Brazil) in the proportion defined by Lima et al., (2018). The inoculum
was prepared mixing equal amounts of UASB sludge and FBM in order to obtain an initial
concentration of 0.05 g VSS gTS-1. Then, the inoculum was pre-incubated for four days
at 35 ± 1°C in order to offset the methane production from the biodegradation of any
residual organic matter (Baêta, Lima, Adarme, et al., 2016).
Quantification of biogas produced in the anaerobic co-digestion experiments was
made by daily recording the cumulative pressure in each glass bottle (Manometer®,
model PM-9100HA). Biogas composition was evaluated by gas chromatography as
described in the section 5.3.10.1. At the end of BMP experiments, liquid samples were
collected from the glass bottles and filtered on nitrocellulose membranes (0.45 µm,
44
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Unifil) to allow the measurement of dissolved COD and volatile fatty acids (VFA) by
standard protocols (see section 5.3.10). BMP experiments were stopped when the
methane production was smaller than 5% of their total accumulation (Baêta, Lima, Filho,
et al., 2016). All results from BMP experiments were normalized considering standard
temperature and pressure (STP) conditions (273.15 K and 105 Pa), and dividing such
value by the dissolved COD removed (COD) at each glass bottle.

5.3.5 Design of Experiments and Statistical Analysis

For optimizing the anaerobic co-digestion of substrates, a Doehlert experimental


design (DED) was used. The independent variables evaluated were the volumetric
proportion of substrates (HH and vinasse) in the mixture at a range of 100-0% (v v-1) and
the F/M ratio at a range of 0.05-0.7 (g COD g VSS-1). The dependent variable was the
methane yield (NmL CH4 g CODr-1) and the results were analyzed with the Statistica®
software (Statsoft Inc, version 10) with a 95% significance level. The matrix of
experiments is presented in Table 5.1.

Table 5.1 Doehlert experimental design for anaerobic co-digestion of vinasse and
hemicellulose hydrolysates (HH)

Assay Volumetric mixture F/M ratio


proportion (g COD g VSS-1)
HH-Vinasse
1 100-0% (-1) 0.375 (0)
2 75-25% (-0,5) 0.700 (0.886)
3 0-100% (1) 0.375 (0)
4 25-75 % (0,5) 0.050 (-0.886)
5 75-25% (-0,5) 0.050 (-0.886)
6 25-75% (0,5) 0.700 (0.886)
7 50-50 % (0) 0.375(0)
8 50-50 % (0) 0.375(0)
9 50-50 % (0) 0.375(0)

45
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

5.3.6 Effect of Yeast Extract Addition on AcD

The effect of yeast extract (YE) – which might be industrially obtained from
discharged or spent yeasts – addition on anaerobic co-digestion of substrates was
evaluated keeping the F/M ratio at 0.375 g COD g VSS-1 and testing three different values
of HH-to-vinasse volumetric proportion (25-75%, 50-50% and 100-0%). These
experiments were performed in triplicate with three different amounts of YE (0.5, 1.0 and
1.5 g L-1) as reported by Baêta et al., (2012) and the resulting methane yield was estimated
as described before.

5.3.7 Effect of Sugarcane Bagasse fly ash addition on AcD

The effect of sugarcane bagasse fly ash (SBFA) addition on anaerobic co-
digestion of the substrates was evaluated using the conditions previously optimized (HH-
to-Vinasse volumetric proportion; F/M ratio; YE amount) and three SBFA dosages were
tested (15, 45 and 75 g L-1) in accordance with Romero-Güiza et al., (2016) and
guaranteeing that amount generated per harvest is enough for full scale implementations.
The experiments were performed in triplicate and the methane yield was measured as
described before.

5.3.8 Modeling the Methane Production in AcD Experiments

The modified Gompertz model (Eq. (1)) and the two-phase exponential model
(Eq. (2)) (Lima et al., 2018) were used in attempt to predict the biogas production in the
anaerobic co-digestion experiments.

  e  
P = P0 exp − exp  m (  − t ) + 1  (1)
  P0  

P = P01[1 − exp(−k1t )] + P02 [1 − exp(−k2t )] (2)

46
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

where P0, P01, P02 are methane production potentials (NmL CH4 g CODr-1), k1 (d-1) and
k2 (d-1) are kinetic rate constants for two-phase exponential model, µm is the maximum
methane production rate (NmL CH4 g CODr-1d-1), λ is the lag phase time (d) and e is the
Euler’s constant (≈ 2.718). The selection and validity of the model was made by
evaluating the determination coefficient (R2), root-mean-square error (RMSE),
normalized root-mean square error (NRMSE) and Akaike information criterion (AIC), as
detailed by (Lima et al., 2018).

5.3.9 Evaluation of Toxicants Adsorption onto SBFA

The role of SBFA as an adsorbent of toxic furfural compounds was evaluated. The
adsorption capacity (q) of the SBFA was determined using two different HH-to-vinasse
volumetric mixture proportions, i.e. 25-75% and 100-0%. Batch adsorption experimental
conditions were those used in the anaerobic co-digestion experiments (T = 35°C, 180 rpm,
pH = 7.2). The effect of SBFA dosage on the removal of COD and anaerobic digestion
inhibiting compounds (HMF and FF) was evaluated by adding 50, 100, 300, 750,
1000,1300, 1600, 2000, 2600 and 3000 mg of SBFA to 125 mL Erlenmeyer flasks
containing 50 mL of HH-to-vinasse in the volumetric proportions aforementioned. All
flasks were incubated for 24 h which was the time previously determined to reach the
equilibrium. After 24 h of contact aliquots of the supernatant were sampled, filtered and
analyzed for COD, FF, HMF as detailed in section 5.3.10. SBFA adsorption capacity for
each analyte was calculated using Eq. (3).

(Ci − Ce )V
q= (3)
wSBFA

where q (mg g-1) is the adsorption capacity at equilibrium, V (L) is the solution
volume, Ci and Ce (both at mg L-1) are the initial and equilibrium concentrations of the
target analytes (adsorbate) and wSBFA (g) is the weight of the SBFA used as adsorbent.

47
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Equilibrium adsorption data were fitted with the Langmuir, Freundlich and Sips
isotherm models using their respective equations as reported by Foo and Hameed, (2010).
BET isotherm for liquid phase adsorption was fitted according to Ebadi et al., (2009). The
selection and validity of the model that best described the adsorption of COD, FF, HMF
were as described in Section 5.3.8.

5.3.10 Analytical Methods

5.3.10.1 Characterization of gas fractions

Biogas composition, in terms of methane and hydrogen (%, v v−1), was


determined by gas chromatography (Shimadzu, model 2014/ TCD) equipped with a
thermal conductivity detector operating at 120°C with a current of 85 mA. Nitrogen was
used as carrier gas at a flowrate of 1.79 mL min−1 and a 5 Å packed capillary column
(30 m × 0.53 mm Restek®) kept at 120 °C was used to separate the biogas components.
An external calibration curve was built with a standard gas mixture
(24.99 mol CH4 mol mixture−1 and 25.40 mol H2 mol mixture−1) for methane and
hydrogen quantification (Lima et al., 2018).

5.3.10.2 Characterization of liquid fractions

The concentration of sugars (cellobiose, glucose, xylose and arabinose), organic


acids (acetic, formic, propionic, butyric, isobutyric, valeric and isovaleric acids) and sugar
breakdown products (FF and HMF) was determined by high performance liquid
chromatography (HPLC) using a Shimadzu HPLC system equipped with an Aminex
HPX 87H column (300 × 7.8 mm, Bio-Rad®) at 55°C (Shimadzu, model CTO-30A). A
refractive index (Shimadzu, model RID-6A) detector was used for sugars quantification
and a UV-Vis (Shimadzu SPD-10AV) set at 210 nm and 274 nm was used for organic
acids, FF and HMF detection. Aqueous 5 mmol L-1 H2SO4 solution was used as eluent at
a constant flow rate of 0.6 mL min-1.

48
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

The concentration of xylooligomers (XOS) present in the HH was determined as


described by NREL LAP-015 (HH-to-water ratio of 1:10, 4 wt.% H2SO4 at 121 °C for 1
h) using an acidic treatment of a HH sample. The content of XOS was then estimated
from the difference in the concentration of monomeric sugars and sugar degradation
products found in the HH before and after the hydrolysis (acidic treatment) step.
The concentrations of total organic carbon (TOC) and total nitrogen (TN) in the
HH and vinasse were determined using a TOC–L CPH/CPN Shimadzu equipment
coupled with a TNM-L module as described by Quintão, (2017). The chemical oxygen
demand (COD) (method 5220D), pH (potentiometric method) and volatile suspended
solids (VSS) (method 2540 G) were determined according to the Standard Methods for
the Examination of Water and Wastewater (APHA, 2005).

The aqueous samples were also characterized by ions chromatography using a


Metrohm equipment (model 930 Compact IC Flex) equipped with METROSEP C4
150/4.0 and METROSEP A Supp 10-150/4.0 columns coupled to a conductivity detector
to quantify cations and anions, respectively. The chromatographic conditions for cations
(K+, Ca2+, Mg2+, Na+ and NH4+) separation employed 0.0017 mol L-1 HNO3 and aqueous
dipicolinic acid (0.0007 mol L-1) solutions as mobile phase which was pumped at a
constant flow rate of 0.9 mL min-1 at 25°C. The conditions for anions (Cl-, NO2-, NO3-,
PO43- and SO42-) separation employed a mobile phase composed by aqueous 0.005 mol L-
1
Na2CO3 solution and aqueous 0.005 mol L-1 NaHCO3 solution pumped at a constant
flow rate of 1.0 mL min-1 at 45°C.

5.3.10.3 Characterization of solid fractions

Samples of SBFA and YE were analyzed by Fourier Transformed Infrared (FTIR)


spectroscopy. About 5 mg of SBFA or YE were mixed with 100 mg of spectroscopy
grade KBr and pressed in a hydraulic press (Pike CrushIR, model 181-1110, Pike
Technologies, Canada) to obtain 13-mm diameter KBr pellets. The spectra were recorded

49
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

on an FTIR spectrometer (ABB Bomen, model MB 3000) at 4 cm-1 resolution, from 500
to 4000 cm-1 with 32 scans.
Samples of SBFA and YE were also analyzed by Inductively Coupled Plasma
Optical Emission (ICP-OES) spectrometry to investigate the presence of metallic
compounds. For this, the samples were acid digested according to the EPA 3050B method
(EPA, 2017) and the resulting liquid fraction was analyzed for the presence of Al, As, Ba,
Be, Ca, Cd, Mg, Mn, Mo, Na, Ni, Pb, Sn, Sr, Ti, V and Zn using an ICP-OES equipment
(Agilent, model 725).
The characterization of undigested samples of SBFA and YE was also carried out
to evaluate the presence of oxides (SiO2, Al2O3, Fe2O3, TiO2, MnO, MgO, CaO, Na2O,
K2O, P2O5). Initially, the loss of ignition was calculated gravimetrically after submitting
the samples to 1050°C for 1 h. Then, 2.5 g of each sample and 1 g of the biding agent
Licowax C Micro powder PM (Amide wax) was used for pellet preparation. An X-ray
fluorescence (XRF) spectrometer (Philips Panalytical, model MagiX) with automatic
sampler PW2540 equipped with 2.4 kW Rh tube target was used for these analyses.
The textural properties of SBFA were also determined by means of N2 physisorption at -
196°C. For this, the sample was degassed at 200°C for 10 h under vacuum prior to the
analysis in a BET Autosorb iQ station 1 analyzer (Quantachrome Instruments, USA).

5.3.11 Energy Balance

Energy balances were calculated for the best conditions considering the amount
of energy consumed per kg of SB (EHP) – equation (4) - in the hydrothermal pretreatment
at the conditions optimized by Baêta et al., (2016a), i.e. T = 183°C, t = 41 min, LSR =
3.94 mL g-1SB. The inlet temperatures of raw SB and water (from the process) were
assumed as 25°C and 95°C, respectively. Such water temperature would be possible
considering the use of engine cooling water and the heat of the flue gas of a combined
heat and power (CHP) system.

EHP = [Cp,SB  (183 − 25) + Cp,H2O  LSR  H2O  (183 − 95)] (4)

50
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

The values of specific heat capacity (Cp,SB) assumed to SB and water were
1.76 × 10-3 MJ (kg°C°)-1 and 4.19 × 10-3 MJ (kg°C°)-1, respectively (Baêta et al., 2016a).
The energy generated in AcD experiments was calculated considering a typical
commercial CHP system as described by Cano et al., (2015). Using the methane yield
from experimental BMP tests, it was possible to estimate the energy production (EP)
according to Eq. (5).

TCOD
EP = BMP   34.5 MJ Nm3CH 4-1 (5)
kg SB

where BMP (Nm³ CH4 kg CODr-1) is the final accumulated methane production,
TCOD is the amount of total chemical oxygen demand recovered after hydrothermal
pretreatment per kg of SB, on a dry-weight basis.
Considering typical efficiencies of the CHP system, the Ep was converted to
thermal (ET) and electric (EE) energies according to Eq (6) and (7).

ET = EP  0.85  0.66 (6)

EE = EP  0.85  0.33 (7)

Where 0.85 is the efficiency of the combustion engine; 0.66 is the conversion
factor of Ep to ET and 0.33 is the conversion factor of Ep to EE considering a conventional
CHP system.
In order to estimate the energy profitability, both thermal (MJ kg SB-1) and
electric (kWh kg SB-1), of the entire process, the difference between the energy consumed
by hydrothermal pretreatment (EHP) and the energy generated by AcD (ET and EE) was
also evaluated.

51
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

RESULTS AND DISCUSSION

5.4.1 Physicochemical Characterization of Liquid and Solid Fractions

Data of physicochemical characterization of vinasse, HH, YE and SBFA is


presented in Table 5.2. As can be seen in Table 5.2, the amount of reducing sugars as well
as formic acid in HH is significantly higher in comparison to vinasse. This was expected
as the continuous fermentation system at the ethanol industrial plant is designed to obtain
a maximum conversion of hexoses into 1G ethanol. In its turn, the high sugar content
(mainly pentose) in HH is due to the thermal treatment performed on the SB which is
devised to release them by autohydrolysis, as discussed in the section 2.1.2.
Table 5.2 also shows that the chemical composition of HH in relation to nitrogen
content may hamper its direct use for biogas production, since this element is unbalanced
and may require supplementation for the best performance of AD. On the other hand,
vinasse presents a nitrogen amount about 3.4 times higher than HH, suggesting that this
sugarcane biorefinery waste could be used to balance the C/N ratio when HH is used to
biogas production, which have a great influence in the anaerobic digestion (Passos et al.,
2018).
In this study the main constituents of vinasse (Table 5.2) were found in
concentrations that are within the range reported by other authors (Fuess et al., 2018;
Leite et al., 2015). Due to its chemical characteristics, i.e. presence of plant nutrients, and
its relatively low cost, vinasse has been widely used in the irrigation of sugarcane
cultivated areas. However, the high COD and inorganics of vinasse lead to some
environmental issues when this management is employed without caution. The presence
of toxic compounds is also relevant, causing potential impacts on soil and water resources,
as well as the accumulation of inorganics in soil, which may lead to its salinization
(Hoarau, J. et al., 2018). Moreover, as seen in Table 5.2, the vinasse used in this study
had Ba in a high concentration (23.1 mg L-1), which may cause problems to subsoil

52
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

(Abreu et al., 2012). Christofoletti et al., (2013) reported a Ba concentration of 0.41 mg L-


1
in the sugarcane vinasse studied, which is much lower than the one found in this study.

Table 5.2 Physicochemical characterization of vinasse, hemicelluloses hydrolysate, yeast


extract and sugarcane bagasse fly ash

Liquid fraction Solid fraction


Hemicelluloses
Parameters Yeast
Vinasse Hydrolysate
Extract (YE)*
(HH)
COD/(g L-1) 18.07 ± 0.92 71.17 ± 1.76 -
TOC/(g L-1) 8.57 ± 0.41 35.49 ± 1.88 0.40 ± 1.2×10-2
Total N/(g L-1)) 0.41 ± 0.02 0.12 ± 0.005 0.70 ± 2.8×10-2
TS/(g L-1) 23.01 ± 0.05 63.75 ± 0.58 -
VSS/(g L-1) 17.16 ± 0.06 61.25 ± 0.59 -
Cellobiose/(g L-1) - - -
Glucose/(g L-1) - 2.65 ± 0.022 -
Xylose/(g L-1) 1.55 ± 2.09×10-2 15.31 ± 0.34 -
0.86 ±
Arabinose/(g L-1) - -
9.49×10-2
Xylooligomers/(g L-1) - 20.87 ± 0.17 -
Formic acid/(g L-1) 2.97 ± 0.28 10.49 ± 0.26 -
Acetic acid/(g L-1) - 11.92 ± 0.24 -
0.58 ±
HMF/(g L-1) - -
3.31×10-3
1.33 ±
FF/(g L-1) - -
1.16×10-2
0.13 ±
PO43-/(g L-1) 2.51 ± 0.11 -
7.65×10-3
NO3-/(g L-1) 0.31 ± 2.79×10-2 - -
0.21 ±
SO42-/(g L-1) 1.99 ± 7.96×10-2 -
1.26×10-2
NH4+/(g L-1) 0.02 ± 1.20×10-3 - -
K+/(g L-1) 2.24 ± 8.96×10-2 0.39 ±3.12×10-2 11.37 ± 0.71
- -1 -2 -3
Cl /(g L ) 1.41 ± 8.12×10 0.07 ± 4.20×10 53.50 ± 2.14
2+ -1 -3 -3
Ca /(g L ) 0.12 ± 7.20×10 0.06 ± 3.1×10 1.91 ± 0.11
2+ -1 -2 -3
Mg /(g L ) 0.23 ± 1.84×10 0.10 ± 6.1×10 -
2+ -1 -3
Ba /(g L ) 0.02 ± 1.1×10 - -

53
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Liquid fraction Solid fraction


Hemicelluloses
Parameters Yeast
Vinasse Hydrolysate
Extract (YE)*
(HH)
Zn2+/(g L-1) - - 0.13 ± 7.81×10-3
Br-/(g L-1) - - 0.05 ± 3.1×10-3
Total Fe/(g L-1) - - 0.12 ± 9.12×10-3
Sc3+/(g L-1) - - 1.32 ± 0.11
*
Measured using solutions of 1.5 and 1.0 gL-1

Table 5.3 Physicochemical characterization of sugarcane bagasse fly ash by XRF

Solid fraction
Parameters Sugarcane bagasse fly ash
(SBFA)
SiO2/(%) 25.04
Al2O3/(%) 7.84

Fe2O3/(%) 7.83
TiO/(%) 1.51
MnO/(%) 0.17
MgO/(%) 2.54
CaO/(%) 3.65
Na2O/(%) 0.05
K2O/(%) 4.54
P2O5/(%) 2.37
Loss of 43.7
ignition/(%)

Data on characterization of YE used in this study is shown in the Table 5.2 and it
can be seen that it has a variety of mineral micronutrients required by AcD
microorganisms in their metabolic pathways. In fact, YE is used as a nutritional
supplement in animal feed formulations because of its high protein and high B-complex
vitamins concentrations, besides providing an excellent amino acids balance (Bergmann
et al., 2018).

54
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

FTIR spectrum of YE presented in Figure 5.1 shows the presence of nitrogen


functional groups as confirmed by the bands of the IR spectrum related to amino acids
containing carboxylate and ammonium groups. The bands at 1585 and 1406 cm-1 are
related to asymmetric and symmetrical stretching of the carboxylate group, whereas the
bands at 1607 and 1517 cm-1 are related to asymmetric and symmetric bending N-H of
primary amine salt, and the band at 3296 cm-1 is related to ammonium ion. Usually,
commercial YE samples have protein content higher than 40% (Bergmann et al., 2018;
Jalles Machado, 2018) which allows this by-product to be used as a nitrogen source for
the AcD of substrates lacking this nutrient. This is the case for HH, and we hoped that its
co-digestion with YE increases the performance of anaerobic microorganisms. It is
noteworthy that although the addition of yeast extract may increase the costs of the
anaerobic digestion process, some researchers showed it is possible to obtain yeast extract
from spent yeasts using low-cost methods that guarantee its technical and economic
feasibility for these kind of applications (Victral et al., 2016).

Figure 5.1 Typical FTIR spectra of the yeast extract sample used in this study.

SBFA characterization revealed that it has a heterogeneous composition, with


particles of different shapes and sizes. SBFA chemical composition can vary with the
burning process parameters, such as temperature and time, as well as sugarcane variety
(Rodríguez-díaz et al., 2015). As can be seen in Table 5.3, the SBFA sample analyzed by
XRF had an ignition calcination loss of approximately 43.7%. The amount of non-
55
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

oxidized carbon in ashes depends on the combustion process efficiency and SB humidity.
Thus, when the SBFA has a large amount of non-oxidized carbon it is an indicative of
low process efficiency (Chingono et al., 2018). Such remaining carbon content, which
was not oxidized to CO or CO2, may confer the SBFA hydrophobic properties which are
suitable for adsorbing organic compounds from the HH such as lignin fragments, furans
and organic acids (Balakrishnan e Batra, 2011).
SFBA textural properties revealed a pore size distribution in a range of 21.81-
88.25 Å and a surface area of 222.8 m2 g-1. According to the International Union of Pure
and Applied Chemistry (IUPAC), SBFA could be classified as a mesoporous material
(20-130 Å) .This mesoporous character offers advantages for the adsorption of medium-
size organic molecules such as FF, HMF, organic acids and phenol derivatives
(Rodríguez-díaz et al., 2015).
Silica (SiO2) is the second major component of SBFA (25.04%). It mainly comes
from sand and quartzite minerals that are residues from sugarcane bagasse mechanical
crop and remains after the washing steps employed in the 1G ethanol production process
(Cacuro, T A e Waldman, 2015). The presence of this constituent was also confirmed in
the IR spectrum presented in the Figure 5.2. In this Figure it is seen bands at 1100 and
800 cm-1 related to symmetric Si-O-Si stretching and the band at 770 cm-1 related to
asymmetric Si-O-Si stretching which are typical of silicate network (Barbieri et al., 2013).
In addition, the band at 1100 cm-1 is also related to -C-O-H stretching and O-H
deformation. Hydroxyl and carboxyl groups were also identified in SBFA by the presence
of bands at 3400 cm-1 related to O-H stretching and 1690 cm-1 related to C=O stretching
in carboxylic acids (Rodríguez-díaz et al., 2015).

56
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 5.2 FTIR spectra for sugarcane bagasse fly ash

In addition to silica, alkaline oxides such as K2O (4.54%), CaO (3.65%) and MgO
(2.54%) are also present in SBFA. These compounds are responsible for increasing the
alkalinity in aqueous medium, thereby contributing to buffer the aqueous solution in AD
systems. This helps to neutralize the volatile organic acids produced during AD
controlling the solution pH, which may enhance the performance of the methanogenic
archaea and contributes to the increase of methane production (Chernicharo, 2007). On
the other hand, excess of these oxides can cause biological inhibition due to increased
cations concentration and also by forming precipitates which reduce the bioreactor useful
volume. Then, it is important to optimize the ash dosage in the AD to maximize methane
production, thereby minimizing inhibition and reducing the costs related to alkalinity
supplementation. The latter represents one of the economic limitations for the industrial
implementation of AD technologies (Fuess, L T et al., 2017; Janke, L, Leite, A. F., et al.,
2016).
The presence of compounds of interest, such as metals (e.g. Cu, Mg, Zn) that can
act as enzymatic cofactors in anaerobic metabolism and hence could favor the
performance and growth of anaerobic microorganisms, was also identified by ICP-OES
(Table 5.4). As it is widely known, some metallic elements, when present in high
concentrations, can cause inhibitory effects, lower microbial growth and decrease in

57
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

process performance. This is another reason for optimizing the dosage of SBFA in the
AcD of vinasse, HH and YE.

Table 5.4 Analysis by ICP OES for sugarcane bagasse ash digested.

mg kg-1
Element
Ba 104.22 ± 5.74
Cr 49.55 ± 3.67
Cu 35.13 ± 1.78
Ni 9.18 ± 0.48
Sc 5.08 ± 0.32
Sr 93.20 ± 4.93
Th 5.59 ± 0.21
V 58.32 ± 3.08
Y 7.88 ± 0.48
Zn 97.84 ± 5.63
Mn 530.64 ± 30.18
Al 20581.00 ± 1224.21
Fe 17173.78 ± 1003.15
Ca 13975.88 ± 702.36
K 13918.60 ± 676.68
Mg 5913.10 ± 356.91
Na 204.38 ± 19.30
Zr 2.77 ± 0.07
P 7102.72 ± 379.69
S 787.30 ± 6.54
Si 111.16 ± 0.34
Ti 1365.49 ± 76.82

5.4.2 Anaerobic co-digestion

5.4.2.1 Effect of vinasse and HH mixture on methane production

The AcD experiments of vinasse and HH without the addition of a nutrient


solution were performed according to DED experiments for the methane production
optimization (see Table 5.1). The Pareto’s chart of the standardized effects is presented
in Figure 5.3 where it is seen that the HH-to-vinasse volumetric mixture ratio was the
variable that most influenced the methane yield, having a significant negative effect (-
43.26). This means that mixtures composed of lower ratio of HH/vinasse resulted in

58
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

higher methane yields. This result can be explained by the dilution effect, caused by
vinasse, of the inhibitory compounds present in HH (e.g. FF plus HMF). In the case of
experiments 1 and 6 the concentration of total inhibitory compounds FF and HMF were
191 and 47.8 mgL-1 respectively.
Figure 5.3 shows that the F/M ratio also had a significant negative effect (-4.80),
i.e., the methane yield was higher when lower levels of food were available to
microorganisms. This behavior was also observed by Slimane et al., (2014) and
Eskicioglu and Ghorbani, (2011), who used wastewater from slaughterhouse and vinasse
as substrates for AD, respectively. One hypothesis is that the excess of microorganisms
in relation to the food available (low F/M ratio) forces the enzymatic apparatus of starving
microorganisms to deal with and degrade even more recalcitrant compounds (e.g. lignin
and its derivatives), therefore leading to a higher methane yield.
BMP values obtained from the DED experiments are presented in Figure 5.4.
Comparing the experiments with the same HH-to-vinasse volumetric mixture ratio or F/M
ratio it is possible to note the influence of the substrate composition or the inoculum
amount. For experiments 2 and 5, it is clear that a HH-to-vinasse volumetric mixture ratio
of 75-25% combined with higher (0.7) and lower (0.05) level of F/M ratio resulted in
higher and lower methane production, respectively. For experiments 6 and 4, a HH-to-
vinasse volumetric mixture ratio of 25-75% combined with a higher (0.7) and a lower
(0.05) level of F/M ratio resulted in lower and higher BMP values, respectively. However,
the BMP values for experiments 6 and 4 were higher than those obtained for experiments 2 and
5.

59
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 5.3 Pareto’s chart of the standardized effects for anaerobic co-digestion of HH and
vinasse

For experiment 1, a F/M ratio of 0.375 g COD g VSS-1 combined with a HH-to-
vinasse volumetric mixture ratio of 100-0% resulted in the lowest BMP value. For
experiment 3, a F/M ratio of 0.375 g COD g VSS-1 combined with a HH-to-vinasse
volumetric mixture ratio of 0-100% resulted in the second highest BMP value, thereby
evidencing that HH anaerobic degradation is not straightforward. This confirms the
hypothesis related to the toxicity of furans (FF plus HMF) for anaerobic microorganisms
as reported by Barakat et al., (2012). In the present study, the concentration of such
inhibitors was 1.91 g L-1 in HH and below the limit of detection in vinasse.

60
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 5.4 Biochemical methane potential for different mixtures of HH-to-vinasse and
F/M ratio

Among all DED experiments, the experiment 6 led to the highest specific methane
production, i.e. 0.145 Nm3 CH4 kg CODr-1. This was probably related to the high dilution
of furans as well as to vinasse composition, which contains high amounts of nutrients
(mainly N, P, K) that could be used for cell growth and production of essential enzymes
by the anaerobic consortia.
Figure 5.5 shows the cumulative methane production for experiment 6 (25-75%
HH-to-vinasse volumetric mixture ratio) whose data was used to test the kinetic models.
Both kinetic models tested, i.e. modified Gompertz and two-phase exponential, fit the
cumulative methane production data, as seen in Table 5.5.
The two-phase exponential kinetic model was the one that best described the
methane production based on the values of error functions (R2, RMSE, NRMSE and AIC)
(Table 5.5). An explanation for this “two-phase” behavior in methane production is the
adaptation of the microorganisms to the presence of inhibitory compounds, mainly from
HH. Li et al., (2018) reported that the use of xylan mixtures, cellulose and lignin for
61
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

methane production resulted in lag phases that lasted for 0.4 to 1.9 days. Another hypothesis
to explain this behavior is the difficulty of anaerobic microorganisms to hydrolyze HH
xylooligomers (XOS) (Table 5.2) into monosaccharides (Baêta, Lima, Adarme, et al., 2016).
XOS present in HH represents almost 30% of the total COD and microorganisms of the
anaerobic consortium that are not adapted to this substrate have to synthesize hydrolytic
enzymes to breakdown these compounds. Therefore, XOS hydrolysis may have been the
limiting-step for the anaerobic degradation of HH with a negative impact on methane
production (Pérez et al., 2002). In addition, the BMP experiments were monitored for 60 days,
which can be considered a long time for liquid-phase AD systems.

Figure 5.5 Fitting (line (two-phase model) and dashed line(Gompertz model)) of the
experimental data (o) for the best condition of BMP

62
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental Engineering – PROAMB

Table 5.5 Kinetic constants for methane production for best anaerobic co-digestion process using substrates from sugarcane biorefinery.
Model Parameters Best conditions for anaerobic co-digestion processes
HH-Vin HH-YE HH-Vin-YE HH-Vin-YE HH-YE-SBFA HH-Vin-YE-SBFA
(25%-75%) (100 %-1.5 gL-1) (50%-50%-1.5 gL-1) (25%-75%-1.0 gL-1) (100 %-1.5 gL-1 -45 gL-1) (25%-75%-1.0 gL-1-15 gL-1)
Modified PO 133.05 81.32 204.07 205.03 272.51 319.97
Gompertz µm 5.81 3.41 10.53 11.58 18.76 13.84
λ 4.44 9.05 5.11 5.25 5.89 6.80
R2 0.97 0.94 0.98 0.98 0.99 0.98
RMSE 8.07 6.24 8.80 9.69 8.09 13.73
NRMSE 5.58 7.13 4.20 4.45 3.02 4.91
AIC 252.46 211.09 227.89 264.91 181.70 226.08
Two-phase PO1 9.15 33.90 8.96 0.88 72.93 142.38
exponential P02 142.54 70.27 200.39 248.68 227.45 187.17
k1 0.038 0.024 0.032 0.039 0.045 0.014
k2 0.039 0.025 0.029 0.038 0.046 0.012
R2 0.97 0.95 0.95 0.94 0.94 0.96
RMSE 12.06 10.02 20.66 21.09 12.43 17.57
NRMSE 8.34 11.46 9.86 9.68 8.27 9.86
AIC 301.87 266.17 316.89 355.08 286.5 276.61
HH: Hemicellulosic hydrolysate ,Vin: Vinasse, YE: yeast extract, SCBA: sugarcane bagasse ash, PO: max production of CH4 (NmL CH4 g COD-
1
),k1,2: kinetic constants (day-1),µm: maximum methane production rate (NmL CH4 day-1 g COD-1),λ: lag phase (day),RMSE: Root Mean Square
Error (NmL CH4 g COD-1), NRMSE: Normalized Root Mean Square Error, AIC : Akaike Information Criterion,R2: coefficient of determination

63
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

For experiments 3 and 6, the COD-to-SO42- ratio was 9.09 and 20.2, respectively,
These values and their respective methane yields are in good agreement with data reported
by Kiyuna et al., (2017), who confirmed the negative influence of sulfate on the biogas
production when vinasse is used as substrate. It is seen that when there was a higher
amount of sulphate in relation to COD (experiment 3) the methane production was about
17% lower than the best scenario found for biogas production (experiment 6).
The best experimental condition to be fixed at the subsequent experiments, which
aimed to evaluate the effect of YE and SBFA addition on methane production, were
obtained from the response surface (Figure 5.6) and the second-degree polynomial
equation adjusted to the experimental data (R2 = 0.93, R2adj = 0.89, p-value = 0.06). The
best condition to maximize methane production was the one that employed a HH-to-
vinasse volumetric mixture ratio of 25-75% and a F/M ratio of 0.375 g COD g VSS-1.

Figure 5.6 Response surface plot for BMP experiments according to the Doehlert
experimental design

64
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

According to the technology employed in Brazil to process sugarcane bagasse, it


would be possible to produce, from 1 ton of sugarcane, about 90 L of ethanol, 280 kg SB
(with 50% moisture) and from 900 to 1350 L of vinasse (Soccol et al., 2016). The
experimental conditions employed in the thermal treatment of SB (T = 183°C, t = 41 min,
LSR = 3.94 L kgSB-1 dry basis) would lead to the production of 249 L of HH. In other
words, this amount of HH would be produced by processing only 50% of the raw SB
available at the sugarcane biorefinery. Thus, it is highlighted that the optimized values of
methane production reported in this study offer a realistic technological route for the final
disposal of vinasse. This would demand the use of 57% of 1G vinasse for AcD with HH,
thereby leaving a vinasse surplus that could still be used for fertigation.

5.4.2.2 Effect of YE addition on the methane production

Figure 5.7 shows the BMP values and the volatile fatty acids (VFAs) produced
for the 3 mixtures composition of interest, i.e. HH-to-vinasse of 100-0%, 50-50% and 25-
75%. Considering the experiment 10 (Table 5.6), which evaluated pure HH as the
substrate and three different YE concentrations (0.5, 1.0 and 1.5 g L-1), it is seen an
improvement in the anaerobic biodegradability with the YE addition, as more methane
was produced in comparison to experiment 1 (Figure 5.4). This improvement in methane
production can be explained by the decrease in the C/N ratio with YE addition, as more
methane was produced per unit of COD removed (0.090 Nm3 CH4 kg CODr-1), as well as
to the addition of vitamins and cofactors (e.g. riboflavin) that constitute the YE. It is
noteworthy that such methane production is considered low when compared to the
theoretical estimated value (0.3546 Nm3 CH4 kg CODr-1). An explanation for this might
be the inhibitory effect caused by the presence of furans in the HH; and the VFA
accumulation in the assays 10 and 12 seems to support this hypothesis. Figure 5.7 shows
that VFA concentration was higher (~750 mg VFA L-1) in the assay 10, which had a
higher amount of HH in mixture, when compared to the assay 12 (~250 mg VFA L-1)

65
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

which employed a HH to vinasse proportion of 25-75%. The main VFA accumulated was
the propionic acid (≈ 550 mg L-1), indicating that a kinetic/thermodynamic imbalance
occurred due to non-consumption of this acid by the slow-growth acetogenic
microorganisms (Tabatabaei e Ghanavati, 2018).

Table 5.6 Experimental design for anaerobic co-digestion of vinasse, hemicellulose


hydrolysates (HH), yeast extract and sugarcane bagasse ashes

Volumetric mixture F/M ratio Yeast Sugarcane


Assays proportion HH-to extract bagasse fly ash
Vinasse (g COD g (gL-1) (gL-1)
VSS-1)
0.5
10 100-0% 0.375 1.0 ----
1.5
0.5 ----
11 50-50% 0.375 1.0
1.5
0.5 ----
12 25-75% 0.375 1.0
1.5
15
13 25-75% 0.375 1.0 45
75
15
14 100-0% 0.375 1.5 45
75

66
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 5.7 Methane production (open bar) and VFA(filled bar) (sum of formic acid, acetic
acid, propionic acid, butyric acid, isobutyric acid, valeric acid and isovaleric acid
concentrations) accumulation (line and dot) for anaerobic co-digestion in the presence of
yeast extract at three different conditions.

For experiment 11 (Table 5.6) (50-50% HH-to-vinasse), the best methane


production was obtained when the YE concentration was 1.5 g L-1
(0.210 Nm3 CH4 kgCODr-1), and in such condition there was a dilution effect of HH,
which certainly improved the biodegradability of the mixed substrate. This is confirmed
by the lower amounts of VFA accumulated at the end of the experiment 11 in comparison
to the experiment 10 with pure HH (Figure 5.7). The kinetic parameters adjusted for
methane production in the experiment 11 with 1.5 g L-1 YE (Table 5.5) confirmed a
decrease in the lag phase (5.11 days) and an increase in the maximum methane production
potential in comparison to the condition with no HH dilution (0.090 Nm3 CH4 kg CODr-
1
), which exhibited a lag phase of 8.95 days.

67
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

The highest methane production (0.217 Nm3 CH4 kg CODr-1) was achieved in the
experiment 12 (25-75% HH-to-vinasse). However, in such experiment the highest biogas
production was observed when the YE concentration of 1 g L-1 was used (C/N=18.91)
whereas using YE of 1.5 gL-1, C/N was 12.30. These data confirm that exists a combined
effect of HH dilution and C/N ratio, considering that both vinasse and YE have macro-
and micro-nutrients that are essential for the AD process. In addition, as reported in
Section 5.4.1, the organic components of YE are mainly amino acids and peptides. This
is in good agreement with their metabolic roles in transporting metal nutrients via specific
membrane-bound transport systems. Some authors (Gonzalez-Gil et al., (2003), Minh et
al., (2016)) believe that trace metals (e.g. Ni, Co) pass through the biological membrane
in the form of lipophilic metal-amino acid complexes, therefore the supplementation of
amino acids and peptides contributes to increasing the bioavailability of such nutrients.

The values of VFA accumulated (Figure 5.7) at the end of the experiment 12
(Table 5.6) are similar to the ones observed in assay 11 and lower than those measured in
assay 10. This behavior strengthens the hypothesis that HH is less biodegradable and
contains toxicants that need to be diluted when compare to vinasse. When AcD was
planned with 100% of HH (assay 10) - the most adverse condition – the highest YE
concentration (1.5 g/L) enhanced biogas production and helped to decrease the VFA
accumulation. On the other hand, during the conditions where vinasse was co-digested
with HH (assays 11 and 12), the presence of yeast extract showed no apparent influence
on VFA accumulation (Figure 5.7). This seems to indicate that YE is more important in
systems that contains toxic compounds and/or a higher proportion of more recalcitrant
substrates. On the other hand, the C/N ratio for the systems that used vinasse was in the
range from 12.30 (HH-to-vinasse 25-75%, 1.5 gL-1 ) to 32.90 (HH-to-vinasse 50-50%,
1.5 gL-1 ), which is considered the optimal range (Caillet et al., 2016) , while for the
system using pure HH the values ranged between 66.92 and 26.29
A kinetic analysis of methane production was carried out for the best conditions,
attained in the experiments 10 (1.5 g L-1 YE), 11 (1.5 g L-1 YE) and 12 (1.0 g L-1 YE).
68
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

The kinetic parameters estimated by fitting the kinetic models to the experimental data
are presented in Table 5.5. For the experiment 10 (pure HH as substrate), the anaerobic
biodegradability was estimated at 28.2%, which is higher than the biodegradability for
the same condition without YE addition (7.14% in the experiment 1, Section 5.4.2.1). The
two-phase exponential model better explained this behavior because this model considers
that biogas production can occur in two or more phases, disregarding the interaction
between them (Lima et al., 2018). For experiments 11 and 12, the modified Gompertz
model better explained the behavior of methane production since the multiple adaptation
phases no longer exist (Table 5.5). As previously discussed, this seems to be related to
the mixture of vinasse and YE, which enhanced the C/N ratio and improved the
availability of essential nutrients for the anaerobic consortia.

5.4.2.3 Effect of SBFA addition on the methane production

Figure 5.8 shows the results of methane production and accumulated VFA for
experiments 13 and 14, which were performed with mixed substrates composed of 25-
75% HH-to-vinasse with 1 g L-1 YE and 100-0% HH-to-vinasse with 1.5 g L-1 YE,
respectively. Data presented in Figure 5.8 when compared to the results previously
discussed show that there was an increase in the biodegradability of the substrates with
SBFA addition for both experiments 13 and 14. For experiment 14 with pure HH, YE
(1.5 g L-1) and 45 g L-1 SBFA, the accumulated methane production increased about
300% compared to experiments without the SBFA addition (Figure 5.4 (experiment 1)
and Figure 5.7 (experiment 10)).
The experiment 14 exhibited the second highest methane production
(268 NmL CH4 g CODr-1) combined with a low VFA accumulation (≈ 720 mg L-1) and
these results are better than those presented by Baêta et al., (2016a), who reported a VFA
concentration of 1043 mg L-1 and an accumulated methane production of
0.173 Nm3 CH4 kg CODr-1 for AD of pure HH with the addition of a solution containing

69
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

macro- and micro-nutrients. In addition, it is noteworthy that the use of SBFA allowed a
significant reduction of about 50% in the times for methane stabilization, as the BMP
bottles were monitored for 34 days in this study instead of 60 days (Baêta et al., 2016).
The results presented here are in good agreement with those reported by Zhang et al.,
(2018) for carbon-based materials, such as activated carbon and biochar obtained from
citrus peel waste and sawdust.

Figure 5.8 Accumulated methane production (open bar) and VFA (filled bar) (sum of
formic acid, acetic acid, propionic acid, butyric acid, isobutyric acid, valeric acid and
isovaleric acid concentrations) accumulation for two different anaerobic co-digestion
conditions (25-75% HH-to-Vinasse, 1.0 g L-1 YE, 15 g L-1 SBFA and 100-0% HH-to-
Vinasse, 1.5 g L-1 YE, 45 g L-1 SBFA).

Three hypotheses could be used to explain the good results obtained with SBFA
addition. The first one is related to the SBFA capacity to act as an adsorbent material of
inhibitory compounds and will be discussed in detail in Section 5.4.3. The second one is
related to the positive effects to the anaerobic digestion, as reported by different authors
(Chernicharo, 2007; Romero-Güiza et al., 2016), caused by the addition of macro- and
micro-nutrients provided by the SBFA.
The high concentration of soluble trace metals bioavailable in SBFA and/or
vinasse in experiments 13 and 14 (0.52-1.58 mg Cu L-1,0.13-0.41 mg Ni L-1,77.28-
70
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

231.84 mg Fe L-1, 1.46-4.40 mg Zn L-1) might be explained by chelation/complexation


effects with organic ligands, such as soluble microbial products (SMPs), extracellular
polymeric substances (EPS) and YE constituents that were present in these AcD
experiments (Minh et al., 2016).
The third hypothesis is related to the alkalinizing effect provided by the SBFA
addition, as this material contains alkaline (e.g. K2O) and alkaline earth (e.g. CaO, MgO)
metal oxides (Table 5.3), which benefits anaerobic digestion by maintaining the pH at the
ideal range for microbial growth (Wang et al., 2017).
The experiments 13 (25-75% HH-to-vinasse and 1 g L-1 YE) and 14 (with pure
HH and 1.5 g L-1 YE) that were planned with a SBFA concentration of 75 g L-1 probably
led to an inhibition effect caused by the high concentration of K (2991.8 and 2825 mg L-
1
), Ca (1965.4 and 1957.9 mg L-1) and Mg (1165.8 and 1151.1 mg L-1), respectively.
These results are in good agreement with the concentration values of cations that cause
inhibition (400 to 28934 mgL-1 for K; 300 to 8000 mgL-1 for Ca and < 730 mgL-1 for Mg),
as reported by Chernicharo, (2007) and Romero-Güiza et al., (2016).
Table 5.5 shows the parameters estimated by fitting the kinetic models to the data
produced by experiments 13 and 14. According to the AIC and NRMSE values the
modified Gompertz model fit the experimental data better than the two-phase exponential
model. This implies that adaptation multiple stages did not take place during the
experiments, which might have been due to the good conditions provided by SBFA
addition which enhanced methane yield.
For experiment 13 (25-75 % HH-to-vinasse and 1.0 g L-1 YE), the effect of SBFA
addition was not so remarkable for a SBFA concentration of 45 g L-1. This was probably
because this mixture already had a better C/N ratio compared to experiment 14 (pure HH,
1.5 g L-1 YE and 45 g L-1 SBFA). For VFA accumulated in the experimental conditions
evaluated (Figure 5.8), it was observed that their concentrations were lower in comparison
to experiments 9 and 12 that used only YE as substrate, which confirms the better
nutritional balance and the alkalinizing effect provided by the SBFA addition. Moreover,

71
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

this behavior can also be explained by the presence of carbon material in SBFA
composition, which promotes interspecies electron transfer (IET) allowing the syntrophic
growth of anaerobic consortia, as described by Fagbohungbe et al., (2017) and Tan et al.,
(2015). Indeed, carbon-based materials such as graphite, graphene, biochar, activated
carbon and carbon nanotubes have been applied to enhance AD performance, as reported
by Zhang et al., (2018) and Baêta et al., (2013).
Despite the fact the addition of such materials improves AD by promoting IET,
increasing enzymatic activity and raising biogas yield, these materials add a significant
cost to the process (Luz et al., 2018). The use of SBFA might overcome this issue since
it is a low-cost material that contains non-oxidized carbon which can help AD as observed
in this study. In addition, the use of SBFA in AD might be a good solution for ash disposal
according to environmental regulations (Rodríguez-díaz et al., 2015; Sudibyo et al., 2018;
Fito et al., 2017). Currently one part of ashes from boilers are used as alkali agent and
fertilizer for soil and its excess must be disposal to the landfill.

5.4.3 Mechanism of SBFA Adsorption during AcD.

To confirm the hypothesis that SBFA played a role on the adsorption of inhibitory
compounds (FF and HMF) during AcD, adsorption experiments were performed for the
best conditions according to BMP experiments (experiment 13: 25-75% HH-to-vinasse,
1.0 g L-1 YE, 15 g L-1 SBFA and experiment 14: pure HH, 1.5 g L-1 YE, 45 g L-1 SBFA).
As can be seen in Table 5.7, for experiment 13 the HMF adsorption was best fit
by the Freundlich isotherm (smaller values of R2, RMSE, NRMSE and AIC), exhibiting a
maximum adsorption capacity (Qmax) of 61.71 mg HMF g SBFA-1. As the adsorbent
weight used in this experiment was 0.9 g, it was possible to remove 55.53 mg of HMF,
which is a higher amount than that initially inoculated (0.87 mg HMF or 14.5 mg HMF L-
1
). The value of the Freundlich constant was estimated at 0.93, suggesting that HMF
adsorption on the SBFA had a low intensity. The Freundlich model assumes that the

72
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

adsorption is not limited to a monolayer, with a non-uniform distribution of adsorption


sites along the adsorbent surface and the presence of adsorption sites exhibiting different
affinities for the adsorbates (Chingono et al., 2018). In the case of HMF a multilayer
adsorption onto SBFA might occur due to ‘pi stacking’, which is the interaction between
neighbor molecules having unsaturated (π) bonds.

The textural properties of SBFA (Section 5.4.1) are important features that
promoted the high removal efficiencies of inhibitory compounds. As SBFA contains a
high amount of non-oxidized carbon, the adsorption of hydrophobic organic molecules
was favored. The main interactions of hydrophobic organic molecules with SBFA
containing heteroatoms are probably by π-stacking (for phenolic compounds), dipole-
dipole (mainly for neutral organic acids), ion-dipole (mainly for ionized organic
compounds) (Tan et al., 2015).
For experiment 14, from the Qmax values for FF and HMF (Table 5.7), it was
possible to estimate that the maximum amount of toxicants that could be adsorbed onto
the SBFA would be 52.38 mg for FF and 10.69 mg for HMF. These calculations indicate
it would be possible to remove 7.98 mg FF and 3.48 mg HMF (for 0.9 g SBFA) from the
anaerobic medium, which is highly when compared to initial concentration.
The values of the Langmuir binding constants (b) suggest that the affinity of the
adsorption surface sites for HMF is higher than for FF, which indicates that in a mixture
of adsorbates HMF would be probably removed first from the solution (Foo e Hameed,
2010). In addition, the values of the Sips isotherm constant (n) were higher than unity,
indicating that these adsorption systems are heterogeneous. This confirms that the
Freundlich model best described the adsorption of FF and HMF onto SBFA.
As far as COD adsorption onto SBFA is concerned, the experimental data could
only be fitted by BET isotherm, as shown in Figure 5.9. These adsorption isotherms could
be classified as type III (refers to multilayer adsorption by weak interactions with low
energy between adsorbed molecules and the adsorbent having macropore) and the

73
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

parameters estimated by fitting BET isotherm to the experimental data are presented in
Table 5.7.

Table 5.7 Isotherm parameters estimated by the models for experiments 13 (25-75% HH-
to-vinasse, 1.0 g L-1 YE) and 14 (pure HH, 1.5 g L-1 YE) using SBFA as adsorbent at
concentrations of 1.5 to 60 g L-1.

HH HH-Vin
Isotherm
Parameter (100%) (25%-75%)
Model
COD FF HMF COD HMF
Qmax (mg g )-1
- 19.40 3.96 - 17.83
b (L mg-1) - 0.057 1.21 - 0.51
RL - 0.11 0.013 - 0.11
Langmuir R2 - 0.93 0.85 - 0.95
RMSE - 3.52 0.52 - 1.42
NRMSE - 12.51 14.30 - 11.16
AIC - 27.24 10.10 - 8.79
KF ((mg g-1)/(L mg-1)1/n) - 4.62 2.98 - 3.48
n - 3.70 14.39 - 0.93
R2 - 0.91 0.85 - 0.97
Freundlich
RMSE - 3.45 0.51 - 1.14
NRMSE - 11.75 14.29 - 8.95
AIC - 26.55 10.09 - 4.38
4.68
Qmax (mg g-1) - 21.243 - 25.66
1.07E-
b (L mg-1) - 0.0046 - 0.29
06
n - 2.70 - 0.57
10.62
Sips R2 - 0.94 - 0.87
0.86
RMSE - 3.39 - 1.14
0.81
NRMSE - 11.41 - 8.95
11.31
AIC - 28.23 - 6.38
12.09
- -
Ks (L mg-1) 3.39e-6 3.07e-6 -
- -
KL (L mg-1) 1.15e-4 2.58e-4 -
- -
Qmax (mg g-1) 115.13 793.37 -
BET - -
R2 0.97 0.96 -
- -
RMSE 36.83 44.41 -
- -
NRMSE 5.21 5.92 -

74
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 5.9 a)BET isotherm fitted (full line) to experimental data (circles) of COD
adsorption using a mixture of 25-75 % of HH-Vinasse and 1.0 gL-1 of Yeast extract and
SBFA as adsorbent at concentrations of 1.5 to 60 g L-1; b) BET adsorption isotherm for
experimental data obtained using a mixture of 100-0 % of HH-Vinasse and 1.5 gL-1 of
Yeast extract

For both adsorption isotherms (experiments 13 and 14) at low COD


concentrations the adsorption was unfavorable. However, as COD concentration
increased the adsorption became favorable. The Qmax value for experiment 13 (793.37 mg
COD g SBFA-1) was higher than that for experiment 14 (115.13 mg COD g SBFA-1), and
the equilibrium constants for first and consecutive layers (Ks, KL) indicate a good binding
intensity, favoring the adsorption process. These results are in good agreement with those
reported by Ebadi et al., (2009), who studied the adsorption of phenolic compounds over
carbonaceous materials. Probably, the COD adsorption isotherm profiles observed in this
study are due to the attractive lateral interactions between adsorbed molecules. According
to Crittenden and Thomas (1998) this occurs when the COD concentration increases to a
limit value where these interactions predominate.
Therefore, it is concluded that the unburned carbon fraction (43.7% ignition loss)
of the SBFA contributes to decrease the inhibition of anaerobic microorganisms by the
toxic compounds present in vinasse and HH. In addition, the SBFA addition favored the

75
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

AD process, providing nutritional supplements and helping with the acclimation and
immobilization of bacterial cells (Fagbohungbe et al., 2017).

5.4.4 Energy Analysis

The results of the energy balance for the best experimental conditions are
summarized in Figure 5.10. As can be seen in Figure 5.10, surplus energy was obtained
for two AcD scenarios that exhibited the highest methane production employing vinasse,
HH, YE (1.0 and 1.5 g L-1) and SBFA in the pre-defined optimized conditions.
The energy balance for experimental conditions that used the highest SBFA
concentration (75 g L-1) for both pure HH and 25-75% HH-to-vinasse were negative,
showing that the inhibitory effects of dissolved cations, such as K, Ca and Mg and/or the
presence of high amounts of FF and HMF decreased the performance of anaerobic
digestion. The best scenarios with energy surplus were those with pure HH with 45 g L-1
of SBFA and 1.5 g L-1 of YE (0.34 MJ kg SB-1) and 25-75% HH-to-vinasse with 15 g L-
1
of SBFA and 1.5 g L-1 of YE (0.42 MJ kg SB-1). The energy surplus for pure HH
obtained in the present study is much better than that reported by Baêta et al., (2016a),
who reported an energy deficit of 0.58 MJ kg SB-1, during anaerobic digestion of pure
HH supplemented with a synthetic solution of macro- and micronutrients.
Figure 5.11 shows a flowchart of the process proposed in the present study for an
integrated 1G/2G or autonomous mill to maximize energy generation from AcD,
considering a CHP system for conditions of the assay 13 (25 % HH - 75% Vinasse) with
1.0 g.L-1 of YE and 15 g.L-1 of SBFA. Considering the boundary conditions described in
the Section 5.4.2.1 and the conditions for 2G ethanol production described by Baêta et
al., (2016b), i.e. weight loss of 29.5% after SB hydrothermal pretreatment (HP); an
enzymatic hydrolysis efficiency of 62.18% (using SB pretreated by HP); an yield of
0.42 g ethanol g glucose-1 in yeast fermentation; and an ethanol density of 789 g L-1; it is
possible to obtain 1.35 MJ kg SB-1 of thermal energy and 0.202 kWh kg SB-1 of electric
energy.

76
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

However, considering that the thermal energy required for the pretreatment (1.66
MJ.kg SB-1) was higher than thermal energy generated in CHP systems (1.35 MJ kg SB-
1
), part of the thermal energy necessary to complete the SB pretreatment (0.31 MJ kg SB-
1
) can be provided from surplus electric energy generated by a CHP system (0.202
kWh kg SB-1). Thus, the net energy generated by AcD system would be able to supply all
energy expended in the sugarcane bagasse pretreatment with a surplus of
0.12 kWh kg SB-1 of electric energy. The results obtained in this scenario are in good
agreement with those reported by Rodriguez et al., (2018), who also evaluated the
potential energy that could be generate from AD of vinasse as a single substrate within
the context of an integrated 1G/2G mill using simulation tools and found that is possible
obtain 1.57 MJ kgBC-1.

Figure 5.10 Energy balance for the process proposed using AcD of HH and vinasse with
addition of YE (1.0 and 1.5 g L-1) and SBFA (15, 45 and 75 g L-1) for two proportions
of substrates.

77
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 5.11 Flowchart of the process proposed for managing sugarcane and its wastes at
an 1G/2G ethanol production process in a biorefinery concept.

78
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

6 CONTINUOUS BIOGAS PRODUCTION BY ANAEROBIC CO-DIGESTION


OF SUGARCANE BYPRODUCTS: ANALYSIS OF REACTOR
CONFIGURATIONS

ABSTRACT
In this chapter we present data that aimed to address the objectives 3 and 4 (section
4.2) and to test the hypothesis CH2 (section 3.2) of this thesis. This study considered the
application of anaerobic co-digestion (AcD) using two mesophilic reactors fed
continuously with hemicelluloses hydrolysate and vinasse. A single stage hybrid
anaerobic reactor (HAR) was compared with a two-stage acidogenic-methanogenic
system formed by an acidogenic structured-bed reactor (ASTBR) followed by an UASB
methanogenic reactor. Increasing organic loading rate (OLR) (from 0.9 to 10.8 g COD L-
1
d-1) were applied to these systems by fixing the hydraulic retention time (HRT) in the
reactors (17.5 h in HAR; 6 h in the ASTBR and 19.9 h in the UASB) and changing the
influent COD. The results showed the feasibility of applying the two-stage system
(ASTBR/UASB) to treat the HH-vinasse mixture, leading to a global COD removal
higher than 80% and methane yield of 0.245 NL CH4 g CODr-1. In its turn, the single
stage (HAR) system attained 65% of COD removal and 0.205 NL CH4 g CODr-1.
Microbial community analyses of sludge collected from the ASTBR at different
operational conditions (OLR = 0.9 g COD L-1 d-1 (phase I) and OLR = 10.8 g COD L-1 d-
1
(phase II)) revealed the main species were Clostridium beijerinckii (61.9%) and
Desulfovibrio desulfuricans (19.1%) for phase I; and Lactobacillus casei (66.3 %) and
Lactococcus lactis (22.4 %) for phase II. Regarding the UASB reactor, the main identified
species were Longilinea arvoryzae (15.5%), Methanosaeta concilii (19.9 %) and
Syntrophus aciditrophicus (15.4%) at phase I; and Longilinea arvoryzae (31.7%),
Methanosaeta concilii (10.6 %) and Pleomorphomonas oryzae (10.1%) at phase II. In its
turn, the main species identified in the HAR sludge were Bacteroides graminisolvens
(11.3%), Desulfovibrio desulfuricans (12.6 %) and Methanosaeta concilii (16.4%.) Both
systems (single HAR and ASTBR-UASB) exhibited a stable long-term operation (240
days) with low VFA accumulation (average of 550 mg/L at HAR and 625 mg/L at
UASB).
79
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

INTRODUCTION
Biofuels and bioenergy are terms normally associated with fossil-free systems that
must guarantee mainly no competition with food production, low environmental impacts
(i.e. water consumption, land use) and a positive net balance of energy (generated minus
spent). In this sense, the sugarcane biorefinery context offers different options for using
its byproducts combined with green technologies to obtain energy and value-added
products. The six main byproducts (or ‘wastes’ in some cases) obtained during 1G/2G
ethanol production are ashes, bagasse, pentoses liquor, filter cake, molasses and vinasse.
Some of these byproducts have interesting physicochemical characteristics, as described
in chapter 5, which allow to explore different technological routes for their better use.
Anaerobic digestion may be considered the primary alternative for managing such
‘wastes’ in sugarcane biorefineries (Cortez, Baldassin e Almeida, de, 2020a).
Consequently, different authors have studied some of these byproducts (on their own,
without mixing them) as substrates for methane production (Janke et al., 2015; López
González, Pereda Reyes e Romero Romero, 2017; Parsaee, Kiani Deh Kiani e Karimi,
2019; Ribeiro et al., 2017b) with varying and controversial outputs. Indeed, the direct
utilization of these substrates is difficult because of nutritional imbalance, presence of
inhibitory compounds, lack of diversified/adapted microorganisms, amongst other
operational/environmental factors. Thus, co-digestion processes have been recommended
to overcome these drawbacks (Hagos et al., 2017).
For industrial biogas production, the operation of continuously fed anaerobic
digestion reactors is highly recommended since it helps to keep a sustainable supply of
biogas and is relatively cheaper to treat larger flowrates. In addition, laboratory scale
batch reactors cannot reflect the performance of mass transfer and biogas production that
is observed in continuously fed systems (Huang et al., 2017). Usually, a variety of reactor
configurations (CSTR, UASB, AnMBR, EGSB, fluidized and packed bed reactors)
operating as single or two-stage systems have been used for anaerobic digestion of
individual sugarcane byproducts (Fuess, L T et al., 2018; Hagos et al., 2017; Hoarau, J et
al., 2018; Marafon et al., 2020; Surendra et al., 2015). It is important to highlight that at
a single stage reactor, all biochemical reactions occurs simultaneously in one vessel

80
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

whereas in a two-stage system the hydrolytic-acidogenic stage is spatially separated from


the methanogenic one; in this way acidogenic microbes could be stimulated to produce
more enzymes, enhancing their ability to hydrolyze and convert sugars and complex
organic molecules into simpler compounds that could be used as substrates in the
methanogenic stage (Gunes et al., 2019; Hagos et al., 2017). However, inappropriate
selection of co-substrates proportions as well as co-substrate compositions and operating
conditions can lead to process instability and significant reduction of methane production
in reactors continuously fed (Xie et al., 2016).
Based on the above, the aim of this chapter was to evaluate and compare the effect
of reactor configurations (single and two-stage systems) and applied organic loading rate
(OLR) during continuous methane production using AcD of vinasse, hemicelluloses
hydrolysate, yeast extract and sugarcane bagasse fly ashes. It is hoped this study can offer
important information, within the sugarcane biorefinery context, for biogas production at
1G-2G ethanol production process, from the point of view of AD application and reactors
operation.

MATERIAL AND METHODS

6.3.1 Feedstocks for anaerobic co-digestion (substrates and inoculum)

Hemicelluloses hydrolysate (HH) was obtained by hydrothermal treatment (HT)


of sugarcane bagasse (SB) using the optimized conditions (T=183°C, reaction
time=41 min and liquid-to-solid ratio of 3.94 mL/g SB) defined by Baêta et al., (2016a).
SB was provided by Jatiboca Sugar and Ethanol Plant (Ponte Nova, Brazil) and the
sample was from the 2017/2018 harvest. Vinasse was sampled from a first-generation
(1G) ethanol distillation system at Agropéu Sugar and Ethanol Plant (Pompéu, MG,
Brazil). SB fly ash (SBFA) was sampled after SB combustion in a high-pressure boiler
and was provided by the Minas Gerais Sugarcane Industries Association (SIAMIG, MG,
Brazil). The feed flow was a volumetric mixture of HH-to-Vinasse in a fixed proportion
of 25-75 %, and was properly diluted with tap water to yield the desired OLR (Table 6.1)
for each operational phase (I, II, III). Psychochemical characteristics of HH, vinasse, yeast
81
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

extract and SBFA are shown in Table 5.2. The inoculum used was a mixture of fresh
bovine manure (FBM) - sampled from local cow farms - and sewage sludge - sampled
from a pilot-scale mesophilic upflow anaerobic sludge blanket (UASB) reactor fed with
sieved sewage at the Centre for Research and Training on Sanitation (CePTS),
UFMG/Copasa (Arrudas Wastewater Treatment Plant, Belo Horizonte, MG, Brazil). The
inoculum was prepared by mixing equal amounts of UASB sludge and FBM in order to
give an initial concentration of 0.5 g VSS gTS-1 (Lima et al., 2018).

6.3.2 Experimental setup and operating conditions

The experimental two-stage system consisted in an acidogenic structured bed


reactor (ASTBR) followed by a methanogenic UASB reactor. The ASTBR had a useful
volume of 1.85 L and 0.370 L of headspace; and was filled with low-density polyethylene
(LDPE) as support material (Figure 6.1). The UASB reactor had a 2.3 L of effective
volume and 0.450 L of headspace (Figure 6.1). In its turn, the experimental single stage
system consisted in a new hybrid anaerobic reactor (HAR) which was basically an UASB
with a structured bed ( Figure 6.2). The HAR had a working volume of 1.45L and a
headspace of 0.350 L.
All anaerobic reactors were built with rigid polyvinyl chloride (PVC) tubes and
valves. In order to maintain the desired temperature of 35°C, a W1209 self-contained
temperature controller module was used for heating the reactor with a resistance heating
cord of 158 𝑜ℎ𝑚𝑚-1.
The ASTBR reactor was self-inoculated by pumping the feed solution with pH
4.5 and recirculating it for approximately 8h. The methanogenic reactors (UASB and
HAR) were inoculated with the mixture described in section 6.3.1 ensuring a volatile solid
(TVS) concentration inside the reactors of 10gL-1.
Sugarcane bagasse fly ashes and yeast extract were added to each methanogenic
reactor for attaining a concentration of 15 gL-1 and 1.5 gL-1 respectively, in accordance
with the optimal conditions described before (section 5.4.2.3). The ASTBR feed solution
was adjusted to give an initial pH value of 5.0 – 5.3. The UASB reactor was fed with the
effluent of acidogenic reactor after adjusting its pH to 7. The HAR influent also had its

82
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

pH adjusted to 7. Sodium hydroxide (40 gL-1) was used to adjust the pH of all feed
solutions.
Influent was pumped into the reactors from their bottom using peristaltic pumps
(Milan, model BP600, with two channels). The operational conditions (OLR, HRT, pH)
for the described systems as well as the phase operations (I, II and III) studied are
summarized in the Table 6.1.

Table 6.1 Operational conditions for single and two-stage anaerobic reactors
Reactor Reactor Phase I Phase II Phase III
configuration type
OLR: 1.25 g OLR: 2.50 g CODL- OLR: 6.05 g CODL-
CODL-1d-1 1 -1
d 1 -1
d
HRT: 0.25 d HRT: 0.25 d HRT: 0.25 d
ASTBR
Operational time: Operational time: 60 Operational time: 69
90 d d d
pH: 5.5 pH: 5.5 pH: 5.5
Two-stage
OLR: 0.94 g OLR:1.61 g CODL- OLR: 3.74 g CODL-
CODL-1d-1 1 -1
d 1 -1
d
HRT: 0.83 d HRT: 0.83 d HRT: 0.83 d
UASB
Operational time: Operational time: 60 Operational time: 69
90 d d d
pH: 7 pH: 7 pH: 7
OLR: 1.39 g OLR:2.38 g CODL- OLR: 4.81 g CODL-
CODL-1d-1 1 -1
d 1 -1
d
HRT: 0.78 d HRT: 0.78 d HRT: 0.78 d
Single stage HAR
Operational time: Operational time: 60 Operational time: 78
87 d d d
pH: 7 pH: 7 pH: 7

83
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 6.1 Schematic diagram of two stage anaerobic system consisting in acidogenic
structured bed reactor (ASTBR) and upflow anaerobic sludge blanket (UASB)

Figure 6.2 Schematic diagram of single stage anaerobic system using a hybrid reactor
(HAR)

84
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

6.3.3 Performance evaluation: analytical methods

The response variables used to assess the performance of the methanogenic


reactors included COD removal (CODr, in %), biogas flow rate (mL d-1), volumetric
methane production rate (VMPR, in LCH4 L-1 d-1), methane yield (YM, in NmL CH4
gCODr -1). Therefore, the following section focuses on the methods applied to anaerobic
reactors monitoring.

6.3.3.1 Characterization of gas fractions

Biogas flowrate was measured using gas meters (Ritter Milligascounter Type
MGC-1 PMMA, Dr.-Ing. Ritter Apparatebau GMBH & Co. KG, Bochum, Germany)
coupled to the reactors headspace (Figures 6.1 and 6.2). Biogas composition in terms of
methane was determined by chromatography as described in section 5.3.10.1.

6.3.3.2 Characterization of liquid fractions

The concentration of sugars (cellobiose, glucose, xylose and arabinose), organic


acids (acetic, formic, propionic, butyric, isobutyric, valeric and isovaleric acids) and sugar
breakdown products (FF and HMF) in influent and effluents from each anaerobic reactor
were determined by high performance liquid chromatography (HPLC) using a Shimadzu
HPLC system equipped with an Aminex HPX 87H column (300 × 7.8 mm, Bio-Rad®) at
55°C (Shimadzu, model CTO-30A). A refractive index (Shimadzu, model RID-6A)
detector was used for sugars quantification and a UV-Vis (Shimadzu SPD-10AV) set at
210 nm for organic acids and at 274 nm for FF and HMF detection. Aqueous 5 mmol L-
1
H2SO4 solution was used as eluent at a constant flowrate of 0.6 mL min-1.
The chemical oxygen demand (COD) (method 5220D), pH (potentiometric
method) and volatile suspended solids (VSS) (method 2540 G) were determined
according to the Standard Methods for the Examination of Water and Wastewater
(APHA, 2005). Total phenolic compounds were quantified spectrophotometrically by the
Folin–Ciocalteu method (Molina-Cortés et al., 2019). A sample aliquot of 1.0 mL was

85
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

mixed with 9 mL of acetone (70%) and ultrasound for 20 min at 4°C in order to extract
the analyte. After that, the sample was centrifuged (3000 rpm for 15 min) so that 150 µL
of supernatant was mixed with 520 µL of Folin-Ciocalteu reagent (a solution of
phosphomolybdate and phosphotungstate), 1.25 mL of Na2CO3 (7 % wv-1) and 350 µL
of distillated water. After 35 minutes of incubation in the dark, the amount of total
polyphenols was measured by absorbance at 725 nm using distilled water as reference
cell (zero adjustment). A calibration curve was built for phenolic compounds using tanic
acid as a standard. A standard stock solution of 100 mgL-1 was properly diluted to yield
a calibration curve in the range of 22.026 to 0.881 mgL-1.

The aqueous samples were also characterized by ions chromatography using a


Metrohm equipment (model 930 Compact IC Flex) equipped with METROSEP C4
150/4.0 and METROSEP A Supp 10-150/4.0 columns coupled to a conductivity detector
to quantify cations and anions, respectively. The chromatographic conditions for cations
(K+, Ca2+, Mg2+, Na+ and NH4+) separation employed 0.0017 mol L-1 HNO3 and aqueous
dipicolinic acid (0.0007 mol L-1) solutions as mobile phase which was pumped at a
constant flowrate of 0.9 mL min-1 at 25°C. The conditions for anions (Cl-, NO2-, NO3-,
PO43- and SO42-) separation employed a mobile phase composed by aqueous 0.005 mol L-
1
Na2CO3 solution and aqueous 0.005 mol L-1 NaHCO3 solution pumped at a constant
flowrate of 1.0 mL min-1 at 45°C.

6.3.3.3 Characterization of solid products

The analyses of solid samples, collected from the bottom of anaerobic reactors
and corresponding to sludge (ASTBR), sludge+SBFA (UASB; HAR), as well as SBFA
were taken on the 150th day of operation. About 1.5 g of each material was placed on
glass slides which were dried at 45 °C for 4 h. Characterization of all solids was performed
by scanning electron microscopy coupled to energy dispersive X-ray spectroscopy (SEM-
EDX) (Tescan, model Vega 3).

86
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

6.3.3.4 Metagenomic characterization of reactors sludge.

Microbial samples were obtained on days 90th (operational phase I) and 230th
(operational phase III) from both two-stage (ASTBR+S-UASB) and single stage (HAR)
anaerobic systems. The OLR at the moment of sampling was that defined at Table 6.3 for
each configuration and operational phases I and III. The sludge samples for microbial
analysis were immediately stored at -20°C until DNA extraction was carried out. The
protocol used for DNA extraction, PCR reactions conditions, primers used and clustered
sequences processed are descried and detailed in Christoff et al., (2017).

6.3.3.5 Kinetic analysis

A kinetic analysis was performed for single and two-stage anaerobic systems to
determine kinetic parameters regarding COD consumption and methane production.
MatLab software was used to adjust kinetic parameters allowing verifying which model
best fit the experimental data gathered at different operation conditions. For this purpose,
two kinetic models were tested: First-order (eq. 8) and Monod (eq. 9). Moreover, the mass
balance in a transient state was described by eq. 10, assuming that the ASTBR, UASB
and HAR reactors behaved as continuously stirred tank reactors in series (CSTR), as
found by the hydrodynamic tests (see next section).
−rCOD = k1  COD (8)

max  COD
−rCOD = (9)
K S + COD

dCOD
V = Q  CODi − Q  COD − rCODV  X o (10)
dt

where V is the reactor volume (L); Q is the influent or effluent flowrate (Ld-1);
CODi is the influent COD rate (gL-1d-1); COD is the effluent COD rate (gL-1d-1); r is the
COD degradation rate (gL-1d-1); k1 is the first-order kinetic constant (d-1); Xo is the initial
microorganism concentration (estimated as VSS, mg VSSL-1); µmax is the maximum
specific microbial growth (gL-1d-1); and Ks is the saturation constant (gL-1).

87
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

6.3.4 Hydrodynamic analysis

The hydrodynamic behavior of the acidogenic and methanogenic reactors was


studied using the residence time distribution curve (RTD) technique. The RTD was
experimentally measured through the pulse-type stimulus-response assay, using a mass
of 0.5 g of lithium chloride (LiCl) as a tracer. The concentration of the reactor was
measured at the reactor effluent outlet, by collecting samples every hour during a period
equivalent to 3 HRTs for each reactor. The samples collected were filtered through a 0.45
μm cellulose acetate membrane so that lithium could be determined by an ion
chromatography (Metrohm) equipped with a 732 - IC detector and a Metrosep A SUPPP
5 - 250 (Metrohm) column. The mathematical models and methodology for finding the
RTD parameters (average residence time, axial dispersion number, number of CSTR
reactors in series, percentage of dead zones) were calculated in accordance with
Levenspiel, (2012).

RESULTS AND DISCUSSION

6.4.1 Organic matter removal at different organic loading rates

6.4.1.1 Single stage system

For the single stage system, the hybrid anaerobic reactor (HAR) was inoculated
with the sludge mixture described in section 6.3.1 (10 g VSL-1), sugarcane bagasse fly
ash (15 gL-1) and yeast extract (1.5 gL-1) in accordance with the best conditions obtained
for batch operation, as described in section 5.4.2. The system was fed with a mixture of
HH-to-Vinasse (25-75%) and operated during 240 days in three periods under different
OLR (Table 6.1). OLR was progressively increased by reducing dilution and increasing
influent concentration. HAR performance throughout its operation is shown in Figure 6.3.
As can be seen, during the first 50 days the COD removal efficiency changed from 83.5%
88
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

to 37.3%, suggesting the anaerobic inoculum was adapting to the environment and toxic
and/or recalcitrant compounds present in HH and vinasse (Huang et al., 2017).
Subsequently, between days 60 and 88, it was observed a stabilization of COD removal
efficiency which reached 82.9%; and these results are in good agreement with those
reported by Rajagopal et al.,(2010) who found COD removal of 84% using hybrid
anaerobic reactor fed with vinasse with an HRT of 25h and recirculating system.
After the first phase, the OLR was gradually increased from 1.39 to 2.38 g COD
L-1 d-1 (phase II) and then to 4.81 g COD L-1 d-1 (phase III). The HAR was operated for
almost 70 days and 80 days at phases II and III, respectively. As expected, the efficiency
of the anaerobic process in terms of COD removal decreased with the OLR increase. The
lower COD removals observed for phases II (62.1%) and III (52.3%) indicates a possible
inhibition by toxic compounds, such phenolic fractions and furans, which are produced
during HT of SB and, hence, are present in HH. To investigate this hypothesis, data of
furan and phenolic compounds should be addressed.
In the case of furans, two reasons could explain their removal during AD. The first
one is related to the role of SBFA as adsorbent, as discussed in the section 5.4.3 which
showed that HMF and FF could be completely adsorbed since they were not found in
anaerobic effluents of each operational phase. The second reason is related to the ability
of sulfate-reducing bacteria (SRB) to completely convert furans to acetate, as elucidated
by Boopathy, Bokang e Daniels, (1993) and Wierckx et al., (2011).
By its turn, phenolic fractions were detected and measured in anaerobic effluents,
and corresponded to 24.87 and 94.39 mg total phenols L-1 for phases II and III,
respectively. Is important to note that low molecular weight phenolic compounds (e.g.
lignin fractions) are considered more toxic to microorganisms than high molecular weight
ones (Klinke, Thomsen e Ahring, 2004). In this sense, the results obtained in this study
are confirmed by authors like Barakat et al., (2012) and Li et al., (2018) who show that
when phenolic derivatives are used in anaerobic digestion, the measured biochemical
methane potential (BMP) are only 7 to 25.9% of the theoretical values, therefore, lower
COD removal efficiencies are expected.

89
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Another important characteristic related to COD removal is the reactor


configuration. Normally, high-rate anaerobic systems have the ability to effectively
separate HRT and solid retention time (SRT) so that an efficient reactor design (i.e. lower
volume) is attained. Thereby, the HAR studied here has similar behavior compared to the
configurations reported by Hoarau, J et al., (2018) who showed that continuously fed
systems such as Anaerobic Fluidized Bed Reactor (AFBR), Biphasic charcoal packed
Anaerobic Reactor(BcpAR) and Upflow Fixed Film Column Reactor (UFFCR) reached
COD removals from 64 to 88 % for OLRs varying from 11.33 to 30 g COD L-1 d-1.
However, it is important to highlight that the HRT used by Hoarau, J. et al., (2018)
(3.5<HRT<15 days) were greater than those reported by this study (HRT of 0.78 day;
2.38 g COD L-1 d-1). In this sense, the HAR used in this study could process nearly 4.5
times the amount of raw materials (HH and vinasse) used, reaching comparable OLR
between similar reactor configurations. Moreover, is important to note that the systems
described (AFBR, BcpAR and UFFCR) used vinasse as unique substrate while the HAR
system was fed with byproducts from 1G and 2G ethanol production process, which can
be more recalcitrant.

Figure 6.3 COD behavior in HAR along the three operational phases with varying OLR.

90
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

6.4.1.2 Two-stage system

COD removal profile for acidogenic reactor ASTBR is shown in Figure 6.4a for
three operational phases which lasted 240 days. As can be seen, the COD removal at the
ASTBR reached 31.1% and 28.5% during the first and second phase for an OLR applied
of 3.75 g COD L-1 d-1and 5.65 g COD L-1 d-1, respectively. The values obtained are
consistent with the values typically associated with the operation of acidogenic reactors
which normally do not remove COD, instead convert complex organic matter in organic
acids by fermentative reactions without any net loss of electrons to common electron
acceptors (e.g. CO2). This process often generates H2 and large amount of small chain
volatile fatty acids (VFA) that can be used at the methanogenic stage (Ferraz Júnior,
Etchebehere e Zaiat, 2015; Fuess, L.T., Kiyuna, et al., 2017).
On the other hand, during phase III (OLR= 10.81 g COD L-1 d-1), COD removal
decreased to 10.2%, and this was probably related to high levels of total phenolic
compounds in the effluent (140.89 mgL-1) which might have inhibited acidogenic
microorganisms (Koyama et al., 2017). It is important to note that typical start-up
conditions for fixed biomass reactors are based on a progressive increase of the OLR by
means of HRT reduction while keeping the COD concentration constant (Toledo-
Cervantes et al., 2018). However, the strategy adopted in this study was different, and
based on the reduction of the dilution factor, in order to evaluate the effect of toxic species
on the anaerobic co-digestion process. During phase III, the dilution factor employed was
10, thereby indicating that the dilution effect offered by the vinasse was not sufficient to
reduce the effect of the toxic compounds present in HH. Thus, from a practical point of
view, it would be necessary to use alternative flows (i.e. sewage) that allows enhance the
performance of AcD process.
Subsequently to the acidogenic stage carried out in the ASTBR system, the pH of
the acidogenic effluent was adjusted from 5.7 to 7.0 in order to be fed to the methanogenic
stage (UASB reactor). Figure 6.4b illustrates the performance of COD removal
throughout UASB reactor operation at the different applied OLR. The lower values of
OLRs applied to the methanogenic reactor were a consequence of the COD removal in
the ASTBR. As can be seen, after reactor stabilization (final period of each phase) the
91
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

mean values attained for COD removal were 80.51%, 87.60% and 73.36% for phases I,
II and III respectively. Such removal efficiencies are comparable with those reported for
the anaerobic digestion of vinasse using an UASB-PBR operated with a HRT of 2 d
(86.7%) (Cabrera-Díaz et al., 2017); a PBR+UASB with a HRT of 0.83 d (76%) and a
PBR-STBR operated with the same HRT of 0.83 d (82%) (Fuess, L.T., Kiyuna, et al.,
2017).
Although, the increase in OLR applied to the ASTBR can promote a slight
accumulation of VFAs in the UASB reactor, the results have shown a robust performance
of the two-stage anaerobic system. The structured bed configuration, which allowed
retaining a high amount of biomass in the reactor, enabled a global COD removal of
72.5%. Moreover, the UASB reactor was inoculated with SBFA which was previously
shown to contribute to anaerobic digestion by adsorbing toxic compounds. Thus, an
adsorption effect of SBFA over the residual toxic compounds (sulfate and phenolics
fraction) from the acidogenic stage seemed to have enhanced the removal and degradation
of organic matter.

Figure 6.4 COD profile at influent and effluent of ASTBR (a) and UASB (b) reactors
operated at different OLRs

92
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

6.4.2 Biogas production and methane yield

6.4.2.1 Single stage system

Biogas production and methane yield in HAR is showed in Figure 6.5 for the
different OLRs tested (1.39, 2.38 and 4.81 g COD L-1 d-1). The average methane yield
was 170.24, 227.63 and 209.45 NmLCH4gCODr-1 for phases I, II and III respectively.
This values are lower than those reported by Rajagopal et al., (2010) and Cabrera-Díaz et
al., (2016) whose yields varied from 330 to 340 mL CH4 g CODr-1 at OLRs ranging from
5 to 11.8 g COD L-1 d-1. This difference might be partially explained due to the formation
of granules in the HAR which floated and had less contact with the substrate. For some
reason which is not yet clear, the granules formed in HAR and collected at the end of
phases I and III had gas trapped inside them, as illustrated in Figure 6.6.
At this point, is important to highlight that anaerobic sludge granulation is
desirable since the granules are normally denser than ordinary flocs, thereby allowing
higher upflow velocities and lower HRT in anaerobic reactors. However, granulation is
quite complex and is affected by many physicochemical and operational parameters.
Although several hypotheses have been continuously made regarding the granulation
process in anaerobic digesters, it is still not clearly understood. Some recent studies have
reported the use of precursor substances such as Ca+2 ; Fe+2 and Mg+2 ions to accelerate
sludge granulation. (Faria et al., 2019).Therefore, in this study, these ions could be
provided by SBFA in accordance with its psychochemical characteristics, as discussed in
section 5.4.1.
The average sulfate removal observed in HAR during phases I, II and III were
95.3%, 92.1% and 69.8% respectively, showing that for a COD/SO4-2 ratio that was kept
nearly constant, the biogas production is related to the carbon source used due to the
inhibitory compounds added or generated during AD. Thus, specific operational
conditions such as HH and vinasse characteristics as well as reactor type can change this
behavior (Cetecioglu et al., 2019). On the other hand, the average removal of phenolic
compounds obtained for phases I, II and III corresponded to 89.5%, 70.1% and 36.4%
respectively. Therefore, phenol degradation seemed to be facilitated in anaerobic systems

93
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

characterized by the concomitant establishment of methanogenic and sulfate-reducing


conditions (Godoi, de et al., 2019).

Figure 6.5 Biogas production and methane yield for three different OLR in HAR.

Figure 6.6 Granules sampled from HAR at two different OLR corresponding with phase
I and phase III

6.4.2.2 Two-stage system

In this section, it is discussed the biogas production and methane yield in two-
stage anaerobic system (ASTBR+UASB). Normally, it is well known that decoupling the

94
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

acidogenic and methanogenic stages of provides a better environment for the anaerobic
digestion, thereby enhancing the reactor stability and biogas production (Solera, Romero
e Sales, 2002). Initially, it is important to note that hydrogen production in acidogenic
stage was not addressed in this work, since its main purpose was to hydrolyze the complex
organic matter (e.g. xylolygomers) and maximize the VFA production to easy methane
formation in the methanogenic reactor. Additionally, the acidogenic phase was set in an
attempt to remove toxic compounds, such as HMF and FF as well as sulfate which is an
electron acceptor that competes for substrate and decreases methane formation in
anaerobic reactors.
Figure 6.7a shows biogas production and methane yield in the ASTBR during the
240 days of its operation at three operation phases. In relation to biogas production, as
expected, only low production of biogas and methane yield was obtained (<0.180 L d−1,
25 mL CH4 g CODr -1) in the acidogenic step, hence indicating an effective development
of acidogenic bacteria biofilm and a low abundance of methanogenic microorganisms.
On the other hand, although the operational parameters (↓ HRT and ↑ OLR) of ASTBR
can contribute to this behavior, the presence of some methane in the acidogenic step could
be explained by homoacetogenic processes where microorganisms consume the H2
produced to reduce CO2 and produce methane (Saady, 2013). Moreover, nutrient
deficiency in the ASTBR might favor the dominance of homoacetogenic bacteria which
severely decreases the H2 yield (Ferraz Júnior, Etchebehere e Zaiat, 2015). .

95
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 6.7 Biogas production and methane yield for three different OLR in (a) ASTBR
system and (b) two stage anaerobic system.

The results obtained for biogas production and methane yield in the UASB reactor
(methanogenic stage) are presented in Figure 6.7b. As can be seen, for phase I (OLR= 1.23
g COD L-1 d-1), II (OLR= 2.25 g COD L-1 d-1) and III (OLR= 4.51 g COD L-1 d-1) it was
possible to attain methane yields of 221.8, 242.2 and 193.3 mL CH4 g CODr-1 for each
phase, respectively. The better results obtained in this study (phase II- 242.2 mL CH4 g
CODr-1) is a little lower ( ≈280 mL CH4 g CODr-1 ) than those reported in previous studies
for the treatment of vinasse in UASB reactors (Cabrera-Díaz et al., 2017; Fuess, L.T.,
Kiyuna, et al., 2017). Methane yield depends on substrate composition, operational
conditions provided to the reactor, and also on the synergistic effects caused by substrate
mixing (Pereda-Reyes, Díaz Pagés e Horvart, 2016). Therefore, although the result of
methane yield in the two stage system (UASB reactor) is similar to that obtained in the
one-stage system (HAR), the global OLR used is higher (2.89 g COD L-1 d-1), so it would
be possible to process a higher amount of substrate in the two-stage system, thereby
leading to more methane and energy.

6.4.3 Volatile fatty acid analysis

6.4.3.1 VFA profile in the acidogenic reactor (ASTBR)

Total VFA profile and its relation with COD removal in the ASTBR is show in
Figure 6.8. During phases I, II and III the total VFA accumulated in the acidogenic reactor
averaged of 692.5, 911.7 and 2329.2 mgL-1 which corresponded to 81.1, 69.4 and 86.75%
of the soluble COD from the acidogenic reactor, respectively. Moreover, the main acids
present in the effluent of ASTBR were acetic and propionic. Then, total VFAs
concentration in the effluent increased throughout the reactor operation (Figure 6.8),
resulting in lower COD removals, and these results are in good agreement with those
reported by (Cabrera-Díaz et al., 2017) who used UASB as acidogenic reactor, HRT of 2
days and vinasse as substrate. It is important to highlight that the major presence of acetic
acid in the effluent during the operational phases seems to be directly related to the
96
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

mechanisms involved in the removal of sulfates, furans and phenolic compounds (Godoi,
de et al., 2019; Wierckx et al., 2011). This is because some microorganisms, such as the
SRB, which normally predominate in acidogenic reactors due to the shorter HRT, have
the ability to use a wide range of organic compounds as electron donors or carbon source,
including hydrocarbons and aromatic compounds (Boopathy, Bokang e Daniels, 1993;
Liamleam e Annachhatre, 2007). The use of such toxic compounds by SRB normally
generates acetate and helps to biodetoxify the effluent before the methanogenic step.
However, an excess of acetate may cause inhibition due to the kinetic mismatch between
producers and consumers of such substrate. Therefore, finding optimal concentrations of
furans, sulfates and phenols allows to optimize the production of methane in reactor after
the acidogenic stage.

Figure 6.8 Total volatile fatty acids (VFA) profile in mg/L (line and square) and is
mgCOD/L (dashed line) accumulated in the ASTBR along the three operational phases.

6.4.3.2 VFA profile in methanogenic reactors of single (HAR) and two-stage (UASB)
systems

The stability of anaerobic digestion process can be evaluated as a function of the


total VFA that accumulates in methanogenic reactors. As can be seen in Figure 6.9a, the

97
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

increase in VFA concentration in HAR as the OLR is increased indicates an incomplete


conversion of VFA to the final targeted end product (methane), which probably occurred
due to phenols toxicity as previously discussed (see sections 6.4.1 and 6.4.2). In addition,
during phase III, the allegedly inhibitory VFA concentration (900 mgL-1 according to
(Wang et al., 2009)) was nearly exceeded, which might explain the decrease in methane
yield. On the other hand, the volumetric methane production rate increased in HAR with
the increase of the applied OLR reaching 0.8 L CH4 L-1 d-1 in phase III. These results are
lower than those reported by Cabrera-Díaz, A et al., (2016) (1.13 L CH4 L-1 d-1) and
Acharya, Mohana e Madamwar, (2008) (1.44 L CH4 L-1 d-1) who used similar reactor
configurations (fixed bed reactor) but fed the reactors with vinasse as the main substrate.
Figure 6.9b also shows the VFA profile during the three operation phases for
reactor UASB. The total and the individual VFA concentration tended to increase with
the increase of the OLR as reported by Cabrera-Díaz et al., (2017). The main VFAs that
accumulated in the UASB reactor were acetic and propionic acid; their concentration
were in equal proportion and always remained below 600 mgL-1 when the reactor
stabilized at a given operational phase.
Regarding the volumetric methane production rate, the results presented here
corroborate what was previously discussed for the operational reactor stability and
organic matter removal, indicating the better performance of UASB compared to HAR.
The better condition offered to the UASB reactor led to a volumetric methane production
rate of ~0.9 CH4 L-1 d-1, which is in good agreement with the value reported by Fuess,
L.T., Kiyuna, et al.,(2017) who reached 1.04 CH4 L-1 d-1 in a UASB reactor fed with
vinasse at an OLR of 15 g COD L-1 d-1 and thermophilic conditions.

98
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 6.9 Total volatile fatty acids accumulated (line and square) as function of
volumetric methane production rate (dashed line) in HAR (a) and UASB (b) reactors.

6.4.4 Microbial community analysis at different OLR

6.4.4.1 Single stage system

High-rate methanogenic reactors transform organic pollutants into methane


thereby removing COD from the liquid phase. Monitoring microbial community
dynamics is very helpful for predicting and explaining functional changes that occurred
in the anaerobic reactors when they were fed with HH and vinasse under different OLR
conditions. To reveal the microbial shift in response to increasing OLR for HAR,
qualified reads were analyzed at species level (Figure 6.10). Four mainly phylogenetic
groups related to phyla emerged in all samples. These phyla contained Bacteroidetes,
Proteobacteria, Chloroflexi and Firmicutes. The main species which had with a relative
abundance higher than 6% and were identified in HAR when the OLR was 1.38 g COD
L-1 d-1 are: Bacteroides graminisolvens (11.3%), Desulfovibrio desulfuricans (12.6 %)
Pleomorphomonas oryzae (24.8%), Campylobacter sp.(9.1%), Clostridium beijerinckii
(6.3%), Enterobacter aerogenes (6.9%) and Raoultella planticola (9.1%). These species
can utilize several carbohydrates to produce organic acids (e.g., lactate, butyrate,
propionate, acetate and formate) and ethanol (Nishiyama et al., 2009; Xie e Yokota,
2005; Zhang, Ban e Li, 2018). On the other hand, the abundance of Methanobacterium
palustre and Methanosaeta concilii was less than 1.5%.
99
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Desulfovibrio desulfuricans is related to hydrogen producing acetogenic activity.


In low-sulfate media, they could degrade ethanol and lactate to produce acetate, H2, and
CO2 through synergistic association with H2-utilizing methanogens (Zhang, Ban e Li,
2018). In addition, this specie can also transform furans and phenolic compounds in
acetate (Wierckx et al., 2011), thus, its presence in HAR supports the performance and
behavior of such reactor as discussed in sections 6.4.1, 6.4.2 and 6.4.3.

Figure 6.10 Relative abundance of microbial communities at species level for HAR in
two different OLR

When the OLR was increased to 4.81 g COD L-1 d-1 there was a change in the
microbial profile of HAR. As can be seen in Figure 6.10 new relative abundances
appeared for the following species: Desulfovibrio burkinensis (31.7%), Longilinea
arvoryzae (10.5%), Citrobacter freundii (8.6%), Methanobacterium palustre (4.7%) and
Methanosaeta concilii (10.8%). Methanosaeta concilii is a specialist in utilizing acetate
and is identified as the dominant acetotrophic methanogen in various anaerobic reactors
with granular sludge (Sariaslani e Gadd, 2016). This behavior allows enhancing the
methanogenic activity since granular cultures provide a closer spatial microbial proximity
compared to suspended cultures (Faria et al., 2019). In this way, the presence of

100
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Methanosaeta concilii confirms the formation of granules observed during HAR


throughout its operational time.
A Venn diagram (Figure 6.11) was built for evaluating species distribution
amongst samples collected at different OLR conditions. It is seen in Figure 6.11 that
there are 10 common species (58.8%) and that 5 species (corresponding with 29.1%)
changed by increasing the OLR. These results are in agreement with that reported by
Muñoz-Páez et al., (2020) who showed the effect of increasing the concentration of
lignocellulosic hydrolysates on the community of an anaerobic reactor.

Figure 6.11 Venn diagram of species identified in HAR submitted to two different OLR

6.4.4.2 Two-stage system

The two stage AcD allows the selection of different microorganisms and
enhancement of their activity in each reactor. This contributed to increase the stability of
the whole process by controlling the acidification phase in the first reactor and avoiding
the overloading and inhibition of the methanogenic population in the second reactor
(Hagos et al., 2017). To elucidate the microbial shift in response to increasing OLR for
the system ASTBR-UASB, qualified reads were analyzed at species level for each

101
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

reactor. Figure 6.12a reveals that the main species (abundance higher than 6%) that were
present in the ASTBR operating under an OLR of 3.75 g COD L-1 d-1 were Clostridium
beijerinckii (61.9%) and Desulfovibrio desulfuricans (19.1%). Clostridium beijerinckii
can generate products of industrial interest, including H2, using glucose, galactose,
mannose, and xylose as substrate (Fonseca et al., 2019). In addition, during their
exponential growth phase, these bacteria excrete acetate, butyrate, H2, and CO2.
Desulfovibrio desulfuricans can also produce hydrogen and acetate, and help in the
degradation of phenolic compounds as discussed in section 6.4.4.1. Thus, the relative
abundance of this species might explain the VFA profile (predominance of acetate and
butyrate) and the low COD removal in the acidogenic reactor.
Afterwards, when the OLR applied to the ASTBR was increased to 10.8 g COD
L-1 d-1 , there was a change in the microbial profile observed during phase I. As can be
seen in Figure 6.12 the increase in the OLR led to the selection of the following species
(and relative abundances): Lactobacillus casei (66.3 %) and Lactococcus lactis (22.4 %).
These results indicate that lactate, a common product of acidic fermentation, might be a
key intermediate of anaerobic digestion of HH + vinasse at higher OLR. Lactate has not
been observed in higher amounts in the HPLC chromatograms, and this might be related
to its effective utilization by AD microorganisms, as observed by different authors (Fuess,
Lucas Tadeu, Ferraz, et al., 2018). Nevertheless, anaerobic lactate oxidation and lactate
oxidizers, as a physiological group in methane-yielding microbial communities, have not
received enough attention in the context of acetogenesis, which is an essential step of
anaerobic digestion (Detman et al., 2018). In this sense, the identified species in ASTBR
correlate well with the results obtained for VFA profiles (acetic and propionic acid
accumulation, mainly) and reactors performance, since the fermentation of lactate can
follow a metabolic pathway (lactate → propionate + acetate) that has already reported for
anaerobic reactors treating vinasse (Fuess, Lucas Tadeu, Ferraz, et al., (2018)).
It is shown in Figure 6.12b the species identified in the methanogenic UASB
reactor at two different OLRs. When the OLR of 1.23 g COD L-1 d-1 was applied, the
main species, with relative abundance higher than 6%, identified were: Longilinea
arvoryzae (15.5%), Methanosaeta concilii (19.9 %), Syntrophus aciditrophicus (15.4%).

102
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

These species have indeed been identified in anaerobic reactors that used vinasse as
unique substrate (Godoi, de et al., 2019). It is important to highlight that each species has
specific features, i.e. Longilinea arvoryzae has acetate; lactate and H2 as the main
fermentation byproducts(McIlroy et al., 2017); Methanosaeta concilii uses acetate as its
sole energy and carbon source and Syntrophus aciditrophicus is characterized for having
strict anaerobes able to degrade benzoate and VFAs in syntrophy with hydrogenotrophic
organisms (Godoi, de et al., 2018). In this manner, the results presented in sections 6.4.1,
6.4.2 and 6.4.3, regarding methane yield, COD removal and phenol removal in the UASB
reactors correlates well with the phylogenetic analysis presented here.

Figure 6.12 a)The Relative abundance of microbial communities at species level for
ASTBR anaerobic system in two different OLR b)The Relative abundance of microbial
communities at species level for S-UASB anaerobic system in two different OLR

Afterwards, when the OLR applied to the UASB reactor was increased to 4.51 g
COD L-1 d-1 the main species identified were: Longilinea arvoryzae (31.7%),
Methanosaeta concilii (10.6 %) and Pleomorphomonas oryzae (10.1%). Likewise, a
Venn diagram (Figure 6.13) was constructed for evaluating the distribution of species
amongst the single and two-stage anaerobic systems. It is seen in Figure 6.13 that there
were not common species (0%) in the reactors that comprise the two-stage system (ASBR
and UASB), hence confirming a complete phase separation. On the other hand, the single
stage reactor (HAR) had 17 species of which 9 are common to the two-stage system
reactors, mainly to the UASB reactor (7 species).
103
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 6.13 Venn diagram of species identified at the single (HAR) and two-stage
(ASTBR+UASB) anaerobic systems

6.4.5 Hydrodynamics and kinetic analysis

Reactor performance predictions must account for two aspects: i) the rate at which
fluid composition changes, which depends on the kinetic variables; ii) how the fluid flows
through the reactor, which depends on hydrodynamics parameters. Hydrodynamic
parameters obtained for the single and two-stage anaerobic systems are presented in Table
6.2. In the case of ASTBR, the experimental HRT obtained was 8.72 h, slightly higher
than the theoretical value (8 h) calculated by dividing the reactor useful volume by the
influent flowrate. This usually indicates an occurrence of short circuits, which could be
explained by the disposal of the polyurethane foams in the structured bed used. By fitting
the experimental data to the model of ‘N ideal reactors in series’ provided an axial
104
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

dispersion value (D μL-1) of 0.15 and led to a representation of 3.77 CSTRs. From the
practical point of view, it means that the ASTBR behaved as four CSTR arranged in
series. The calculated dead volume was close to zero, indicating that practically the entire
volume of the reactor contributed to the flow regime. Finally, the percentage of plug flow
was nearly 20% while the complete mixture responded to the main flow regime (80%).
Regarding the UASB reactor, the adjustment to the model of ‘N ideal reactors in
series’ provided an axial dispersion value (D μL-1) of 0.32 and corresponded to 2.3
CSTRs. In addition, 90% of stirred flow was present in the UASB, and this completely
mixed flow behavior might have been achieved by scaling the diameter of the reactor
inlet zone (bottom part as shown in Figure 6.1). This allowed the existence of higher
upflow velocities which improves mass transfer and enhance the reactor performance. In
addition, the higher gas production – in relation to the ASTBR – might also have
contributed to the higher mixture in the UASB.
As far as the single stage reactor (HAR) is concerned, it is seen that the adjusted
model of ‘N ideal reactors in series’ led to 1.94 CSTRs with an axial dispersion value (D
μL-1) of 0.91. In addition, the HAR exhibited 85.5% of stirred flow, confirming that
biogas production and scaling the diameter of inlet zone play an import role in improving
mixing, thereby contributing to better mass transfer coefficients.

Table 6.2 Hydrodynamic parameters obtained for single and two-stage anaerobic systems
Model Two-stage Single-stage
parameters ASTBR UASB HAR
Actual HRT (h) 8.72 18.01 16.89
Theoretical HRT (h) 8 18.4 18
Variance ( σ2) 20.17 144.06 50.35
Ideal CSTR in series 3.77 2.25 1.94
Normalized variance (σθ) 0.26 0.44 0.51
Dispersion number 0.15 0.32 0.91
% Plug flow 18.25 11.72 14.48
% Stirred flow 81.74 88.28 85.52
% Dead Volume 1.15 x10-9 2.07 5.91

Once the hydrodynamic characteristics have been defined, the next step is to
establish the kinetic behaviour of organic matter removal. Table 6.3 compiles kinetic

105
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

parameters of models (first order and Monod) that adjusted to the experimental data -
gathered at three different OLR – of COD removal.

Table 6.3 - Kinetic parameters for COD removal in single and two-stage anaerobic
reactors at different operational phases.

Reactor Substrate Parameters Phase I Phase II Phase III


configurati removal
on model
HAR First order Ss (g.L-1) 0.232 0.917 2.155
(single K (d-1) 0.379 0.376 6.173
stage) R2 0.921 0.93 0.5917
RMSE 0.056 0.049 0.450
Monod Ks (g.L-1) 28.093 51.701 15.638
µmax (d-1) 0.445 0.1789 0.0363
R2 0.676 0.552 0.442
RMSE 0.207 0.155 0.519
ASTBR First order Ss (g.L-1) 0.752 1.141 2.894
(two-stage: k(d-1) 0.0481 0.166 3.943
acidogenic) R2 0.645 0.854 0.792
RMSE 0.035 0.086 0.108
Monod µmax (d-1) 3.950 3.564 9.058
Ks (g.L-1) 0.0123 0.0135 0.0051
R2 0.586 0.611 0.542
RMSE 0.096 0.198 0.233
UASB First order Ss (g.L-1) 0.161 0.0746 1.075
(two-stage: k (d-1) 0.0625 0.0455 0.115
methanoge R2 0.941 0.954 0.892
nic) RMSE 0.051 0.078 0.351
Monod µmax (d-1) 0.128 0.120 0.172
Ks (g.L-1) 0.00190 0.00202 0.0011
R2 0.873 0.946 0.774
RMSE 0.087 0.076 0.472

For the ASTBR, UASB and HAR systems, the model that best fit the experimental
data was the ‘first order model’. It is seen that for all reactors, the reaction velocity
constant (k, (d-1)) increased when the applied OLR was increased. This can be explained
by the fact the increase in OLR inputs a higher amount of organic matter to the reactor,
which enhances the rate of degradation in a first order kinetics. In addition, when the
substrate concentration is increased, the possibility of having higher concentrations of
106
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

inhibitory compounds slows the rate of substrate degradation by the microbial consortium
(López, Passeggi e Borzacconi, 2015).

107
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

7 TECHNO-ECONOMIC ANALYSIS OF ANAEROBIC CO-DIGESTION OF


SUGARCANE BIOREFINERY WASTES

ABSTRACT

In this chapter we present data that aimed to address the objective 5 (section 4.2 )
and to test the central hypothesis (section 3.1) of this thesis. This study considered
evaluate the potential energy generated from AcD applied to vinasse and hemicelluloses
hydrolysate within the context of an integrated 1G2G sugarcane biorefinery. Technical
data from chapter 6 were based the scenario modeling and assessment, including
economic and indicators to compare the studied alternatives. Economic indicators showed
the unviability of 1G2G sugarcane mill including AD unit when considering the current
technologies for 2G ethanol production in view of their high capital and operational costs;
however, fixing the experimental methane yield (0.245 Nm3 kg COD-1 r ) obtained for
two-stage system and using 50% of bagasse surplus was possible achieve the internal rate
of return (IRR), return on investment (ROI), and payback period of 21.8%, 59.50% and
10.55 years respectively. On the other hand, when experimental methane yield was
increased in 10% for the same 50% of surplus bagasse used, was possible achieve the
internal rate of return (IRR), return on investment (ROI), and payback period of
26%,89.05% and 5.36 years, respectively. Thus, the results evidenced the importance of
investment in R&D especially in 2G ethanol production (available bagasse) but also in
the anaerobic co-digestion of byproducts to reach the viability of this business.

108
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

INTRODUCTION
The technological development of new integrated industrial process frequently
requires efficient results from scientific and technological research. Alternatively, cleaner
energy production from lignocellulosic materials has been identified as one of the ways
through which environmental impacts, dependence on fossil fuels can be
reduced(Ayodele, Alsaffar and Mustapa, 2020). In this context, ethanol from sugarcane
has an important role in the future energy supply due to Brazil occupies a privileged
position, leading its production (Matos, de, Santos and Eichler, 2020). One of the main
advantages of producing ethanol from sugarcane is the complete use of biomass, i.e. the
bagasse that remains after milling for broth extraction steps, is usually burned in CHP
system to produce electric energy for the industrial operations, which is approximately
30 kWh per ton of cane (Haandel, Van e Lier, Van, 2015). The bagasse required for
energy production depends strongly on the efficiency of the power generation, which in
turn, depends on the steam pressure. Thus, the required bagasse fraction to be burnt for
energy generation ranges from 78% to 30% (Cortez, Baldassin e Almeida, de, 2020b;
Haandel, Van e Lier, Van, 2015). In this sense, excess energy generated can be feed into
the public electricity grid, allowing economic value generation. However, an alternative
of the surplus bagasse, is the use of this fraction for the second generation (2G) ethanol
production. According to Junqueira et al., (2017), the stage of 2G technology allows
increasing the ethanol produced as well as the global sustainability of the biorefinery.
Nevertheless, even today, some authors like Santos et al., (2020b) and Hemansi et al.,
(2019) define that the cost of pretreatment and the energy demanded for this one is still
the main economic obstacle to the commercialization of bioconversion technology of
lignocellulosic biomass in ethanol.
On the other hand, as has been reported in chapters 5 and 6 of this thesis, AcD
process applied to byproducts (e.g. vinasse, hemicelluloses hydrolysate, surplus yeast
among others) generated during 1G and 2G ethanol production could goes beyond its
environmental suitability within the bioethanol production chain, being a potential source
for energy production on the sugarcane mills (Rodriguez, Manochio e Moraes, 2018).
Different authors have evaluated the inclusion of anaerobic digestion of vinasse, as unique

109
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

substrate, in sugarcane biorefineries using modeling, simulation and optimization tools in


order to verify its economic and technical feasibility. In general, the results have shown
that successful energetic integration using AD is directly related to factors like: public
environmental policies, competitive AD technological packages and the final use for the
biogas produced (Fuess e Zaiat, 2018; Longati et al., 2019; Moraes et al., 2014).
Based on the above, the aim of this chapter was to evaluate the techno-economic
feasibility of energy generated from AcD applied to vinasse and HH, in order to know if
anaerobic digestion can be a final step for closing cycle within the context of an integrated
1G2G sugarcane biorefinery.

METHODOLOGY

7.3.1 Basic inputs and scenario description

The techno-economic assessment considered performance data from small-


medium sized Brazilian sugarcane biorefineries, based on a milling capacity of 1.2
million tons of sugarcane in a 200-day harvesting period. Vinasse flow rates from
annexed plant producing sugar from juice and ethanol from both the juice and molasses,
biorefineries were used as a reference in the estimates for biogas and electricity or
biomethane production. Performance data for the AcD, i.e., organic matter removal levels
and methane production yields, were obtained from experimental results developed in
chapter 6 considering the better condition for each of both single and two-stage
mesophilic systems and are showed in Table 7.2. The proposed model was developed in
Aspen Plus, for analyze the feasibility HT pretreatment stage and Excel software through
structured simulations in order to analyze biogas stage (Biogas tool (Danish Energy
agency e Tech, 2019)). The model was based on a simulated flowchart to evaluate two
scenarios based on quantity of bagasse use against the base scenario, all described as
follow: a) 2G + AcD: three different fractions of bagasse (20,50 and 80%) from 1G
ethanol production are used for 2G ethanol production and taking account the energy
from biogas generated through AcD of vinasse and HH. In this scenario, 100% of
digestate is disposed to fertigation (
110
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

). The main input data regarding the sugarcane biorefinery are presented in Table
7.1 b) 2G+EnAcD : three different fractions of bagasse (20,50 and 80%) from 1G ethanol
production are used for 2G ethanol production, taking account an increase of 10% in
methane yield, obtaining the energy from biogas generated through AcD of vinasse and
HH. In this scenario, 100% of digestate is disposed to fertigation (
).
For the scenario including 2G ethanol production, the distilleries were considered
integrated to 1G ethanol production. The configuration of 2G process was based on
hydrothermal pretreatment in accordance with conditions defined by (Baêta, Lima, Filho,
et al., 2016) and showed in Table 7.1.

Figure 7.1 Flowchart of 2G+AcD scenario

111
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Table 7.1 - Data of 1G2G sugarcane ethanol production used in the simulations,
considering annexed distilleries

Sugarcane Input data Values Reference


biorefinery
Milling capacity (million-ton of cane per season) 1.2 [1]
Harvest period (d) 200 [3]
1G ethanol Bagasse dry base (kg per ton of cane) 140 [3]
production Ethanol Yield (L of EtOH per ton of cane) 70.6 [1]
process Vinasse production (L of vinasse per L of EtOH) 12 [1]
Vinasse flow rate (m3.d-1) 5176.6 [1]
Energy produced and exported (MW) 51705.01 [1]
COD vinasse (g.L-1) 18.07 Measured
Hydrothermal pretreatment
Temperature (°C) 183 [2]
2G ethanol Liquid-to-solid ratio (L.kgSB-1) 3.94 [2]
production Reaction time (h) 0.72 [2]
process Weight loss (%) 29.50 [2]
Cellulose content in pretreated bagasse (%) 59.25 [2]
Enzymatic hydrolysis efficiency (%) 62.18 [2]
Fermentation yield (g of ethanol per g of glucose) 0.42 [2]
COD Hemicellulose Hydrolysate (g.L-1) 71.17 Measured
[1]:(Agropéu 2017, 2019), [2]: (Baêta, Lima, Filho, et al., 2016), [3]:(Bonomi et al., 2016)

Table 7.2 – Anaerobic Co-digestion system parameters for simulation

Anaerobic co- Reactor Parameter Value*


digestion system configuration
Two-stage ASTBR OLR (kg COD.m−3.d-1) 5.65
HRT (d) 0.25-0.33
COD removal (%) 25-28
Methane yield (Nm3 kgCODr-1) 0.025
pH (effluent) 5.5
UASB OLR (kg COD.m−3.d-1) 2.21
HRT (d) 0.83
COD removal (%) 80-87
Methane yield (Nm3 kgCODr-1) 0.245
pH (effluent) 8.1
Single stage HAR OLR (kg COD.m−3.d-1) 2.38
HRT (d) 0.78
COD removal (%) 58.3-62.1
Methane yield (Nm3 kgCODr-1) 0.205
pH (effluent) 8.3
*all values were obtained in this study
112
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

7.3.2 Estimates and calculations

2S
Basic calculations included the determination of the Energy from two-stage ( ECH 4

1S
) and single stage anaerobic system ( ECH 4
) in MJ kg SB-1 using eq. (11) and (12), in which

the terms YM , MFR, CODMix , ERASTBR ,ERUASB, ERHAR, LHVCH4 and BFR are the
methane yield (Nm3 kg COD-1), mixture flow rate (m3 day-1), COD concentration of
mixture for codigestion (kg COD m-3 ),efficiency COD removal for ASTBR reactor (%),
efficiency COD removal for UASB reactor, efficiency COD removal for HAR, low
heating value of methane (MJ Nm-3 CH4 ) and bagasse flowrate (kg SB day-1)
respectively. The energy used during hydrothermal treatment in MJ kg SB-1 is calculated
by eq (13). The inlet temperatures of raw SB and water (from the process) were assumed
as 25°C and 95°C, respectively. The values of specific heat capacity (Cp,SB) assumed to
SB and water were 1.76 × 10-3 MJ (kg°C°)-1 and 4.19 × 10-3 MJ (kg°C°)-1, respectively
(Baêta et al., 2016a).

MJ YM  MFR  CODMix  (1 − ERASTBR )  ERUASB  LHVCH 4


2S
ECH ( )=
4
kg SB BFR (11)

MJ YM  MFR  CODMix  ERHAR  LHVCH 4


1S
ECH ( )= (12)
4
kg SB BFR
MJ
EHT ( ) = [Cp,SB  (183 − 25) + Cp,H2O  LSR  H2O  (183 − 95)] (13)
kg SB

7.3.3 Economic assessment methodology

The economic assessment was performed considering only the effects of the
construction and operation of the anaerobic digestion plant as well as CHP system for all
process configurations described previously (section 7.3.1). In other words, the feasibility
of adding the AD sector to an already existing process unit was assessed. Hence, it was
considered that the biorefineries already existed and were operating normally prior to the

113
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

installation of the AD unit, which is considered as a process improvement –actually, that


would be the case in most practical situations in the sector.
Capital costs (CAPEX) of the anaerobic digestion plant, corresponding to UASB
reactors, CHP system and Heat exchanger were estimated based on correlations described
in BiogasTool (Danish Energy agency e Tech, 2019). The costs for the anaerobic reactors
and CHP units included the cost of equipment, project and fees associated with the
investment cost. The operating cost (OPEX) was based on Fuess e Zaiat, (2018).
Electricity and biogas market prices were set as 88.39 USD MWh-1 (EPE,2019) and 0.42
USD Nm−3 (Bonomi et al., 2016), respectively. The final cost for the installed plant
included equipment and installation, civil construction, electricity installation,
instrumentation, land, and engineering. Results were reported in terms of production
costs, the internal rate of return (IRR) for achieving net present value NPV = 0, and NPV
assuming IRR = 12%. Thus, the minimum acceptable rate of return (MARR) was
assumed as 12%/year. project lifetime was 20 years, depreciation (linear)of 10 years.

RESULTS AND DISCUSSION

7.4.1 Energy Assessment

It is important to highlight that AcD, when implemented within the concept of


sugarcane biorefineries, has two main objectives: i) to improve the energy balance of the
process by using the biogas energy to pay for the pretreatment step; ii) reduce the
environmental impacts associated with the emission of greenhouse gases and with land
use in relation to ethanol yield.
The energetic potentials of methane obtained through the two tested
configurations (single and two-stage) and based in three different proportions of bagasse
used for 1G-2G ethanol production are presented in Figure 7.2. As can be seen, when de
YM are fixed for each anaerobic system (0.205 Nm3 CH4 per g CODr for single stage and
0.245 Nm3 CH4 per g CODr for two-stage systems) the energy produced increase until the
amount of energy spent on pretreatment is almost reached. In the case of using 50% of
the bagasse for 2G ethanol production, the two-stage system cover approximately 94.4%
of the energy spent in the hydrothermal pretreatment. In this case there would be an excess
114
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

of energy from burning the other half of bagasse, which is enough to fulfill the energy
requirements of 1G production process. In addition, this condition allows the production
of 11.2 L of ethanol per ton of cane processed.
These results show that implementing 1G-2G the integration of ethanol production
process would increase its yield by 15.7 %. In this sense, the results presented here are in
accordance with that reported by Rodriguez, Manochio e Moraes, (2018) who found an
increment of 15% on 2G ethanol production when AD is used. On the other hand,
Macrelli, Mogensen e Zacchi, (2012) achieved 16 % increase of ethanol productivity
when combining 2G and 1G ethanol production without mixing the material streams and
considering the cogeneration of solid residues from 2G ethanol production (lignin and
non-hydrolyzed cellulose) as well as biogas from AD of pentose liquor.

Figure 7.2 Energy production for three scenarios of surplus bagasse usage for 2G ethanol
production and biogas production at single- or two-stage anaerobic systems. E BB –
energy obtained by bagasse burning; E HT – energy spent on hydrothermal pretreatment;
E 1Stage – energy obtained as biogas in the HAR reactor; E 2Stage – energy obtained as
biogas in the ASTBR-UASB system.

A conservative increase of 10% in the methane yield (single stage: 0.225 Nm3 CH4
g CODr -1 and two-stage: 0.297 Nm3 CH4 g CODr -1) was simulated for the three scenarios
of surplus bagasse use (20, 50 and 80 %) in order to explore its effect on the energy
balance. Increasing methane yield could be done by, for instance, optimizing reactor

115
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

configuration and that would be feasible in larger scale reactors devised to AcD the
sugarcane biorefinery byproducts. The results of such simulation are shown in Figure 7.3,
and in all cases it was possible to observe that the energy produced by the AcD in the
two-stage system was enough to pay for the energy spent in the pre-treatment step, with
an average surplus of 5.6% for all three case scenarios. Thus, this confirms that anaerobic
systems with sequential acidogenic-methanogenic stages enhances the energy balance in
the context of sugarcane biorefinery (Fuess, L T et al., 2018) since they lead to better
performance in terms of methane production.
It is important to mention that these results were found considering a conservative
increase of 10% on the methane yield, due to the fact that 1G-2G integrated plants,
sugarcane bagasse are subjected to harsher processing steps to depolymerize the structural
polysaccharides. These processes result in side reaction products that could potentially be
inhibitory to microbial growth due to the presence of lignin and lignin related compounds
such as phenolic compounds (Longati et al., 2019), which are inhibitory for anaerobic
digestion, as shown in chapter 6.

Figure 7.3 Energy production for three scenarios of surplus bagasse usage for 2G ethanol
production and biogas production at single- or two-stage anaerobic systems considering
an increase of 10% in methane yield. E BB – energy obtained by bagasse burning; E HT
– energy spent on hydrothermal pretreatment; E 1Stage – energy obtained as biogas in
the HAR reactor; E 2Stage – energy obtained as biogas in the ASTBR-UASB system.

116
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

7.4.2 Economic Assessment

The economic performance of two-stage AcD of sugarcane vinasse and HH when


fixing the methane yield is presented Table 7.3, which shows estimate values of NPV,
IRR, ROI and payback period for each scenario of bagasse usage tested. Considering a
given type of biorefinery, i.e., annexed, it can be observed that all three scenarios are
economically viable options that would guarantee the profitability of AcD units. Although
OPEX + CAPEX associated with the pretreatment unit, calculated as USD 36.31 Million,
was comparable with the value (USD 52.1 Million) reported by Albarelli, Ensinas e
Silva,(2014), this biorefinery unit was not considered within economical analysis due to
the high costs.
Table 7.3 shows that in the case scenario of using 80% of surplus bagasse for 1G-
2G ethanol production the payback time would be 5.57 years with high NPV, IRR and
ROI values. It is important to note that these results are the best option if the sugarcane
biorefinery has an efficient unit of power generation since it would be possible to supply
the energy needed to process 1 ton of sugarcane by burning just 20% of the spent bagasse
(Haandel, Van e Lier, Van, 2015). Likewise, burning the cellu-lignin fraction generated
during 2G ethanol process, would also contribute to enhance economical indices and
feasibility of thermal pretreatment. This has been proposed by some authors (Longati et
al., 2019; Rodriguez, Manochio e Moraes, 2018), hence confirming that this option is one
of the most feasible opportunities to integrate 1G-2G ethanol production.
The results related to economic indicators (NPV = USD 5.59 million, IRR=21.8%
and 10.55 years of payback time) obtained here when considering the use of only 50 %
of bagasse are in good agreement with those reported by Fuess e Zaiat, (2018) who
considered a two-stage anaerobic system for an annexed distillery with vinasse as unique
substrate and found the following values: NPV = USD 6 million, IRR=12% and 13 years
of payback time.

117
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Table 7.3 – Economical analysis for each scenario of surplus bagasse used at an integrated
1G-2G plant and AcD with a fixed methane yield of 0.245 Nm3 kg CODr -1

Economic indicators % of bagasse used at 1G-2G processes


20 % BC 50 % BC 80 % BC
NPV (millions of USD) 1.49 5.59 11.22
IRR (%) 17. 61 21.88 26.16
ROI (%) 32.98 59.50 86.86
Payback time (years) 9.64 10.55 5.57

Figure 7.4 presents the incremental cash flow analysis for the scenarios including
the AcD unit in the sugarcane mills when operating using 20%, 50% and 80% of bagasse
surplus. The improvements foreseen for the technologies of 2G ethanol production and
efficient power systems had an important impact on the economic viability of the assessed
scenarios and reveals that their current operational costs make the 1G-2G sugarcane mill
with AcD unit an unfeasible business. The results evidence the importance of investing
in research and development, especially in 2G ethanol production but also in AcD of
vinasse or biorefinery byproducts, as corroborated by (Rodriguez, Manochio e Moraes,
2018).
In order to verify the hypothesis related to more efficient anaerobic systems, a
conservative increase of 10% in in the values of methane yield reported in chapter 6 for
the two-stage configuration was simulated. The economical indices and incremental cash
flow analysis for this case scenario are shown in
Table 7.4 and Figure 7.5, respectively. The results confirm that with this relatively
low increment on the methane yield it was possible to get NPV= USD 9.04 million,
IRR=26.5% and 5.36 years of payback time when just 50% of surplus bagasse is used for
2G ethanol process.

118
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 7.4 Incremental cash flow analysis for the scenarios of 2G ethanol production
integrated to biogas production from vinasse and HH AcD according to bagasse use: a)
20% of bagasse for 2G ethanol), b) 50% of bagasse for 2G ethanol c) 80% of bagasse for
2G ethanol

Table 7.4 – Economical analysis for each scenario of surplus bagasse use at an integrated
1G-2G plant and AcD with a fixed methane yield of 0.301 Nm3 kg CODr -1
Economic indicators % of bagasse used in 1G2G process
20 % BC 50 % BC 80 % BC
NPV (Million $US) 2.76 9.04 17.04
IRR (%) 21.2 26.5 32.6
ROI (%) 56.81 89.05 123.3
Payback time (years) 7.05 5.36 4.21

119
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Figure 7.5 Incremental cash flow analysis for the scenarios of 2G ethanol production
integrated to biogas production from vinasse and HH AcD according to bagasse use and
considering a methane yield increase of 10%: a) 20% of bagasse for 2G ethanol), b) 50%
of bagasse for 2G ethanol c) 80% of bagasse for 2G ethanol

Finally, a direct comparison of the obtained results with the literature is not
straightforward since there is a large variability in the premises adopted for technical
solutions in each study (Longati et al., 2019) as well as lack of experimental data for AcD
under the conditions tested in this thesis. However, it is important to mention that the
results obtained here, with their own experimental base, are a good approximation that
could be used as starting point to continue exploring the technological diversity of
anaerobic digestion to achieve the integration of sugarcane biorefinery from a technical
and economic point of view.

120
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

8 CONCLUSIONS

Experimental data and results discussed in Chapter 5 allowed to confirm the


complementary hypothesis 1 and 3, as well as to address the specific objectives 1 and 2,
since the mixture of sugarcane biorefinery substrates and/or wastes (hemicelluloses
hydrolysate (HH), vinasse, yeast extract (YE) and fly ash (SBFA)) improved the
anaerobic co-digestion (AcD) and increased methane production, providing an energy
surplus for the process. These results were due to the combined effect of C/N
enhancement, nutrients supplementation, alkalinity and trace metals addition, as well as
the role of SBFA as adsorbent of toxic compounds, mainly FF and HMF. AcD of 25–
75% HH-to-vinasse at 0.375 F/M, 1 g L−1 YE and 15 g L−1 SBFA led to a biodegradability
of 79%; whereas the use of HH with 45 g L−1 SBFA and 1.5 g L−1 YE resulted in
biodegradability of 76%. It is important to highlight that the amounts of vinasse, HH, YE
and SBFA used in the experiments were set to be realistic from a technological point of
view, i.e. in accordance with actual capacities and characteristics of sugarcane mills.
The results presented in chapter 6 allow confirm the complementary hypothesis 2
as well as the specific objective 4, as mesophilic reactors fed continuously with
hemicelluloses hydrolysate and vinasse and assembled in two configurations (single or
two-stage systems) were compared. The single stage hybrid anaerobic reactor (HAR)
performed worst than the two-stage acidogenic-methanogenic system formed by an
acidogenic structured-bed reactor (ASTBR) followed by an UASB methanogenic reactor.
It has been shown that the use of ASTBR-UASB to treat the HH-vinasse mixture led to a
global COD removal higher than 80% and methane yield of 0.245 NL CH4 g CODr-1.
Microbial community analyses of sludge collected at different operational stages
confirmed there was a physical separation in the microbiota of ASTBR and UASB
reactors. In ASTBR the main species identified were Clostridium beijerinckii (61.9%)
and Desulfovibrio desulfuricans (19.1%) during the lowest ORL (3.75 g COD L-1 d-1 )
tested; and Lactobacillus casei (66.3 %) and Lactococcus lactis (22.4 %) when the OLR
was increased to 10.81 g COD L-1 d-1 Y. Regarding the UASB reactor, the main identified
species were Longilinea arvoryzae (15.5%), Methanosaeta concilii (19.9 %) and
Syntrophus aciditrophicus (15.4%) at the lowest OLR; and Longilinea arvoryzae
121
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

(31.7%), Methanosaeta concilii (10.6 %) and Pleomorphomonas oryzae (10.1%) at the


higher OLR. In its turn, the main species identified in the HAR sludge were Bacteroides
graminisolvens (11.3%), Desulfovibrio desulfuricans (12.6 %) and Methanosaeta concilii
(16.4%.) at 1.38 g COD L-1 d-1.
Finally, the results presented in Chapter 7 aimed to verify the economic and
technical feasibility of anaerobic digestion processes investigated, as well as to test this
thesis central hypothesis and to address the specific objective 5. The economic indicators
showed that the integration of 1G-2G sugarcane mill including an AD unit is not viable
when considering the current technologies for 2G ethanol production due to the high
capital and operational costs involved. By fixing the experimental methane yield at 0.245
Nm3 kg CODr-1, which was obtained for the two-stage system, and considering a 50% of
bagasse surplus, it was possible to achieve values of internal rate of return (IRR), return
on investment (ROI), and payback period of 21.8%, 59.50% and 10.55 years respectively.
The use of biogas energy allows, from an energetic point of view, to pay for the
sugarcane pretreatment, but it was not possible to support the CAPEX and OPEX related
to the hydrothermal pretreatment. These results confirm that sugarcane biorefinery needs
different products with high added values in order to guarantee the integration of 1G-2G
ethanol production.

122
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

9 FUTURE WORKS

Technical and economic feasibility of the AcD processes of biogas production


have been investigated, especially within the last five years indicating that there is a
potential to further increase the biogas production using co-digestion of different
byproducts from sugarcane biorefinery. However, there are different kinds of challenges
during the development of the AcD process at both pilot scale and industry level which
could be addressed in further research:
1) how seasonal changes of feedstocks affect its biodegradability and methane
potential?
2) designing new anaerobic reactors, aimed at optimizing and guaranteeing short
HRT and adequate CRT, would result in higher methane yields during the AcD of
sugarcane biorefinery by-products/wastes?
3) can robust mathematical models be devised to integrate 1G-2G ethanol
production and allowing to insert common design factors between AcD, use of lignin
fractions derived from the process, as well as pretreatments conditions with low fixed and
operational cost?

123
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

10 REFERENCES

ABREU, C. A.; CANTONI, M.; COSCIONE, A. R.; PAZ-FERREIRO, J. Organic Matter


and Barium Absorption by Plant Species Grown in an Area Polluted with Scrap Metal Residue.
Applied and Environmental Soil Science, v. 2012, p. 476821, 2012.

ACHARYA, B. K.; MOHANA, S.; MADAMWAR, D. Anaerobic treatment of distillery


spent wash - A study on upflow anaerobic fixed film bioreactor. Bioresource Technology, v. 99,
n. 11, p. 4621–4626, 2008.

AGROPÉU 2017. Agropéu Agro Industrial Pompéu S/APompeu,MG, 2019.


Disponível em: <http://www.agropeu.com.br/site/>. Acesso em: 15 nov. 2019

AKRAM, A.; STUCKEY, D. C. Flux and performance improvement in a submerged


anaerobic membrane bioreactor (SAMBR) using powdered activated carbon (PAC). Process
Biochemistry, v. 43, n. 1, p. 93–102, 2008.

ALBANEZ, R.; LOVATO, G.; RATUSZNEI, S. M.; ZAIAT, M.; RODRIGUES, J. A.


D. Feasibility of biohydrogen production by Co-digestion of vinasse (Sugarcane stillage) and
Molasses in an Ansbbr. Brazilian Journal of Chemical Engineering, v. 35, n. 1, p. 27–41, 2018.

ALBARELLI, J. Q. Produção de açúcar e etanol de primeira e segunda geração:


simulação, integração energética e análise econômica. [s.l.] Universidade Estadual de
Campinas-UNICAMP, 2013.

ALBARELLI, J. Q.; ENSINAS, A. V.; SILVA, M. A. A new proposal of cellulosic


ethanol to boost sugarcane biorefineries: Techno-economic evaluation. International Journal of
Chemical Engineering, v. 2014, 2014.

APHA. Standard Methods for the Examination of Water and Wastewater. [s.l.]
American Public Health Association, 2005.

AQUINO, S.; CHERNICHARO, C. Acúmulo de ácidos graxos voláteis (AGVs) em


reatores anaeróbios sob estresse: causas e estratégias de controle. Engenharia Sanitária e
Ambiental, v. 10, n. 2, p. 152–161, 2005.

124
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

AQUINO, S. F.; ARAÚJO, J. C.; PASSOS, F.; CURTIS, T. P.; FORESTI, E.


Fundamentals of anaerobic sewage treatment. In: CHERNICHARO, C. A. DE L.; BRESSANI-
RIBEIRO, T. (Eds.). . Anaerobic Reactors for Sewage Treatment: Design, Construction and
Operation. 1st. ed. [s.l.] IWA, 2019. p. 25–29.

AYODELE, B. V.; ALSAFFAR, M. A.; MUSTAPA, S. I. An overview of integration


opportunities for sustainable bioethanol production from first- and second-generation sugar-based
feedstocks. Journal of Cleaner Production, v. 245, p. 118857, 2020.

BAÊTA, B. E. L. Aproveitamento energético a partir da digestão anaeróbia de


hidrolisado hemicelulósico gerado pelo pré-tratamento por auto-hidrólise do bagaço de
cana-de-açúcar considerando o contexto da biorrefinaria. [s.l.] Universidade Federal de Ouro
Preto-UFOP, 2016.

BAÊTA, B. E. L.; AQUINO, S. F.; SILVA, S. Q.; RABELO, C. A. Anaerobic


degradation of azo dye Drimaren blue HFRL in UASB reactor in the presence of yeast extract a
source of carbon and redox mediator. Biodegradation, v. 23, n. 2, p. 199–208, 2012.

BAÊTA, B. E. L.; LIMA, D. R. S.; ADARME, O. F. H.; GURGEL, L. V. A.; AQUINO,


S. F. Optimization of sugarcane bagasse autohydrolysis for methane production from
hemicellulose hydrolyzates in a biorefinery concept. Bioresource Technology, v. 200, p. 137–
146, 2016.

BAÊTA, B. E. L.; LIMA, D. R. S.; FILHO, J. G. B.; ADARME, O. F. H.; GURGEL, L.


V. A.; AQUINO, S. F. Evaluation of hydrogen and methane production from sugarcane bagasse
hemicellulose hydrolysates by two-stage anaerobic digestion process. Bioresource Technology,
v. 218, p. 436–446, 2016.

BAÊTA, B. E. L.; LUNA, H. J.; SANSON, A. L.; SILVA, S. Q.; AQUINO, S. F.


Degradation of a model azo dye in submerged anaerobic membrane bioreactor (SAMBR)
operated with powdered activated carbon (PAC). Journal of Environmental Management, v.
128, p. 462–470, 2013.

BAGUDO, B. U.; DANGOGGO, S. M.; HASSAN, L. G.; GARBA, B. Influence of


catalyst ( Yeast ) on the Biomethanization of Selected Organic Waste Materials Department of
Pure and Applied Chemistry Usmanu Danfodiyo University Sokoto . Energy Commission of

125
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Nigeria Abuja . ES products + E. v. 18, p. 209–216, 2010.

BAHIA, M.; PASSOS, F.; ADARME, O. F. H.; AQUINO, S. F.; SILVA, S. Q.


Anaerobic-Aerobic Combined System for the Biological Treatment of Azo Dye Solution using
Residual Yeast. Water Environment Research, v. 90, n. 8, p. 729–737, 2018.

BALAKRISHNAN, M.; BATRA, V. S. Valorization of solid waste in sugar factories


with possible applications in India : A review. Journal of Environmental Management, v. 92,
n. 11, p. 2886–2891, 2011.

BARAKAT, A.; MONLAU, F.; STEYER, J. P.; CARRERE, H. Effect of lignin-derived


and furan compounds found in lignocellulosic hydrolysates on biomethane production.
Bioresource Technology, v. 104, p. 90–99, 2012.

BARBIERI, L.; ANDREOLA, F.; LANCELLOTTI, I.; TAURINO, R. Management of


agricultural biomass wastes: Preliminary study on characterization and valorisation in clay matrix
bricks. Waste Management, v. 33, n. 11, p. 2307–2315, 2013.

BARROS, V. G. D.; DUDA, R. M.; VANTINI, J. D. S.; OMORI, W. P.; FERRO, M. I.


T.; OLIVEIRA, R. A. D. Improved methane production from sugarcane vinasse with filter cake
in thermophilic UASB reactors, with predominance of Methanothermobacter and Methanosarcina
archaea and Thermotogae bacteria. Bioresource Technology, v. 244, p. 371–381, 2017.

BARUAH, J.; NATH, B. K.; SHARMA, R.; KUMAR, S.; DEKA, R. C.; BARUAH, D.
C.; KALITA, E. Recent Trends in the Pretreatment of Lignocellulosic Biomass for Value-Added
Products. Frontiers in Energy Research, v. 6, n. December, p. 1–19, 2018.

BERGMANN, J. C.; TRICHEZ, D.; SALLET, L. P.; PAULA E SILVA, F. C. DE;


ALMEIDA, J. R. M. Technological Advancements in 1G Ethanol Production and Recovery of
By-Products Based on the Biorefinery Concept. In: Advances in Sugarcane Biorefinery. [s.l:
s.n.]. p. 73–95.

BONCZ, M. A.; FORMAGINI, E. L.; SANTOS, L. D. S.; MARQUES, R. D.; PAULO,


P. L. Application of urea dosing for alkalinity supply during anaerobic digestion of vinasse.
Water Science and Technology, v. 66, n. 11, p. 2453–2460, 2012.

BONOMI, A.; CAVALETT, O.; CUNHA, M. P. DA; LIMA, M. A. Virtual Biorefinery

126
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

An Optimization Strategy for Renewable Carbon Valorization. 1st. ed. [s.l.] Springer, 2016.

BOOPATHY, R.; BOKANG, H.; DANIELS, L. Biotransformation of furfural and 5-


hydroxymethyl furfural by enteric bacteria. Journal of Industrial Microbiology, v. 11, n. 3, p.
147–150, 1993.

CABRERA-DÍAZ, A.; PEREDA-REYES, I.; DUEÑAS-MORENO, J.; VÉLIZ-


LORENZO, E.; DÍAZ-MARRERO, M. A.; MENÉNDEZ-GUTIÉRREZ, C. L.; OLIVA-
MERENCIO, D.; ZAIAT, M. Combined treatment of vinasse by an upflow anaerobic filter-
reactor and ozonation process. Brazilian Journal of Chemical Engineering, v. 33, n. 4, p. 753–
762, 2016.

CABRERA-DÍAZ, A; PEREDA-REYES, I.; DUEÑAS-MORENO, J.; VÉLIZ-


LORENZO, E.; DÍAZ-MARRERO, M. A.; MENÉNDEZ-GUTIÉRREZ, C. L.; OLIVA-
MERENCIO, D.; ZAIAT, M. Combined treatment of vinasse by an upflow anaerobic filter-
reactor and ozonation process. Brazilian Journal of Chemical Engineering, v. 33, n. 4, p. 753–
762, 2016.

CABRERA-DÍAZ, A.; PEREDA-REYES, I.; OLIVA-MERENCIO, D.; LEBRERO, R.;


ZAIAT, M. Anaerobic Digestion of Sugarcane Vinasse Through a Methanogenic UASB Reactor
Followed by a Packed Bed Reactor. Applied Biochemistry and Biotechnology, v. 183, n. 4, p.
1127–1145, 2017.

CACURO, T A; WALDMAN, W. R. Fly-ash from biomass burning: Applications and


potentialities . Revista Virtual de Quimica, v. 7, n. 6, p. 2154–2165, 2015.

CACURO, THIAGO A.; WALDMAN, W. R. Cinzas da queima de biomassa: aplicações


e potencialidades. Revista Virtual de Quimica, v. 7, n. 6, p. 2154–2165, 2015.

CADETE, R. M.; MELO-CHEAB, M. A.; DUSSÁN, K. J.; RODRIGUES, R. C. L. B.;


SILVA, S. S. DA; GOMES, F. C. O.; ROSA, C. A. Production of bioethanol in sugarcane bagasse
hemicellulosic hydrolysate by Scheffersomyces parashehatae, Scheffersomyces illinoinensis and
Spathaspora arborariae isolated from Brazilian ecosystems. Journal of Applied Microbiology,
v. 123, n. 5, p. 1203–1213, 2017.

CAILLET, H.; LEBON, E.; AKINLABI, E.; MADYIRA, D.; ADELARD, L. Influence
of inoculum to substrate ratio on methane production in biochemical methane potential (BMP)
127
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

tests of sugarcane distillery waste water. Procedia Manufacturing, p. 259–264, 2016.

CAMMAROTA, M. C.; CAMPORESE, S. E. F.; DA, C. G. A.; DE, C. A. L.; GROPOSO,


S. C. J.; MACHADO, D. C. A.; MARTINS, B. L. F.; MELO, S. A. L. M.; NOGUEIRA, M. D.
Method for producing energy-rich gases from lignocellulosic material streamsGoogle
Patents, , 2012. Disponível em: <http://www.google.com/patents/WO2012003556A1?cl=en>

CANO, R.; PÉREZ-ELVIRA, S. I.; FDZ.-POLANCO, F. Energy feasibility study of


sludge pretreatments : A review. Applied Energy, v. 149, p. 176–185, 2015.

CESARO, A.; BELGIORNO, V. Combined Biogas and Bioethanol Production:


Opportunities and Challenges for Industrial Application. Energies, v. 8, n. 8, p. 8121–8144, 2015.

CETECIOGLU, Z.; DOLFING, J.; TAYLOR, J.; PURDY, K. J.; EYICE, Ö. COD/sulfate
ratio does not affect the methane yield and microbial diversity in anaerobic digesters. Water
Research, v. 155, p. 444–454, 2019.

CHERNICHARO, C. A. DE L. Anaerobic Reactors. 1st. ed. London: IWA, 2007. v. 04

CHINGONO, K. E.; SANGANYADO, E.; BERE, E.; YALALA, B. Adsorption of


sugarcane vinasse effluent on bagasse fly ash: A parametric and kinetic study. Journal of
Environmental Management, v. 224, n. July, p. 182–190, 2018.

CHRISTOFF, A. P.; FERNANDA, A.; SEREIA, R.; RODRIGUES, D.; LUCIO, R.;
MORAES, V. DE; FELIPE, L.; OLIVEIRA, V. DE. Bacterial identification through accurate
library preparation and high-throughput sequencing. Neoprospecta Microbiome Technologies,
n. May 2017, p. 1–5, 2017.

CHRISTOFOLETTI, C. A.; ESCHER, J. P.; CORREIA, J. E.; MARINHO, J. F. U.;


FONTANETTI, C. S. Sugarcane vinasse: Environmental implications of its use. Waste
Management, v. 33, n. 12, p. 2752–2761, 2013.

COLTURATO, L. F. DE D. B. Dessulfuração de biogás da metanização da vinhaça:


uma nova abordagem para remoção de altas concentrações de H2S. [s.l.] Universidade
Federal de Minas Gerais, 2015.

COMPANHIA NACIONAL DE ABASTECIMENTO (CONAB). Acompanhamento


da safra brasileira de cana-de-açúcar V.4-Safra 2017/18 N.4. Brasília: [s.n.].
128
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

___. Acompanhamento da safra brasileira de cana-de-açúcar V.6-Safra 2019/20 N.3.


Brasilia: [s.n.].

CORTEZ, L. A. B.; BALDASSIN, R.; ALMEIDA, E. DE. Energy from sugarcane. In:
Sugarcane Biorefinery, Technology and Perspectives. [s.l: s.n.]. p. 117–139.

___. Energy from sugarcane. In: Sugarcane Biorefinery, Technology and Perspectives.
[s.l: s.n.]. p. 117–139.

CRITTENDEN, B.; THOMAS, W. J. Adsorption Technology & Design. First ed. Great
Britain: Butterworth-Heinemann, 1998.

DAI, X.; CHEN, S.; XUE, Y.; DAI, L.; LI, N.; TAKAHASHI, J.; ZHAO, W. Hygienic
treatment and energy recovery of dead animals by high solid co-digestion with vinasse under
mesophilic condition: feasibility study. Journal of Hazardous Materials, v. 297, p. 320–328,
2015.

DANISH ENERGY AGENCY; TECH, I. Biogas Tool: Calculating costs and benefits
of biogas production in Mexico 2018-2019. Mexico: [s.n.].

DELGADO, F.; EVANGELISTA, M.; ROITMAN, T.; QUINTELLA DE


VASCONCELLOS, C. O. BIOCOMBUSTÍVEIS. Rio de Janeiro, Brasil: [s.n.].

DETMAN, A.; MIELECKI, D.; PLEŚNIAK, Ł.; BUCHA, M.; JANIGA, M.;
MATYASIK, I.; CHOJNACKA, A.; JȨDRYSEK, M. O.; BŁASZCZYK, M. K.; SIKORA, A.
Methane-yielding microbial communities processing lactate-rich substrates: A piece of the
anaerobic digestion puzzle. Biotechnology for Biofuels, v. 11, n. 1, p. 1–18, 2018.

DIAS, M. O. S.; JUNQUEIRA, T. L.; CAVALETT, O.; PAVANELLO, L. G.; CUNHA,


M. P.; JESUS, C. D. F.; MACIEL FILHO, R.; BONOMI, A. Biorefineries for the production of
first and second generation ethanol and electricity from sugarcane. Applied Energy, v. 109, p.
72–78, 2013.

EBADI, A.; MOHAMMADZADEH, J. S. S.; KHUDIEV, A. What is the correct form of


BET isotherm for modeling liquid phase adsorption ? Adsorption, n. 15, p. 65–73, 2009.

EGGERT, H.; GREAKER, M. Promoting Second Generation Biofuels: Does the First
Generation Pave the Road? Energies, v. 7, n. 7, p. 4430–4445, 2014.
129
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

EPA, U. S. E. P. A. Selected Analytical Methods for Environmental Remediation and


Recovery (SAM) 2017. Cincinnati,OH: [s.n.].

ESKICIOGLU, C.; GHORBANI, M. Effect of inoculum/substrate ratio on mesophilic


anaerobic digestion of bioethanol plant whole stillage in batch mode. Process Biochemistry, v.
46, n. 8, p. 1682–1687, 2011.

ESPAÑA-GAMBOA, E.; MIJANGOS-CORTES, J.; BARAHONA-PEREZ, L.;


DOMINGUEZ-MALDONADO, J.; HERNANDEZ-ZARATE, G.; ALZATE-GAVIRIA, L.;
HERNÁNDEZ-ZARATE, G.; ALZATE-GAVIRIA, L. Vinasses: Characterization and
treatments. Waste Management and Research, v. 29, n. 12, p. 1235–1250, 2011.

FAGBOHUNGBE, M. O.; HERBERT, B. M. J.; HURST, L.; IBETO, C. N.; LI, H.;
USMANI, S. Q.; SEMPLE, K. T. The challenges of anaerobic digestion and the role of biochar
in optimizing anaerobic digestion. Waste Management, v. 61, p. 236–249, 2017.

FARIA, C. V.; SOUZA, D. F.; PONTES, T. M.; AMARAL, M. C. S.; FONSECA, F. V.


Strategies of anaerobic sludge granulation in an EGSB reactor. Journal of Environmental
Management, v. 244, n. May, p. 69–76, 2019.

FERRAZ JÚNIOR, A. D.; ETCHEBEHERE, C.; ZAIAT, M. High organic loading rate
on thermophilic hydrogen production and metagenomic study at an anaerobic packed-bed reactor
treating a residual liquid stream of a Brazilian biorefinery. Bioresource Technology, v. 186, p.
81–88, 2015.

FERREIRA, P. P. L.; BRAGA, R. M.; TEODORO, N. M. A.; MELO, V. R. M.; MELO,


D. M. A.; MELO, M. A. F. Adsorption of Cu2+ and Cr3+ in waste water using bagasse fly ash |
Adsorção de Cu2+ e Cr3+ em efluentes líquidos utilizando a cinza do bagaço da cana-de-açúcar.
Ceramica, v. 61, n. 360, p. 435–441, 2015.

FITO, J.; TEFERA, N.; HULLE, S. W. H. VAN. Adsorption of distillery spent wash on
activated bagasse fly ash: Kinetics and thermodynamics. Journal of Environmental Chemical
Engineering, v. 5, n. 6, p. 5381–5388, 2017.

FONSECA, B. C.; RIAÑO-PACHÓN, D. M.; GUAZZARONI, M. E.; REGINATTO, V.


Genome sequence of the H2-producing clostridium beijerinckii strain Br21 isolated from a
sugarcane vinasse treatment plant. Genetics and Molecular Biology, v. 42, n. 1, p. 139–144,
130
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

2019.

FOO, K. Y.; HAMEED, B. H. Insights into the modeling of adsorption isotherm systems.
Chemical Engineering Journal, v. 156, p. 2–10, 2010.

FUESS, L.T.; ARAUJO JR, M. M.; GARCIA, M. L.; ZAIAT, M. Designing full-scale
biodigestion plants for the treatment of vinasse in sugarcane biorefineries: How phase separation
and alkalinization impact biogas and electricity production costs? Chemical Engineering
Research and Design, v. 119, p. 209–220, 2017.

FUESS, L T; ARAÚJO JR, M. M.; GARCIA, M. L.; ZAIAT, M. Designing full-scale


biodigestion plants for the treatment of vinasse in sugarcane biorefineries: How phase separation
and alkalinization impact biogas and electricity production costs? Chemical Engineering
Research and Design, v. 119, p. 209–220, 2017.

FUESS, LUCAS TADEU; FERRAZ, A. D. N.; MACHADO, C. B.; ZAIAT, M.


Temporal dynamics and metabolic correlation between lactate-producing and hydrogen-
producing bacteria in sugarcane vinasse dark fermentation: The key role of lactate. Bioresource
Technology, v. 247, p. 426–433, 2018.

FUESS, L. T.; GARCIA, M. L.; ZAIAT, M. Seasonal characterization of sugarcane


vinasse: Assessing environmental impacts from fertirrigation and the bioenergy recovery
potential through biodigestion. Science of the Total Environment, v. 634, p. 29–40, 2018.

FUESS, L.T.; KIYUNA, L. S. M.; FERRAZ, A. D. N.; PERSINOTI, G. F.; SQUINA, F.


M.; GARCIA, M. L.; ZAIAT, M. Thermophilic two-phase anaerobic digestion using an
innovative fixed-bed reactor for enhanced organic matter removal and bioenergy recovery from
sugarcane vinasse. Applied Energy, v. 189, p. 480–491, 2017.

FUESS, LUCAS TADEU; KLEIN, B. C.; CHAGAS, M. F.; ALVES FERREIRA


REZENDE, M. C.; GARCIA, M. L.; BONOMI, A.; ZAIAT, M. Diversifying the technological
strategies for recovering bioenergy from the two-phase anaerobic digestion of sugarcane vinasse:
An integrated techno-economic and environmental approach. Renewable Energy, v. 122, p. 674–
687, 2018.

FUESS, L T; KLEIN, B. C.; CHAGAS, M. F.; ALVES FERREIRA REZENDE, M. C.;


GARCIA, M. L.; BONOMI, A.; ZAIAT, M. Diversifying the technological strategies for
131
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

recovering bioenergy from the two-phase anaerobic digestion of sugarcane vinasse: An integrated
techno-economic and environmental approach. Renewable Energy, v. 122, p. 674–687, 2018.

FUESS, L. T.; ZAIAT, M. Economics of anaerobic digestion for processing sugarcane


vinasse: Applying sensitivity analysis to increase process profitability in diversified biogas
applications. Process Safety and Environmental Protection, v. 115, p. 27–37, 2018.

GARCÍA-DEPRAECT, O.; GÓMEZ-ROMERO, J.; LEÓN-BECERRIL, E.; LÓPEZ-


LÓPEZ, A. A novel biohydrogen production process: Co-digestion of vinasse and Nejayote as
complex raw substrates using a robust inoculum. International Journal of Hydrogen Energy,
v. 42, n. 9, p. 5820–5831, 2017.

GIL, A.; SILES, J. A.; MÁRQUEZ, P.; GUTIÉRREZ, M. C.; MARTÍN, M. A.


Optimizing the selection of organic waste for biomethanization. Environmental Technology
(United Kingdom), p. 1–12, 2017.

GODOI, L. A. G. DE; FUESS, L. T.; DELFORNO, T. P.; FORESTI, E.; DAMIANOVIC,


M. H. R. Z. Characterizing phenol-removing consortia under methanogenic and sulfate-reducing
conditions: potential metabolic pathways. Environmental Technology (United Kingdom), p.
1–11, 2018.

GODOI, L. A. G. DE; FUESS, L. T.; DELFORNO, T. P.; FORESTI, E.; DAMIANOVIC,


M. H. R. Z. Characterizing phenol-removing consortia under methanogenic and sulfate-reducing
conditions: potential metabolic pathways. Environmental Technology (United Kingdom), v.
40, n. 24, p. 3216–3226, 2019.

GONZALEZ-GIL, G.; JANSEN, S.; ZANDVOORT, M. H.; LEEUWEN, H. P. VAN.


Effect of Yeast Extract on Speciation and Bioavailability of Nickel and Cobalt in Anaerobic
Bioreactors. Biotechnol Bioeng, v. 82, p. 134–142, 2003.

GUNES, B.; STOKES, J.; DAVIS, P.; CONNOLLY, C.; LAWLER, J. Pre-treatments to
enhance biogas yield and quality from anaerobic digestion of whiskey distillery and brewery
wastes: A review. Renewable and Sustainable Energy Reviews, v. 113, n. February, p. 109281,
2019.

HAANDEL, A. VAN; KATO, M. T.; CAVALCANTI, P. F. F.; FLORENCIO, L.


Anaerobic reactor design concepts for the treatment of domestic wastewater. Reviews in
132
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Environmental Science and Biotechnology, v. 5, n. 1, p. 21–38, 2006.

HAANDEL, A. VAN; LIER, J. B. VAN. Role of Anaerobic Digestion in Increasing the


Energy Efficiency and Energy Output of Sugar Cane Distilleries. In: Anaerobic Biotechnology
Environmental Protection and Resource Recovery. First ed. [s.l.] Imperial College Press,
2015. p. 421.

HAGOS, K.; ZONG, J.; LI, D.; LIU, C.; LU, X. Anaerobic co-digestion process for
biogas production: Progress, challenges and perspectives. Renewable and Sustainable Energy
Reviews, v. 76, n. March 2016, p. 1485–1496, 2017.

HEMANSI; GUPTA, R.; YADAV, G.; KUMAR, G.; YADAV, A.; SAINI, J. K.;
KUHAD, R. C. Second Generation Bioethanol Production: The State of Art. In: SRIVASTAVA,
N.; SRIVASTAVA, M.; MISHRA, P. K.; UPADHYAY, S. N.; RAMTEKE, P. W.; GUPTA, V.
K. (Eds.). . Sustainable Approaches for Biofuels Production Technologies: From Current
Status to Practical Implementation. Cham: Springer International Publishing, 2019. p. 121–
146.

HOARAU, J; CARO, Y.; GRONDIN, I.; PETIT, T. Sugarcane vinasse processing:


Toward a status shift from waste to valuable resource. A review. Journal of Water Process
Engineering, v. 24, p. 11–25, 2018.

HOARAU, J.; CARO, Y.; GRONDIN, I.; PETIT, T. Sugarcane vinasse processing:
Toward a status shift from waste to valuable resource. A review. Journal of Water Process
Engineering, v. 24, n. May, p. 11–25, 2018.

HU, A. Y.; STUCKEY, D. C. Activated Carbon Addition to a Submerged Anaerobic


Membrane Bioreactor: Effect on Performance, Transmembrane Pressure, and Flux. Journal of
Environmental Engineering, v. 133, n. 1, p. 73–80, 2007.

HUANG, C.; GUO, H. J.; WANG, C.; XIONG, L.; LUO, M. T.; CHEN, X. F.; ZHANG,
H. R.; LI, H. L.; CHEN, X. DE. Efficient continuous biogas production using lignocellulosic
hydrolysates as substrate: A semi-pilot scale long-term study. Energy Conversion and
Management, v. 151, n. 2, p. 53–62, 2017.

IQBAL SYAICHURROZI, B.; SUMARDIONO, S. Kinetic Model of Biogas Yield


Production from Vinasse at Various Initial pH: Comparison between Modified Gompertz Model
133
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

and First Order Kinetic Model. Research Journal of Applied Sciences, Engineering and
Technology, v. 7, n. 13, p. 2798–2805, 2014.

JALLES MACHADO. Yeast. Disponível em:


<https://www.jallesmachado.com/en/products/yeast>. Acesso em: 7 mar. 2018.

JANKE, L; LEITE, A.; BATISTA, K.; WEINRICH, S.; STRÄUBER, H.; NIKOLAUSZ,
M.; NELLES, M.; STINNER, W. Optimization of hydrolysis and volatile fatty acids production
from sugarcane filter cake: Effects of urea supplementation and sodium hydroxide pretreatment.
Bioresource Technology, v. 199, p. 235–244, 2016.

JANKE, L.; LEITE, A. F.; BATISTA, K.; SILVA, W.; NIKOLAUSZ, M.; NELLES, M.;
STINNER, W. Enhancing biogas production from vinasse in sugarcane biorefineries: Effects of
urea and trace elements supplementation on process performance and stability. Bioresource
Technology, v. 217, p. 10–20, 2016.

JANKE, L; LEITE, A. F.; BATISTA, K.; SILVA, W.; NIKOLAUSZ, M.; NELLES, M.;
STINNER, W. Enhancing biogas production from vinasse in sugarcane biorefineries: Effects of
urea and trace elements supplementation on process performance and stability. Bioresource
Technology, v. 217, p. 10–20, 2016.

JANKE, L.; LEITE, A.; NIKOLAUSZ, M.; SCHMIDT, T.; LIEBETRAU, J.; NELLES,
M.; STINNER, W. Biogas Production from Sugarcane Waste: Assessment on Kinetic Challenges
for Process Designing. International journal of molecular sciences, v. 16, n. 9, p. 20685–20703,
2015.

JOELSSON, E.; ERDEI, B.; GALBE, M.; WALLBERG, O. Techno-economic


evaluation of integrated first- and second-generation ethanol production from grain and straw.
Biotechnology for Biofuels, v. 9, n. 1, p. 1–16, 2016.

JUNQUEIRA, T. L.; CHAGAS, M. F.; GOUVEIA, V. L. R.; REZENDE, M. C. A. F.;


WATANABE, M. D. B.; JESUS, C. D. F.; CAVALETT, O.; MILANEZ, A. Y.; BONOMI, A.
Techno-economic analysis and climate change impacts of sugarcane biorefineries considering
different time horizons. Biotechnology for Biofuels, v. 10, n. 1, p. 1–12, 2017.

KAPARAJU, P.; SERRANO, M.; THOMSEN, A. B.; KONGJAN, P.; ANGELIDAKI,


I. Bioethanol, biohydrogen and biogas production from wheat straw in a biorefinery concept.
134
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Bioresource Technology, v. 100, n. 9, p. 2562–2568, 2009.

KIYUNA, L. S. M.; FUESS, L. T.; ZAIAT, M. Unraveling the influence of the


COD/sulfate ratio on organic matter removal and methane production from the biodigestion of
sugarcane vinasse. Bioresource Technology, v. 232, p. 103–112, 2017.

KLINKE, H. B.; THOMSEN, A. B.; AHRING, B. K. Inhibition of ethanol-producing


yeast and bacteria by degradation products produced during pre-treatment of biomass. Applied
Microbiology and Biotechnology, v. 66, n. 1, p. 10–26, 2004.

KOYAMA, M.; YAMAMOTO, S.; ISHIKAWA, K.; BAN, S.; TODA, T. Inhibition of
anaerobic digestion by dissolved lignin derived from alkaline pre-treatment of an aquatic
macrophyte. Chemical Engineering Journal, v. 311, p. 55–62, 2017.

LEITE, A. F.; JANKE, L.; HARMS, H.; ZANG, J. W.; FONSECA-ZANG, W. A.;
STINNER, W.; NIKOLAUSZ, M. Assessment of the Variations in Characteristics and Methane
Potential of Major Waste Products from the Brazilian Bioethanol Industry along an Operating
Season. Energy and Fuels, v. 29, n. 7, p. 4022–4029, 2015.

LEITE, A. F.; JANKE, L.; LV, Z.; HARMS, H.; RICHNOW, H.-H.; NIKOLAUSZ, M.
Improved monitoring of semi-continuous anaerobic digestion of sugarcane waste: Effects of
increasing organic loading rate on methanogenic community dynamics. International Journal
of Molecular Sciences, v. 16, n. 10, p. 23210–23226, 2015.

LEVENSPIEL, O. Tracer Technology Modeling the flow of fluids. [s.l.] Springer,


2012.

LI, W.; KHALID, H.; ZHU, Z.; ZHANG, R.; LIU, G.; CHEN, C.; THORIN, E. Methane
production through anaerobic digestion: Participation and digestion characteristics of cellulose,
hemicellulose and lignin. Applied Energy, v. 226, n. May, p. 1219–1228, 2018.

LIAMLEAM, W.; ANNACHHATRE, A. P. Electron donors for biological sulfate


reduction. Biotechnology Advances, v. 25, n. 5, p. 452–463, 2007.

LIMA, D. R. S.; ADARME, O. F. H.; BAÊTA, B. E. L.; GURGEL, L. V. A.; AQUINO,


S. F. DE. Influence of different thermal pretreatments and inoculum selection on the
biomethanation of sugarcane bagasse by solid-state anaerobic digestion: A kinetic analysis.

135
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Industrial Crops and Products, v. 111, p. 684–693, 2018.

LONGATI, A. A.; LINO, A. R. A.; GIORDANO, R. C.; FURLAN, F. F.; CRUZ, A. J.


G. Biogas Production from Anaerobic Digestion of Vinasse in Sugarcane Biorefinery: A Techno-
economic and Environmental Analysis. Waste and Biomass Valorization, n. 0123456789, 2019.

LÓPEZ GONZÁLEZ, L. M.; PEREDA REYES, I.; ROMERO ROMERO, O. Anaerobic


co-digestion of sugarcane press mud with vinasse on methane yield. Waste Management, v. 68,
p. 139–145, 2017.

LÓPEZ, I.; PASSEGGI, M.; BORZACCONI, L. Validation of a simple kinetic modelling


approach for agro-industrial waste anaerobic digesters. Chemical Engineering Journal, v. 262,
p. 509–516, 2015.

LOVATO, G.; ALBANEZ, R.; TRIVELONI, M.; RATUSZNEI, S. M.; RODRIGUES,


J. A. D. Methane Production by Co-Digesting Vinasse and Whey in an AnSBBR: Effect of
Mixture Ratio and Feed Strategy. Applied Biochemistry and Biotechnology, p. 1–19, 2018.

LUZ, F. C.; CORDINER, S.; MANNI, A.; MULONE, V.; ROCCO, V. Biochar
characteristics and early applications in anaerobic digestion-a review. Journal of Environmental
Chemical Engineering, v. 6, n. 2, p. 2892–2909, 2018.

MACRELLI, S.; MOGENSEN, J.; ZACCHI, G. Techno-economic evaluation of 2nd


generation bioethanol production from sugar cane bagasse and leaves integrated with the sugar-
based ethanol process. Biotechnology for Biofuels, v. 5, n. 1, p. 22, 2012a.

___. Techno-economic evaluation of 2 nd generation bioethanol production from sugar


cane bagasse and leaves integrated with the sugar-based ethanol process. Biotechnology for
Biofuels, v. 5, p. 1–18, 2012b.

MAISON, M.; NAVA, F.; PELÁEZ, C. Manual de Estado del Arte de la Co-digestión
Anaerobia de Residuos Ganaderos y Agroindustriales. España: [s.n.].

MANOCHIO, C. Integração energética da produção de biogás em biorrefinarias de


cana-de-açúcar integradas de 1a e 2a geração. [s.l.] Universidade Federal de Alfenas, 2015.

MARAFON, A. C.; SALOMON, K. R.; AMORIM, E. L. C.; PEITER, F. S. Use of


sugarcane vinasse to biogas, bioenergy, and biofertilizer production. In: Sugarcane Biorefinery,
136
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

Technology and Perspectives. [s.l.] Elsevier Inc., 2020. p. 179–194.

MARTINEZ-JIMENEZ, F. D.; MACIE PINTO, M. P.; MUDHOO, A.; ALENCAR


NEVES, T. DE; ROSTAGNO, M. A.; FORSTER-CARNEIRO, T. Influence of ultrasound
irradiation pre-treatment in biohythane generation from the thermophilic anaerobic co-digestion
of sugar production residues. Journal of Environmental Chemical Engineering, v. 5, n. 4, p.
3749–3758, 2017.

MARTINEZ, E. J.; JOSE, M.; OTERO, M.; GOMEZ, X. Enhancing anaerobic digestion
of poultry blood using activated carbon. p. 297–307, 2017.

MATHEW, A. K.; ABRAHAM, A.; MALLAPUREDDY, K. K.; SUKUMARAN, R. K.


Lignocellulosic Biorefinery Wastes , or Resources ? In: Waste Biorefinery. 1st Editio ed. [s.l.]
Elsevier B.V., 2018. p. 267–297.

MATOS, M. DE; SANTOS, F.; EICHLER, P. Sugarcane world scenario. In: Sugarcane
Biorefinery, Technology and Perspectives. [s.l: s.n.]. p. 1–19.

MCILROY, S. J.; KIRKEGAARD, R. H.; DUEHOLM, M. S.; FERNANDO, E.;


KARST, S. M.; ALBERTSEN, M.; NIELSEN, P. H. Culture-independent analyses reveal novel
anaerolineaceae as abundant primary fermenters in anaerobic digesters treating waste activated
sludge. Frontiers in Microbiology, v. 8, n. JUN, 2017.

MINH, P.; KETHEESAN, B.; YAN, Z.; STUCKEY, D. Trace metal speciation and
bioavailability in anaerobic digestion : A review. Biotechnology Advances, v. 34, n. 2, p. 122–
136, 2016.

MOLINA-CORTÉS, A.; SÁNCHEZ-MOTTA, T.; TOBAR-TOSSE, F.; QUIMBAYA,


M. Spectrophotometric Estimation of Total Phenolic Content and Antioxidant Capacity of
Molasses and Vinasses Generated from the Sugarcane Industry. Waste and Biomass
Valorization, n. 0123456789, 2019.

MOORE, C. C. S.; NOGUEIRA, A. R.; KULAY, L. Environmental and energy


assessment of the substitution of chemical fertilizers for industrial wastes of ethanol production
in sugarcane cultivation in Brazil. International Journal of Life Cycle Assessment, v. 22, n. 4,
p. 628–643, 2017.

137
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

MORAES, B. S.; JUNQUEIRA, T. L.; PAVANELLO, L. G.; CAVALETT, O.;


MANTELATTO, P. E.; BONOMI, A.; ZAIAT, M. Anaerobic digestion of vinasse from
sugarcane biorefineries in Brazil from energy, environmental, and economic perspectives: Profit
or expense? Applied Energy, v. 113, p. 825–835, 2014.

MORAES, B. S.; TRIOLO, J. M.; LECONA, V. P.; ZAIAT, M.; SOMMER, S. G. Biogas
production within the bioethanol production chain: Use of co-substrates for anaerobic digestion
of sugar beet vinasse. Bioresource Technology, v. 190, p. 227–234, 2015.

MORAES, B. S.; ZAIAT, M.; BONOMI, A. Anaerobic digestion of vinasse from


sugarcane ethanol production in Brazil: Challenges and perspectives. Renewable and
Sustainable Energy Reviews, v. 44, p. 888–903, 2015a.

___. Anaerobic digestion of vinasse from sugarcane ethanol production in Brazil:


Challenges and perspectives. Renewable and Sustainable Energy Reviews, v. 44, p. 888–903,
2015b.

MUÑOZ-PÁEZ, K. M.; ALVARADO-MICHI, E. L.; MORENO-ANDRADE, I.;


BUITRÓN, G.; VALDEZ-VAZQUEZ, I. Comparison of suspended and granular cell anaerobic
bioreactors for hydrogen production from acid agave bagasse hydrolyzates. International
Journal of Hydrogen Energy, v. 45, n. 1, p. 275–285, 2020.

NÁTHIA-NEVES, G.; NEVES, T. A.; BERNI, M.; DRAGONE, G.; MUSSATTO, S. I.;
FORSTER-CARNEIRO, T. Start-up phase of a two-stage anaerobic co-digestion process:
Hydrogen and methane production from food waste and vinasse from ethanol industry. Biofuel
Research Journal, v. 5, n. 2, p. 813–820, 2018.

NAVANEETHAN, N.; TOPCZEWSKI, P.; ROYER, S.; ZITOMER, D. Blending


anaerobic co-digestates: Synergism and economics. Water Science and Technology, v. 63, n.
12, p. 2916–2922, 2011.

NISHIYAMA, T.; UEKI, A.; KAKU, N.; WATANABE, K.; UEKI, K. Bacteroides
graminisolvens sp. nov., a xylanolytic anaerobe isolated from a methanogenic reactor treating
cattle waste. International Journal of Systematic and Evolutionary Microbiology, v. 59, n. 8,
p. 1901–1907, 2009.

O’HARA, I. M.; MUNDREE, S. G. Sugarcane-based biofuels and bioproducts. [s.l:


138
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

s.n.].

OMC | Estadísticas del comercio internacional, 2010. Disponível em:


<https://www.wto.org/spanish/res_s/statis_s/its2010_s/its10_toc_s.htm>. Acesso em: 5 maio.
2015.

PARSAEE, M.; KIANI DEH KIANI, M.; KARIMI, K. A review of biogas production
from sugarcane vinasse. Biomass and Bioenergy, v. 122, n. January, p. 117–125, 2019.

PASSOS, F.; CORDEIRO, P. H. M.; BAETA, B. E. L.; AQUINO, S. F. DE; PEREZ-


ELVIRA, S. I. Anaerobic co-digestion of coffee husks and microalgal biomass after thermal
hydrolysis. Bioresource Technology, v. 253, p. 49–54, 2018.

PEREDA-REYES, I.; DÍAZ PAGÉS, J.; HORVART, S. I. Anaerobic Biodegradation of


Solid Substrates from Agroindustrial Activities — Slaughterhouse Wastes and Agrowastes.
Intech, v. i, p. 13, 2016.

PÉREZ, J.; MUÑOZ-DORADO, J.; LA RUBIA, T. DE; MARTÍNEZ, J. Biodegradation


and biological treatments of cellulose, hemicellulose and lignin: An overview. International
Microbiology, v. 5, n. 2, p. 53–63, 2002.

PEREZ, M.; RODRIGUEZ-CANO, R.; ROMERO, L. I.; SALES, D. Anaerobic


thermophilic digestion of cutting oil wastewater: Effect of co-substrate. Biochemical
Engineering Journal, v. 29, n. 3, p. 250–257, 2006.

PINTO, M. P. M.; MUDHOO, A.; ALENCAR NEVES, T. DE; BERNI, M. D.;


FORSTER-CARNEIRO, T. Co–digestion of coffee residues and sugarcane vinasse for
biohythane generation. Journal of Environmental Chemical Engineering, v. 6, n. 1, p. 146–
155, 2018.

QUINTÃO, F. J. O. Caracterização dos produtos de degradação dos fármacos


metformina, enalapril, captopril e ranitidina, por espectrometria de massas de alta
resolução e avaliação da toxicidade após processos oxidativos avançados e cloração. [s.l.]
Universidade Federal de Ouro Preto, 2017.

RABELO, S. C.; CARRERE, H.; MACIEL FILHO, R.; COSTA, A C. Production of


bioethanol, methane and heat from sugarcane bagasse in a biorefinery concept. Bioresource

139
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

technology, v. 102, n. 17, p. 7887–95, set. 2011.

RAJAGOPAL, R.; MEHROTRA, I.; KUMAR, P.; TORRIJOS, M. Evaluation of a hybrid


upflow anaerobic sludge-filter bed reactor: Effect of the proportion of packing medium on
performance. Water Science and Technology, v. 61, n. 6, p. 1441–1450, 2010.

REBELATO, M. G.; RODRIGUES, A. .; THOMAZ, A. . DE B.; SARAN, L. M.;


MADALENO, L. .; OLIVEIRA, O. . DE. Developing an index to assess human toxicity potential
of sugarcane industry. Journal of Cleaner Production, v. 209, p. 1274–1284, 2019.

RIBEIRO, F. R.; PASSOS, F.; GURGEL, L. V. A.; BAÊTA, B. E. L.; AQUINO, S. F.


DE. Anaerobic digestion of hemicellulose hydrolysate produced after hydrothermal pretreatment
of sugarcane bagasse in UASB reactor. Science of the Total Environment, v. 584–585, p. 1108–
1113, 2017a.

___. Anaerobic digestion of hemicellulose hydrolysate produced after hydrothermal


pretreatment of sugarcane bagasse in UASB reactor. Science of The Total Environment, v. 584–
585, p. 1108–1113, 2017b.

ROCHA-MENESES, L.; RAUD, M.; ORUPÕLD, K.; KIKAS, T. Second-generation


bioethanol production : A review of strategies for waste valorisation. Agronomy Research, v.
15, n. 3, p. 830–847, 2017.

RODRÍGUEZ-DÍAZ, J. M.; PRIETO, J. O.; BRAVO, L. R.; SILVA, M. G. C. DA;


SILVA, V. . DA; ARTEAGA-PÉREZ, L. E. Comprehensive Characterization of Sugarcane
Bagasse Ash for Its Use as an Adsorbent. Bioenergy Research, 2015.

RODRÍGUEZ, L.; VILLASEÑOR, J.; FERNÁNDEZ, F. J.; BUENDÍA, I. M. Anaerobic


co-digestion of winery wastewater. Water Science and Technology, v. 56, n. 2, p. 49–54, 2007.

RODRIGUEZ, R. P.; MANOCHIO, C.; MORAES, B. D. S. Energy Integration of Biogas


Production in an Integrated 1G2G Sugarcane Biorefinery : Modeling and Simulation. Bioenergy
Research, 2018.

ROMERO-GÜIZA, M. S.; VILA, J.; MATA-ALVAREZ, J.; CHIMENOS, J. M.;


ASTALS, S. The role of additives on anaerobic digestion: A review. Renewable and Sustainable
Energy Reviews, v. 58, p. 1486–1499, 2016.

140
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

RUIZ, H. A.; THOMSEN, M. H.; TRAJANO, H. L. Hydrothermal Processing in


Biorefineries: Production of Bioethanol and High Added-Value Compounds of Second and
Third Generation Biomass. [s.l.] Springer, 2017.

SAADY, N. M. C. Homoacetogenesis during hydrogen production by mixed cultures


dark fermentation: Unresolved challenge. International Journal of Hydrogen Energy, v. 38, n.
30, p. 13172–13191, 2013.

SAFARI, A.; KARIMI, K.; SHAFIEI, M. Integrated ethanol and biogas production from
pinewood. RSC Advances, v. 6, n. 43, p. 36441–36449, 2016.

SALES, A.; LIMA, S. A. Use of Brazilian sugarcane bagasse ash in concrete as sand
replacement. Waste Management, v. 30, n. 6, p. 1114–1122, 2010.

SANTOS, F.; EICHLER, P.; QUEIROZ, J. H. DE; GOMES, F. Production of second-


generation ethanol from sugarcane. In: Sugarcane Biorefinery, Technology and Perspectives.
[s.l: s.n.]. p. 195–228.

___. Production of second-generation ethanol from sugarcane. In: Sugarcane


Biorefinery, Technology and Perspectives. [s.l: s.n.]. p. 195–228.

SANTOS, L. C. D.; ADARME, O. F. H.; BAÊTA, B. E. L.; GURGEL, L. V. A.;


AQUINO, S. F. D. Production of biogas (methane and hydrogen) from anaerobic digestion of
hemicellulosic hydrolysate generated in the oxidative pretreatment of coffee husks. Bioresource
Technology, v. 263, p. 601–612, 2018.

SANTOS, O. S.; MENDONÇA, A. G. R.; SANTOS, J. C. C.; SILVA, A. P. B.; COSTA,


S. S. L.; OLIVEIRA, L. C.; CARMO, J. B.; BOTERO, W. G. The use of sugar and alcohol
industry waste in the adsorption of potentially toxic metals. Environmental Technology (United
Kingdom), v. 37, n. 2, p. 279–291, 2016.

SANTOS, P. S.; ZAIAT, M.; OLLER DO NASCIMENTO, C. A.; FUESS, L. T. Does


sugarcane vinasse composition variability affect the bioenergy yield in anaerobic systems? A dual
kinetic-energetic assessment. Journal of Cleaner Production, v. 240, p. 118005, 2019.

SARIASLANI, S.; GADD, G. M. ADVANCES APPLIED MICROBIOLOGY. 1st


Editio ed. [s.l.] Elsevier Inc., 2016.

141
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

SARIPAN, A. F.; REUNGSANG, A. Biohydrogen production by


thermoanaerobacterium thermosaccharolyticum KKU-ED1: Culture conditions optimization
using mixed xylose/arabinose as substrate. Electronic Journal of Biotechnology, v. 16, n. 1,
2013.

SHAH, F. A.; MAHMOOD, Q.; RASHID, N.; PERVEZ, A.; RAJA, I. A.; SHAH, M. M.
Co-digestion, pretreatment and digester design for enhanced methanogenesis. Renewable and
Sustainable Energy Reviews, v. 42, p. 627–642, 2015.

SIDDIQUE, N. I.; WAHID, Z. A. Achievements and perspectives of anaerobic co-


digestion: a review. Journal of Cleaner Production, v. 194, p. 359–374, 2018.

SILVA, C. E. F.; ABUD, A. K. S. Influence of manure concentration as inoculum in


anaerobic digestion of vinasse . Acta Scientiarum - Biological Sciences, v. 39, n. 2, p. 173–180,
2017.

SILVA, C.; PACHECO, R.; ARCENTALES, D.; SANTOS, F. Sustainability of


sugarcane for energy purposes. In: Sugarcane Biorefinery, Technology and Perspectives. [s.l:
s.n.]. p. 89–102.

SLIMANE, K.; FATHYA, S.; ASSIA, K.; HAMZA, M. Influence of inoculums/substrate


ratios (ISRs) on the mesophilic anaerobic digestion of slaughterhouse waste in batch mode:
Process stability and biogas production. Energy Procedia, v. 50, p. 57–63, 2014.

SOCCOL, C. R.; BRAR, S. K.; FAULDS, C.; RAMOS, L. P. Green Fuels Technology
Biofuels. 1st. ed. [s.l.] Springer, 2016.

SOLERA, R.; ROMERO, L. I.; SALES, D. The evolution of biomass in a two-phase


anaerobic treatment process during start-up. Chemical and Biochemical Engineering
Quarterly, v. 16, n. 1, p. 25–29, 2002.

SUDIBYO, H.; SHABRINA, Z. L.; WONDAH, H. R.; HASTUTI, R. T.; HALIM, L.;
PURNOMO, C. W.; BUDHIJANTO, W. Anaerobic digestion of landfill leachate with natural
zeolite and sugarcane bagasse fly ash as the microbial immobilization media in packed bed
reactor. Acta Polytechnica, v. 58, n. 1, p. 57–68, 2018.

SURENDRA, K. C.; SAWATDEENARUNAT, C.; SHRESTHA, S.; SUNG, S.;

142
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

KHANAL, S. K. Anaerobic Digestion-Based Biorefinery for Bioenergy and Biobased Products.


Industrial Biotechnology, v. 11, n. 2, p. 103–112, 2015.

TABATABAEI, M.; GHANAVATI, H. Biogas: Fundamentals,Process and


Operation. 1st. ed. [s.l.] Springer, 2018.

TAN, X.; LIU, Y.; ZENG, G.; WANG, X.; HU, X.; GU, Y. Application of biochar for
the removal of pollutants from aqueous solutions. CHEMOSPHERE, v. 125, p. 70–85, 2015.

TEIXEIRA, S. R.; MAGALHÃES, R. S.; ARENALES, A.; SOUZA, A. E.; ROMERO,


M.; RINCÓN, J. M. Valorization of sugarcane bagasse ash: Producing glass-ceramic materials.
Journal of Environmental Management, v. 134, p. 15–19, 2014.

TOLEDO-CERVANTES, A.; GUEVARA-SANTOS, N.; ARREOLA-VARGAS, J.;


SNELL-CASTRO, R.; MÉNDEZ-ACOSTA, H. O. Performance and microbial dynamics in
packed-bed reactors during the long-term two-stage anaerobic treatment of tequila vinasses.
Biochemical Engineering Journal, v. 138, p. 12–20, 2018.

UELLENDAHL, H.; AHRING, B. K. Anaerobic digestion as final step of a cellulosic


ethanol biorefinery: Biogas production from fermentation effluent in a UASB reactor - Pilot-scale
results. Biotechnology and Bioengineering, v. 107, n. 1, p. 59–64, 2010.

VALINHAS, R. V; PANTOJA, L. A.; MAIA, A. C. F.; NELSON, D. L.; CORRESP, A.


S. S.; MIGUEL, M. G. C. P. Xylose fermentation to ethanol by new Galactomyces geotrichum
and Candida akabanensis strains. PeerJ-The Journal of Life and Environmental Sciences, p.
1–16, 2018.

VAZ, S. Sugarcane-Biorefinery. In: Advances in Biochemical


Engineering/Biotechnology. Berlin: Springer Berlin Heidelberg, 2017. .

VICTRAL, D. M.; AQUINO, S. F.; SILVA, S. Q.; BAÊTA, B. E. L. Application of


residual yeast as a source of redox mediators for the anaerobic decolorization of a model azo dye.
Brazilian Journal of Chemical Engineering, v. 33, n. 4, p. 705–711, 2016.

WANG, D.; AI, J.; SHEN, F.; YANG, G.; ZHANG, Y.; DENG, S.; ZHANG, J.; ZENG,
Y.; SONG, C. Improving anaerobic digestion of easy-acidification substrates by promoting
buffering capacity using biochar derived from vermicompost. Bioresource Technology, v. 227,

143
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

p. 286–296, 2017.

WANG, W.; XIE, L.; CHEN, J.; LUO, G.; ZHOU, Q. Biohydrogen and methane
production by co-digestion of cassava stillage and excess sludge under thermophilic condition.
Bioresource Technology, v. 102, n. 4, p. 3833–3839, 2011.

WANG, W.; XIE, L.; LUO, G.; ZHOU, Q. Enhanced fermentative hydrogen production
from cassava stillage by co-digestion : The effects of different co-substrates. International
Journal of Hydrogen Energy, v. 38, n. 17, p. 6980–6988, 2013.

WANG, W.; XIE, L.; LUO, G.; ZHOU, Q.; LU, Q. Optimization of biohydrogen and
methane recovery within a cassava ethanol wastewater / waste integrated management system.
BIORESOURCE TECHNOLOGY, v. 120, p. 165–172, 2012.

WANG, Y.; ZHANG, Y.; WANG, J.; MENG, L. Effects of volatile fatty acid
concentrations on methane yield and methanogenic bacteria. Biomass and Bioenergy, v. 33, n.
5, p. 848–853, 2009.

WERTZ, J.-L.; DELEU, M.; COPPÉE, S.; RICHEL, A. Hemicelluloses and lignin in
Biorefineries. 1st Editio ed. Boca Raton,FL: CRC Press, Taylor & Francis Group, 2018.

WESTERHOLM, M.; HANSSON, M.; SCHNÜRER, A. Improved biogas production


from whole stillage by co-digestion with cattle manure. Bioresource Technology, v. 114, p. 314–
319, 2012.

WIERCKX, N.; KOOPMAN, F.; RUIJSSENAARS, H. J.; WINDE, J. H. DE. Microbial


degradation of furanic compounds: Biochemistry, genetics, and impact. Applied Microbiology
and Biotechnology, v. 92, n. 6, p. 1095–1105, 2011.

XIE, C. H.; YOKOTA, A. Pleomorphomonas oryzae gen. nov., sp. nov., a nitrogen-fixing
bacterium isolated from paddy soil of Oryza sativa. International Journal of Systematic and
Evolutionary Microbiology, v. 55, n. 3, p. 1233–1237, 2005.

XIE, S.; HAI, F. I.; ZHAN, X.; GUO, W.; NGO, H. H.; PRICE, W. E.; NGHIEM, L. D.
Anaerobic co-digestion: A critical review of mathematical modelling for performance
optimization. Bioresource Technology, v. 222, p. 498–512, 2016.

XIN, Y.; SHEN, J.; WANG, B.; ZHU, X.; WANG, Y.; WANG, G.; CHEN, Q. Effects of
144
MINISTÉRIO DA EDUCAÇÃO
Universidade Federal de Ouro Preto
Graduate Program in Environmental
Engineering – PROAMB

vinasse and silica mud on the performance of thermophilic fermentation of brewery sludge from
brewery waste substrates. Environment Protection Engineering, v. 44, n. 2, p. 119–127, 2017.

ZHANG, J.; ZHAO, W.; ZHANG, H.; WANG, Z.; FAN, C.; ZANG, L. Recent
achievements in enhancing anaerobic digestion with carbon- based functional materials.
Bioresource Technology, v. 266, n. July, p. 555–567, 2018.

ZHANG, L.; BAN, Q.; LI, J. Microbial community dynamics at high organic loading
rates revealed by pyrosequencing during sugar refinery wastewater treatment in a UASB reactor.
Frontiers of Environmental Science and Engineering, v. 12, n. 4, p. 1–12, 2018.

ZHENG, Y.; ZHAO, J.; XU, F.; LI, Y. Pretreatment of lignocellulosic biomass for
enhanced biogas production. Progress in Energy and Combustion Science, v. 42, p. 35–53, jun.
2014.

ZIEMIŃSKI, K.; KOWALSKA-WENTEL, M. Effect of enzymatic pretreatment on


anaerobic co-digestion of sugar beet pulp silage and vinasse. Bioresource Technology, v. 180,
p. 274–280, 2015.

145

You might also like